C.
Sousa
a,
X.
Dong
b,
R.
Broer
c,
T. P.
Straatsma
bd and
C.
de Graaf
*ef
aDepartament de Ciència de Materials i Química Física and Institut de Química Teòrica i Computacional, Universitat de Barcelona, 08028 Barcelona, Spain
bNational Center for Computational Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37831-6373, USA
cZernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
dDepartment of Chemistry and Biochemistry, University of Alabama, Tuscaloosa, AL 35487-0336, USA
eDepartament de Química Física i Inorgànica, Universitat Rovira i Virgili, 43007 Tarragona, Spain. E-mail: coen.degraaf@urv.cat
fICREA, Pg. Lluís Companys 23, Barcelona, Spain
First published on 30th June 2025
The non-orthogonal configuration interaction with fragments (NOCI-F) approach is extended opening the possibility to study intramolecular processes and materials with covalent or ionic lattices. So far, NOCI-F has been applied to study intermolecular energy and electron transfer employing ensembles of fragments that do not have atoms or bonds in common. The here presented approach divides the target system into two overlapping fragments that share one or more atoms and/or one or more bonds. After the construction of a collection of (multiconfigurational) fragment wave functions in a state specific optimization procedure, the fragment wave functions are combined to form many-electron basis functions for the non-orthogonal configuration interaction of the whole system. The orbitals in the overlapping fragment are defined by a corresponding orbital transformation of the fragment orbitals through a singular value decomposition. The overlapping fragments approach is first illustrated for a model system and then used to highlight some possible applications of NOCI with overlapping fragments. The results of excited state diffusion in transition metal oxide, intramolecular singlet fission and magnetic interactions in organic biradicals and ionic compounds are discussed.
Based on Fermi's golden rule, many models are being used to quantify the rates of the different transfer processes.9 The electronic coupling between initial and final states can be estimated from Frontier molecular orbital models such as the one developed by Michl and co-workers,1,10 the dimer projection method,11 the analytic overlap method,12 and its extension to machine-learning.13 Next, these couplings can also be obtained from supramolecular total energy calculations, including unrestricted Hartree–Fock,14,15 multi-configurational wave functions approaches,16–19 ionization and electron affinity coupled cluster,20 time-dependent density functional theory (TD-DFT), quasi-particle energy DFT,21 and semi-empirical methods such as density matrix tight-binding.22 Furthermore, computational approaches have been extensively applied that rely on diabatization procedures among which the generalized Mulliken–Hush method23 is one of the reference methods. Mulliken–Hush is based on diabatization through diagonalization of the dipole moment operator, but other choices of operators can also be used.24
Diabatization schemes based on the locality of the orbitals find their origin in the proposal of Cederbaum and co-workers of block diagonalization of the Hamiltonian.25 This has been implemented in constrained DFT,26 fragment-orbital DFT,27,28 projector operator DFT29 and diabatization using absolute localized orbitals.30
Many of the above mentioned schemes aim at the calculation of the coupling between initial and final states in processes where electron or holes are transferred. Energy transfer can be captured in models such as the Frenkel–Davydov exciton model31,32 and its ab initio implementation by Herbert et al.33 Other widely used models are based on the so-called Dexter and Förster couplings34,35 and the transition dipole coupling model derived from it.36,37
Non-orthogonal configuration interaction with fragments (NOCI-F)38–40 is designed to study the electronic structure of systems containing discrete fragments found in molecular crystals or (self-assembled) stacks of molecules and can be used to calculate electronic couplings without making any ad hoc assumption about the character of the initial or final state. It can be applied to study both energy and electron or hole transport processes.
Starting with an ensemble of fragments, in most cases a collection of molecules, NOCI-F first optimizes a set of multiconfigurational wave functions for different electronic states of each member of the ensemble. Typically, these are complete active space self-consistent field wave functions, but they can be of any type. The state specific optimization of the fragment states ensures a full account of the orbital relaxation, which can have a large impact on the quality of the ensemble wave function when fragment states are considered whose dominant electronic configurations are different. Once all the fragment states are generated, NOCI-F constructs the many-electron basis functions (MEBFs) of the whole ensemble as anti-symmetrized spin-adapted product of the fragment wave functions. In this way, we obtain an optimal (diabatic) representation of the electronic states involved in the energy and/or charge transfer processes among the fragments in the ensemble.
This procedure is to some extent similar to the active space decomposition method of Parker,41,42 with the major difference that NOCI-F uses different orbitals for different configurations. This implies that the MEBFs and their orbitals are mutually non-orthogonal, which requires an extra computational effort for the calculation of the Hamiltonian. The upside is that the resulting wave function is very compact (the number of MEBFs is typically on the order of ten) and yet fully accounts for non-dynamic correlation and orbital relaxation. Dynamic correlation effects can be included through the procedure described in ref. 40. Other fragment based approaches to calculate interactions between monomers are the symmetry adapted perturbation theory43–45 and the decomposition analysis of Lin and Corminboeuf.46 Furthermore, we mention the recently published local active space state interaction singles (LASSIS) of Hermes and co-workers as an interesting approach to partition a system in fragments.47
In the last few years, NOCI-F is applied to study intermolecular energy and electron transfer processes such as exciton diffusion, electron and hole mobility, transformation of an excited singlet state into two singlet-coupled triplet states on neighboring molecules, etc.40,48–51 Here, we present an extension of the NOCI-F method to tackle materials that do not have a molecular crystal and to study of intramolecular processes. With this extension that is based on earlier single-configuration work of Stoyanova et al.,52 NOCI-F can be used to study processes such as intramolecular (excited state) charge transfer, magnetic coupling and excited state dispersion in transition metal compounds, or electron and hole mobility in polymeric materials. In short, the applicability window of the NOCI-F approach is vastly extended.
Once the two fragments of the system are properly defined, multiconfigurational wave functions are separately optimized for the relevant electronic states of each fragment. The list of non-orthogonal fragment states typically include the ground state of the fragment, local excited states, and in case charge transfer effects can be expected to play a role in the system under study, one also includes cationic and anionic fragment states in the procedure. The so obtained fragment wave functions include the most important static electron correlation and a full orbital relaxation, that is, the orbitals in which the electronic states are expressed are optimally adapted to describe the respective states. Furthermore, the effect of dynamic correlation can be included by estimating the energy correction through multiconfigurational second-order perturbation theory approaches such as CASPT2 or NEVPT2. These correlation energy corrections on the fragment states are used later on in the procedure by shifting the diagonal matrix elements of the Hamiltonian of the full system as described in ref. 40.
The fragment states cannot be merged into MEBFs in the same way as is done in NOCI-F studies with ensembles of molecules that do not share atoms or bonds. In those cases, the MEBFs are directly constructed as anti-symmetrized spin-adapted products of the fragment wave functions. Now, additional steps need to be carried out to take into account the common part of the two fragments. To avoid the double counting of the electrons and nuclei in the overlapping part, not only electrons and nuclei need to be removed but also the set of orbitals of the whole system obtained by superimposing the orbitals of the two fragments must be reduced. However, the natural orbitals of the fragments have typically a rather delocalized character, and hence, it is impossible to unambiguously identify orbitals as belonging to the overlapping part of the system. This identification can efficiently be done by performing a corresponding orbital transformation of the orbitals of the two fragments. First, the overlap matrix (SABMO) of the occupied orbitals of the fragments A and B is constructed. In case of multiconfigurational fragment wave functions, the active orbitals are excluded to maintain the unitary character of the transformation. By performing a singular value decomposition (SVD) on the overlap matrix
| SABMO = UλVT | (1) |
The last step towards the final wave functions and energy is the actual non-orthogonal configuration interaction calculation on the whole system. The Hamiltonian and overlap matrix elements among the MEBFs constructed in the previous step are computed using the GNOME algorithm53 implemented in the massively parallel and GPU-accelerated GronOR code.39 Two thresholds have to be set, τMO and τCI. The first fixes the size of the common molecular orbital basis while the second controls the number of determinant pairs included in the construction of the NOCI matrix elements. The default values of 10−4 and 10−5 for τMO and τCI have been shown to give accurate results in previous applications of the NOCI-F approach.40,48,49 Additionally, computational time can be saved by freezing the core orbitals, thus reducing the NOCI computation. The matrix elements give access to the electronic coupling between the states represented by the MEBFs and solving the eigenvalue problem results in NOCI energies and wave functions. The dynamic correlation energy correction can be applied in this stage of the procedure by transforming the Hamiltonian to an orthogonal MEBF basis, applying the correlation correction to the diagonal matrix elements and transforming back to the original non-orthogonal MEBFs. The shift applied is the sum of the correlation energy corrections of the fragment wave functions that constitute the MEBF. This neglects interfragment dynamic correlation, which is assumed to be small based on the understanding that dynamic electron correlation is short-ranged. The wall clock time of preparing the MEBFs from the (overlapping) fragment wave functions is negligible in comparison to the actual NOCI-F calculation, as it only involves an SVD and the construction of the CI expansion for each MEBF. The calculation of the Hamiltonian and overlap matrices spanned by the MEBFs requires an SVD, solving an eigenvalue problem and processing the list of two-electron integrals for each bra-ket determinant pair.39,53
The ground state wave functions of acetic acid (fragment A) and alanine (fragment B) are optimized by the complete active space self-consistent filed (CASSCF) approach with an active space of two orbitals and two electrons for A, and 4 orbitals and 4 electrons for B. Fig. 3 and 4 contain the active orbitals and some of the inactive ones. Notice that the inactive orbitals are delocalized over the whole fragment and there is no way to isolate the electrons of the overlapping part. In this case the active orbitals are basically localized on the non-overlapping part of the fragments, only the second active orbital of acetic acid has a small amplitude on the hydrogen atoms of the CH2 bridge. This is however not a problem, the active orbitals can extend into the overlapping region.
![]() | ||
| Fig. 3 Some inactive and the CASSCF(2,2) natural orbitals of acetic acid. ei and ni are the orbital energies and the natural occupation numbers of the inactive and active orbitals, respectively. | ||
![]() | ||
| Fig. 4 Some inactive and the CASSCF(4,4) natural orbitals of alanine. ei and ni are the orbital energies and the natural occupation numbers of the inactive and active orbitals, respectively. | ||
The overlap matrix of the natural orbitals of the two fragments is calculated in the geometry of the full system with a dummy hydrogen placed at the average position of two hydrogens that were introduced upon splitting the molecule in two (grey atoms in Fig. 2). Note that this is only necessary when extra atoms are required for a correct description of the fragments. For example, in case of embedded cluster calculations of ionic materials (see right panel of Fig. 1), the overlapping fragments approach can be employed without extra atoms.
The corresponding orbital transformation is performed in two parts. First, the deep core orbitals (C,N,O-1s) are transformed to localize the C-1s orbital of the overlapping CH2 group, and subsequently, the corresponding orbitals are constructed for the remaining inactive orbitals. Using this two-step procedure, a well-defined set of core orbitals is maintained, which can be eliminated from the final NOCI calculation. The first two rows of Fig. 5 represent the five corresponding orbitals of acetic acid and alanine with the largest singular values. A sharp drop in the singular values is observed for the next pair of corresponding orbitals, λ6 = 0.039. These orbitals (and the rest) do not show any contribution of the CH2 group. As mentioned before, the in-phase combinations of the five corresponding orbital pairs are used to describe the CH2 group of the full system, as can be seen in the last row of Fig. 5. These orbitals do no longer count with the basis functions of the dummy hydrogen atom and have been renormalized. The out-of phase combinations are removed and all other transformed orbitals are added to the set of molecular orbitals of the aspartic acid molecule after removing the basis functions of the dummy hydrogen atom and subsequent renormalization. At this point, the overlapping fragments approach can perform a Löwdin orthogonalization of the inactive orbitals, but this is optional.
![]() | ||
| Fig. 5 From top to bottom: the five corresponding orbitals with the largest singular value λi of acetic acid, alanine and their in-phase combination in aspartic acid. | ||
Table 1 compares the energies of several electronic structure approaches for the aspartic acid molecule. Fully optimizing the CASSCF(6,6) wave function provides the lowest energy. Excluding the fragment-to-fragment excitations in a general active space (GAS) SCF approach does not significantly increase the energy. The CASCI(6,6) with the molecular orbitals generated with the overlapping fragments approach described above give as expected a higher total energy. OpenMolcas54 was used to calculate these energies. This program orthogonalizes the orbitals through the Gram-Schmidt approach, and hence, it is not surprising that the NOCI energy obtained with the Löwdin orthogonalized OVLF orbitals, NOCI(ortho), is practically the same. The only difference with the CASCI(6,6) is the fact that the NOCI active orbitals are not orthogonal to the inactive ones and that no fragment-to-fragment excitations are considered. Finally, the Table lists the NOCI energies obtained using the non-orthogonal orbitals. These lie approximately half-way the CASSCF and CASCI results. In the calculation with τCI = 10−6 only those bra-ket determinant pairs were considered in the evaluation of the NOCI energy whose absolute product of fragment CI coefficients is larger than 10−6. This reduces the number of determinant pairs by a factor of 5.5, but does not significantly affect the final energy, in line with earlier studies of the NOCI-F approach for non-overlapping fragments. Although a single total ground state energy cannot provide a definite answer, the non-orthogonal OVLF orbitals seem to give a reasonable set of MOs to describe the molecules constructed from two overlapping fragments.
| Method | Total energy |
|---|---|
| CASSCF(6,6) | −510.04063973 |
| GASSCF(4,4;2,2) | −510.04060509 |
| CASCI(6,6) | −509.86414831 |
| NOCI (ortho) | −509.86421967 |
| NOCI (τCI = 10−6) | −509.96299424 |
| NOCI (τCI = 0) | −509.96298675 |
ε(k) = ε0 − 2t cos(kx) | (2) |
NOCI-F and NOCI-F with overlapping fragments can be used as alternative making use of many-electron functions contrary to the tight-binding and DIPRO parametrizations. Here, we report the hopping of an excess electron between two neighboring FeO6 octahedra in CaFeO3 as shown in Fig. 6. The electronic structure of the ground state has been analyzed by some of us ref. 59 and 60, showing that the electronic configurations arising from oxygen to iron excitations are dominant in the many-electron wave function. This results in an effective 3d5 configurations for Fe, in line with the regular (albeit tilted) octahedra in the crystal.
![]() | ||
| Fig. 6 Ball and stick representation of two neighboring octahedra in the a–b plane of the CaFeO3 high temperature crystal phase.61,62 Red spheres are oxygen, blue is calcium and the central gold-like spheres are iron. | ||
Since the FeO6 octahedron is only slightly distorted, the electronic states in Table 2 are labeled by the symbols of the Oh symmetry group. The two components of the 5Eg state are separated in energy by no more than 0.04 eV. Adding an extra electron to the FeO6 octahedron lead to a sextet spin state of A1g symmetry. The two fragments used in this study are the left and right FeO6 octahedra of Fig. 6 with the central oxygen shared by both fragments.
| Hopping | Doublet | Quartet | Sextet | Octet | Decaplet |
|---|---|---|---|---|---|
| a5Eg–6A1g → 6A1g–a5Eg | 74.7 | 149.0 | 223.9 | 298.0 | 371.7 |
| a5Eg–6A1g → 6A1g–b5Eg | 124.8 | 250.2 | 375.8 | 498.4 | 621.5 |
| b5Eg–6A1g → 6A1g–a5Eg | 3.4 | 7.2 | 10.2 | 12.5 | 15.4 |
| b5Eg–6A1g → 6A1g–b5Eg | 16.3 | 33.4 | 50.0 | 64.8 | 80.8 |
With these three electronic states, there are four possible transfer pathways for the excess electron. The 6A1g state is either combined with the a5Eg or b5Eg state on the second fragment and electron transfer leads to one of the 5Eg states on the first fragment combined with the symmetric sextet state on the second fragment. The two fragment spin states can be coupled to doublet, quartet, sextet, octet and decaplet and the results in the table show that the coupling between the initial and final states strongly depends on this spin coupling. In fact, these estimates follow very closely the Anderson and Hasegawa model for double exchange63
![]() | (3) |
![]() | (4) |
Table 3 reports the relative energies of the fragments states. Apart from the first excited singlet (S1) and the lowest triplet state (T1), it also reports the energies of the cationic (D+) and the anionic (D−) doublet states. These doublets are combined to form the MEBFs representing the charge transfer (CT) states. It is clear from the relative energies that this covalently-linked naphthalene dimer is not a real candidate for showing singlet fission as the energy of the singlet coupled double triplet MEBF has a relative energy that is more than 1 eV higher than the singlet excited MEBF. Note that the energy of the T1T1 MEBFs is about 0.25 eV lower than two times the energy of the fragment T1 state, indicating a significant binding energy of the two triplet excitons.
| State | CASSCF | CASPT2 | Ref. weight | MEBF | NOCI-F |
|---|---|---|---|---|---|
| S0 | 0.00 | 0.00 | 0.620 | S0S0 | 0.00 |
| S1 | 5.64 | 4.63 | 0.613 | S1S0, S0S1 | 4.51 |
| T1 | 3.60 | 3.25 | 0.603 | T1T1 | 6.27 |
| D+ | 7.17 | 7.71 | 0.620 | D+D−, D−D+ | 5.42 |
| D− | 2.41 | 1.06 | 0.610 |
The actual NOCI calculation (i.e., solving the generalized eigenvalue problem defined by the NOCI Hamiltonian and overlap matrices) leads to six adiabatic states, whose energies and MEBF expansion coefficients are given in Table 4. The final NOCI wave functions show that the singlet exciton is delocalized over the two fragments with the in-phase combination slightly lower in energy than the out-of-phase combination. The coupling between the two local singlet excitons is 17 meV. The CT states show a similar delocalization over the two fragments, the coupling between the two CT MEBFs is 25 meV.
| NOCI-F wave functions | ||||||
|---|---|---|---|---|---|---|
| Ψ 1 | Ψ 2 | Ψ 3 | Ψ 4 | Ψ 5 | Ψ 6 | |
| ΔE | 0.00 | 4.48 | 4.52 | 5.40 | 5.46 | 6.27 |
| S0S0 | −1.000 | 0.000 | −0.011 | −0.001 | −0.003 | −0.002 |
| S1S0 | 0.016 | −0.733 | −0.680 | 0.005 | 0.002 | −0.002 |
| S0S1 | −0.016 | −0.680 | 0.733 | 0.000 | 0.005 | 0.002 |
| T1T1 | 0.001 | 0.000 | 0.003 | 0.002 | −0.010 | −1.000 |
| D+D− | −0.002 | 0.006 | 0.001 | 0.761 | 0.649 | −0.006 |
| D−D+ | 0.001 | −0.002 | 0.003 | 0.649 | −0.761 | 0.010 |
Finally, we mention that the coupling between the S1S0–S0S1 and the T1T1 MEBFs is 5.3 meV. This coupling includes the effect of the CT states, and corresponds to the so-called CT-enhanced coupling. Unlike many other cases, this coupling is almost identical to the direct coupling due to the very weak interaction of the singlet excitonic state and the double triplet state with the CT states, on the order of 5 meV. These interactions were calculated to be one or two orders of magnitude larger in previous studies on crystalline tetracene and pentacene, causing a significant enhancement of the singlet fission couplings when CT states were taken into account. We expect a similar effect in these norbornyl-bridged dimers for the larger acenes, which are currently under study.
The magnetic coupling strength between the two fragments, A and B, JAB, is estimated from the isotropic Heisenberg Hamiltonian
| Ĥ = −JABŜA·ŜB | (5) |
![]() | ||
| Fig. 8 Ball and stick representation of the simplest Ullman's biradical. Dark spheres are carbon atoms, blue is nitrogen, red is oxygen and light spheres represent hydrogen. | ||
In order to apply the NOCI-F method, the biradical molecule at the experimental geometry is divided in two nitronylnitroxide radical fragments with the dangling bond of the central C atom saturated by a H atom at a distance of 1.05 Å. This extra H atom will be removed when superimposing the two fragments. To ensure a correct description of the fragment wave function, a CAS(3,3) including 3π orbitals and 3 electrons has been considered. For each radical, six electronic states have been computed: the doublet ground state, D0, two anionic states in which the nitronylnitroxide radical wins one electron that can be coupled either to a singlet, S−, or a triplet state, T−, the two equivalent cationic states, with one electron less in the active space, S+ and T+, and a quartet state, Q0, corresponding to the maximum spin multiplicity of the radical. The energy of the singlet and triplet states have been computed by NOCI calculations in two ways. First, as a direct singlet or triplet coupling between the D0 states of each radical, constructing the corresponding D0D0 MEBF. Alternatively, the non-orthogonal Hamiltonian is increased by taking into account the charge transfer states between the two radicals. That leads to a 5 × 5 Hamiltonian including the D0D0, S−S+, S+S−, T−T+ and T+T− MEBFs for the singlet state and a 9 × 9 Hamiltonian based on the D0D0, S−T+, T+S−, T−S+, S+T−, T−T+, T+T−, Q0D0, D0Q0 MEBFs for the triplet state.
The values of the magnetic coupling, J, for the Ullman biradical, computed as the energy difference between the singlet and triplet states, are disclosed in Table 5. Various values of the common molecular orbital basis set threshold, τMO, have been considered, ranging from 10−5 to 10−7. Computing the singlet and triplet states from the direct D0D0 coupling (GS only in Table 5) leads to an antiferromagnetic coupling (negative J value) which slightly decreases when tightening the τMO threshold. Including the charge transfer states between the two fragments (see entry GS + CT in Table 5) increases the value of J by around 10–11 cm−1, meaning that for this particular system, charge transfer effects do not play an important role. The effect of the remaining dynamic electron correlation on the magnetic coupling has been considered by shifting the diagonal matrix elements by a dynamic correlation energy correction (DCEC).40 This correction has been computed by CASPT2 calculations, using the standard zeroth-order Hamiltonian (with an IPEA shift of 0.25 a.u.), on all the fragment states. Including the effect of the dynamical electron correlation only increases the J value by 6 cm−1, see +DCEC results in Table 5. This is in accordance with previous results by Angeli et al.,84 pointing out that for the magnetic coupling of the Ullman's biradical, dynamic electron correlation only plays a minor role.
| τ MO | 10−5 | 10−6 | 10−7 |
|---|---|---|---|
| GS only | −175 | −165 | −159 |
| GS + CT | −185 | −176 | −170 |
| + DCEC | −191 | −182 | −176 |
| Frozen core | |||
| GS only | −229 | −181 | −155 |
| GS + CT | −242 | −195 | −169 |
| + DCEC | −249 | −201 | −175 |
In order to discern the effect in the magnetic coupling of freezing the core orbitals, similar calculations have been performed now freezing the 1s core orbitals of all C, N and O atoms. Results are collected in the lower part of Table 5. It can be observed that the frozen core approach leads to larger absolute values of J, specially for a τMO of 10−5. Strengthening the τMO threshold decreases the J values in a way that both, frozen core and no frozen, practically coincide for a τMO threshold of 10−7. In the frozen core NOCI-F calculations, as found when considering all the orbitals, both the charge transfer and dynamical electron correlation effects are small. The magnitude of the magnetic coupling obtained by these calculations, −176 cm−1, is quite far from the experimentally measured value of −311 cm−1. Here it is important to note that the system is divided in two fragments by the C–C bond that connects the two radicals. In both fragments, this bond is modeled by a simpler C–H bond. This relatively simple approach to split the molecule into two fragments results in a crude a representation of the connection. Given that in the biradical systems like this one the magnetic coupling occurs mainly due to a through-bond coupling, it is not completely unexpected to recover only part of the coupling. Other possibilities of defining the overlapping fragments are being explored at present. A promising approach is to saturate each fragment with a CH3 group instead of a mere hydrogen atom. In this way the central C–C bond is much better described and magnetic couplings calculated through the fragments approach are likely to increase.
A second aspect that has been analyzed is the influence of the distance of the saturating H atom to the central C of the radical fragment. From the initial distance of 1.05 Å, a somewhat larger, 1.15 Å, and shorter, 0.95 Å, distances have been considered. Results reported in Table 6 show very small changes in the magnitude of the magnetic coupling, of 8 cm−1, when varying the distance by 0.2 Å. Again, the small effect of the inclusion of charge transfer and dynamic electron correlation effects for this particular system becomes evident.
| d(C−H) | 0.95 | 1.05 | 1.15 |
|---|---|---|---|
| GS only | −179 | −175 | −171 |
| GS + CT | −194 | −185 | −187 |
| + DCEC | −197 | −191 | −189 |
A second example of magnetic interactions in biradicals is given in Table 7. This concerns the interaction between the spin moments on the NO groups bridged by a phenyl group as shown in Fig. 9. The interaction leads to antiferromagnetic coupling when the second NO group is in the para position (left) and ferromagnetic coupling for the meta position (middle). The fragments are phenyl nitroxides (right), which are overlapping through the phenyl group. Previous computational studies on these biradicals estimated the couplings to be around −2000 cm−1 for the para compound and near +700 cm−1 for the molecule with the nitroxides in meta.,85 being the latter in good agreement with experimental estimates.86 To the best of our knowledge, there are no experimental estimates for the para molecule.
| Meta | Para | |
|---|---|---|
| GS only | 70 | −181 |
| GS + LE | 73 | −181 |
| GS + CT | 264 | −792 |
| All states | 269 | −969 |
The following fragment states were considered: the doublet ground and first excited states (D0 and D1), the singlet and triplet cationic states (S+ and T+) and the singlet and triplet anionic states (S− and T−). These states were combined in three types of MEBFs for the NOCI: first, the ground state (GS) of the phenylene dinitroxides (D0D0); second, the local excited states (LE, D1D0 and D0D1); and third, the charge transfer states (CT). For the singlet state of the whole molecule we consider S+S−, S−S+ T+T− and T−T+ and for the triplet state, we include the S+T−, S−T+, T+S−, T+T−, T−S+, and T−T+ CT states.
Only considering the neutral doublet states of the two fragments, the character of the coupling is correctly reproduced, although the strength of the coupling is significantly underestimated. Adding the local excited doublet states does not change the coupling, while including the charge transfer states in the NOCI calculation strongly increases the coupling between the two spin moments. The final estimates based on a NOCI among the GS, LE and CT states results in a moderate coupling of around −1000 cm−1 for para-perylene dinitroxide and 270 cm−1 for the meta variant, corresponding to an underestimation of the coupling by approximately a factor of two, similar to what is observed in the study of Ullman's biradical as commented before in this section.
Ab initio calculations based on a wave function approach and with density functional theory were both able to reproduce the experimental coupling with reasonable precision.89–91 NOCI calculations for the magnetic coupling in these materials have also been reported. The study of van Oosten et al. in a series of copper oxides revealed the importance of incorporating charge transfer states expressed in their own optimal set of orbitals in the description of the electronic structure.92,93 The separate optimization of charge transfer states is much more complicated in systems with larger magnetic moments. The larger number of unpaired electrons translates in the presence of several electronic states arising from transitions within the partially filled TM-3d shell intercalated between the magnetic ground state and the target charge transfer states. The procedure for obtaining a state specific description as described in ref. 94 and 95 could be applied but are extremely tedious and not very practical for routine application. Other algorithms such as root selection based on overlap are not always successful, often the character of the target state is lost during the optimization process. It is in this line that the NOCI-F with overlapping fragments may be useful to construct charge transfer states (and other relevant electronic states) and estimate the coupling through the NOCI approach applied previously for the cuprates.
Table 8 summarizes the NOCI results considering different collections of singlet, triplet and quintet states of an embedded Ni2O11 cluster. The wave functions of the Ni2O11 cluster are obtained by superimposing the wave functions of the NiO6 fragments, similar to the study on CaFeO3 of Section 3.2. The fragment states that were considered are the 3A2g ground state, the 2Eg states with one electron less and one electron more to construct the CT states and the lowest excited singlet state 1Eg. The latter, so-called non-Hund state, is not considered in the Heisenberg Hamiltonian and was demonstrated to be responsible for deviations in the regular energy splitting of the different spin states predicted by the Heisenberg Hamiltonian.96,97
| J ST | J TQ | |
|---|---|---|
| GS only | −0.2 | −0.2 |
| GS + NH | −0.5 | −0.1 |
| GS + NH + CT | −4.4 | −4.7 |
| + DCEC | −7.8 | −8.2 |
Only considering the ground state wave functions of the fragments leads to three virtually degenerate spin states. The addition of the non-Hund (NH) states does not change the picture drastically, only a small deviation to the regular splitting is observed. A more important increase in the coupling is observed when the CT states are included in the NOCI calculation and applying the dynamic correlation energy correction for the fragment states further increases the coupling. Unfortunately, the final result of ∼−8 meV only reproduces 50% of the experimental coupling, similar to what was observed for the Ullman biradical.
Among the several applications that can be envisaged, we have first studied the electron transfer in the ionic CaFeO3 lattice. The NOCI-F results closely follow the predictions of the Anderson–Hasegawa model63 and the expression for double exchange derived by Girerd.64 A clear advantage of the overlapping fragment approach is the fact that the fragments can be defined such that the site symmetry of the lattice is reproduced, which is generally not possible when the wave functions are directly optimized for a two-center cluster containing the metal centers involved in the electron transfer. The use of fragments with the correct site symmetry makes it possible to study effects such as orbital ordering, essential in many materials, for example the series of doped manganites La1−xCaxMnO3.100 The method is of course not restricted to the here studied electron transfer, it can also be used to calculate the diffusion of electronically excited states and the mobility of hole states, among other many-electron processes. By calculating the electronic coupling between initial and final states in different crystal directions, a many-electron band structure can be calculated as explained in ref. 51.
Whereas intermolecular singlet fission has been studied rather intensively, intramolecular singlet fission has important advantages. Choosing wisely the linker between the two parts of the molecule where the singlet exciton transforms into a singlet-coupled double triplet state can increase the singlet fission coupling significantly, or favor the triplet separation process. The here presented OVLF approach lies the foundations for the NOCI study of these processes, which is currently in progress. As proof-of-principle, we have studied the singlet fission in the norbornyl-bridged naphthalene dimer. The coupling in this covalently linked dimer is somewhat on the weak side, but is expected to be significantly larger in systems with larger acenes, used in experimental studies.
The underestimation of the magnetic coupling parameters observed for the organic biradicals and the transition metal oxide studied here can for a large part be ascribed to shortcomings in the description of the region that connects the two fragments. For these systems, the through-bond interactions are dominant compared to the through-space interactions, while the opposite holds for the molecular materials containing organic radicals studied in ref. 40. In the latter case, NOCI-F predicts very precisely the magnetic coupling parameters.
In Ullman's biradical, the central C–C bond originates from the corresponding orbital transformation mostly involving the two C–H orbitals of the fragments. The possibility of using a methyl group to saturate the broken bond is currently under investigation. For the other two systems, the situation is somewhat different. The two fragments do share the group that connects the two magnetic centers; in NiO the central oxygen of the Ni2O11 cluster, and the phenyl group in the phenylene dinitroxides. Their role in the magnetic coupling is however rather passive, the active orbitals of the fragments are basically localized on the magnetic center and extend very little onto the part of the systems that connects them. An improvement can possibly be obtained by defining a third fragment containing the connecting atoms. Fragment states should be constructed for this fragment with the orbitals that play a role in connecting the magnetic centers in the active space. These active orbitals, the π-orbitals of the phenylene group and the O-2pσ orbital for the Ni2O11 cluster, would then also be considered in the MEBF construction following the multi fragment approach101 implemented in GronOR.
Notice that the energy differences that play a role in the determination of the magnetic coupling strength are very small and that small deviations in the description of the coupling of the two fragments can have a substantial effect on the coupling. This effect is less dramatic for other applications of the OVLF approach.
A possible application of NOCI with overlapping fragments that has not been considered here is the intramolecular charge transfer, either thermally induced102 or after populating electronically excited states.103 In both cases, there are no major methodological problems to apply the OVLF approach and extract the electronic coupling of initial and final states of the electron transfer process.
The overlapping fragments approach is not restricted to systems that can be divided in two fragments, but can also applied to systems with more fragments. For example, in an ABCD supersystem one can first account for the interaction between subsystems A and B with C and D treated as spectators. Next, fragment B and C are coupled and finally C is coupled to D. The total NOCI matrices of the ABCD systems are then constructed from the Hamiltonian and overlap matrices of the three dimer fragments. This can of course be generalized to trimers (ABC with D as spectator and BCD with A as spectator) if second neighbour interactions are expected to be important.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5cp01700a |
| This journal is © the Owner Societies 2025 |