Next Article in Journal
Manipulation of Time- and Frequency-Domain Dynamics by Magnon-Magnon Coupling in Synthetic Antiferromagnets
Next Article in Special Issue
The Role of Magnetic Dipole—Dipole Coupling in Quantum Single-Molecule Toroics
Previous Article in Journal
Continuous Flow Labeling and In-Line Magnetic Separation of Cells
Previous Article in Special Issue
Predicting Pt-195 NMR Chemical Shift and 1J(195Pt-31P) Coupling Constant for Pt(0) Complexes Using the NMR-DKH Basis Sets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Microscopic Mechanisms Involved in Superexchange

by
Jacques Curély
Laboratoire Ondes et Matière d’Aquitaine, UMR 5798, University of Bordeaux, 351 Cours de la Libération, CEDEX, 33405 Talence, France
Dedicated to Professor Roland Georges.
Magnetochemistry 2022, 8(1), 6; https://doi.org/10.3390/magnetochemistry8010006
Submission received: 27 October 2021 / Revised: 17 December 2021 / Accepted: 27 December 2021 / Published: 30 December 2021

Abstract

:
In earlier work, we previously established a formalism that allows to express the exchange energy J vs. fundamental molecular integrals without crystal field, for a fragment A–X–B, where A and B are 3d1 ions and X is a closed-shell diamagnetic ligand. In this article, we recall this formalism and give a physical interpretation: we may rigorously predict the ferromagnetic (J < 0) or antiferromagnetic (J > 0) character of the isotropic (Heisenberg) spin-spin exchange coupling. We generalize our results to ndm ions (3 ≤ n ≤ 5, 1 ≤ m ≤ 10). By introducing a crystal field we show that, starting from an isotropic (Heisenberg) exchange coupling when there is no crystal field, the appearance of a crystal field induces an anisotropy of exchange coupling, thus leading to a z-z (Ising-like) coupling or a x-y one. Finally, we discuss the effects of a weak crystal field magnitude (3d ions) compared to a stronger (4d ions) and even stronger one (5d ions). In the last step, we are then able to write the corresponding Hamiltonian exchange as a spin-spin one.

1. Introduction

It was necessary to wait until the end of the 1950s to achieve a good understanding of superexchange interactions, when Anderson first proposed the theory of coupling between identical ions, characterized by a 3dm electronic configuration without orbital degeneracy (m = 1) [1,2], later generalized to m > 1. In this case, the exchange Hamiltonian is of the Heisenberg–Dirac type: Js1.s2. Anderson’s publication has become the starting point for generalizations, notably with the introduction of orbital degeneracy [2,3,4,5,6,7,8,9].
We have proposed a first generalization of the various mechanisms involved in superexchange for identical 3d1 ions: the exchange energy constant J has been expressed vs. fundamental molecular integrals, characterizing each of the σ–type bonds created by the presence of a diamagnetic ligand and a similar (or different) magnetic ion [10].
What is new in this paper? So far, no physical interpretation has been brought. In this article, we give the physical interpretation, and we generalize to ions ndm (3 ≤ n ≤ 5, 1 ≤ m ≤ 10). In addition, we rigorously show for the first time that the Hamiltonian is given by Jsi.sj, where J is the “exchange constant”. The key physical points are as follows:
  • J is expressed vs. fundamental molecular integrals in the absence of a crystal field, uniquely, for the sake of simplicity; we show that the introduction of a crystal field may be achieved very easily, thus allowing us to discuss further the notion of anisotropic couplings;
  • For the first time, we may rigorously predict the ferromagnetic (J < 0) or antiferromagnetic (J > 0) character of spin-spin couplings whereas, so far, we have dealt with empirical rules, i.e., the Goodenough–Kanamori rules published between the middle of the 1950s and the beginning of the 1960s [11,12,13,14].
At that time, the magnetic compounds of interest were essentially the oxides and fluorides of transition elements, as well as “natural” garnets, notably ferrites. The Goodenough–Kanamori rules gave satisfactory results regarding the sign of J but there were some failures. Indeed, although these rules can predict ferromagnetic and antiferromagnetic arrangements, they do not give the magnitude of J. Experimental works have pointed out that superexchange could also lead to ferromagnetic spin arrangements but with a magnitude of J not predicted by Anderson’s model, which mainly predicted antiferromagnetic arrangements [11]. As a result, the first improvements to Anderson’s model have consisted of taking into account polarization effects [3], thus leading researchers to introduce perturbation expansions of the initial model; however, this turned out to be insufficient.
In addition, at the beginning of the 1990s, new magnetic compounds that did not exist in nature were synthetized [15,16,17,18]. These “synthetized” compounds were in opposition to the “natural” ones encountered until they were characterized by the introduction of organic ligands of variable length between magnetic cations. Their introduction deeply undermined the previous interpretation of the J sign.
This is the reason for which:
  • Here, without a crystal field, we deal with a theoretical model, from which we derive the conclusion that when coulombic interactions are dominant, our model follows Hund’s rule and we explain why couplings are automatically ferromagnetic; when coulombic interactions are no longer dominant, our model is equivalent to the molecular orbital one and couplings are always antiferromagnetic (except in a particular case, where couplings are ferromagnetic but present a small absolute value of J);
  • By introducing the notion of a crystal field, we discuss how passing from an isotropic (Heisenberg) coupling to an anisotropic one (z-z, i.e., an Ising-like coupling or an x-y one); in addition, from the theoretical expression of J, we may also predict the ferromagnetic (J < 0) or antiferromagnetic (J > 0) character of spin-spin couplings as in the absence of a crystal field, which is the key finding of this article.
In addition, in each case, the model allows us to express the magnitude of J for any sign. In this review article, we do not consider the polarization effect of the involved bonds.
As a result, the paper is ordered as follows. Section 2 is devoted to the microscopic mechanisms involved in superexchange. We consider the most general case of two different magnetic 3d1 ions, A and B, characterized by σ–type bonds on each side of the diamagnetic bridge, X. The cationic orbitals are of a d-type for A and B, whereas that of the diamagnetic ligand X is of the s- or p-type. We notably define and justify the general assumptions used for developing our theoretical model.
Under these conditions, we show that this general treatment allows the researcher to calculate superexchange interaction within the isolated fragment A–X–B through the construction of the intermediate cationic states, which leads to the determination of the collective state. The matrix associated with the corresponding Hamiltonian is logically derived, as well as the energy spectrum. By doing so, the corresponding exchange energy J is expressed vs. the key molecular integrals characterizing the two σ bonds of A–X–B.
In Section 3, we give a full closed-form expression for J vs. key molecular integrals as well as physical interpretations. As a result, it becomes possible to predict and interpret the sign of J as well as its magnitude.
Finally, we introduce the important concept of crystal field theory. We do not discuss the effect of the crystal field in the splitting of orbital degeneracy. We exclusively discuss its effect on the nature of exchange energy J involved in the coupling of first-nearest spin neighbors. We show that, starting from an isotropic (Heisenberg) exchange coupling when there is no crystal field, the appearance of a crystal field induces an anisotropy of exchange coupling thus leading to a z-z (Ising-like) coupling or a x-y one. For the sake of simplicity, we do not consider the other cases. Thus, we only discuss the effects of a weak crystal field magnitude (3d ions) compared to a stronger (4d ions) and even stronger one (5d ions).
In a final step, without a crystal field (isotropic couplings) or with a crystal field (anisotropic couplings), we are then able to write the corresponding Hamiltonian exchange as a spin-spin exchange.

2. Microscopic Mechanisms Involved in Superexchange

2.1. Basic Physical Considerations

2.1.1. Generalities and Hund’s Rules

When considering lattices composed of transition ions, there is a more or less important effect coming from the electrostatic potential due to the ionic environment: this effect is well known as that of the “crystal field”. It, finally, describes the coulombic interactions inside the ionic cage in which the transition ion is inserted. When considering the electronic shells responsible for the magnetic properties, d orbitals are the most external shells of the ion. As a result, d orbitals are more “sensitive” to an external potential i.e., the crystal field.
Thus, for 3d orbitals, we deal with a (more or less) weak crystal field. Oppositely the crystal field is strong when dealing with 4d and 5d orbitals. This can be simply explained by the fact that, when passing from a 3d ion to a 5d one, the mean radius of the ion increases, whereas the same surrounding cage under consideration, in which the ion is inserted, shows similar dimensions. As a result, the magnitude of coulombic interactions also increases. These findings allow us to directly lead to the determination of adequate orbitals describing the ion’s magnetic properties.
Hund’s rules allow the determination of the orbital and spin momenta characterizing the ion’s ground state. In fact, these rules try to explain how Coulomb repulsion and Pauli’s exclusion principle must be considered simultaneously [10]. If S = Smax, L = Lmax (for S = Smax) compatible with the exclusion principle are the respective values of the spin and orbit momenta along a z-axis of reference, the total momentum is J = L + S. If the external electronic shell is at most half-filled (L and S are antiparallel) J = |L − S|. If the shell is more than half-filled (L and S parallel) J = L + S (the unoccupied orbitals being considered as holes).
In Figure 1, we have reported the values of spin and orbital momenta characterizing a 3d, 4d or 5d external shell. The simplest situation is the electronic configuration 3d1 (ions V4+ and Ti3+), where a single orbital is filled. Thus, according to Hund’s rules, the orbital momentum vanishes (L = 0, S = 5/2) at mid-filling of the 3d shell (case, 3d5, the ions, Fe3+ and Mn2+). As a result, the configuration 3d6 (the ion, Fe2+, L = 2, S = 2) is equivalent to the case 3d4 (the ions Mn3+ and Cr2+). Similarly, the configuration 3d1 (ions V4+ and Ti3+, L = 2, S = 1/2) is equivalent to the case 3d9 (ion Cu2+), and so on.
It is well known that for transition ions considered in the ground state, the orbital momentum is quenched. As a result, the magnetic properties of such a kind of ion are exclusively due to its spin momentum. Thus, when considering the formal coupling of two spins, each one belonging to a magnetic center, one can speak about direct exchange. If these ions are isolated, we deal with an isotropic (Heisenberg) coupling. If a diamagnetic ligand, X, is inserted between two magnetic centers, A and B, the entity A–X–B also shows magnetic properties. In that case, we have an indirect exchange between A and B through X, also known as superexchange.
In this article, in a first step, we exclusively consider the “toy model” of two magnetic centers, A and B, i.e., two transition ions, 3d1, characterized by a single spin, ½, and coupled through a diamagnetic ligand X without a crystal field. Then we introduce a crystal field contribution. As it is one of the main sources of anisotropy for exchange, we shall discuss its influence on the nature of anisotropic couplings (z-zIsing-typeor x-y couplings).
In a second step, we generalize the corresponding results in the case of 3dm ions (1 < m ≤ 10), i.e., ions such as S > ½. In both cases, we consider a situation in which there is no crystal field and one for which a crystal field is introduced.

2.1.2. The First “Historical” Model Proposed by Anderson for Superexchange

In order to explain the magnetic properties of a wide class of materials, such as oxides and fluorides of transition elements and also including “natural” garnets (notably ferrites), Anderson proposed a model explaining the “indirect” exchange phenomenon arising between magnetic centers, in spite of the fact that they were surrounded by non-magnetic entities: this is the superexchange phenomenon [1,2,3,4]. According to Anderson [4], “superexchange acquired its name because of the relatively large distances over which the exchange effect was often found to act between ions, radicals or molecules”.
However, he knew that superexchange never occurs in a classical ferromagnet, i.e., a medium in which the magnetic centers are characterized by a spin quantum number, S, such as S ≥ 5/2, even if it is diluted inside a paramagnetic metal where there are long distances between the ferromagnetic and paramagnetic entities characterized by d orbitals. Why? Taking into account this important experimental observation, Anderson concluded that the origin of this phenomenon was purely quantum because any classical interpretation failed.
As a result, in a first step, Anderson made the following “reasonable” assumptions:
  • The direct overlap of the involved wave functions characterizing the pair of magnetic sites A and B separated by the non-magnetic ligand X vanishes;
  • The ligand wave function is weakly modified by the presence of magnetic ions;
  • This modification confers a magnetic character that is the origin of the exchange interactions between the pair of magnetic ions through the non-magnetic ligand.
In a second step, Anderson considered the simplest modification for describing the unavoidable exchange between magnetic sites A and B through the non-magnetic ligand X: the transfer of one electron of the ligand X characterized by a full s- or p-external shell into the d-external shell of the magnetic ion A (or B) (cf. Figure 2). This assumption was based on the following results:
  • Experimental measurements confirmed the transfer while examining the hyperfine interaction between the ligand nuclear spin and that of the magnetic ion;
  • It has been graphically demonstrated that the ligand wave function is partially magnetic with the expected degrees;
  • The electronic transfer of the up (or down) spin of ligand X to the empty left (or right) d orbital must remain ballistic, i.e., it conserves the spin so that it leads to an antiferromagnetic coupling;
  • The comparison between experimental results obtained for a diluted or a concentrated sample of the same compound shows that the electronic transfer is weak so that the involved wave functions are weakly disturbed [19,20,21,22,23,24,25].
As a result, if ψionic(Stot) and ψexcited(Stot) are the respective wave functions of the starting ionic configuration and the excited one, the final state of the entity A–X–A is described via the global wave function ψ(Stot) = aψionic(Stot) + bψexcited(Stot), where a and b are small.
Anderson has defined the electronic transfer by the integral b m m ( τ ) = d r φ m ( r n ) × ( T + V ) φ m ( r ( n + τ ) ) , m and m’ labeling the corresponding involved orbitals, with one electron per d orbital. H = T + V is the electronic Hamiltonian, where T is the kinetic energy and V, the potential one. The variables r and n allow us to define the position of the concerned electron(s) and the τ’s represent the fundamental translations of the lattice. Due to lattice periodicity, the function φm is a Wannier function. This contribution has been called kinetic exchange because, during the formation of the weak chemical bond between A and X, the antiferromagnetic coupling of electronic spins is characterized by a gain in kinetic energy. If U is the coulombic repulsion energy, Anderson has defined the corresponding exchange energy as Jm,m (kinetic) = −2b m,m (τ)2/U.
When the unpaired electrons belong to orthogonal orbitals spins are coupled ferromagnetically, the corresponding contribution is given by an integral of the type J m m ( potential ) = d r d r φ m ( r n ) φ m ( r n ) ( e 2 / 4 π ε 0 | r r | ) φ m ( r n ) φ m ( r n ) . It is positive because the electrons “try” to avoid themselves through the Fermi hole that appears when spins are parallel (this hole becomes less repulsive when electrons are “closer”). Anderson called this contribution potential exchange.
The total exchange is finally Jmm = Jmm(potential) + Jmm(kinetic), with the conventional writing of an exchange Hamiltonian, 2 J m , m s A m . s B m for a couple of d bands (m,m’), with one d band per ion [2,3,24].
Finally, one may search for the physical origin of the electronic transfer occurring between species A (or B) and X. It is correlated to the creation of a hole in the external electronic shell that the electron has left behind, so that we have to consider an electron-hole couple when describing the electronic transfer. At this stage, we may guess, from now on, that coulombic interactions are going to play a fundamental role. Under these conditions, we deal with a Mott-Wannier exciton because it must occur not only on the whole dimer A–X–B but also over a long distance in a crystal composed of unit cells A–X–B.

2.2. Starting Assumptions

2.2.1. Assumption 1

In this section, we detail the general formalism allowing us to describe the superexchange phenomenon for a centrosymmetric entity AXB, where A and B are two magnetic sites and X is a closed-shell diamagnetic ligand (with one unpaired electron labeled 1, belonging to A, and electron 2 for the one transferred from X to A, for the fragment A–X). In our case, as we deal with a linear entity, A–X–B, we have no Jahn–Teller effect [26,27] and no polarization of the bonds A–X and X–B.
This model is built without the presence of a crystal field and exclusively concerns the case of 3d ions (3d1 in a first step, then 3dm ions, 1 < m ≤ 10). Regarding the compounds whose magnetic properties may be described, we have “natural” compounds, such as oxides, fluorides, and garnets (notably ferrites), that we call “Class I compounds”. In these compounds, the average size of the ligand orbital (p orbital) is close to that of the magnetic cation (d orbital). With the appearance of new “synthetized” magnetic compounds at the beginning of the 1990s, the magnetic cations are well separated by more or less long organic ligands [15,16,17,18]. In other words, the average size of a cation orbital is plainly lower than that of a ligand. We call these compounds “Class II compounds”.
From a physical point of view, for “Class I compounds”, the coulombic interactions between magnetic cations are dominant but in the case of “Class II compounds”, their influence becomes negligible. In both cases, we shall fix the common physical limit that allows the separation of both magnetic regimes.
As a result, the fragment A–X is characterized by the following two-particle Hamiltonian:
H AX = T 1 + V 1 + T 2 + V 2 + e 2 4 π ε 0 r 12 ,   r 12 = | r 1 r 2 | .
T i = 2 i 2 / 2 m e is the electronic kinetic energy operator, where me is the electron mass. The potential operators V1 = V(r1) and V2 = V(r2) include all the nucleus and extra electron contributions to the Coulomb field acting on electrons 1 and 2, involved in the σ bond of the fragment A–X. The unique electron 1 comes from the 3d1 external shell of the magnetic ion A, and electron 2 comes from the full external electronic shell of ligand X. As a result, we work within the framework of the Hartree–Fock approximation, i.e., the action of extra electrons over electrons 1 and 2 is taken into account through a mean-field approximation.
A similar Hamiltonian HXB may be written for the fragment X–B (with HAX = HXB when A = B). The electron coming from the external shell of X is labeled 3, and the one coming from B is labeled 4. All the physical quantities derived from the Hamiltonian HAX (respectively, HXB) are labeled QAX for part A–X (respectively, QXB for part X–B).

2.2.2. Assumption 2

The potential operator V(r) commutes with any symmetry operator O whose action is described by the properties of a double group G with the identity operation Magnetochemistry 08 00006 i001. As we deal with a two-electron problem on both sides of ligand X, separately, the two low-lying states of the bonds A–X and X–B are a spin-singlet and a spin-triplet. Indeed, if the spin wave function χ(s1,s2) =|s1,s2> =|s1>⊗|s2> = | s 1 z >⊗| s 2 z > = | s 1 z , s 2 z > describes the spin states, we have four possible pairings: |↑↑>,|↑↓>, |↓↑> and |↓↓>. If S = s1 + s2, S z = s 1 z + s 2 z we have two classes of possibilities for writing the spin wave function:
  • |0, 0> = 1 2 (|↑↓> − |↓↑>), S = 0 (singlet state),
  • | 1 , 1 > = | > | 1 , 0 > = 1 2 ( | > + | > ) , | 1 , 1 > = | > }     S = 1 (triplet state).
The singlet state is odd with respect to the interchange of s1 and s2, while the triplet state is even.

2.2.3. Assumption 3

If Φa(r1), respectively, Φb(r2) describes the eigenstate of the Hamiltonian H1 = T1 = p 1 2 /2me, where me is the electron mass, respectively, H2 = T2 = p 2 2 /2me, the secular equation det(HE Magnetochemistry 08 00006 i001) = 0 (where 1 is the identity matrix and H = T1 + T2 + U(r1,r2)) can be solved (cf. Equation (1)). The solutions are spatially symmetric and antisymmetric wave functions, i.e., Φ S / A ( r 1 , r 2 ) = ( Φ a ( r 1 ) Φ b ( r 2 ) ± Φ a ( r 2 ) Φ b ( r 1 ) ) / 2 , with <ΦS(r1,r2)|ΦA(r1,r2)> = 0 by construction.

2.2.4. Assumption 4

As there is no spin-orbit coupling, the collective wave function describing the coupling of a pair of electrons is as appears in Ψ(u1,u2) = Φ(r1,r2)χ(s1,s2), where ui = (ri,si). When combining the respective parity properties of functions Φ(r1,r2) and χ(s1,s2) with Pauli’s exclusion principle, we must have:
  • S = 0 χ(s1,s2) = |s1,s2> odd, Φ(r1,r2) = ΦS(r1,r2) even,
  • S = 1 χ(s1,s2) = |s1,s2> even, Φ(r1,r2) = ΦA(r1,r2) odd.
It means that a singlet state (S = 0) is a non-magnetic state (i.e., a purely diamagnetic state) whereas the triplet state (S = 1) is a magnetic one.

2.2.5. Assumption 5

If Γ1/2 is the representation of a spin ½, and if coupling a pair of these spins, we then have to consider the operation Γ1/2 ⊗ Γ1/2 = Γ0 ⊕ Γ1, where ⊕ is the direct sum symbol. Γ0 and Γ1 are the corresponding irreducible representatives (irrep) with dim Γ0 = 1, dim Γ1 = 3 and dim (Γ1/2 ⊗ Γ1/2) = 4.

2.2.6. Assumption 6

If considering the energy levels of the entity A-X-B they are sufficiently close in energy to be both populated at room temperature. In Figure 3 we have reported the radial behavior of V(r) as well as that of energy for the AXB entity. The three corresponding atomic orbitals which are magnetic orbitals [28,29,30] are ΦA and ΦB centered on A and B, respectively, and ΦX centered on the bridge, X. The corresponding states are |A>, |B> and |X>. ΦA, ΦB and ΦX are assumed to be real and are considered as starting (non-disturbed) wave functions i.e., free atomic wave functions that give a spatial description of each of the states |A>, |B> or |X>. In the present case, ΦA and ΦB are cationic d-orbitals and ΦX is an anionic (s or p) orbital.

2.2.7. Assumption 7

The following treatment is based on the symmetry group of the entity under consideration. The determination of this group is essential to find the relevant atomic orbitals which are permitted to combine in order to form “molecular” orbitals. Let us recall that the electrons that are important are those belonging to the valence shell of A, B and X, respectively.

2.2.8. Assumption 8

We have the following properties based on the symmetry problem, notably on the fact that A and B are far apart and without interaction.
  • The states are normalized but not orthogonal (except |A> and |B>):
<A|A> = 1, <B|B> = 1 , <X|X> = 1, <A|B> = <B|A> = 0.
As |A> and |B> are orthogonal (i.e., without overlap) there is no π bond between A and B (but this could be taken into account in the present model). As noted after Equation (1), only two unpaired electrons belonging, respectively, to cations A and B participate in the creation of a bond on each side of the central ligand X, the other electrons being considered as passive. This is the active-electron approximation, which is plainly justified from an experimental point of view.
  • The overlap between A and X on the one hand, and X and B on the other one, are defined as follows:
<A|X> = sA, <X|B> = sB, sA > 0, sB > 0.
In the particular case where A = B (degenerate case), then sA = sB = s > 0.
  • Both magnetic sites A and B have a cationic energy level higher than the anionic one, as is generally the case for transition metal compounds; the energy difference between A and X levels (respectively, X and B levels) is 2δAEA, with EA > 0 (respectively, 2δBEB, with EB > 0), so that we have for the fragment A–X linked to X–B:
<A|(T1 + V1) |A> = − (1 − δA)EA, <B|(T4 + V4)|B> = − (1 − δB)EB,
<X|(T2 + V2)|X> = − (1 + δA)EA, EA > 0, <X|(T3 + V3)|X> = − (1 + δB)EB, EB > 0¸ A ≠ B,
<A|(T1 + V1)|A> = <B|(T4 + V4)|B> = − (1 − δ)E,
<X|(T2 + V2)|X> = <X|(T3 + V3)|X> = − (1 + δ)E, E > 0, A = B.
δA < 1 for fragment A–X (respectively, δB for fragment X−B) but is not necessarily small. δi > 0 (with i = A or B) is a very common scenario, and δi < 0 corresponds to the particular case of the dihydrogen molecule (with A = B, X being absent in that particular case).
A similar remark may be made for δAB, which is defined by an analogy with respect to δA and characterizes the link between fragments A–X and X–B. In Figure 3, we also note that, due to the stabilization of AXB during the creation of the bonds A–X and X–B, we do have −(1 + δA)EA = −(1 + δB)EB. Of course, when A = B, δA = δAB = δB.
  • The transfer integrals between |A> and |X> on the one hand, and |X> and |B> on the other hand, are given by:
<A|(T2 + V2)|X> = − tAXEA, <X|(T3 + V3)|B> = − tXBEB, <A|(Ti + V i)|B> = 0, i = 1,4.
If A = B, tAX = tXB > 0, but when A ≠ B, tAXtXB > 0. The previous equation, stating that there is no transfer between A and B but exclusively between A and X or X and B, is a consequence of the condition imposed by Equation (2). However, as previously noted, the case <A|(T1 + V1)|B> ≠ 0 may be introduced in a more general model, without difficulty, notably by the bias of an π orbital between A and B.
  • sA, tAX and tXB are small compared to unity, and tAX or tXB is mainly related to the potential interaction between the anion and the cation, so that:
tAXEA  sAΔA < sA(1 − δA) EA,
where Δ′A is as defined in Figure 3. A similar relationship may be written for X−B, with Δ′B also defined in Figure 3.

2.3. Expression of the Intermediate “Cationic” States

In this subsection, we exclusively consider the orbital (radial) part of the total wave function. As there is no direct overlap, i.e., no direct transfer between A and B, any exchange interaction between electrons belonging to the magnetic sites A and B automatically occurs through the anionic intermediate bridge, X. We use the starting rules suggested by Anderson [10]:
  • The first rule consists of treating the extra electrons in terms of the simple one-electron Hartree–Fock functions;
  • The second rule is treating them as excitations of a many-body system; this operation is achieved while keeping a constant value for the total spin involved, Stot; this leads us to consider an ionic part for centers A and B and an excited one for the ligand X;
  • As we deal with weak energies involved in the process of excitation, the orbital part ψ(Stot) may be written as the following hybridization: ψ(Stot) = aψionic(Stot) + bψexcited(Stot), where the coefficients a and b must remain small.
We diagonalize the one-particle Hamiltonian Ti + Vi (i = 1,2) in the reduced basis {|A>, |X>} for the fragment A–X (respectively, Ti + Vi (i = 3,4) in the reduced basis {|X>, |B>} for the fragment X−B, separately). The goal of such an operation is to obtain the new cationic (antibonding) normalized eigenstates | A > and | B >, such as:
| A > = ( 1 α A ) | A > + β A | X > , | B > = β B | X > + ( 1 α B ) | B >
where αi and βi (with i = A or B) are real numbers. As we deal with a weak chemical bond between A and X (respectively, between X and B), βi must remain small, and αi, close to unity. In addition, if using the normalization condition <A|A> = <B|B> = 1, as well as <A|X> = sA, <X|B> = sB (sA > 0, sB > 0), the new normalization condition < A | A > = 1 and < B | B > = 1 leads to the following equation:
( 1 α i ) 2 + 2 β i s i ( 1 α i ) + β i 2 1 = 0 , i = A   or   B ,
characterized by the solution 1 – αi = – βisi + 1 β i 2 + ( β i s i ) 2 (with i = A or B). If defining the new direct overlap S, we derive owing to Equation (2):
S = < A | B > = < B | A > = β B s A ( 1 α A ) + β A s B ( 1 α B ) + β A β B > 0   A     B , S = 2 β s ( 1 α ) + β 2 > 0 ,   A = B .
Setting:
S A = 2 β A s A ( 1 β A ) + β A 2 > 0 ,   S B = 2 β B s B ( 1 β B ) + β B 2 > 0 ,  
where SA and SB are functionals of the various overlaps corresponding to the bonds A–X and X−B, respectively, we have:
S = 1 2 β Β β A S A + β A β B S B
and, by reporting in Equation (8), we have:
1 α A = ± 1 S A ,   1 α B = ± 1 S B ,   A     B , 1 α = ± 1 S ,   A = B .  
From now on, the + sign is conventionally adopted. The new energy is:
E A = < A | ( T 1 + V 1 | A ) > ,   E A B = < B | ( T 1 + V 1 ) | B > ,   A B , E = E A = E A B = A | ( T 1 + V 1 ) | A ,   A = B .
Thus, if A ≠ B (general case), the system is non-degenerate ( E A | E > B) but, if A = B, we are dealing with a degenerate state ( E = E A = E B). Introducing the definition of | A > and | B > given by Equation (7), we may write, owing to Equations (4) and (5):
E A = [ ( 1 α A ) 2 ( 1 δ A ) + β A 2 ( 1 + δ A ) ] E A 2 β A ( 1 α A ) t A X E A , E A B = [ ( 1 α B ) 2 ( 1 δ A B ) E B + β B 2 ( 1 + δ A ) E A ] 2 β B ( 1 α B ) t X B E B   , A     B , E = E A = E A B = [ ( 1 α ) 2 ( 1 δ ) + β 2 ( 1 + δ ) ] E 2 β ( 1 α ) t E ,   A = B .
The undetermined parameter, βi (with i =A or B), may now be chosen so that E i is minimum, i.e., owing to Equation (14), the equation ∂ E i/∂βI = 0, after a few calculations, yields:
β A = ± ( 1 α A ) = ± 1 S A ,   β B = ± ( 1 α B ) = ± 1 S B
where SA and SB are given by Equation (10). Thus, βi and 1 − αi (with i = A or B) may show the same sign or opposite signs: if βi > 0 (respectively, βi < 0) the state | A > or | B > will be represented by a spatially symmetric wave function (respectively, spatially antisymmetric). In addition, as βi is small, 1 − αi is also small, and αi is close to unity, as expected. Then, using the particular value of βi given by Equation (15), the ground state energy is:
E A = 2   ( 1 S A )   ( 1 t A X ) E A , E A B = ( 1 S B )   [ ( 1 δ A B ) E B + ( 1 + δ A ) E A 2 t X B E B ]   ,   A     B , E = E A = E A B = 2 ( 1 S )   ( 1 t ) E A ,   A = B
with t = tAX = tXB. The sign ± comes from that of β. We finally define the transfer integral T AB as:
T A B = < A | ( T 1 + V 1 ) | B > = < B | ( T 1 + V 1 ) | A .
Proceeding as for S and E , we derive:
T A B = β B ( 1 α A ) t A X E A     β A ( 1 α B ) t X B E B β A β B ( 1 + δ A ) E A ,   A     B ,   T = 2 β ( 1 α ) t E β 2 ( 1 + δ ) E ,   A = B ,
with T = T AB = T BA. When A ≠ B, we always have T AB = T BA on the condition that (1 + δA)EA = (1 + δB)EB (cf. Figure 3). This illustrates the principle of indistinguishability of electrons: the electronic transfer may indifferently occur not only from X to A (case 1) but also from X to B (case 2), with the same physical effect. However, in both cases, we deal with the coupling of 2 electrons on the side of A or on the side of B.
In the particular case where there is no superexchange, we must have S = < A | B > = < B | A > = 0, so that αA = αB = 0, βA = βB = 0. From Equations (14) and (18), we derive, as expected, E A = ( 1 δ A )   E A , E B = E A B = ( 1 δ B )   E B , T AB = 0 (no transfer between A and B).
Then, using the particular value of βi, (with i = A or B) given by Equation (15), we have:
T A B = 1 S A 1 S B [ ( 1 + δ A ) E A ± t A X E A ± t X B E B ]   ,   A     B , T = ( 1 S ) ( 1 + δ ± 2 t ) E   ,   A = B ,
where the + sign holds for βi > 0 and – holds for βi < 0. As Ei > 0, 1 − Si > 0, 1 + δi > 0 whatever the sign of δi EAEB > 0, and |t|<< 1, T AB (if A ≠ B) or T (if A = B) is negative. Thus, before constructing the collective states, it is clear that SA, E A and T AB (respectively, SB, E B and T BA) appear as the basic parameters of the bond A–X (respectively, X−B) and finally characterize the collective states of AXB.

2.4. Construction of the Collective States

In this part, we wish to construct the wave function of the collective state describing the entity AXB. In the case of a fermionic assembly of N particles, the collective wave function must appear as an antisymmetric combination of current terms Ψ p 1 ( u 1 ) Ψ p 2 ( u 2 ) Ψ p N ( u N ) , where ui = (ri,si) and p1, p2, …, pN characterize the states occupied by these fermions. It is given by the following Slater determinant:
Ψ ( u 1 , u 2 , , u N ) = 1 N ! Ψ p 1 ( u 1 ) Ψ p 1 ( u 2 ) Ψ p 1 ( u N ) Ψ p 2 ( u 1 ) Ψ p 2 ( u 2 ) Ψ p 2 ( u N ) Ψ p N ( u 1 ) Ψ p N ( u 2 ) Ψ p N ( u N ) .
Thus, the permutation of two particles is assumed by exchanging two columns, so that the determinant changes the sign. In addition, if any couple of numbers, pi and pj, is as pi = pj, the determinant shows two identical lines and vanishes, in agreement with Pauli’s principle, which states that two identic fermions cannot occupy the same state.
The “cationic” states | A > and | B > defined in the previous subsection may now give rise to four “cationic” spin-orbital states | A ,+>, | A ,−>, | B ,+> and | B ,−>, from which we may construct, in a first step, four “molecular” states characterizing the entity AXB obeying a G symmetry group, thus conditioning the nature of the involved “molecular” orbitals.
Notably, owing to their behavior under the interchange of | A ,σ> and | B ,σ>, σ = ±1, we call them “gerade” (unchanged) or “ungerade” (sign change), labeled g and u, respectively:
| g , σ > = 1 2 ( 1 + S ) ( | A , σ   > + | B , σ   > ) ,   | u , σ > = 1 2 ( 1 S ) ( | A , σ   > | B , σ   > ) ,   σ = ± 1
where the overlap S is given by Equation (9). We here transpose the problem of sign change that was previously examined when interchanging 2 coupled fermions. Thus the property “gerade” (unchanged) is analogous to the triplet character when coupling two electrons whereas the property ”ungerade” (sign change) is analogous to the singlet character.
The coefficients [2(1 ± S)]−1/2 are self-evident normalizing factors and σ = ± recalls the nature of the corresponding ½ spin state (“up” or “down”). At this point, due to the orthogonality condition < X ,σ| Y ,σ’ > = δXY δσσ’ (with X = A or B ), now, we must have <X,σ|X,σ’ > = δσσ’, with X = g or u. Then it is easily shown that the related energies are:
E A g = < g σ ( T 1 + V 1 ) g σ > = E A + 2 T A B + E A B 2 ( 1 + S ) δ σ , σ , E A u = < u σ ( T 1 + V 1 ) u σ > = E A 2 T A B + E A B 2 ( 1 S ) δ σ , σ ,   A     B , E A g = < g σ ( T 1 + V 1 ) g σ > = E + T 1 + S δ σ , σ , E A u = < u σ ( T 1 + V 1 ) u σ > = E T 1 S δ σ , σ ,   A = B ,
where E and T are defined by Equations (13) and (17), respectively, and the difference of energy between the “gerade” and “ungerade” states is:
E A g E A u = 2 T A B S ( E A + E A B ) 1 S 2 ,   A     B , E A g E A u = 2 T S E 1 S 2 ,   A = B ,
that is, as βi, si (with i = A or B), t, tAX, tXB (A = B or A ≠ B) and S are small:
E A g E A u 2 T A B S ( E A + E A B ) 2 T A B , A     B , E A g E A u 2 ( T S E ) 2 β 2 ( 1 + δ ) E   2 T , A = B .
Thus, E Ag E Au is independent of the sign of βi, as expected, and remains very small. In the particular case where there is no superexchange, we recall that S = 0 and T = 0, so that E Ag = E Au.
From a quantum point of view, with two pairs of electrons and four available spin orbitals |g,±> and |u,±>, 24 determinantal collective states may then be built on each side of the fragment A–X–B. However, due to Pauli’s exclusion principle, coupled to one of indistinguishability, when dealing with 4 fermions coupled in pairs (one per centers A and B, 2 for X with one electron possibly transferred to A, and one possibly transferred to B), the number of states reduces to 4 2 = 6 .
It means that the collective wave function must automatically be composed of 6 different states, each one itself being a 2 × 2 Slater determinant or a linear combination of these determinants.
As a result, let us label |XS,SZ > the collective states. We have X = U (ungerade) or X = G (gerade) if referring to the symmetry of the orbital part with respect to the interchange of | A > and | B >; S and Sz describe the total spin configuration. We define the collective basic state |X,σ; Y,σ’> as the corresponding Slater determinant:
X , σ ; Y , σ = 1 2 X ( r 1 ) σ ( s 1 ) Y ( r 1 ) σ ( s 1 ) X ( r 2 ) σ ( s 2 ) Y ( r 2 ) σ ( s 2 ) .
Clearly, it is immediately apparent that a combination of a g-type orbital and a u-type one allows us to obtain a U-type collective state, whereas combining two g- or two u-orbitals gives rise to a G-type orbital. As a result, we now may build the six following collective states:
| U 1 , 1 > = | u , + ; g , + > , | U 1 , 0 > = 1 2 ( | u , + ;   g , > + | u , ;   g , + > ) , | U 1 , 1 > = | u , ;   g , > , | U 0 , 0 > = 1 2 ( | u , + ;   g , > | u , ;   g , + > ) , | G 0 , 0 g > = | g , + ;   g , > , | G 0 , 0 u > = | u , + ;   u , > .
It is very easily checked that these functions are orthogonal by construction. Their spin characters may be simply verified by applying to each of the previous collective states the convenient total spin operator. Then, introducing the expressions previously obtained for the molecular spin-orbital states (cf Equations (21) and (25)), we may express |U11>,|U10>,|U1-1>,|U00>,| G 0 , 0 g > and | G 0 , 0 u > on the basis of the following Slater determinants for the fragment A–X−B:
| U 1 , 1 > = 1 1 S 2 | A , + ; B , + >   , | U 1 , 0 > = 1 2 ( 1 S 2 ) ( | A , ; B , + > + | A , + ; B , > ) , | U 1 1 > = 1 1 S 2 | A , ; B , >   , | U 0 , 0 > = 1 2 ( 1 S 2 ) ( | A , + ; A , > | B , + ; B , > ) , | G 0 , 0 g > = 1 2 ( 1 + S ) ( | A , + ; A , > + | A , + ; B , > + | B , + ; A , > + | B , + ; B , > ) , | G 0 , 0 u > = 1 2 ( 1 S ) ( | A , + ; A , > | A , + ; B , > | B , + ; A , > + | B , + ; B , > ) .
Concerning these latter states, it is useful to notice that they may be also expressed by means of the polar and non-polar, normalized (but non-strictly orthogonal) states | G 0 , 0 p > and | G 0 , 0 n p >, respectively:
| G 0 , 0 g > = 1 + S 2 2 ( 1 + S ) ( | G 0 , 0 p > + | G 0 , 0 n p > ) , | G 0 , 0 u > = 1 + S 2 2 ( 1 S ) ( | G 0 , 0 p > | G 0 , 0 n p > )   ,
with:
| G 0 , 0 p > = 1 2 ( 1 + S 2 ) ( | A , + ; A , > + | B , + ; B , > ) ,   | G 0 , 0 n p > = 1 2 ( 1 + S 2 )   ( | A , + ; B , > + | B , + ; A , > ) .

2.5. The Hamiltonian Matrix and Energy Spectrum

Now, we calculate the elements of the Hamiltonian matrix in the new basis {| G 0 , 0 g >, | G 0 , 0 u >, |U0,0>, |U1,1>, |U1,0>, |U1,−1>}. The non-vanishing terms are those existing between states belonging to the same irreducible representation (irrep) of the orbital G and spin R symmetry groups so that the final group is G R . As a result, one may expect:
  • Diagonal and off-diagonal terms between | G 0 , 0 g > and | G 0 , 0 u >;
  • Only diagonal terms for the states |US,SZ > with S = 0 (Sz = 0) and S = 1 (Sz = 0, ±1);
  • All the diagonal terms of the states |U1,SZ > are equal because we deal with the irrep Γ1⊗Γ3,u. Under these conditions, the Hamiltonian matrix is:
H = E 0 G g K 0 0 0 0 K E 0 G u 0 0 0 0 0 0 E 0 U 0 0 0 0 0 0 E 1 U 0 0 0 0 0 0 E 1 U 0 0 0 0 0 0 E 1 U
with:
E 0 G g = < G 0 , 0 g | H | G 0 , 0 g > ,   K = < G 0 , 0 g | H | G 0 , 0 u > = < G 0 , 0 u | H | G 0 , 0 g > ,   E 0 G u = < G 0 , 0 u | H | G 0 , 0 u > , E 0 U = <   U 0 , 0 | H | U 0 , 0 > ,   E 1 U = <   U 1 , 1 | H | U 1 , 1 > = <   U 1 , 0 | H | U 1 , 0 > = <   U 1 , 1 | H | U U 1 , 1 > .
For the sake of simplicity, we restrict the results to the fragment A1−X−A2 (A = B). Then, we define the following quantities:
U A = < A 1 | < A 1 | U 1 , 2 | A 1 > | A 1 > ,   U B = < A 2 | < A 2 | U 1 , 2 | A 2 > | A 2 > = U A = U ,   A = B C = < A 1 | A 2 | U 1 , 2 | A 2 > | A 1 > , γ 1 = < A 1 | < A 2 | U 1 , 2 | A 1 > | A 2 > , γ 2 = < A 1 | < A 1 | U 1 , 2 | A 1 > | A 2 >
with:
< W | < X | U 1 , 2 | Y > | Z > = d r 1 d r 2 Φ W ( r 1 ) Φ X ( r 2 ) e 2 4 π ε 0 r 12 Φ Y ( r 2 ) Φ Z ( r 1 )
where r12 =|r1r2|, the electron labeled 1 is coming from A, the one labeled 2, from X. The physical meaning of the parameters U, C, γ1 and γ2 is simply the following one:
  • U is the Coulomb energy for an electron pair occupying the same site;
  • C is the Coulomb energy for two electrons occupying neighboring sites;
  • γ1 is the Coulomb self-energy of the exchange charge distribution −eΦA(rB(r) and is, thus, referred to as the exchange integral;
  • γ2 appears as the Coulomb energy between the exchange charge distribution and an electron charge localized on one site. γ2 is a transfer integral between two cationic orbitals, resulting from the effective coulombic potential created by the charge of another electron involved in the secular problem;
  • When there is no superexchange, i.e., no exchange between A and B through X, we have γ1 ≠ 0 (the exchange charge distribution is restricted to the bond between A and X, X and B), γ2 = 0 as there are no more cationic orbitals and U ≠ 0, C ≠ 0 (the Coulomb energy for two electrons is restricted to first neighboring sites: A and X or X and B).
When A ≠ B, UAUB. C, γ1 and γ2 conserve the same definition but not the same value as in the case where A = B.
Under these conditions, the matrix elements given by Equation (31) may easily be calculated:
E 0 G g = 2 E A g + U + C + 2 γ 1 + 4 γ 2 2 ( 1 + S ) 2 , E 0 G u = 2 E A u + U + C + 2 γ 1 4 γ 2 2 ( 1 S ) 2 ,
K = U C 2 ( 1 S 2 ) ,
E 0 U = E A g + E A u + U γ 1 1 S 2 , E 1 U = E A g + E A u + C γ 1 1 S 2 ,   A = B ,
where E Ag, E Au are given by Equation (22) for A = B or A ≠ B and U, C, γ1 and γ2 by Equation (32). When A ≠ B U is replaced by (UA + UB)/2, the contribution of C, γ1 and γ2 is unchanged, though showing a different value than in the case where A = B. In addition, by diagonalizing the upper 2 × 2 matrix in Equation (30), we have the following eigenvalues:
E 0 , 0 ± = 1 2 E 0 G g + E 0 G u ± ( E 0 G g E 0 G u ) 2 + 4 K 2
as well as the diagonal energy terms Ep and Enp for | G 0 , 0 p > and | G 0 , 0 n p >, respectively:
E p = 2 ( E + S T ) + U + γ 1 1 + S 2 , E n p = 2 ( E + S T ) + C + γ 1 1 + S 2 .

3. Physical Interpretation

3.1. Expression of Jm,m′

As just seen in the previous subsection, when comparing the formal coupling of two electrons initially isolated, here we deal with:
  • The states |US,Sz > with S = 1 (Sz = 0, ±1) that are associated with a “triplet state”, characterized by the eigenvalues E 0 U and E 1 U (three-times degenerated); and
  • The states | G 0 , 0 g > and | G 0 , 0 u > , with S = 0, are associated with a “singlet state”, characterized by the eigenvalues E 0 , 0 ± .
At this stage, when calling ES,0 and ET,0 the low-energy levels of the singlet and triplet spectra, respectively, it is worth recalling that the exchange energy may be defined according to two conventional writings:
  • Jm,m′ = ES,0ET,0 with the corresponding Hamiltonian exchange Hex = −Jm,m′s1.s2 (convention 1); in that case, J < 0 corresponds to an antiferromagnetic arrangement, with ET,0 > ES,0, whereas J > 0 corresponds to a ferromagnetic one, with ET,0 < ES,0, where m and m′ are the name of d bands located on each side of the ligand X.
  • Jm,m′ = ET,0ES,0 with the corresponding Hamiltonian exchange Hex = Jm,m′s1.s2 (convention 2); in that case, J > 0 corresponds to an antiferromagnetic arrangement, with ET,0 > ES,0, whereas J < 0 corresponds to a ferromagnetic one, also with ET,0 < ES,0.
As often used by theoreticians, we here adopt convention 2. The spin-spin Hamiltonian Jsi.sj, here J = ET,0ES,0, is submitted according to a couple of conditions:
  • J << Δ
  • kBT << Δ
where T is the absolute temperature, kB, the Boltzmann’s constant, and Δ = ES,1ES,0; ES,1 is the first excited-state energy in the singlet spectrum.
E 0 , 0 (respectively, E 1 U ) is the lowest energy of the singlet (respectively, triplet) spectrum (cf Figure 4). Under these conditions, we have:
J m , m = E 1 U E 0 , 0
where E 0 , 0 and E 1 U are given by Equations (34)–(37). The resulting energy level scheme is reported in Figure 4. As a result, a general theoretical expression of Jm,m may be written for the fragment A–X−A:
J m , m = C γ 1 1 S 2 1 + S 2 U + C + 2 γ 1 8 S γ 2 2 1 S 2 2 + E A g E A u S ( U + C + 2 γ 1 ) 2 1 + S 2 γ 2 1 S 2 2 2 + U C 2 2 .
In the physical case of a small overlap S, the previous equation reduces to:
J m , m = 2 γ 1 U C 2 + E A g E A u S U + C + 2 γ 1 + 2 γ 2 2 + U C 2 2 , E A g E A u 2 ( T S E ) 2 T ,   S   < <   1
where E , T , U, C, γ1 and γ2 are, respectively, given by Equations (16), (19) and (32). E A g E A u 2 T represents the electronic energy transfer between the magnetic cations and the non-magnetic ligand, in the case of weak overlap (S << 1).
Now, the physical criterion that allows us to describe compounds for which coulombic interactions are dominant naturally appears, i.e., E A g E A u 2 T < U , and compounds for which coulombic interactions become negligible, E A g E A u 2 T > U . However, we shall see that this first classification must also be slightly refined.
At this stage, it is worth noting that, in the absence of superexchange, there is no electronic transfer T = 0 and E A g E A u T = 0 = 0 . Let Jm,m′,0 be the corresponding exchange energy, always given by Equation (41), in which T = 0, U ≠ 0, C ≠ 0, γ1 ≠ 0 but γ2 = 0. We retrieve from Equation (41) that Jm,m′,0 = −2γ1, as expected. We just have an exchange between A and X on the one hand, and X and B on the other hand, without a connection between A and B.
In a previous paper [31], we evaluated the exchange energy magnitude between two atoms. We have shown that, in the case of a small overlap between A and X, and, if adapting to our conventional writing for Hamiltonian (here Hex = Js1.s2), the exchange energy magnitude is J ≈ −(juS2), where u is the Hartree term (direct term) and j the Fock term (exchange term), respectively, defined as:
u = dr 1 dr 2 Φ A ( r 1 ) 2 e 2 4 π ε 0 r 12 Φ X ( r 2 ) 2 , j = d r 1 d r 2 Φ A ( r 1 ) Φ X ( r 2 ) e 2 4 π ε 0 r 12 Φ X ( r 1 ) Φ A ( r 2 )
and similar expressions when A is replaced by B. If S << 1, on which we focus here, JAX = JXBj. The magnitude of JAX (respectively, JXB) reduces to the Fock (exchange) term, as expected, so that for the two bonds, A–X and X−B, considered separately, we have J0 ≈ −2j. If comparing with Equations (32) and (33), we have j ≈ γ1 and u = U so that finally we retrieve:
Jm,m,0 = J0 ≈ −2γ1.
thus, validating Equation (41).
We derive the difference of exchange energy between the superexchange (J) and no superexchange (Jm,m′,0 = J0):
J m , m J m , m , 0 = U C 2 + E A g E A u S U + C + 2 γ 1 + 2 γ 2 2 + U C 2 2 > 0 , S < < 2 .
However, it is worth writing Jm,m for the most commonly encountered physical case. Indeed, from the definitions of γ2, C, γ1 and U, given by Equation (32), we have the following classification:
γ2 << C ≈ γ1 << U
where the corresponding “physical” values range from tenths of an eV to a few eV. In addition, from a mathematical point of view, this classification is also due to the respective values of r1 and r2 (cf. Equations (32) and (33)) appearing in the arguments of the functions, giving the spatial behavior of the involved atomic orbitals.
Under these conditions, i.e., in the case of weak overlap, S << 1 on which we focus:
J m , m J m , m , 0 = U C 2 + E A g E A u 2 + U C 2 2 > 0 .
As a result, the physical discussion exclusively concerns the classification of the coulombic terms U, C and γ1 with respect to the difference | E Ag E Au|, which is very small, γ2 being strongly negligible. Thus, two kinds of situations may occur: E A g E A u 2 T < < U or E A g E A u 2 T > > U .
As a result, we derive the universal relationships if S << 1:
J m , m J m , m , 0 = 2 E A g E A u U C 2 8 T 2 ( U C ) 2 > > 1 ,   E A g E A u 2 T < < U , J m , m J m , m , 0 = E A g E A u 1 U C 2 E A g E A u 2 T 1 U C 4 T 2 T , E A g E A u 2 T > > U .
We conclude that it is the electronic transfer that enforces the superexchange.
  • Case 1:  E A g E A u 2 T < < U (see Figure 4).
The coulombic interactions, U, dominate the term E A g E A u 2 T , which represents the total electronic energy transfer between A and X, X and B. We are dealing with Class I compounds. For such materials, the size of the X-ligand orbital (s- or p-like) has the same order of magnitude as that of the A(B)-magnetic cation (d-like). We then have:
J m , m , 1 2 γ 1 + 8 T 2 ( U C ) 2 < 0
Jm,m′,1 may take values of order, varying between one eV to tenths of eV. We deal with ferromagnetic couplings. Thus, the coulombic interaction favors a ferromagnetic coupling. This is due to a subtle mechanism that is a consequence of the first Hund’s rule. This aspect is detailed in Section 3.2.1.
  • Case 2:  E A g E A u 2 T > > U (see Figure 5).
The contribution of coulombic interactions is now negligible vs. the small term E A g E A u 2 T , which represents the electronic transfer energy. When E A g E A u 2 T > > 2 S U + C + 2 γ 1 , we then have:
J m , m , 2 E A g E A u 1 U C 2 E A g E A u 2 T 1 U C 4 T 2 T > 0
Jm,m’,2 shows a small value in eV, as suggested by Anderson’s model. We are dealing with antiferromagnetic couplings. This mechanism is detailed in Section 3.2.2. because we now deal with a molecular orbital model. We may deal with Class I compounds (S << 1). However, this is always the case with Class II compounds because, due to the more or less important length of ligand X, the involved electrons show a very small coulombic interaction magnitude. E A g E A u 2 T + S E is greater now, with S < 1.
When E A g E A u < < 2 S U + C + 2 γ 1 | E Ag E Au| is smaller and we again find case 1: the couplings are now ferromagnetic, but the corresponding value of J is very small.
Here, there is an apparent contradiction with the experimental results. This would mean that there are no ferromagnets for Class II compounds characterized by a stronger magnitude of J. In a previous paper ([10] and the references therein), we explained the strategy for obtaining molecular organic ferromagnets, as well as the main models illustrating the charge transfer process. All these approaches are characterized by the introduction of excited (anomalously polarized) configurations in the Hamiltonian basis, similar to the polar states introduced in our model. In fact, they do not drastically differ from that scheme for the energy-lowering term 4(γ2 + T )2/(UC), involving the transfer integral T .

3.2. Physical Comments Regarding the Sign of J

3.2.1. Hund’s First Rule

Let us recall that the states | A > and | B > defined by Equation (7) are solutions of the same secular problem. We have a weak overlap, S . The eigenvalues of the 2 × 2 matrix appearing in the Hamiltonian expression given by Equations (30) and (34)–(37) may be written as:
E 0 , 0 + 2 E + U + γ 1 + 4 γ 2 2 U C , E 0 , 0 2 E + C + γ 1 4 γ 2 2 U C ,   S   < <   1
because E Ag + E Au = 2 E , E being given by Equation (16). For an atom, γ2 = 0 and, for a molecule or a polyatomic ion, γ2 ≠ 0. These eigenvalues must be compared to E 0 U and E 1 U :
E 0 U = 2 E + U γ 1 , E 1 U = 2 E + C γ 1 ,   S   < <   1 .
The order of magnitude of the coulombic terms has been pointed out in Equation (42). The ground level is the triplet state ET,0 and is stabilized by the factor 2γ1 (few eV for intra-atomic exchanges) with respect to the first excited level, ES,0, even in the presence of orbital degeneracy (cf. Figure 4). The ferromagnetic interaction based on the Coulomb exchange integral γ1 is called the Heisenberg exchange.
As the Pauli exclusion principle stipulates that two electrons cannot occupy the same position, when characterized by the same spin number, the corresponding wave function vanishes due to its property of antisymmetry. In other words, this function shows smooth variations, as requested for avoiding too large a kinetic energy.
Consequently, this keeps weak values when two electrons showing parallel spins are not too greatly separated. The probability density shows the same property, so that a hole appears in the up (down) spin distribution near one electron of the couple: this is the Fermi hole. In other words, the larger distance thus maintained between them allows us to explain that they are characterized by lower Coulomb energy compared to the scenario of antiparallel electronic spins.

3.2.2. Molecular Orbital Model

Here, the Coulomb interaction is small and, thus, plays a secondary role. The energy level scheme described in Figure 4 is replaced by that of Figure 5. Now, the ground state is given by one of the singlets |g,+; g,−> or |u,+; u,−>, if taking into account their relative stabilities (G-type collective state). As a result, the first excited state is the triplet | U 1 , S z >, with Sz = 0, ± 1.
This scenario occurs when the overlap between neighboring site orbitals is large (but S < 1), as it notably appears for C−C covalent bonds, for instance. Even in the case of orbital degeneracy, the low-lying state |g,σ> or |u,σ> is always available for the electronic pair, which may then couple in order to form a singlet or a triplet state. Finally, the first Hund’s rule informs us that the triplet is more stable, as often occurs.

3.3. Superexchange Hamiltonian for a 3d1 ion. Generalization to an ndm ion (3 ≤ n ≤ 5, 1 ≤ m ≤ 10)

Anderson has shown that, for d bands labeled m and m’, characterizing each ion on both sides of ligand X, the corresponding Hamiltonian may be written:
H = m , m , n , τ 2 b m m ( τ ) 2 U s n m . s n + τ m + C s t
where the τ’s represent the fundamental translations of the lattice. bm,m(τ) is the transfer integral between d bands labeled m and m’; (m,m’) is the corresponding pair of involved bands; U > 0 is the Coulomb repulsion energy. Anderson has called this contribution the “kinetic” part of the exchange Jm,m((kinetic). The “potential” part of the exchange, Jm,m(potential), is defined at the end of Section 2.1.2, and the total exchange is given by Jm,m = Jm,m((kinetic) + Jm,m(potential).
The formalism that describes the superexchange phenomenon, according to Anderson, exclusively reduces the Hamiltonian to the transfer integral, i.e., the kinetic contribution. In our model, we have seen that we have a more general formalism. We wish to show that the Hamiltonian may be written in the formula:
H = K s 1 m . s 2 m ,   s 1 m . s 2 m = 1 4 P T 3 P S .
where PT and PS are projectors:
Magnetochemistry 08 00006 i002
acting in the triplet and singlet subspaces, respectively. K is a constant by which to determine values.
After a few algebraic calculations, it is easy to show that the projectors have the following properties:
Magnetochemistry 08 00006 i003
where Xt is a transposed matrix, and Magnetochemistry 08 00006 i001 is the 4 × 4 identity matrix. Under these conditions, it is easy to show:
PT −3PS = (PTPS)(PT − 3PS) − 6PS
so that:
H = K 4 ( P T P S ) ( P T 3 P S ) 6 P S .
Let |Ψ> be the generic collective wave function and let us introduce the projector |Ψ>< Ψ| = 1 on the basis vectors. We then have:
H Ψ = K 4 ( P T P S ) Ψ Ψ ( P T 3 P S ) 6 P S Ψ
with:
Ψ ( P T 3 P S ) = ( P T t 3 P S t ) Ψ = ( P T 3 P S ) Ψ
due to Equation (50), and the fact that |Ψ> = <Ψ| because |Ψ> is real. Finally, using the fact that <Ψ|Ψ> = 1, we may finally write:
H Ψ = K 4 ( P T P S ) Ψ ( P T 3 P S ) 6 P S Ψ .
According to the Hamilton principle, which states that in the true evolution of a physical system, described by the generalized coordinates q(t), the action S must be minimal with respect to q(t). The action S may be expressed with the Lagrangian system, L q ( t ) , q ( t ) , where q ( t ) = d q ( t ) / d t and, after an adequate Legendre transform, with the Hamiltonian system H. As a result, minimizing S means that, at the equilibrium, H must be minimized. In gauge field theories, notably applied in particle physics, this process is called a minimal coupling scheme. Under these conditions, the equilibrium of any dynamic situation is described by the following Hamiltonian:
min H Ψ = min K 4 ( P T P S ) Ψ ( P T 3 P S ) 6 P S Ψ
where the operator min(x) exclusively applies on a number x (for the Hamiltonian applied to a vector |Ψ>, it concerns the result of the corresponding application). We then have:
min K 4 ( P T P S ) Ψ = 1 4 ( E T , 0 m , m E S , 0 m , m ) Ψ = J m , m Ψ ,   min P S Ψ = E S , 0 m , m Ψ
where E T , 0 m , m and E S , 0 m , m are the lowest energies of the triplet and singlet states, respectively, for d bands m and m’, characterizing each ion located on each side of the ligand, X. We recognize J m , m = E T , 0 m , m E S , 0 m , m . As a result, at the equilibrium, we have:
H e q Ψ = min H Ψ = J m , m 4 Ψ ( P T 3 P S ) 3 2 E S , 0 m , m Ψ ,   K = 1 .
Due to Equations (54) and (49), we then have:
Ψ ( P T 3 P S ) = Ψ ( P T 3 P S ) = ( P T t 3 P S t ) Ψ = ( P T 3 P S ) Ψ
so that finally, for an ion 3d1 characterized by bands m and m’ involved on each side of ligand X:
H e q = J m , m 4 ( P T 3 P S ) 3 2 E S , 0 m , m 1 = J m , m s 1 m s 2 m 3 2 E S , 0 m , m 1 ( ion   3 d 1 )
where Jm,m is given by Equation (40).
When dealing with several d bands located on each side of ligand X, we suppose that these bands are independent, i.e., non-interacting between themselves through the coulombic processes. We simply have:
H = ( m , m ) J m , m s 1 m s 2 m 3 2 E S , 0 m , m 1 ( ion   n d m ,   3   n   5 ,   1 m 5 ) .
Indeed, as seen on Figure 1, for ions nd1 to nd5, we have different half-filled electronic shells, so that the value of the couple (m,m’) varies between 1 and 5. If this value is greater than 5, only the half-filled shells intervene in the calculation of Jm,m so that an ion of nd6 (L = 2, S = 2), nd7 (L = 3, S = 3/2), nd8 (L = 3, S = 1) and nd9 (L = 2, S = 1/2) is treated as an ion nd4, nd3, nd2, and nd1, respectively.

3.4. Introduction of Crystal Field Theory

3.4.1. Expression of Jm,m; Physical Discussion of the Crystal Field Effect

The crystal field theory was notably introduced by Van Vleck in the 1930s [32]. When considering the ions of the cage in which each magnetic cation is inserted, we observe that each ion, k, can be described by a unique charge, qk, due to the fact that there are very weak overlaps between first-nearest neighbors. We do not discuss in this paper the symmetry properties of the cage and the correlative simplifications that may occur.
In this subsection, we will only discuss the effect of the crystal field on the nature of exchange energy, J, involved in the coupling of first-nearest spin neighbors. We show that, starting from an isotropic (Heisenberg) exchange coupling when there is no crystal field, the appearance of a crystal field induces an anisotropy of exchange coupling, thus leading to a z-z (Ising-like) coupling or an x-y one. For the sake of simplicity, we do not consider the other cases, but the corresponding problem is solvable.
Thus, we only discuss the effects of a weak crystal field magnitude (3d ions) to a stronger (4d ions) and then stronger one (5d ions). Let us recall that this effect is due to the fact that the mean radius of the transition ion increases when passing from 3d ions to 5d ions, whereas the dimensions of the surrounding cage in which these ions are inserted roughly show the same size (we exclude the scenario of distortions that might occur).
However, we do not discuss the effect of the crystal field in the splitting of orbital degeneracy.
Let O be an arbitrary origin, OX1, OX2 and OX3, three orthonormal axes and r = (x1, x2, x3) the Cartesian coordinates of any point M bearing an electron (with the usual correspondence, x1 = x, x2 = y, x3 = z). As a result, the potential describing the crystal field effect may be written as:
V C F ( r ) = k q k e 4 π ε 0 R k r ,   q k e   >   0   or   q k e   <   0
where Rk = (X1,X2,X3) is the position vector associated with each ion k of the surrounding cage (Rk = Cst). Considering the two cases, qke > 0 or qke < 0, means that we envisage the possibility not only of an electrically neutral entity constituted by the cage and the inserted cation A or B but also that the set cage + cation may show a positive or a negative electrical charge. Finally, we may write the full electrostatic contributions of the crystal field for the fragment A–X (electron 1 coming from cation A and electron 2 coming from anion X):
V A X C F ( r ) = k q k e 4 π ε 0 1 R k r 1 + 1 R k r 2
and, for the fragment X−B (electron 3 coming from anion X and electron 4 coming from cation B):
V X B C F ( r ) = k q k e 4 π ε 0 1 R k r 3 + 1 R k r 4
for similar insertion cages (with A = B or, exceptionally, for A ≠ B). Note that it is possible to consider different insertion cages, i.e., cages showing different geometrical properties. In this respect, we must add that this reasoning is valid for Class I compounds characterized by a ligand orbital length of the same order as cationic orbital length. However, for Class II compounds characterized by a ligand orbital length that is plainly greater than the cationic one, it has no physical sense by which to consider crystal field effects. The latter may only come from the electrostatic environment of the cation itself, which may be inserted within a crystallographic site.
As a result, the full contribution of the crystal field for the fragment A–X–B is:
U anis = V A X C F ( r ) + V X B C F ( r ) = k q k e 4 π ε 0 i = 1 4 1 R k r i ,   A B .
When A = B, on which we focus for the sake of simplicity, we have:
U anis = 2 V A X C F ( r ) = 2 k q k e 4 π ε 0 i = 1 2 1 R k r i ,   A = B .
Under these conditions, only the coulombic contributions defined by Equation (32) and called Xiso in the absence of a crystal field are altered. We then have:
U anis = U A anis = A 1 A 1 U 12 2 k i = 1 2 q k e 4 π ε 0 R k r i A 1 A 1 = U A iso + U A C F , C anis = A 1 A 2 U 12 2 k i = 1 2 q k e 4 π ε 0 R k r i A 2 A 1 = C iso + U C C F , γ 1 anis = A 1 A 2 U 12 2 k i = 1 2 q k e 4 π ε 0 R k r i A 1 A 2 = γ 1 iso + U γ 1 C F , γ 2 anis = A 1 A 1 U 12 2 k i = 1 2 q k e 4 π ε 0 R k r i A 1 A 2 = γ 2 iso + U γ 2 C F ,   A = B .
As previously noted, for Class II compounds characterized by long ligand chains, there is no crystal field and factor 2 must be replaced by unity in Equation (63). In addition, in Appendix A, we have shown that the following physical parameters T and V, E A and T , E Ag and E Au, respectively, as given by Equations (1), (16), (18) and (22), the coulombic terms U A iso , C iso , γ 1 iso and γ 2 iso , given by Equation (32), and the coulombic contributions of crystal field U A C F , U C C F , U γ 1 C F and U γ 2 C F defined in Equation (63) can be labeled by the generic term Yk and decomposed as a sum of their respective Cartesian components along the axes OX1, OX2 and OX3, called Y k ; x i (as a scalar is a zero-rank tensor):
Y k = i = 1 3 Y k ; x i , Y k ; x i = Y k 3
because, for each ion, k, we deal with a spherical symmetry (cf. Equations (A6)–(A10)). However, for a given similar xi-component Y k x i , components Y k characterized by different k values are not equal (except when the site in which the cation is inserted shows particular symmetry properties). Due to the definitions of U A C F , U C C F , U γ 1 C F and U γ 2 C F (cf Equation (63)), we may give the same mathematical argument regarding the respective values of r and Rk appearing in the argument of the atomic orbitals (cf Equation (A12) in Appendix A). We have (cf. Equation (42)):
U γ 2 C F < <   U C C F   U γ 1 C F < < U A C F
and a similar inequality with the Cartesian components, for a given k.
The Hamiltonian matrix, defined by Equation (30) in the absence of a crystal field, now appears as:
H = i = 1 3 H x i anis , H x i anis = E 0 , x i G g , anis K x i anis 0 0 0 0 K x i anis E 0 , x i G u , anis 0 0 0 0 0 0 E 0 , x i U , anis 0 0 0 0 0 0 E 1 , x i U , anis 0 0 0 0 0 0 E 1 , x i U , anis 0 0 0 0 0 0 E 1 , x i U , anis
with:
E 0 , x i G g , anis = E 0 , x i G g , iso + U A , x i C F + U C , x i C F + 2 U γ 1 , x i C F + 4 U γ 2 , x i C F 2 ( 1 + S ) 2 , E 0 , x i G u , anis = E 0 , x i G u , iso + U A , x i C F + U C , x i C F + 2 U γ 1 , x i C F 4 U γ 2 , x i C F 2 ( 1 S ) 2 , K x i anis = K x i iso + U A , x i C F U C , x i C F 2 ( 1 S 2 ) , E 0 , x i U , anis = E 0 , x i U , iso + U A , x i C F U γ 1 , x i C F 1 S 2 ,   E 1 , x i U , anis = E 1 , x i U , iso + U C , x i C F U γ 1 , x i C F 1 S 2 ,   A = B ,
where the corresponding isotropic component, for each of the quantities above, is given by Equation (34). Similar remarks, given after Equation (36) and concerning the scenario where A ≠ B, remain valid here. If diagonalizing the upper 2 × 2 matrix in Equation (66), we have the following eigenvalues:
E 0 , 0 , x i ± , anis = 1 2 E 0 , x i G g , anis + E 0 , x i G u , anis ± ( E 0 , x i G g , anis E 0 , x i G u , anis ) 2 + 4 K x i anis 2 .
As a result, for the fragment A–X−A, the general theoretical expression of J may be written from Equation (40), where each physical quantity is replaced by the same one in the presence of a crystal field and given by Equation (67):
J m , m = C anis γ 1 anis 1 S 2 1 + S 2 U anis + C anis + 2 γ 1 anis 8 S γ 2 anis 2 1 S 2 2 + E A g anis E A u anis S ( U anis + C anis + 2 γ 1 anis ) 2 1 + S 2 γ 2 anis 1 S 2 2 2 + U anis C anis 2 2 .
In the physical case of a small overlap, S, on which we focus, the previous equation reduces to:
J m , m = 2 γ 1 anis U anis C anis 2 + E A g anis E A u anis S U anis + C anis + 2 γ 1 anis + 2 γ 2 anis 2 + U anis C anis 2 2 , S   < <   1
where all the physical quantities can be expressed along the axes OX1, OX2 and OX3, according to Equation (64).
As discussed in the scenario where there is no crystal field after Equation (41), using similar reasoning, it becomes easy to show that the difference of exchange energy between superexchange (Jm,m) and no superexchange (Jm,m’,0) is:
J m , m J m , m , 0 = U anis C anis 2 + E A g anis E A u anis S U anis + C anis + 2 γ 1 anis + 2 γ 2 anis 2 + U anis C anis 2 2 > 0 , S   < <   1 .
As in the absence of a crystal field, we have to compare the magnitude of the coulombic interactions U anis to the quantity E A g anis E A u anis = E A g iso E A u iso = E A g E A u = 2 T .
As a result, we may also derive the universal relationships if S << 1:
J m , m J m , m , 0 = 2 E A g E A u U anis C anis 2 8 T 2 U anis C anis 2 > > 1 , E A g E A u 2 T < < U anis , J m , m J m , m , 0 = E A g E A u 1 U anis C anis 2 E A g E A u 2 T 1 U anis C anis 4 T 2 T , E A g E A u 2 T > > U anis .
This means that the appearance of a crystal field also strengthens the superexchange.
  • Case 1:  E A g anis E A u anis = E A g iso E A u iso 2 T < < U anis .
The coulombic interactions, including those coming from the crystal field, dominate the term E A g anis E A u anis = E A g iso E A u iso 2 T . We then have:
J m , m , 1 anis 2 γ 1 anis 2 γ 1 iso + U γ 1 C F J m , m , 1 iso 2 U γ 1 C F .
with:
J m , m , 1 anis = i = 1 3 J m , m , x i , 1 anis = J m , m , 1 x x + J m , m , 1 y y + J m , m , 1 z z , U γ 1 , x C F U γ 1 , y C F U γ 1 , z C F .
As discussed, when there is no crystal field, we also deal here with Class I compounds. For such materials, the size of the X-ligand orbital (s- or p-like) has the same order of magnitude as that of A(B)-magnetic cation (d-like). However, we may also have Class II compounds (the X-ligand length is plainly greater than in the cation), thus imposing a weak crystal ferromagnetic contribution in absolute value (see below).
If U γ 1 C F > 0 (qke < 0), as γ 1 iso > 0 and J m , m , 1 iso = 2 γ 1 iso < 0 (cf. Equation (43)), we always have J m , m , 1 anis < 0 according to Equation (71), and all the coulombic interactions favor a ferromagnetic coupling. This effect is enhanced by the crystal field (CF) contribution, independently of its magnitude. This is also due to a subtle mechanism that is a consequence of the first Hund’s rule, in spite of the presence of the crystal field. This aspect has been detailed in Section 3.2.1. This allows us now to justify the fact that Class I and Class II compounds enter this case.
If U γ 1 C F < 0 (qke > 0), we have two possibilities, when examining Equation (71):
  • γ 1 iso > > 2 U γ 1 C F > 0 ; we deal with a weak CF contribution (in the case of 3dm ions); J m , m , 1 anis 2 γ 1 iso < 0 ; the surrounding cage is mainly characterized by an important geometrical size: we may deal with Class I compounds. This is also the case when using organic ligands whose long length may be adapted to the magnetic system that one wishes to build up [15,16,17,18]; this is a good way to obtain isotropic (Heisenberg) spin-spin couplings for Class II compounds; we always have ferromagnetic couplings, including in the particular case of γ 1 iso 2 U γ 1 C F > 0 so that, finally, the ferromagnetic coupling is strongly enhanced and J m , m , 1 anis 4 γ 1 iso 4 U γ 1 C F < 0 (in the case of ions 4dm and 5dm);
  • γ 1 iso < < 2 U γ 1 C F > 0 ; we deal with a strong CF contribution (case of 5dm ions); now we have J m , m , 1 anis 2 U γ 1 C F > 0 ; antiferromagnetic couplings are favored and this only concerns Class I compounds.
Now we examine the influence of the crystal field over the initial isotropic character of exchange coupling. As previously announced at the beginning of the present subsection, we restrict our study to those cases where the axis OX3 = Oz is favored (with respect to OX1 = Ox and OX2 = Oy) or the axes OX1 = Ox and OX2 = Oy are equally favored by the crystal field (with respect to OX3 = Oz). Then, we have:
J m , m , 1 anis J m , m , 1 z z if   U γ 1 , x C F U γ 1 , y C F < < U γ 1 , z C F ( z - z   coupling ;   here   Ising   coupling   as   s = 1 / 2 ) , J m , m , 1 anis J m , m , 1 x x + J m , m , 1 y y if   U γ 1 , x C F U γ 1 , y C F > > U γ 1 , z C F ( x - y   coupling ) .  
We can have J m , m , 1 anis < 0 or J m , m , 1 anis > 0 . This is in the case of 4d and 5d ions.
  • Case 2:  E A g anis E A u anis = E A g iso E A u iso 2 T > > U anis
The contribution of coulombic interactions is weak vs. the small term| E Ag E Au|, which is insensitive to the crystal field by definition. This is in the case of 3d ions.
We have the same discussion as in the isotropic case, i.e., without a crystal field. When E A g iso E A u iso > > 2 S U + C + 2 γ 1 , we have:
J m , m , 2 E A g iso E A u iso > 0 .
We always deal with antiferromagnetic couplings independently of the crystal field. This mechanism is detailed in Section 3.2.2 because we now deal with a molecular orbital model.
When E A g iso E A u iso < < 2 S U anis + C anis + 2 γ 1 anis E A g iso E A u iso is smaller, we find again Case 1: the couplings are now ferromagnetic, but the corresponding value of J is very small.
As discussed, in the absence of a crystal field we deal with Class I compounds (S << 1). However, this is always the case with Class II compounds because, due to a more or less important length of ligand X, the involved electrons show a very small coulombic interaction magnitude. E A g E A u 2 T + S E is greater, now S < 1. Similar remarks may be made for molecular organic ferromagnets.

3.4.2. Expression of the Hamiltonian

We wish to show that the Hamiltonian may be written in the form of:
H = ( m , m ) J m , m x x s 1 x , m . s 2 x , m + J m , m x y s 1 y , m . s 2 y , m + J m , m z z s 1 z , m . s 2 z , m .
By an analogy with the isotropic case, we define the following projectors:
s 1 x , m . s 2 x , m + s 1 y , m . s 2 y , m = 1 4 P T 3 P S x y ,   s 1 z , m . s 2 z , m = 1 4 P T 3 P S z z ,
with the properties:
P T 3 P S x y + P T 3 P S z z = P T 3 P S ,   P T 3 P S x y . P T 3 P S z z = 0
where PT and PS are projectors (defined by Equation (49)) and are acting in the triplet and singlet subspaces, respectively. P T 3 P S x y is the projector acting in the subspace corresponding to the x-y plane and P T 3 P S z z acts along the z-axis.
Similar reasoning performed for the isotropic case, i.e., without a crystal field (cf. Section 3.3) allows one to directly write in the presence of a crystal field:
H e q , x y = J m , m x y 4 ( P T 3 P S ) x y 3 2 E S , 0 x y ; m , m I = J m , m x y s 1 x , m s 2 x , m + s 1 y , m s 2 y , m 3 2 E S , 0 x y ; m , m I x - y   plane   favored   by   CF   ( ion   nd 1 ,   3     n     5 )
and:
H e q , z z = J m , m z z 4 ( P T 3 P S ) z z 3 2 E S , 0 z z ; m , m 1 = J m , m z z s 1 z , m s 2 z , m 3 2 E S , 0 z z ; m , m 1 z - axis   favored   by   CF   ( ion   nd 1 ,   3 n 5 )
where J m , m x y and J m , m z z are given by Equations (71)–(74).
When dealing with several d bands, located on each side of ligand X, we again suppose that these bands are independent, i.e., non-interacting between themselves through coulombic processes. We simply have:
H = ( m , m ) J m , m x y s 1 x , m s 2 x , m + s 1 y , m s 2 y , m 3 2 E S , 0 x y ; m , m 1 + J m , m z z s 1 z , m s 2 z , m 3 2 E S , 0 z z ; m , m 1 ( ion   nd m ,   3 5 ,   1 5 ) .

4. Conclusions

In this article, we have developed a general model allowing us to describe the underlying microscopic mechanisms leading to superexchange in terms of the isolated centrosymmetric fragment A–X–B, where A and B are magnetic sites bearing d orbitals and X is a diamagnetic ligand involving an s or p orbital. We have considered not only the degenerate case (A = B) but also the general one (A ≠ B). The orbitals describing the states |A> and |B> do not overlap but the important scenario of overlap through a π bond may also be taken into account.
The energy spectrum has been constructed. Notably, from the single-triplet splitting, a closed-form expression of the exchange energy J has been expressed vs. key molecular integrals, for a given 3d1 ion. For the first time, this result has allowed us to predict the sign of J and its magnitude: when coulombic interactions are dominant, our model follows Hund’s rule and we explain why couplings are automatically ferromagnetic (J < 0); when coulombic interactions are no longer dominant, our model is equivalent to the molecular orbital one and couplings are always antiferromagnetic (J > 0), except in one particular case where couplings are ferromagnetic but with a small absolute value of J.
Owing to this general formalism, we have shown for the first time that the corresponding Hamiltonian may be written using the formula Js1.s2. The model has been easily generalized in the case of the ndm ion (3 ≤ n ≤ 5, 1 ≤ m ≤ 10, orbital degeneracy).
Finally, we have introduced crystal field theory and we have shown how we can pass from an isotropic (Heisenberg) coupling (no crystal field, 3d ions) to a stronger (4d ions) and stronger (5d ions) contribution, which introduces an anisotropic exchange coupling. The nature of the spin arrangement has been determined, vs. the crystal field magnitude. We have shown that our formalism also allows us to write a spin-exchange Hamiltonian.
Important generalizations may be made. Notably, spin polarization effects may be introduced for the fragment A–X–B, as well as spin-orbit coupling. In particular, if this latter contribution remains small, the formalism may be slightly altered through a perturbation expansion, the zeroth-order orbitals being replaced by new magnetic orbitals, taking into account the spin-orbit perturbation.
This work has previously been conducted by Moriya, but he used Anderson’s theory as a starting point [33,34]. As a result, one can say that the Dzialoshinskii coupling is a “close parent” of the superexchange coupling [35,36]. Moriya has notably determined the crystal symmetries that must be fulfilled for favoring the establishment of an antisymmetric spin coupling. More recently, Katsnelson et al. also used Anderson’s theory and an exact perturbation expansion of the total energy of weak ferromagnets showing a canted spin structure, with the unique assumption of local Hubbard-type interactions. This allows the expression of the modulus of the Dzialoshinskii–Moriya vector, with a natural separation into spin and orbital contributions [37]. However, these models do not take into account all the key molecular integrals detailed in this paper or allow the expression of superexchange energy J exactly.
The most important key point concerns the possible generalization of our formalism to any kind of molecule or polyatomic ion, opening the possibility of exact calculations of exchange couplings in real molecules such as polymers or biopolymers. It is a promising and important field of research in biology. Finally, the exact knowledge of the energy spectrum may allow others to interpret the experiences of photomagnetism that have recently been published, concerning ions Mo(IV) and W(IV) [38].

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

Appendix A

Let O be an arbitrary origin, OX1, OX2 and OX3 three orthonormal axes and (x1,x2,x3) the coordinates of any point M bearing a mass me and a charge −e. The kinetic energy T, defined by Equation (1), may be written as:
T = 1 2 m e i = 1 3 x i 2 = i = 1 3 T x i , x i = d x i d t ,   r j = O M j = ( x j ; 1 , x j ; 2 , x j ; 3 ) ,   j = 1 , 2  
where the ith-space coordinate of species rj is labeled xj;i for the sake of simplicity. The potential V, also defined in Equation (1), represents the interaction of one electron with respect to the rest of the cation. If dealing with a spherical system, we have:
V = k d r 4 π ε 0 r = k ln r + C s t , r = x 1 2 + x 2 2 + x 3 2
where k = −Qe is the electric charge of cation. We neglect the gravitational potential energy. As we deal with spherical symmetry, we have three equal contributions when expressing V along each axis OXi, i = 1, 3. As a result, after the adequate integration of the first of Equation (A2), V is invariant by the permutation of x1, x2 and x3, and may be artificially written:
V = i = 1 3 V x i , V x i = V 3 .
From Equation (33), the coulombic interactions may be defined by the generic form:
U c = < W | < X | U 1 , 2 | Y > | Z > = d r 1 d r 2 Φ W ( r 1 ) Φ X ( r 2 ) K 4 π ε 0 r 12 Φ Y ( r 2 ) Φ Z ( r 1 )
where r12 = |r1r2|, K = e2 for UA, UB, γ1 and γ2 given by Equation (33); dri represents the elementary volume in spherical coordinates i.e., ri2sinθidθidφi. In all cases, K > 0. r1 and r2 are the position vectors representing electron 1 from cation A, respectively, electron 2 from anion X, for instance. In a spherical coordinate system, we deal with the triplet (rjjj), defined as follows:
θ j = arccos x j ; 3 r j , φ j = arctan x j ; 2 x j ; 1 , r j = i = 1 3 x j ; i 2 ,   j = 1 , 2 .  
1/r12 may be expressed under the symmetric form with respect to the exchange of indices 1 and 2 [39]:
1 r 12 = 4 π r 1 2 + r 2 2 l = 0 + n = l , l + 2 + K n , l r 1 r 2 r 1 2 + r 2 2 n m = l + l Y l , m ( θ 1 , φ 1 ) Y l , m ( θ 2 , φ 2 ) , K n , l = ( 2 n 1 ) ! ! ( n l ) ! ! ( l + n + 1 ) ! ! .
In the previous equation, the notation of the current index of the n-summation means that n only takes l-values with an (l + 2)-step; Yl,m(θ,φ) represents the well-known spherical harmonics [39].
As a result, we have:
U c = K ε 0 d r 1 d r 2 f ( r 1 , r 2 )
with:
f ( r 1 , r 2 ) = l = 0 + n = l , l + 2 + K n , l r 1 r 2 n r 1 2 + r 2 2 n + 1 / 2 m = l + l Y l , m ( θ 1 , φ 1 ) Y l , m ( θ 2 , φ 2 ) × × Φ W ( r 1 ) Φ X ( r 2 ) Φ Y ( r 2 ) Φ Z ( r 1 ) .
The integration over angular variables is easy: each integral is equal to unity with the selection rule l = 0, m = 0. As a result, we finally have the radial contribution:
U c = K ε 0 n = 0 , 2 + K n , 0 d r 1 d r 2 r 1 r 2 n r 1 2 + r 2 2 n + 1 / 2 Φ W ( r 1 ) Φ X ( r 2 ) Φ Y ( r 2 ) Φ Z ( r 1 ) .
In the previous equation, n only takes even values. This complicated integral may be uniquely calculated numerically. In addition, as we deal with the isotropic case of spherical coordinates, Uc may be decomposed into 3 Cartesian components, U c x i , so that:
U c = i = 1 3 U c x i , U c x i = U c 3 .
Thus, any coulombic interaction Uc may be decomposed into a system of Cartesian coordinates along each axis OXi, i = 1,3.
Now let us introduce the notion of crystal field. For the sake of simplicity, we restrict ourselves to the case of a unique cation A or B (respectively, anion X) inserted within a crystallographic site. The reasoning may be extended to all the ions involved in A–X–B. The potential describing the crystal field effect may be written as:
V C F ( r ) = k q k e 4 π ε 0 R k r
where Rk = (X1,X2,X3) is the position vector associated with each ion k (Rk = Cst) and r = (x1,x2,x3) is that of electron belonging to cation A (or B) or anion X. Under these conditions, the corresponding crystal field contribution may be written as:
U c C F = k Φ X ( R k ) Φ Y ( R k ) d r Φ W ( r ) q k e 4 π ε 0 R k r Φ Z ( r ) .
The expansion of |Rkr|−1 has been examined by Hutchings [40] in the assumption r < Rk, near an origin of cubic point symmetry corresponding to the three most-encountered situations (see Figure 2 in [40]):
  • The charges composing the cage are located at the corners of an octahedron (sixfold coordination); in that case, Rk = (a,0,0), (−a,0,0), (0,a,0), (0,−a,0), (0,0,a) and (0,0,−a); Rk = a is the distance of each corner from the origin;
  • The charges are located at the corners of a cube (eightfold coordination); Rk = (a,a,a), (−a,a,a), (a,−a,a), (−a,−a, a), (a,−a,−a), (a,a,−a), (−a, a,−a) and (−a,−a,−a); R k = a 3 ;
  • The charges are located at the corners of a tetrahedron (with two tetrahedra per cube, fourfold coordination); Rk = (a,a,a), (−a,−a,a), (a,−a,−a), (−a,a,−a) for tetrahedron 1 and Rk = (a,−a,a), (−a,a,a), (−a,−a, −a), (a,a,−a) for tetrahedron 2; R k = a 3 .
In a first step, Hutchings has expanded |Rkr|−1 in Cartesian coordinates, with the assumption that r/Rk < 1 [40]. Under these conditions, this author obtains a perturbation series analogous to the famous dipolar expansion. He shows that V C F ( r ) may be written with Cartesian coordinates under the generic form:
V C F ( x 1 , x 2 , x 3 ) = A + V x 1 C F + V x 2 C F + V x 3 C F + V mix ( x 1 , x 2 , x 3 )
A is a constant composed of a product between a numerical factor and a ratio whose numerator is the coordination number, q, and the denominator is the distance of the cage charges from the origin. V x 1 C F (respectively, V x 2 C F and V x 3 C F ) are the polynomial of even powers of exclusive variables x1 (respectively, x2, x3).
V x 1 C F , V x 2 C F and V x 3 C F may be derived from each other through a permutation of variables x1, x2, and x3 but, according to site symmetries, we may have not only V x 1 C F = V x 2 C F = V x 3 C F but also V x 1 C F V x 2 C F V x 3 C F . V m i x ( x 1 , x 2 , x 3 ) is a polynomial characterized by terms composed of products separately involving two variables x1 and x2, x2 and x3, x3 and x1, showing even different powers, but it is impossible to separate variables x1, x2, and x3 (if odd powers intervene in higher-order terms, the final result of integration vanishes). As a result, Equation (A13) may be rewritten as:
V C F ( x 1 , x 2 , x 3 ) = A + V x 1 C F ( x 1 ) + V x 2 C F ( x 2 ) + V x 3 C F ( x 3 ) + V ˜ x 1 , x 2 ( x 1 , x 2 ) + V ˜ x 2 , x 3 ( x 2 , x 3 ) + V ˜ x 3 , x 2 ( x 3 , x 2 ) .
This is a quadratic form. It is then possible to express V C F ( x 1 , x 2 , x 3 ) in a new eigenbasis, under the following form:
V C F ( x ˜ 1 , x ˜ 2 , x ˜ 3 ) = V x ˜ 1 C F + V x ˜ 2 C F + V x ˜ 3 C F .
If inserting this expansion in Equation (A12) and integrating over variables x1, x2 and x3,  U c C F appears under the form:
U c C F = U x ˜ 1 C F + U x ˜ 2 C F + U x ˜ 3 C F ,   U x ˜ 1 C F U x ˜ 2 C F U x ˜ 3 C F .
Hutchings has also calculated the matrix element Uc given by Equation (A12), in which |Rkr|−1 is expanded in terms of spherical harmonics. The matrix element Uc and correlated general rules of calculation may be found from this expansion, using vector coupling coefficients (see [41], p. 37). Hutchings recalls the rules for determining nonzero matrix elements, which follow from the Wigner–Eckart theorem (see [41], p. 75).
The conclusion derived by Hutchings is that, in a final step, Uc may be also expressed owing to Cartesian components. This method allows us to stop the expansion vs. spherical harmonics Yl,m(θ,φ) for small values of l and m, if using symmetries derived from that of the host site, whereas the direct calculation of the integral giving U c C F is longer because it is exclusively ruled by the convergence of the series, giving V C F ( x 1 , x 2 , x 3 ) (cf. Equation (A9)).
As a result, the anisotropic contributions of the crystal field appear as one of the main origins of the anisotropy of exchange.

References

  1. Anderson, P.W. Antiferromagnetism. Theory of superexchange interaction. Phys. Rev. 1950, 79, 350. [Google Scholar] [CrossRef]
  2. Anderson, P.W. New approach to the theory of superexchange interactions. Phys. Rev. 1959, 115, 2. [Google Scholar] [CrossRef]
  3. Anderson, P.W. Theory of magnetic exchange interactions: Exchange in insulators and semiconductors. Solid State Phys. 1963, 14, 99. [Google Scholar]
  4. Anderson, P.W. Magnetism; Rado, G.T., Suhl, H., Eds.; Wiley: New York, NY, USA, 1962; Volume I, p. 25. [Google Scholar]
  5. Kanamori, J. Theory of the magnetic properties of ferrous and cobaltous oxides, I. Prog. Theor. Phys. 1957, 17, 177. [Google Scholar] [CrossRef] [Green Version]
  6. Kanamori, J. Theory of the magnetic properties of ferrous and cobaltous oxides, II. Prog. Theor. Phys. 1957, 17, 197. [Google Scholar] [CrossRef] [Green Version]
  7. Van Vleck, J.H. Note on the use of the Dirac vector model in magnetic materials. Rev. Mat. Fis. Teor. 1962, 14, 189. [Google Scholar]
  8. Levy, P.M. Rare-earth-iron exchange interaction in the garnets. I. Hamiltonian for anisotropic exchange interaction. Phys. Rev. 1964, 135, A155. [Google Scholar] [CrossRef]
  9. Levy, P.M. Rare-earth-iron exchange interaction in the garnets. II. Exchange Potential for Ytterbium. Phys. Rev. 1966, 147, 311. [Google Scholar] [CrossRef]
  10. Curély, J. Magnetic orbitals and mechanisms of exchange. II. Superexchange. Mon. Für Chem. 2005, 136, 1013. [Google Scholar] [CrossRef]
  11. Goodenough, J.B. Direct cation-cation interactions in several oxides. Phys. Rev. 1960, 117, 1442. [Google Scholar] [CrossRef]
  12. Goodenough, J.B. Theory of the role of covalence in the perovskite-type manganites [La,M(II)]MnO3. Phys. Rev. 1955, 100, 564. [Google Scholar] [CrossRef] [Green Version]
  13. Goodenough, J.B. An interpretation of the magnetic properties of the perovskite-type mixed crystals La1−xSrxCoO3−λ. J. Phys. Chem. Solids 1958, 6, 287. [Google Scholar] [CrossRef]
  14. Kanamori, J. Superexchange interaction and symmetry properties of electron orbitals. J. Phys. Chem. Solids 1959, 10, 87. [Google Scholar] [CrossRef]
  15. Escuer, A.; Vicente, R.; Goher, M.A.S.; Mautner, F.A. Synthesis and structural characterization of [Mn(ethyl isonicotinate)2(N3)2]n, a two-Dimensional alternating ferromagnetic-antiferromagnetic compound. Magnetostructural correlations for the end-to-end pseudohalide-manganese system. Inorg. Chem. 1996, 35, 6386. [Google Scholar] [CrossRef]
  16. Escuer, A.; Vicente, R.; Goher, M.A.S.; Mautner, F.A. A new two-dimensional manganese(II)-azide polymer. Synthesis, structure and magnetic properties of [{Mn(minc)2(N3)2}]n (minc = methyl isonicotinate). J. Chem. Soc. Dalton Trans. 1997, 22, 4431. [Google Scholar] [CrossRef]
  17. Goher, M.A.S.; Morsy, A.M.A.-Y.; Mautner, F.A.; Vicente, R.; Escuer, A. Superexchange interactions through quasi-linear end-to-end azido bridges: Structural and magnetic characterisation of a new two-dimensional manganese-azido system [Mn(DENA)2(N3)2]n (DENA = diethylnicotinamide). Eur. J. Inorg. Chem. 2000, 8, 1819. [Google Scholar] [CrossRef]
  18. Escuer, A.; Esteban, J.; Perlepes, S.P.; Stamatatos, T.C. The bridging azido ligand as a central “player” in high-nuclearity 3d-metal cluster chemistry. Coord. Chem. Rev. 2014, 275, 87, and references therein. [Google Scholar] [CrossRef]
  19. Shull, C.G.; Strauser, W.A.; Wollan, E.O. Neutron diffraction by paramagnetic and antiferromagnetic substances. Phys. Rev. 1951, 83, 333. [Google Scholar] [CrossRef]
  20. Tinkham, M. Paramagnetic resonance in dilute iron group fluorides. I. Fluorine hyperfine structure. Proc. R. Soc. 1956, A236, 535. [Google Scholar]
  21. Shulman, R.G.; Jaccarino, V. Effects of superexchange on the nuclear magnetic resonance of MnF2. Phys. Rev. 1956, 103, 1126. [Google Scholar] [CrossRef]
  22. Jaccarino, V.; Shulman, R.G. Observation of nuclear magnetic resonance in antiferromagnetic Mn(F19)2. Phys. Rev. 1957, 107, 1196. [Google Scholar] [CrossRef]
  23. Shulman, R.G.; Jaccarino, V. Nuclear magnetic resonance in paramagnetic MnF2. Phys. Rev. 1957, 108, 1219. [Google Scholar] [CrossRef]
  24. Kondo, J. Band theory of superexchange interaction. Prog. Theor. Phys. 1957, 18, 541. [Google Scholar] [CrossRef] [Green Version]
  25. Blount, E.I. Formalisms of band theory. Solid State Phys. 1962, 13, 305. [Google Scholar]
  26. Jahn, H.A.; Teller, E. Stability of polyatomic molecules in degenerate electronic states-I—Orbital degeneracy. Proc. Roy. Soc. 1937, A161, 220. [Google Scholar]
  27. Bates, C.A. Jahn-Teller effects in paramagnetic crystals. Phys. Rep. 1978, 35, 187. [Google Scholar] [CrossRef]
  28. Kahn, O. Magneto-Structural Correlations in Exchange Coupled Systems; Willett, R.D., Gatleschi, D., Kahn, O., Eds.; NATO ASI Series, Series C: Mathematical and Physical Sciences; Reidel Publishing Company: Dordrecht, The Netherlands, 1985; Volume 140, p. 37. [Google Scholar]
  29. Kahn, O. Molecular Magnetism; VCH Publishers: New York, NY, USA, 1993. [Google Scholar]
  30. Girerd, J.J.; Journaux, Y.; Kahn, O. Natural or orthogonalized magnetic orbitals: Two alternative ways to describe the exchange interaction. Chem. Phys. Lett. 1981, 82, 534. [Google Scholar] [CrossRef]
  31. Curély, J. Magnetic orbitals and mechanisms of exchange. I. Direct exchange. Special Issue. Mon. Für Chem. 2005, 136, 987. [Google Scholar]
  32. Van Vleck, J.H. Theory of the variations in paramagnetic anisotropy among different salts of the iron group. Phys. Rev. 1932, 41, 208. [Google Scholar] [CrossRef]
  33. Moriya, T. Magnetism; Rado, G.T., Suhl, H., Eds.; Wiley: New York, NY, USA, 1962; Volume I, p. 85. [Google Scholar]
  34. Moriya, T. Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 1960, 120, 91. [Google Scholar] [CrossRef]
  35. Dzialoshinskii, I.E. Thermodynamic theory of “weak” ferromagnetism in antiferromagnetic substance. Sov. Phys. JETP 1957, 6, 1259. [Google Scholar]
  36. Dzialoshinskii, I.E. A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics. J. Phys. Chem. Solids 1958, 4, 241. [Google Scholar] [CrossRef]
  37. Katsnelson, M.I.; Kvashnin, Y.O.; Mazurenko, V.V.; Lichtenstein, A.I. Correlated band theory of spin and orbital contributions to Dzialoshinskii-Moriya interactions. Phys. Rev. 2010, 82, 100403. [Google Scholar] [CrossRef] [Green Version]
  38. Qi, X.; Guionneau, P.; Lafon, E.; Pérot, S.; Kaufmann, B.; Mathonière, C. New photomagnetic ionic salts based on [MoIV(CN)8] 4− and [WIV(CN)8]4− anions. Magnetochemistry 2021, 7, 97. [Google Scholar] [CrossRef]
  39. Varshalovich, D.A.; Moskalev, A.N.; Kherkonskii, V.K. Quantum Theory of Angular Momentum; World Scientific: Singapore; Teaneck, NJ, USA; Hong Kong, China, 1988; p. 166. [Google Scholar]
  40. Hutchings, M.T. Point-charge calculations of energy levels of magnetic ions in crystalline electric fields. Solid State Phys. 1964, 16, 227. [Google Scholar]
  41. Edmonds, A.R. Angular Momentum in Quantum Mechanics; Princeton University Press, Ed.; Princeton University Press: Princeton, NJ, USA, 1957. [Google Scholar]
Figure 1. Values of spin S and orbital L momenta of an isolated transition ion, characterized by electronic shells 3dm, 4dm or 5dm (1 ≤ m ≤ 10).
Figure 1. Values of spin S and orbital L momenta of an isolated transition ion, characterized by electronic shells 3dm, 4dm or 5dm (1 ≤ m ≤ 10).
Magnetochemistry 08 00006 g001
Figure 2. “Ground” and “excited” configurations in the original superexchange process for the sequence A–X–A; the electronic configuration of the valence shell has been added for each ion (in our case, A = Mn, X = O, for instance); due to the weak overlap, coefficients a and b are small. In the present case, the electronic transfer is from X to A but it can also occur from X to B, indifferently (here, B = A).
Figure 2. “Ground” and “excited” configurations in the original superexchange process for the sequence A–X–A; the electronic configuration of the valence shell has been added for each ion (in our case, A = Mn, X = O, for instance); due to the weak overlap, coefficients a and b are small. In the present case, the electronic transfer is from X to A but it can also occur from X to B, indifferently (here, B = A).
Magnetochemistry 08 00006 g002
Figure 3. Radial behavior of energy for the centrosymmetric entity AXB in the general case of different magnetic sites (A ≠ B) on both sides of ligand X.
Figure 3. Radial behavior of energy for the centrosymmetric entity AXB in the general case of different magnetic sites (A ≠ B) on both sides of ligand X.
Magnetochemistry 08 00006 g003
Figure 4. Energy level scheme for the AXB centrosymmetric system (with here A = B for the sake of simplicity) in the case of dominant coulombic interactions; the difference | E Ag E Au| has been artificially zoomed for clarity. J is given by Equations (39) and (43).
Figure 4. Energy level scheme for the AXB centrosymmetric system (with here A = B for the sake of simplicity) in the case of dominant coulombic interactions; the difference | E Ag E Au| has been artificially zoomed for clarity. J is given by Equations (39) and (43).
Magnetochemistry 08 00006 g004
Figure 5. Energy level scheme for the AXB centrosymmetric system (with, here, A = B for the sake of simplicity) when Coulomb interactions are negligible (molecular orbital model). J is given by Equations (39) and (44).
Figure 5. Energy level scheme for the AXB centrosymmetric system (with, here, A = B for the sake of simplicity) when Coulomb interactions are negligible (molecular orbital model). J is given by Equations (39) and (44).
Magnetochemistry 08 00006 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Curély, J. The Microscopic Mechanisms Involved in Superexchange. Magnetochemistry 2022, 8, 6. https://doi.org/10.3390/magnetochemistry8010006

AMA Style

Curély J. The Microscopic Mechanisms Involved in Superexchange. Magnetochemistry. 2022; 8(1):6. https://doi.org/10.3390/magnetochemistry8010006

Chicago/Turabian Style

Curély, Jacques. 2022. "The Microscopic Mechanisms Involved in Superexchange" Magnetochemistry 8, no. 1: 6. https://doi.org/10.3390/magnetochemistry8010006

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop