Open Access Article
Robert J. Deeth
Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK. E-mail: r.j.deeth@warwick.ac.uk
First published on 28th November 2025
Ligand field theory (LFT) is generally formulated either as an application of the linear combination of atomic orbitals (LCAO) molecular orbital (MO) model (LFT-MO) or as freely-parameterised crystal field theory with the global crystal field replaced by the local cellular ligand field (CLF) formalism (LFT-CLF). LFT-MO and LFT-CLF are conceptually and numerically different. These differences are highlighted by the LFT-MO concept of an ‘inverted ligand field’ (ILF). Using formally low-spin d8 and d7 ML4 complexes, it is demonstrated that the LFT-MO ILF concept does not account for how the structures and reactivities of these systems change as a function of L or formal metal oxidation state. The LFT-MO overlap picture is an incomplete representation of how the sub-shell d electrons in transition metal complexes actually interact with their surroundings. The LFT-CLF picture of d electrons localised on the metal, but sensitive to the topology of the ligand field potential, VLF, is a better model. However, VLF does not invert. Instead, the ‘internal redox’ chemistry that the ILF concept attempts to rationalise is described via the LFT-CLF d-level breach. Conceptually, a d-level breach occurs when the bonding levels get too high or the d levels get too low. The empty d levels are filled and the integrity of the original dn configuration is compromised. A d-level breach should be abrupt with a significant impact on the geometric and electronic structure. This behaviour is confirmed computationally. The d-level breach is thus a significant descriptor for predicting enhanced ligand electrophilicity while the absence of a breach unambiguously and definitively confirms the dn configuration and metal oxidation state. In contrast, the %d components of canonical LCAO-type MOs used to invoke an ILF are unreliable descriptors and cannot be used to assign oxidation states. In general, ILFs have little chemical relevance but they are important here since they highlight several conceptual and numerical deficiencies of the theory which has underpinned the LFT-MO picture of TM systems for over 60 years.
Of course, in a numerical sense, LFT could never have been considered an application of MO theory. LFT has always focussed on the many-electron, multiplet states associated with d–d spectra and paramagnetism and, in 1957, the software required for an ‘ab initio’ MO-based calculation of multiplet states was still decades away. The only way of tackling this complicated quantum-mechanical problem numerically was via the ligand field Hamiltonian shown in eqn (1):
![]() | (1) |
The confusion concerning the nature of LFT is conceptual. LFT is simply freely-parametrised crystal field theory (CFT)3 and was enormously successful at rationalising the spectral and magnetic properties of coordination complexes. However, its global parameters failed to describe the local σ and π bonding interactions of interest to chemists. For example, in Oh symmetry, there is only one degree of freedom in the eg–t2g splitting of the d orbitals and hence only one parameter, Δoct, can be fitted to experiment (Fig. 1). Independent measures of σ and π interactions cannot be extracted. Moreover, the expected correlation between the ligand charge and the magnitude of Δoct was at odds with the values determined from experimental data. The crystal field model did not explain the spectrochemical series since highly electronegative ligands like F− generated smaller splittings than neutral ligands like carbon monoxide.
![]() | ||
| Fig. 1 Octahedral d-orbital splitting diagram. For CFT, the point charges are ze, 〈r〉 is the averaged d orbital radius and a is the distance between the origin and the point charges. | ||
In contrast, Mulliken's linear combination of atomic orbitals (LCAO) MO model separated the σ-type eg MOs from the π-type t2g MOs. This provided at least a qualitative rationalisation for the spectrochemical series in terms of π-donor ligands pushing the t2g ‘d-type’ MOs up and making Δoct smaller versus π-acceptor ligands which push the ‘d-type’ t2g MOs down and increase Δoct (Fig. 2, left).
![]() | ||
| Fig. 2 How local π-donor and π-acceptor bonding effects on the octahedral d-orbital splitting are used to rationalise the spectrochemical series. LFT-MO model (left) and LFT-CLF model (right). | ||
The cellular ligand field (CLF) model, introduced by Deeth and Gerloch in 19864 and formerly referred to as ligand field version of the angular overlap model,3,5,6 resolved the same issue within the framework of
by reformulating VLF in terms of local, σ- and π-bonding parameters. Thus, by design, the CLF model encapsulates the concept of localised M–L bonding and the spectrochemical series is rationalised in a similar manner to LFT-MO: π donors have positive eπ parameters which tends to reduce Δoct while π acceptors have negative eπ values which tends to increase the d-orbital splitting (Fig. 2, right).
Both models thus account for the trends observed in the spectrochemical series but, crucially, their conceptual and physical bases are completely different. The MO model has explicit bonding with ligand-based orbitals interacting with metal-based orbitals. MO formation depends on their relative phases and overlap. The CLF model has implicit bonding where the ligand-field d electrons are presumed to be localised on the metal centre but are able to respond to the topology of the potential field arising from the bonding electron density. This potential, which has both non-d metal and ligand components, has shape but no phase.
Consequently, in cases where orbital phase is not an issue, such as the conceptual rationale for the spectrochemical series described above, there is no conflict between LFT-MO and LFT-CLF and they present a similar conceptual picture. However, when LFT-MO relies on orbital phase arguments, there is invariably a conceptual clash with LFT-CLF.7–13 ‘Phase-coupled ligators’ is one such case where an MO-based mechanism generates inconsistencies which manifest in overlap-based approaches such as Schäffer and Jørgensen's angular overlap model (AOM) but are absent in the CLF picture.10 The so-called inverted ligand field (ILF) is another such case.
The ILF is an LCAO MO construct involving changes in MO compositions rather than the inversion of the explicit ligand field of the LFT-CLF model which would require a change in field polarity. To highlight this distinction, ILFs will henceforth be referred to as inverted MO sequences (IMOSs).
The composition of MOs depends on the relative energies of the contributing fragment orbitals (Fig. 3, left). In TM systems, if the ligand fragment orbitals lie below the isolated-metal d orbitals, the anti-bonding σ-type MO is mostly metal-based and highest: the ‘ligand field’ is deemed ‘normal’. However, as the energies of the ligand orbitals rise relative to the metal d, the MO compositions change and, ultimately, the bonding MO is mostly metal-based. It is now the lowest energy ‘d’ orbital and the ligand field is deemed to have ‘inverted’. The electrons which started in ligand-based MOs now find themselves in metal-based MOs and the metal is formally reduced.
However, in the LFT-CLF model, the progressive movement of electron density from the ligand to the metal simply leads to a larger CLF parameter value and an increase in the affected d orbital energy (Fig. 3, right). The d-orbital sequence is unaltered and the VLF does not ‘invert’. In addition, if a putative inversion were to occur, the relevant d orbital would have a negative energy (relative to the mean d energy,
, vide infra) and thus the CLF parameter would change sign and the ligand would be deemed an acceptor. Conceptually, achieving this ‘switch’ from ligand-donor to ligand-acceptor behaviour would be abrupt plus this ‘internal electron transfer’ would remove electrons from the bonding levels and place them in the metal d orbitals. The ‘holes’ in the bonding levels would generate Jahn–Teller or pseudo-Jahn–Teller electronic instability and a significant effect on the molecular structure can be anticipated. This contrasts with the LFT-MO picture where it is argued that the change from normal to inverted is smooth with no apparent effect on structure.14
The LFT-MO and LFT-CLF conceptual pictures are thus qualitatively different and hence their relative performance can be tested numerically. Using a series of formally low-spin d7 and d8 complexes, including the archetypal IMOS example [Cu(CF3)4]−, this paper demonstrates how the IMOS picture, which is based on orbital overlap, is unable to account fully for how the structures and reactivities of these systems change as a function of L or formal metal oxidation state. In contrast, a different conceptual mechanism derived from the potential-based LFT-CLF picture – the d-level breach – is much more successful.
The ab initio ligand field theory (aiLFT)17 standard protocol is based on a minimal n electron, 5 orbital (n, 5) active space and employed the DFT-optimised coordinates which, where appropriate, were slightly idealised to the nearest point-group symmetry to aid assignment of the active-space d orbitals. The def2-SVP basis sets were used for all atoms except the metal which retained the def2-TZVP basis.
Briefly, aiLFT is a correlated wavefunction method based on a complete active space self consistent field (CASSCF)18 treatment of static correlation augmented by an n-electron valence state perturbation (NEVPT2)19–22 treatment of dynamic correlation. For a given dn configuration, use of a minimal active space (ms) comprising 5 orbitals and n electrons generates a one-to-one mapping between the CASSCF/NEVPT2 energy levels and the relevant Russell–Saunders ligand field multiplets. The CASSCF/NEVPT2 energy differences become a surrogate for the experimental d–d transitions and can be fitted using the conventional ligand-field Hamiltonian (eqn (1)) to generate ms aiLFT d-orbital energies and interelectron repulsion parameters in a way akin to, but far more refined than, how Tanabe–Sugano diagrams are used. However, aiLFT has limitations. In general, getting enough, and the correct balance between, static and dynamic correlation is challenging and the minimal active space of aiLFT is relatively small. The analyses in this paper thus rely more on qualitative features than quantitative accuracy.
In contrast, Snyder's32 ab initio Hartree–Fock and MP2 calculations on [Cu(CF3)4]− yielded Cu 3d populations much closer to 10 than 8. He therefore reformulated the complex as “a normal cuprate [CF3–Cu–CF3]− in which the copper is further coordinated to a trifluoromethyl anion and cation” (Fig. 5).
![]() | ||
| Fig. 5 Formal structural diagrams for [Cu(CF3)4]−. Left: classical Werner-type structure involving four dative Cu–C bonds from anionic CF3− ligands; Right: snyder's formulation32 comprising a linear d10 [CuI(CF3)2]− fragment plus nucleophilic CF3− and electrophilic CF3+ groups. | ||
IMOSs and their implications for the assignment of formal metal dn numbers and OS became a popular, if somewhat contentious, issue.13,14,23,24,26,31,34,35 As described in the seminal review by Hoffmann et al.,31 the IMOS is a widespread feature of LCAO-type methods including KS DFT. It has also enjoyed strong support from techniques such as X-ray absorption spectroscopy (XAS) to the extent that the ‘experimental’ Cu 3d percentage in the acceptor orbital of nominally CuIII species, including [Cu(CF3)4]−, is so low that the d8 CuIII formulation has been suggested to be largely mythical35 although this assignment for [Cu(CF3)4]− is contradicted by Geoghegan et al.27 who favour the d8 CuIII formulation based on a combination of XAS and valence-to-core X-ray emission spectroscopy (VtC XES) data.
However, while interpretations of spectroscopic data may vary, a significant feature of the conceptual LFT-MO picture of 16-electron planar complexes is that an IMOS is not associated with any effect on the structure. Thus, according to the AOM36 analysis presented by Walroth et al.14 (Fig. 6), a stable, square-planar structure is suggested irrespective of the d count (Fig. 6). The MO diagrams of Fig. 4 support this view. The ‘normal’ and ‘inverted’ forms are unremarkable. Thus, LFT-MO predicts that all low-spin 16-electron species should be planar irrespective of whether they also have an IMOS or not.
![]() | ||
| Fig. 6 AOM analysis for σ-only ML4 systems with ‘normal’ (top) and inverted (bottom) d-orbital sequences (inverted part after Fig. 3 of ref. 14). Ligand levels in red and d levels in blue. MOSE = molecular orbital stabilisation energy. | ||
This assertion can be tested numerically. The need to compute reliable structures precludes the strategy adopted by Hoffmann et al.31 who sought to develop model IMOS systems mainly by manipulating the ligand EHMO Hii parameters. Instead, explicit-electron, KS DFT optimisations were carried out for the series [CuL4]− where the ligand is varied L = F, OH, NH2, CH3, BH2, BeH and Li to mimic the change from the extreme nucleophilicity of F− to the extreme electrophilicity of Li−. KS DFT is exceptionally consistent at generating accurate structures of metal complexes and many functionals could be used including BP8637,38 providing the basis sets are at least double-ζ quality and that an appropriate dispersion correction and a solvation field are included.
Since angle-bending potentials are relatively low,39 metal complexes are quite ‘plastic’ and the optimised structures are sensitive to subtle electronic effects. The BP86 results are shown in Fig. 7.
There is an abrupt change of structure at L = BeH with distorted structures for [Cu(BeH)4]− and [CuLi4]−. All the other systems, including [Cu(CH3)4]− which, like its CF3 analogue, has an IMOS, have square planar ground states consistent with low-spin d8 CuIII. Taken at face value, Fig. 7 demonstrates that the qualitative expectations of LFT-MO are not met although sceptics may point to the small energy differences between the optimised local minimum and the planar excited state or query whether the manner of the distortion, or at what point it occurs, is method dependent. To test the latter proposition, B3LYP and TPSSh calculations were carried out. As described in the SI, the hybrid functional results agree with the BP86-based protocol in terms of where the abrupt change occurs. However, the pseudo-Jahn–Teller nature of breached systems can present multiple, finely-balanced distortion pathways and the end results may differ. Importantly, though, the issue here is whether an abrupt change in structure happens at all. For the moment, therefore, let us proceed on the basis that the geometry change is a real effect and explore how it manifests in the LFT-CLF model and what the further consequences for LFT-MO might be.
The most important effects of the ligand-field potential come from the spatially and energetically nearest sources – i.e. the electron density localised in the M–L bonds. Conceptually, if the potential from the bonding electrons is below that of the lowest, empty d orbital, the integrity of the dn configuration is maintained. However, if the bonding level rises above one or more empty d orbitals, or empty d orbitals descend below the bonding level, those d levels are ‘breached’, filled, and the dn configuration is compromised. Fig. 8 presents a simplified representation of the first ‘d-level breach’ scenario for a low-spin d8 ML4 system, where relatively nucleophilic L is replaced by relatively electropositive L′.
As ligand electropositivity rises, the bonding potential rises, the field strength increases and the d-orbital splitting increases. At some stage, at the prevailing geometry, the bonding potential overwhelms the d orbitals. The left-most panel of Fig. 8 represents the situation just prior to a d-level breach: the bonding potential in ML4 is high but still lower than that of the empty d level and the d8 configuration retains its integrity. Replacing L with L′ raises the bonding potential above the empty d level generating an unstable electronic state (Fig. 8, panel 2). Electrons ‘jump’ from the bonding levels to fill the lower-lying empty d orbital thus ‘breaching’ the d8 configuration and generating a d10 metal centre plus two ‘holes’ in the bonding levels. Since the latter are degenerate in a square planar environment – they are akin to but different from Pauling hybrids – a pseudo Jahn–Teller effect (pJTE) results and the geometry is unstable (Fig. 8, panel 3). Hence, the complex distorts to re-establish a stable ground state structure with a formally reduced metal centre.
The second scenario where metal oxidation results in the empty d level falling below the bonding potential is basically the same mechanism.
The proposed d-level breach concept is an abrupt internal redox process – either the bonding potential is low enough for the integrity of the original dn configuration to be maintained or it is so high that the dn configuration is compromised. In the latter case, punching holes in the bonding potential can be expected to have structural consequences. The geometries in Fig. 7 suggest that only [Cu(BeH)4]− and [CuLi4]− are breached. Since the IMOS criterion of less than 50% 3d character in the acceptor σ-type MO is met already at [Cu(CH3)4]−, the d-level breach structural threshold is clearly higher than the IMOS requirement. This seems reasonable since the empty d orbital which we are trying to breach with a rising bonding potential is being progressively elevated by that approaching potential. This either/or feature of the dn configuration integrity also echoes the ‘winner takes all’ oxidation state localised orbitals method of Gimferrer et al.28
Beyond the breach threshold, the dn configuration will be compromised and no longer applicable. Moreover, if the ‘new’ configuration is not an open d shell, the CLF picture may still be valid but this cannot be tested using a model based on
. The extension of the conceptual LFT-CLF model to breached systems, as well as formally d0 and d10 systems in general, is beyond the scope of the present manuscript. However, for those d1 through d9 systems which fall within the ligand field regime, a successful ms aiLFT17,41–43 calculation provides a ‘first principles’ test of the integrity of the dn configuration. In the LFT-CLF picture, therefore, the dn configuration is much more than the ‘useful heuristic’ suggested by Norman and Pringle.44 We will return to this issue in a future publication.
Meanwhile, for the systems shown in Fig. 7, the d8 aiLFT calculations converge for the square planar systems [CuF4]−, [Cu(OH)4]−, [Cu(NH2)4]− and [Cu(CH3)4]− and fail for the distorted systems [Cu(BeH)4]− and [CuLi4]− (see SI for further details). Interestingly, the model IMOS system, [Rh(AlMe)4]+, (Fig. 9) which, according to Hoffmann et al.,31 has an electronic structure consistent with a formally d10 Rh−1 system, also has a geometrical structure similar to [Cu(BeH)4]− consistent with both systems having d-level breaches. This conclusion is further supported by the failure of the d8 ms aiLFT CASSCF step to converge on five suitable active space d orbitals for [Rh(AlMe)4]+ (see SI).
![]() | ||
| Fig. 9 Model IMOS system proposed by Hoffmann et al.31 The distorted structure and the failure of the CASSCF step in the ms aiLFT calculation confirm this system also has a d-level breach (see SI.). | ||
The CASSCF step for [Cu(BH2)4]− fails too with the CASSCF converging onto three d, one s and one p instead of five mostly d orbitals (see SI), but, given the square-planar structure predicted by DFT, the aiLFT failure is attributed to the shortcomings of the ms aiLFT active space rather than any change in the d8 CuIII LFT-CLF picture. Moreover, the central issue is here not the precise point at which the geometry changes abruptly but the fact that there is an abrupt change at all. Meanwhile, the aiLFT developers are addressing the limitations of the minimal active space45 but a molecular version of the code is not yet available. For the moment, therefore, the presence or absence of a d-level breach is indicated if both the dn integrity, as assessed by ms aiLFT, and the DFT structure are in accord. If not, then the structural evidence may be more significant.
Optimisation of Cu(CF3)4 gives a geometry which is distorted from the approximately planar first-order saddle point of D2d symmetry to a more see-saw structure (Fig. 10, top) of C2v symmetry while the d7 ms aiLFT calculation does not generate a suitable set of active space d orbitals and the aiLFT transformation fails. However, in this case, the structure is not definitive support for a d-level breach since a distortion would also result from removing an electron from the degenerate (in D2d) dxz/dyz orbitals thus generating a Jahn–Teller-active 2E state. This occurs for the isoelectronic NiIII analogue where both D2d and see-saw C2v are local minima on the BP86 energy surface (Fig. 10, middle and bottom and SI) with the latter lower in energy by about 10 kcal mol−1. The d8 NiII analogue has the same structure as the CuIII analogue. Significantly, the d7 NiIII and d8 NiII ms aiLFT calculations succeed (see SI) which confirms that the different behaviours of Cu(CF3)4 versus [Cu(CF3)4]− are due to the change in oxidation state and therefore that the formally CuIV system has a d-level breach. This provides an ideal opportunity to compare the IMOS and d-level breach concepts since all four complexes have IMOSs but only Cu(CF3)4 has a d-level breach.
![]() | ||
| Fig. 10 DFT-optimised local minima for Cu(CF3)4 and its isoelectronic nickel analogue [Ni(CF3)4]−. Cartesian coordinates and energies are given in the SI. | ||
![]() | ||
| Fig. 11 Reaction schemes for elimination of C2F6 (top) and nucleophilic attach by F− (bottom) for [Cu(CF3)4]p systems, p = −1, 0 based on Snyder's formulation. | ||
The results for the four complexes reacting with F− are shown in Fig. 12.
Clearly, the most telling feature is that a low barrier is only calculated for the system which has a d-level breach. Once again, the presence of an IMOS is unimportant while the presence of a d-level breach is highly significant.
On the one hand, the notion that d8 CuIII does not exist is contradicted both by the results reported above as well as the combined XAS and VtC XES study of Geoghegan et al.27 These workers report that the experimental transition energies for the copper N-heterocyclic complexes shown in Fig. 15 are ‘distinctly different’ for each formal oxidation state. The similarities between [CuIII(NHC4)]3+ and [Cu(CF3)4]− lead them to formulate the latter also as a d8 CuIII complex. In agreement with this proposal, aiLFT calculations for [CuII(NHC2)(NCMe)]2+, [CuIII(NHC2)(NCMe)]3+ and [CuIII(NHC4)]3+ support d9, d8 and d8 configurations respectively (see SI).
![]() | ||
| Fig. 15 Structural diagrams for the N-heterocyclic carbene complexes reported in ref. 27. Details of the optimised geometries and aiLFT calculations are in the SI. | ||
On the other hand, the notion that spectroscopy can somehow ‘prove’ an IMOS needs clarification. As described above, the ‘inversion’ actually refers to a change in MO composition. An IMOS is therefore a feature of the (fictitious) orbitals of an LCAO MO model, such as the EHMO and Hartree–Fock models or KS DFT. These methods solve a Roothaan-style equation eqn (2), where H is the relevant Hamiltonian matrix, S the overlap matrix, C the MO coefficients and E the MO energies. Thus, if one makes the (subjective) choice to use this type of model to probe the electronic structures of formally CuIII species, then a strong correlation between the %Cu 3d component of the acceptor orbital and the area of the L2 + L3 XAS absorption is expected. The %Cu 3d contribution for formally CuIII species is generally calculated to be less than 50% so this class of complex satisfies the IMOS condition. However, to then claim that the %Cu 3d contribution to the LUMO of [Cu(CF3)4]− is 35% ‘experimentally’, as made in Fig. 6 of ref. 35 is invalid. Orbitals are model-dependent and are not observable. Furthermore, the physical significance of the parameters of the LCAO approach – i.e. the MO coefficients – is strictly limited by the nature of the model.
| (H − ES)C = 0 | (2) |
The goal of the LCAO MO model exemplified in eqn (2) is to generate as low a total energy as possible and the coefficients are optimised accordingly. Hence, the only valid conclusion from the previous work by DiMucci et al.35 is that the Löwdin %Cu 3d LUMO contribution is associated with the lowest energy solution of their chosen KS DFT methodology (i.e. B3LYP functional with a CP(PPP) basis set on Cu and ZORA-def2-TZVP(-f) bases on all other atoms). Similar results can be anticipated for other functional/basis set combinations. However, since neither the dn number nor the metal oxidation state is properly defined in LCAO MO theory in general, only the trend is important and the specific %Cu 3d values themselves have no additional significance and are unreliable descriptors for assigning oxidation state. The presence (or absence) of an IMOS has no bearing on whether CuIII exists or not.
Of course, comparable constraints apply to the LFT-CLF model. One makes the (subjective) choice to use the LFT-CLF/aiLFT model, and then optimises the CLF parameters/CASSCF coefficients to generate energies of the (fictitious) ligand-field d states which provide the best fit to experimental (or calculated in the case of an aiLFT analysis) d–d spectra and/or magnetic susceptibilities. The crucial difference is that CLF/aiLFT calculations require the dn configuration to be defined at the start. An acceptable fit/viable active space validates that choice and its associated formal OS. Hence, it is legitimate within the LFT-CLF model to conclude that the successful ms aiLFT calculation for [Cu(CF3)4]− is consistent with its formulation as a low-spin d8 CuIII complex.
Likewise, the successful aiLFT transformations for [Cu(CH3)4]− and the complexes displayed in Fig. 16, which includes [Cu(CF3)4]−, means they can all be formulated as d8 species with ‘normal’ d-orbital sequences. Significantly, the same trend is found for the %d of the highest d orbital from ms aiLFT as from the B3LYP LUMO so both LFT and LCAO MO theory generate the same rank order of changing covalency which is encouraging. However, the components from aiLFT are all greater than 50% so, formally, the criterion for inversion is not even met. From the LFT-CLF perspective, d8 CuIII is not mythical.35
![]() | ||
| Fig. 16 Correlation between calculated d component (%) in the highest d orbital of the aiLFT active space and the B3LYP acceptor MO for a selection of formally d8 CuIII complexes. B3LYP data taken from ref. 35 Optimised structures and aiLFT data are given in the SI. | ||
Finally, and for clarity, the absence of a valid link between the metals’ formal oxidation state and the %3d component of a canonical acceptor MO does not imply that the KS DFT electron density distribution is somehow in error. It is simply a question of how this density has been partitioned.
What seems clear is that the LFT-MO and LFT-CLF models have fundamentally different physical and theoretical foundations. Consequently, the conceptual pictures of how the sub-valence-shell d electrons interact with their surroundings are also very different. The LFT-MO picture emphasises explicit orbital mixing and the d electrons/orbitals are often strongly delocalised onto the ligands. In contrast, the d electrons in the LFT-CLF picture are localised on the metal and the bonding is treated implicitly.
In the context of so-called inverted ligand fields, or IMOSs as they have been renamed here, each model's predictions of how the structures and reactivities of formally low-spin d7 and d8 square-planar ML4 complexes should vary as a function of L and/or metal oxidation state are qualitatively different.
In the LFT-MO picture for d8 complexes, the progression from a ‘normal’ d sequence to an ‘inverted’ sequence is smooth and progressive. At the normal MO sequence extreme, the orbitals on L are well below the d levels, ligand–metal donation is weak, dx2−y2 is the highest d level and empty, and the metal is formally d8. At the inverted sequence extreme, the orbitals on L are well above the d levels, ligand–metal donation is strong, dx2−y2 is the lowest d level and filled, and the metal has been formally reduced to d10. Both normal and inverted systems are predicted to have the same geometric structure but the ligands of the inverted form are expected to have enhanced electrophilic character.
In the LFT-CLF conceptual picture, increasing the ligand-to-metal donation simply increases the d-orbital splitting: VLF does not invert. The rising bonding-level potential initially pushes the dx2−y2 orbital up but the integrity of the d8 configuration is maintained. Eventually, however, the bonding potential ‘swamps’ the empty dx2−y2 orbital causing a d-level breach which formally reduces the metal. The breach is abrupt and generates a (pseudo) Jahn–Teller instability in the bonding levels which leads to a geometric distortion. A breached system is also expected to display enhanced electrophilic reactivity of the ligands.
The combination of DFT geometry optimisations and ms aiLFT dn integrity tests for low-spin d7 and d8 ML4 complexes presented here demonstrates that the LFT-MO predictions are not met. For example, there is an abrupt change in structure in the CuIIIL4 series while all four complexes, d8 [NiII(CF3)4]2−, d7 [NiIII(CF3)4]2−, d8 [CuIII(CF3)4]− and d7 CuIV(CF3)4, have IMOSs but only Cu(CF3)4 is breached. The first three complexes have high barriers to reaction with F− while that for Cu(CF3)4 is small. The inevitable conclusion is that an IMOS has little chemical relevance. Further theoretical analysis also argues that calculated %Cu 3d values from canonical DFT MOs have little physical significance and cannot be used to assign formal dn configurations or metal oxidation states since neither quantity is defined in the LCAO MO model. In contrast, the presence of an LFT-CLF d-level breach is highly significant for reactivity while its absence, in conjunction with an appropriate structure, allows for the definitive and unambiguous assignment of the dn configuration and its associated formal oxidation state.
The results presented here illustrate that the orbital-overlap physics which has underpinned the LFT-MO description of metal–ligand bonding for the past 60 years over-emphasises d-orbital delocalisation/covalency. In contrast, the potential-based CLF model with its localised d electrons is a more realistic representation of how ligand-field d electrons actually interact with their surroundings.
This CLF picture has important conceptual and numerical ramifications for the LFT-MO model which extend beyond the present topic of inverted MO sequences. As will be discussed in a future publication, many of the currently widely accepted concepts of the LFT-MO approach need revision. However, it is important to note that the LFT-CLF model cannot be a universal panacea. There are many situations, such as charge transfer spectroscopy or non-innocent ligands, where
simply does not apply but the MO model is useful. The trick will be to acknowledge the strengths and weaknesses of both the CLF and MO approaches and develop a combined conceptual picture which retains their best features whilst avoiding those which lead to conceptual phantoms such as inverted ligand fields.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5dt02371h.
| This journal is © The Royal Society of Chemistry 2026 |