Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Ab initio non-covalent crystal field theory for lanthanide complexes: a multiconfigurational non-orthogonal group function approach

Alessandro Soncini *ab and Matteo Piccardo b
aDipartimento di Scienze Chimiche, Università degli Studi di Padova, Via Marzolo 1, 35131 Padova, Italy. E-mail: alessandro.soncini@unipd.it
bSchool of Chemistry, The University of Melbourne, Australia

Received 1st December 2021 , Accepted 5th July 2022

First published on 8th July 2022


Abstract

We present a multiconfigurational ab initio methodology based on non-orthogonal fragments for the calculation of crystal field energy levels and magnetic properties of lanthanide complexes, implementing a systematic description of non-covalent contributions to metal–ligand bonding. The approach consists of two steps. In the first step, appropriate ab initio wave functions for the various ionic fragments (lanthanide ions and coordinating ligands) are optimized separately, accounting for the influence of the surrounding environment within various approximations. In the second and final step, the scalar relativistic (DKH2) electrostatic Hamiltonian of the whole molecule is represented on the basis of the optimized metal–ligand multiconfigurational non-orthogonal group functions (MC-NOGFs), and reduced to an effective (2J + 1)-dimensional non-orthogonal configuration interaction (CI) problem via Löwdin-partitioning. Within the proposed formalism, the projected non-orthogonal CI Hamiltonian can be expanded to any desired order of perturbation theory in the fragment-localised excitations out of the degenerate space, and its eigenvalues and eigenfunctions provide systematic approximations to the crystal field energies and wave functions. We present here a preliminary implementation of the proposed MC-NOGF method developed for first-order degenerate perturbation theory within our own ab initio code CERES, and compare its performance both with the simpler non-covalent orthogonal ab initio approach, Fragment Ab Initio Model Potential (FAIMP) approximation, and the full CAHF/CASCI-SO method, accounting for metal–ligand covalency in a mean-field manner. We found that the energies and magnetic properties of 44 complexes obtained via an iteratively optimized version of our MC-NOGF first-order non-covalent method compare remarkably well with those obtained using the full CAHF/CASCI-SO method including metal–ligand covalency, thus exposing the predominantly electrostatic character of the metal–ligand interactions, and are superior to those obtained using the FAIMP approach, which in its iteratively optimised variant was believed to date to be the best ab initio description of non-covalent metal–ligand interactions.


Introduction

Trivalent lanthanide ions are extensively employed in many technological applications, ranging from energy production to life sciences.1 While they are certainly fundamental components in many optical applications,2–4 these ions play a very special role in magnetism, thanks to their large magnetic moments and magnetic anisotropy.5–7 At the root of their enhanced magnetic properties are the intrinsically strong electronic correlation and spin–orbit coupling interactions and comparatively weak interactions between the 4f electron shell and its coordination environment, mainly as a result of the 4f shell's radially compact nature, and its consequent efficient shielding from the surrounding electrostatic and chemical environment by the outer 5s25p6 orbital shells.5 The simplest and earliest approach to describe the splitting of a ground (2J + 1)-fold degenerate ground multiplet produced by the surrounding chemical environment is called crystal field theory (CFT).8 CFT describes the splitting of the 4f many-electron wave functions associated with the (2J + 1) ground spin–orbit multiplet of the pristine atomic 4f-shell as operated solely by the static electric fields induced by the charge distribution on the ligands. CFT entails a chemical bonding description which is purely ionic, and it is in fact referred to also as the ionic model.

While clearly CFT is an oversimplified description for quantitative purposes, which even at the phenomenological level is riddled by an over-parameterization problem (in low symmetry the CFT potential consists of 27 free parameters), its heuristic value should not be underestimated, even today within the field of molecular magnetism.

For instance, Rinehart and Long9, discussing the asphericities of the electron density of the magnetic states in trivalent lanthanide ions calculated by Sievers some twenty years earlier10, pointed out how a proper axial (equatorial) distribution of the charge density in the coordinating ligands could be harnessed to synthesise Ln complexes with large magnetic anisotropy, by stabilizing the appropriate oblate (prolate) charge density distribution in that specific ion. This simple but insightful observation, based on the assumption of purely ionic bonding interactions between the metal and the ligand, has proven rather useful in guiding synthetic chemists towards ever more efficient Ln-based single-molecule magnets (SMMs).

These simple electrostatics arguments were also successfully translated into a more quantitative classical electrostatics repulsion energy minimization procedure, which was shown to be capable of semi-quantitatively predicting the specific direction of the magnetic anisotropy axis in many Dy(III) single-molecule magnets,11 a finding that was recently confirmed via X-ray diffraction experiments.12

Despite these useful results, it is known since the 1950s that quantitative descriptions of lanthanide–ligand bonding cannot be achieved without accounting for non-covalent interactions other than pure electrostatics (e.g. induction, contact or dispersion interactions), and for various flavors of so-called covalent contributions (charge transfer excitations of various kinds).13–22

A way to systematically account for all possible metal–ligand electronic interactions consists of ab initio calculations, typically based on molecular orbital (MO) theory, which in recent years has proved to be a reliable tool for the description of low-lying multiplets of lanthanide complexes with good accuracy.23

However, in these methods the interactions between the electrons in the environment and the ion are not treated in a perturbative manner, leading not only to an increased computational cost, but also to a difficult interpretation of the different physical mechanisms involved in the description of metal–ligand bonding.

A recent work presented a study on the role of the electrostatic interactions in the prediction of crystal field levels in lanthanide complexes.21 In that paper, the authors analyzed the ligand field generated by both different atomic charge distributions and a more rigorous approach based on the Fragment Ab Initio Model Potential (FAIMP) method.24–27 Besides the electrostatic interaction, the FAIMP approach allows inclusion of (i) the exchange interaction between the electrons of the ligand and the metal, and (ii) Pauli's repulsion contribution originating from the requirement of orthogonality between the metal and ligand molecular orbitals, which prevents over-delocalization of the electrons of one fragment in the core regions of the other. The authors' conclusions are that the electrostatic approach, even at the ab initio level via the FAIMP approach, does not suffice for an accurate description of the crystal field levels in Ln(III) complexes, and hence that covalency effects must play an important role.

In this work we investigate fragment ab initio modelling of crystal field levels in Ln(III) systems from an innovative point of view. Because it is considered one of the most rigorous non-covalent embedding methods of a metal ion in an ab initio potential created by the ligand, we start by briefly reviewing the FAIMP model as an ab initio electrostatic description. Then we introduce a novel ab initio approach based on the theory of multi-configurational non-orthogonal group functions (MC-NOGF), with the aim of modelling the strongly ionic metal–ligand interactions in terms of a systematic ab initio description based on purely intermolecular non-covalent interactions. This approach allows us to treat the metal–ligand interactions as a perturbation acting on the wave functions of the isolated fragments, which accounts in principle for both purely non-covalent (electrostatics, induction and dispersion) interactions and overlap effects which arise from the non-orthogonal theoretical framework. While in other methods focused on intermolecular interactions, such as Symmetry Adapted Perturbation Theory (SAPT),28–30 the expansions of the intermolecular interaction operator and of the overlap integrals are truncated to some order, the electronic Hamiltonian in MC-NOGF is the full molecular Hamiltonian, and the overlap effects are fully accounted for. Moreover, in the developments the preservation of the orbital localization on the fragments allows for a clear identification and separation of the different molecular interactions, which is not achievable in the conventional non-orthonormal configuration interaction (CI) methods, where usually the non-orthogonality is partially or totally removed before the evaluation of the Hamiltonian operator matrix elements by appropriate orbital mixing (e.g. biorthogonal basis sets).31–34 After presenting the equations for an ab initio multiconfiguration development involving matrix elements between the non-orthogonal sets of orbitals, as we implemented in our own quantum chemistry code CERES35, we study how different approximations to the ligands' wave function perform within the MC-NOGF method in the simulation of the crystal field level and magnetic properties of a set of Ln(III) complexes.

Hamiltonian and product functions

Within the Born–Oppenheimer approximation, the non-relativistic electronic Hamiltonian of NTOT interacting electrons can be defined in atomic units as
 
image file: d1cp05488k-t1.tif(1)
with
 
image file: d1cp05488k-t2.tif(2)
where ∇2(i) = ∂2/∂xi2 + ∂2/∂yi2 + ∂2/∂zi2, V(i) is the electron-nucleus potential, and rij = rirj. Within the group function (GF) approximation,36–43 the total Hamiltonian [script letter H] can be represented on the basis of the antisymmetrized products of the wave functions for the single parts composing the system as
 
image file: d1cp05488k-t3.tif(3)
with ψRr describing the group R, which collects NR electrons in the state r, and [scr A, script letter A] is the partial antisymmetrizer operator, which exchanges the electron coordinates x between two, or more, different groups. The group functions ψRr are individually antisymmetric with respect to the electron exchange
 
ψRr(…,xi,…,xj,…) = −ψRr(…,xj,…,xi,…)(4)
and ortho-normalized
 
image file: d1cp05488k-t4.tif(5)
where δij is the Kronecker delta. The product function Ψκ is fully characterized once the set of states κ ≡ (r, s,…) is made explicit. No inter-molecular correlation is explicitly included in a wave function of the form in eqn (3), since electrons belonging to different groups are assumed to move independently, but a large part of the intra-molecular correlation may be accounted for by admitting some degree of configuration interaction within each electron group.

Strongly orthogonal group functions and the FAIMP approach

A convenient way to handle fragment approaches based on group functions requires the group functions ψRr to fulfil a so-called strong-orthogonality condition38:
 
image file: d1cp05488k-t5.tif(6)
so that integrating over any one variable x1 common to two different group functions makes the integral vanish identically for all values of the other variables xi. This is a much stronger condition than the orthonormality requirement in eqn (5) which is assumed for the different group functions within each group.

It is well known in the literature that when a wave function has the form given in eqn (3) and satisfies the conditions in eqn (4), (5), and (6), the system is group-separable38. This means that the total energy of the system E can be written as (after integration over the spin variables)

 
image file: d1cp05488k-t6.tif(7)
where
 
HR = 〈 ψRr|[script letter H]R|ψRr(8)
 
image file: d1cp05488k-t7.tif(9)
 
image file: d1cp05488k-t8.tif(10)
where PRr is the spinless one-particle density matrix of the function ψRr. [script letter H]R in eqn (8) is the electronic Hamiltonian for the group R:
 
image file: d1cp05488k-t9.tif(11)
where V(i) in h(i) is the potential energy of electron i in the field of the nuclei of the total system.

Let us assume now that ψRr are functions of the sets of mono-electronic spin-orbital functions {ϕRi}, where ϕRi(x) = χRi(r)ηRi with spatial part χRi(r) and spin ηRi. Then eqn (6) is automatically satisfied if we impose the orthonormality conditions:

 
image file: d1cp05488k-t10.tif(12)
 
image file: d1cp05488k-t11.tif(13)

Starting from the total wave function in eqn (3), and defining the groups R, S,… as doubly occupied closed-shell singlet systems described by the functions

 
image file: d1cp05488k-t12.tif(14)
where [scr A, script letter A]R is the antisymmetrizer operator which exchanges all the electron spin–spatial coordinates x within the group R, and det|…| indicates the Slater determinant. Huzinaga and coworkers found that an optimum {ϕRi} set, which minimizes the energy E while holding fixed the {ϕSi} sets with SR, and imposing the conditions in eqn (12) and (13), is given by24–26
 
image file: d1cp05488k-t13.tif(15)
[scr F, script letter F] is the Fock operator defined by
 
image file: d1cp05488k-t14.tif(16)
with heff(i) given by
 
image file: d1cp05488k-t15.tif(17)
where [scr J, script letter J] and [scr K, script letter K] are the Coulomb and exchange operators, respectively, describing the effective field at point i due to the electrons in the groups S:
 
image file: d1cp05488k-t16.tif(18)
 
image file: d1cp05488k-t17.tif(19)
Note that [scr F, script letter F] is formally identical to the Fock operator for the total wave function Ψκ. The projector [scr P, script letter P]S originates from the orthonormality conditions, and reads24–26
 
image file: d1cp05488k-t18.tif(20)
Eqn (15) can be used as the self-consistent field (SCF) equation to determine the best orbitals for the group R interacting with the closed-shell singlets included as one or more groups S, enforcing the orthogonality between the orbitals of all groups. Eqn (15) is the starting point of all the developments that go under the name FAIMP.27,44

Let us now introduce a group M which is better described by a multi-configurational wave function ψMm defined as

 
image file: d1cp05488k-t19.tif(21)
where Cmm is the weight of the m′-th Slater determinant image file: d1cp05488k-t20.tif as given in eqn (14) built from the set of othonormal spin–orbitals {ϕMi}. As usual when dealing with a multi-configurational function, we can distribute the NM electrons into two sets of orbitals: (i) one set collecting NinactM orbitals that are doubly occupied in all the configurations image file: d1cp05488k-t21.tif, called inactive orbitals, and (ii) a second set collecting NactM orbitals which allow for an occupation number from 0 to 2, called active orbitals. Introducing ψMm in the product function Ψκ defined by eqn (3) we obtain
 
image file: d1cp05488k-t22.tif(22)
where now λ ≡ (m′, r, s,…) and κ ≡ (m, r, s,…). The problem of finding an optimal set {ϕRi} for the closed-shell singlet group R, still described by a single Slater determinant wave function ψRr, leads to the Fock operator as defined in eqn (16), but now heff(i) has the form
 
image file: d1cp05488k-t23.tif(23)
where [scr J, script letter J]Sj(i)/[scr J, script letter J]Mj(i) and [scr K, script letter K]Sj(i)/[scr K, script letter K]Mj(i) are defined in eqn (18) and (19), and
 
image file: d1cp05488k-t24.tif(24)
 
image file: d1cp05488k-t25.tif(25)
where ρmjk is the one-electron transition density matrix for ψMm, given by
 
image file: d1cp05488k-t26.tif(26)
which is averaged on Nm states in eqn (23) to account for the energy degeneracy of ψMm functions. In this work the projector operator [scr P, script letter P]M is defined as
 
image file: d1cp05488k-t27.tif(27)

Non-orthogonal group functions

FAIMP is an ab initio rigorous way to introduce the electrostatic interactions and Pauli repulsion between the metal and the environment; however, perhaps not surprisingly, it has shown poor performance in the accurate modelling of the electronic structure of Ln(III) complexes.21 To improve on this model, while keeping on pursuing a fragment ab initio approach to describe the metal–ligand interactions essentially in terms of intermolecular forces, let us start again from eqn (3), fix the sets of spin–orbitals {ϕRi} for all groups to some optimal orbitals obtained for the isolated fragments, and, crucially, relax the strong-orthogonality constraints. Thus we define the wave function for the whole system as
 
image file: d1cp05488k-t28.tif(28)
where Ψκ is a product function as defined in eqn (3), including a few closed-shell singlet groups described by ψRr wave functions given in eqn (14), i.e. the ligands, and one center characterized by the multi-configurational wave function ψMm shown in eqn (21), i.e. the metal. We fix the spin-orbitals entering the latter functions, and the Cmm coefficients entering eqn (21), to the sets obtained by energy minimization on the isolated fragments. Under this requirement, the orthogonality relationships in eqn (12) and (13) read
 
image file: d1cp05488k-t29.tif(29)
 
image file: d1cp05488k-t30.tif(30)
We are then interested in defining the optimal coefficients Dκ entering eqn (28), which can be estimated by variational minimization of the total energy
 
image file: d1cp05488k-t31.tif(31)
or, equivalently, in a matrix form solving the eigenvalue problem
 
HD = EMD(32)
where
 
Mκκ = 〈Ψκ|Ψκ(33)
 
Hκκ = 〈Ψκ|[script letter H]|Ψκ(34)
The dimension of the variational problem in eqn (32) is typically very large. Hence, to reduce the dimensionality, and also to set up an effective ab initio crystal field splitting theory i.e. a degenerate perturbation theory within the lanthanide ion's degenerate ground level, we apply the Löwdin partitioning procedure. This is achieved by partitioning the full space spanned by H, D and M in the two subspaces A (here the (2J + 1)-degenerate space of the lanthanide ground spin–orbit multiplet) and B (the complement excitation space)38
 
image file: d1cp05488k-t32.tif(35)
Thus an effective Hamiltonian [H with combining tilde]AA can be written for the subspace A as
 
[H with combining tilde]AADA = EMAADA(36)
where
 
[H with combining tilde]AA = HAA − (HABEMAB)(HBBEMBB)−1(HBAEMBA)(37)
Note that [H with combining tilde]AA is itself the function of the unknown energy E. Using the matrix identity
 
(XY)−1 = X−1 + X−1Y(XY)−1(38)
and using the fact that, upon iteration of eqn (38), the inverse of the (XY) term can be written as
 
(XY)−1 = X−1 + X−1YX−1+…(39)
the matrix elements of [H with combining tilde]AA have the form
 
image file: d1cp05488k-t33.tif(40)
where eqn (39) has been used together with
 
X = HBBdiagEMBBdiag(41)
 
Y = HBBoff-diagEMBBoff-diag(42)
to make the different expansion terms explicit. This partitioning scheme projects the problem to the NOGFs formed from the metal-based multielectron wave functions belonging to the (2J + 1)-degenerate ground spin–orbit multiplet, defining subspace A, whereas the interaction with the metal-based and ligand-based excitations belonging to subspace B can be accounted for in a perturbative manner, as per eqn (40).

Here we focus our attention on terms up to first-order in the perturbative expansion eqn (40) (i.e. the very first term on the right-hand side of eqn (40)). The energies are then determined by eqn (36) with [H with combining tilde]κκ = Hκκ, and the variational optimization involves the diagonalization of a matrix of dimension (2J + 1) × (2J + 1) once the integrals in eqn (43)–(45) are made explicit.

Implementation in the code CERES: cofactor versus inverse matrix formalism

We describe here the implementation strategy pursued within our own ab initio code CERES35,45 of the non-orthogonal CI problem arising from eqn (36) and (37) represented on the basis of the multiconfigurational non-orthogonal group functions, which is at the heart of the proposed MC-NOGF fragment approach.

From now on we drop the indices (r, s,…) ≡ (0, 0,…) in λ and κ indices for simplification. The explicit form for the normalization factor Mκκ on the Ψκ basis set is given by

 
image file: d1cp05488k-t34.tif(43)
In Hκκ, the one-electron terms have the form
 
image file: d1cp05488k-t35.tif(44)
and two-electron terms
 
image file: d1cp05488k-t36.tif(45)

Defining the overlap matrix S as

 
image file: d1cp05488k-t37.tif(46)
it is well known in the literature that
 
Ψn|Ψn〉 = det(S)(47)
 
image file: d1cp05488k-t38.tif(48)
and
 
image file: d1cp05488k-t39.tif(49)
where det(S) indicates the determinant of S, cof[S](ij) indicates the first-order cofactor arising from S when the row i and the column j are removed, and cof[S](ij|kl) indicates the second-order cofactor produced by further deleting the row k and the column l. If S−1 exists, the cofactors can be expressed using the Jacobi ratio theorem
 
cof[S](ij) = det(S)[S−1]ji(50)
 
cof[S](ij|kl) = det(S)([S−1]ji[S−1]lk − [S−1]li[S−1]jk)(51)
Although this theorem is a powerful tool that makes the calculation of cofactors simpler, the use of eqn (50) and (51) requires a matrix inversion for every couple of determinants in eqn (43)–(45), and can be a non-valid development when the matrix to be inverted is singular and the inverse does not exist.

Löwdin first presented a formula for the matrix elements of the Hamiltonian between Slater determinants expressed in terms of non-orthogonal spin-orbitals,46–49 and after this different methods have been suggested to avoid the need for calculating the inverse, allowing for a rapid evaluation of the cofactors,50–52 but in general they do not explicitly consider that the determinants from the set employed in a calculation may differ only in few spin-orbitals from a reference determinant. Hayes and Stone are the first to emphasize the advantages coming from properly taking this peculiarity into account,53 followed by Figari and Magnasco, who generalized Hayes' results.54 Starting from the latter developments, we present here a slightly modified approach, which allows for both computational efficiency and generality.

We start reordering the spin-orbital in S to obtain the block partitioning

 
image file: d1cp05488k-t40.tif(52)
where A is a symmetric matrix of dimension Ninact × Ninact, which includes the overlap elements between the spin-orbitals that are common to all the product functions, which are the inactive set of spin-orbitals for the metal and the spin-orbitals included in the ψR0 group functions. The C block collects the overlap terms between the Nact orbitals that can vary in the different product functions, which in this paper coincide with the active set of spin–orbitals for the metal. Note that NactNinact. The U and V blocks include the overlap elements between the active and inactive sets of spin-orbitals. While U, V, and C depend on a specific product function pair, A never changes. Then, S−1 can be written by the use of the Woodbury matrix identity as55
 
image file: d1cp05488k-t41.tif(53)
where Z = (CVA−1U), X = A−1U, and Y = VA−1. To avoid problems when Z is singular, we can apply eqn (50) back to the Z matrix's elements:
 
[Z−1]ij = cof[Z](ji)det(Z)−1(54)
or, in the matrix form
 
Z−1 = cof[Z]T[thin space (1/6-em)]det(Z)−1(55)
where cof[Z]T indicates the matrix collecting all the first-order cofactors of Z transpose. Defining the two matrices
 
image file: d1cp05488k-t42.tif(56)
and
 
image file: d1cp05488k-t43.tif(57)
and remembering that
 
det(S) = det(A)det(Z)(58)
eqn (50) and (51) can be written by the use of eqn (53) and (55) as
 
cof[S](ij)det(A)−1 = det(Z)Tji + Wji(59)
 
image file: d1cp05488k-t44.tif(60)
where Kijkl is defined as (all the possible cases under the conditions i < k and j < l in eqn (49))
 
image file: d1cp05488k-t45.tif(61)
with the p, q, r, and s indices running on the Z block. The relationship
 
cof[Z](ij|kl)det(Z) = cof[Z](ij)cof[Z](kl) − cof[Z](il)cof[Z](kj)(62)
has been used to cancel out the 1/det(Z) factor from Kijkl terms, and the permutational symmetry relationships
 
cof[Z](ij|kl) = −cof[Z](ji|kl) = −cof[Z](ij|lk) = cof[Z](ji|lk)(63)
have been used for simplification. We noted that, defining the two matrices
 
image file: d1cp05488k-t46.tif(64)
where I is the identity matrix with rank(I) = rank(C), Zijkl can be written by a unique expression as
 
image file: d1cp05488k-t47.tif(65)
Eqn (59) and (60) are not affected anymore by singularities, and they do not require any conditions but the existence of A−1. Moreover, the need for the calculation of first- and second-order cofactors of Z is not a big computational issue, because of the generally small dimension of this matrix with respect to the dimension of S. Defining det(C) = 1 when rank(C) = 0, note that if rank(C) = 2 only one second-order cofactor for Z exists, and it is 1, while the Kijkl term disappears when rank(C) < 2. Finally, eqn (59) and (60) reduce to just the T terms when rank(C) = 0, as expected. Some more simplifications can be done when making the spin partitioning of S explicit, and by the use of the Cholesky representation of the electron repulsion integrals,56 see the ESI. The equations for the evaluation of the magnetic moment and g-tensor within the MC-NOGF approach are also reported in the ESI.

Computational details

Calculations were carried out on a set of Ln(III) complexes to analyze how the FAIMP and NOGF approximations to the molecular wave function affect the energy gaps within the lowest energy spin–orbit multiplet and the magnetic properties. The results were compared with the energies obtained at the complete active space configuration interaction with spin–orbit (CASCI-SO) level coupled with the configuration averaged Hartree–Fock (CAHF) method.45,57

In the CAHF/CASCI-SO strategy, the set of molecular orbitals are generated at the CAHF level minimizing the average-energy functional represented on the basis of all possible Slater determinants, of any MS quantum number, built up allowing Nact active electrons to be distributed in all the possible ways in Mact active orbitals. Then, CAHF orbitals are used in the CASCI-SO step to construct the representation of the total Hamiltonian of the system, which includes both the Born–Oppenheimer electrostatic and spin–orbit Hamiltonian, still on the basis of all possible Slater determinants in the CAS space of Nact electrons in Mact orbitals. Finally, the total Hamiltonian is diagonalized to obtain the energy levels.

The number Mact = 7 active orbitals has been used for all the Ln(III) ions studied (i.e. 4f atomic shell), and the number Nact = 8, 9, 10, and 11 active electrons has been used for Tb(III), Dy(III), Ho(III), and Er(III) ions, respectively.

Within the FAIMP approach, the crystal field levels and magnetic proprieties are estimated from the wave function of the metal ψMm, optimized at the CAHF/CASCI-SO level in the presence of the ligands described by the FAIMP model as

image file: d1cp05488k-t48.tif
where ψR0, ψS0, and ψT0 indicate the ground state closed shell single Slater determinants of the separated molecules of the ligands optimized at the HF level and ψW0 indicates the ground state closed shell single Slater determinant of the whole group of the ligands optimized at the HF level. image file: d1cp05488k-t49.tif indicates that the functions on the left-hand side affect the function on the right-hand side by the FAIMP model, and image file: d1cp05488k-t50.tif indicates that the functions on both the left- and right-hand sides are adjusted to each other by a self-consistent procedure.

Within the NOGF theory presented in this paper, the crystal field levels and magnetic proprieties are estimated by the variational optimization of the pure electrostatic Hamiltonian on the basis of the product functions in eqn (28), which includes (i) the wave functions of the isolated metal ψMm optimized at the CAHF/CASCI-SO level and (ii) the ground state closed shell single Slater determinants optimized at the HF level describing the ligands as

image file: d1cp05488k-t51.tif
Note that in NOGF3 the functions ψMm included in the product functions are the wave functions of the isolated metal without any influence from the ligands, whereas ψMm used in the self-consistent procedure to adjust ψW0 is discharged after the optimization.

The geometries of the Ln complexes analyzed have been fixed to their experimental crystallographic structures reported in the literature.58–66 We indicate the ligands with acac = acetylacetonate, dppz = dipyridophenazine, dpq = dipyridoquinoxaline, phen = 1,10-phenanthroline, hfac = hexafluoroacetylacetonate, glyme = dimethoxyethane, paah = N-(2-pyridyl)-ketoacetamide, tta = 2-thenoyltrifluoroacetonate, bipy = 2,2′-bipyridine, meOH = methanol, H3 trensal = 2,2′,2′′-tris(salicylideneimido)trimethylamine, and tfpb = 4,4,4-trifluoro-1-phenylbutane-1,3-dionato, see Fig. 1.


image file: d1cp05488k-f1.tif
Fig. 1 Sketch, mark and color code for the molecules studied in this work.

All atoms were described by the ANO-RCC basis set,67 with the contraction [8s7p5d3f2g1h] for Ln atoms, [4s3p2d1f] for C, N, and O atoms coordinating the metal, [3s2p] for C, N, and O atoms not coordinating the metal, and [2s] for H.

Scalar relativistic terms were included in the one-electron part of the electrostatic Hamiltonian in all the HF, CAHF, and CASCI-SO computations, within the second order Douglas–Kroll–Hess (DKH2) approximation.68 In CASCI-SO evaluations, the spin–orbit interaction was included by the atomic approximation to the Breit–Pauli Hamiltonian, in which the two-electron spin–orbit integrals involving the atomic basis functions on multiple centers are discharged. The Cholesky representation of the electron repulsion integrals was used to speed up the calculations, with δ = 10−8.56

All the calculations were performed using the software package CERES, an ab initio quantum chemistry package specifically designed for the calculation of the electronic structure and magnetic properties of lanthanide complexes.

The errors affecting the energy gaps ΔEAPPi = EAPPiEAPP0, calculated within a given spin–orbit multiplet M = (2S + 1)LJ by the use of the approximation APP, were analyzed by the statistical indicators

 
image file: d1cp05488k-t52.tif(66)
 
image file: d1cp05488k-t53.tif(67)
 
image file: d1cp05488k-t54.tif(68)
where ΔEBENCi = EBENCiEBENC0 are the energy gaps calculated by the use of the true HBPSO, and N is the total number of states in the multiplet (minus one for the ground state in the ground spin–orbit multiplet). The results were graphically represented by the use of the probability density of the normal distribution:
 
image file: d1cp05488k-t55.tif(69)

Results

Fig. 2–5 present the crystal field levels within the ground spin–orbit multiplets of 44 different Tb(III)/Dy(III)/Ho(III)/Er(III) complexes estimated by the use of the different methods presented in the previous sections. In the figures, the energy gaps calculated at FAIMP or NOGF levels are reported on the x-axis, while the results obtained from CAHF/CASCI-SO simulations on the integral systems are reported on the y-axis for comparison (see also Tables S1–S4 in the ESI).
image file: d1cp05488k-f2.tif
Fig. 2 Graphical representation of the crystal field levels within the ground spin–orbit multiplet 7F6 of different Tb(III) complexes, estimated at the CAHF/CASCI-SO level on the whole system (y axis) versus the different approximations presented in this work (x axis). The mean μ and standard deviation σ for the %errors affecting the energy gaps are represented by the normal distribution function.

image file: d1cp05488k-f3.tif
Fig. 3 Graphical representation of the crystal field levels within the ground spin–orbit multiplet 6H15/2 of different Dy(III) complexes, estimated at the CAHF/CASCI-SO level on the whole system (y axis) versus the different approximations presented in this work (x axis). The mean μ and standard deviation σ for the %errors affecting the energy gaps are represented by the normal distribution function.

image file: d1cp05488k-f4.tif
Fig. 4 Graphical representation of the crystal field levels within the ground spin–orbit multiplet 5I8 of different Ho(III) complexes, estimated at the CAHF/CASCI-SO level on the whole system (y axis) versus the different approximations presented in this work (x axis). The mean μ and standard deviation σ for the %errors affecting the energy gaps are represented by the normal distribution function.

image file: d1cp05488k-f5.tif
Fig. 5 Graphical representation of the crystal field levels within the ground spin–orbit multiplet 4I15/2 of different Er(III) complexes, estimated at the CAHF/CASCI-SO level on the whole system (y axis) versus the different approximations presented in this work (x axis). The mean μ and standard deviation σ for the %errors affecting the energy gaps are represented by the normal distribution function.

In FAIMP1 and NOGF1 strategies the metal interacts with the electron densities of the separated molecules, which are included in the environment after separate optimisation. For FAIMP1 simulations, the graphs show a general large underestimation of the energy gaps with respect to the CAHF/CASCI-SO results for all the studied ions, with mean percentage errors of about −80%/−90%. These results can be attributed to a poor description of the ligands' density, which, as expected, cannot be described as the simple sum of the densities of the isolated molecules, but the latter must adjust to each other within the complex structure. However, it is noteworthy that generally slightly better results are obtained when the densities of the isolated molecules are treated within the NOGF method (see NOGF1 results). This improvement is especially observed for the Tb-pc2 complex, which is formed by two highly π conjugated pc−2 ligands, where the NOGF1 methodology reproduces the CAHF/CASCI-SO results with a mean %error smaller than −20% and a small standard deviation of about 4%.

A general improvement in the results is obtained when the ligands are described by a single wave function optimized on the whole set of atoms present in the environment, and the presence of a FAIMP model potential describing the metal ion. This approach gives rise to the FAIMP2 and NOGF2 methods. In this case, while the percentage errors affecting the crystal field levels calculated by the FAIMP2 method are still rather large for all the ions, the NOGF2 strategy performs significantly better. In NOGF2 in fact the wavefunction of the ligands and that of the isolated metal are combined together in a single antisymmetrised non-orthogonal product function, which leads to results in rather good agreement with those obtained at the full CAHF/CASCI-SO level on the whole set of systems explored here. This not only proves the importance of including an accurate description for the electron density of the ligands, but also seems to indicate that the rigid interaction (i.e. no orbital relaxation) between the metal density optimized on the isolated ion and a suitable density which accurately describes the atoms in the environment already suffices for the modelling of the crystal field splittings in these systems.

Finally, the third family of approximations explored here is named FAIMP3 and NOGF3, where the densities of the metal and the ligands are iteratively optimized to each other, to fully capture the polarization effects of both ligands and metal electrons, and, consequently, a refined description of electrostatic interactions between metals and ligands. The only family of non-covalent metal–ligand interactions that cannot be captured here is dispersion interactions, since the dynamical correlation is currently not included in our wave function models. We note however that in our proposed NOGF strategy such dispersion interactions can indeed be captured by going in higher order expansion of eqn (40), while this is not possible in the FAIMP approximation.

Despite the significant improvement observed in both FAIMP3 and NOGF3 approaches, note that the FAIMP3 strategy still underestimates the energy gaps. On the other hand, the NOGF3 method shows crystal field gaps almost coincident with the CAHF/CASCI-SO results for the majority of the Tb(III), Dy(III), and Er(III) complexes in all the range of frequencies. Slightly larger discrepancies are shown by paah2-2NO3-meOH ligands, where an average overestimation of the energy gaps of about +40% and +17% is found for Tb(III) and Dy(III) ions, respectively, and an underestimation of about −26% for the Er(III) metal. Differently, the NOGF3 strategy does not lead to a significant improvement of the crystal field energies of all Ho(III) complexes.

We also note that Dy-trensal and Er-trensal show very different behaviours. On the one hand, the Dy(III) system is accurately reproduced by the NOGF3 method both in the low and high energy ranges. On the other hand, NOGF3 for Er-trensal shows quite large discrepancies with the CAHF/CASCI-SO energies, especially for the levels above 400 cm−1, leading to a large underestimation of the NOGF3 crystal field gaps when compared with the fully interacting CAHF/CASCI-SO results.

Besides the crystal field energy spectrum, we also tested the proposed methods for the calculation of the magnetic properties. In particular, in Tables 1 and 2 we report the principal values of the g tensor for the ground and first excited pseudo-Kramers/Kramers doubles (KD) for Tb(III)/Dy(III) complexes, estimated via different approximations. From the tables, the discrepancies between the NOGF3 and CAHF/CASCI-SO results for the ground KDs never exceed 1.5% for all the Tb(III) and Dy(III) systems. A very good agreement is also found in the orientation of the main magnetic axis, where the angle between the axes estimated by the NOGF3 and CAHF/CASCI-SO methods is in all cases smaller than 8 degrees, see Tables 3 and 4. Larger discrepancies are observed for the first excited KDs, where now the g-tensor calculated at the NOGF3 level is affected by errors of about 5%/10%, and the angles of the main magnetic axes are between 5 and 15 degrees, with the exception of acac3-dpq (22°), acac3-dppz (24°), tta3-phen (39°), and hfac3-glyme (21°) ligands for Tb(III), and the tta3bipy ligand for Dy(III).

Table 1 g tensor values for the two lowest lying pseudo-Kramers doublets (KD) within the ground spin–orbit multiplet 7F6 of different Tb(III) complexes
CAHF/CASCI-SO FAIMP1 FAIMP2 FAIMP3 NOGF1 NOGF2 NOGF3
acac3-2H2 O KD1 17.578 +0.143 −0.575 +0.054 +0.006 −0.567 +0.100
KD2 17.061 −2.928 −5.403 −3.722 −3.454 −3.404 −2.489
acac3-dpq KD1 17.335 −0.356 +0.321 +0.455 −0.358 −0.201 +0.072
KD2 13.189 −0.573 +0.687 +1.089 −0.862 −0.102 +0.664
acac3-dppz KD1 17.620 −0.890 +0.077 +0.251 −2.756 −0.061 +0.143
KD2 17.109 −4.348 −3.384 −2.556 −6.072 −2.372 −1.655
acac3-phen KD1 17.463 −0.109 +0.357 +0.392 −0.129 +0.055 +0.187
KD2 14.755 −1.748 −0.304 −0.228 −1.401 −0.574 −0.286
pc2 KD1 17.918 −0.472 +0.020 −0.038 +0.019 +0.020 +0.002
KD2 14.546 +1.543 +0.310 +0.180 +0.358 +0.407 +0.210
paah2-2NO3-meOH KD1 17.772 −0.316 +0.146 +0.157 +0.039 +0.137 +0.149
KD2 15.199 +0.746 −0.392 −0.382 −0.167 −0.144 −0.197
tta3-bipy KD1 17.759 −1.175 +0.128 +0.173 −0.822 +0.027 +0.095
KD2 15.569 −3.282 −0.595 −0.661 −0.583 −1.355 −0.902
tta3-phen KD1 17.608 −0.217 +0.138 +0.255 −0.645 −0.039 +0.097
KD2 13.736 +0.636 +0.424 +0.858 −0.973 −0.421 +0.448
tfpb3-dppz KD1 17.484 −0.285 +0.057 +0.341 −1.459 −0.147 +0.100
KD2 12.229 +1.537 +1.191 +2.290 +0.038 −0.394 +0.996
hfac3-glyme KD1 17.630 −1.596 −0.061 +0.168 −1.774 −0.144 +0.035
KD2 13.774 −2.114 −0.427 +0.574 −2.462 −1.070 −0.029


Table 2 Main g tensor values for the two lowest lying Kramers doublets (KD) within the ground spin–orbit multiplet 6H15/2 of different Dy(III) complexes
CAHF/CASCI-SO FAIMP1 FAIMP2 FAIMP3 NOGF1 NOGF2 NOGF3
acac3-2H2 O KD1 19.471 +0.380 −0.276 +0.054 −0.627 −0.127 +0.009
KD2 15.549 +1.647 −0.467 +0.465 −0.141 −0.564 +0.123
acac3-dpq KD1 19.219 −0.038 +0.459 +0.517 +0.229 +0.152 +0.275
KD2 16.230 +0.288 +0.096 +0.354 +0.000 −0.011 +0.267
acac3-dppz KD1 19.366 +0.311 +0.131 +0.381 −0.817 −0.045 +0.159
KD2 15.369 +1.292 +0.784 +1.460 −2.917 −0.063 +0.690
acac3-phen KD1 19.344 +0.242 +0.310 +0.390 −0.203 +0.053 +0.192
KD2 15.701 +0.308 +0.809 +1.034 +0.247 +0.414 +0.669
pc2 KD1 17.490 +2.413 +2.411 +2.410 +0.993 +0.133 +0.283
KD2 14.690 +2.465 +2.504 +2.496 +5.126 +1.678 +1.657
paah2-2NO3-meOH KD1 19.592 −0.391 +0.273 +0.292 +0.162 +0.138 +0.193
KD2 16.490 +0.636 +0.648 +0.680 +0.584 +0.338 +0.501
tta3-bipy KD1 19.576 −0.224 +0.131 +0.279 −0.616 +0.073 +0.200
KD2 15.102 +3.354 +1.860 +2.120 −0.869 +2.047 +2.495
tta3-phen KD1 19.485 +0.296 +0.263 +0.317 −0.079 +0.130 +0.224
KD2 15.674 +2.944 +1.249 +1.305 +1.226 +1.038 +1.158
tfpb3-dppz KD1 19.296 +0.190 +0.363 +0.504 −0.270 +0.031 +0.207
KD2 14.549 +4.845 +2.015 +2.388 −1.588 +1.190 +1.883
hfac3-glyme KD1 19.450 +0.336 +0.160 +0.293 −0.278 +0.086 +0.230
KD2 15.930 −3.308 +0.548 +0.889 +0.532 +0.522 +0.789


Table 3 Main magnetic axes for the two lowest lying pseudo-Kramers doublets (KD) within the ground spin–orbit multiplet 7F6 of different Tb(III) complexes: angle (in degrees) between the axes estimated by the different methods proposed in this paper and the ab initio computation on the whole systems at the CAHF/CASCI-SO level
FAIMP1 FAIMP2 FAIMP3 NOGF1 NOGF2 NOGF3
acac3-2H2O KD1 75 36 23 81 6 8
KD2 59 27 32 43 37 22
acac3-dpq KD1 45 14 19 26 4 7
KD2 40 5 6 16 10 6
acac3-dppz KD1 89 18 20 66 3 6
KD2 84 39 40 57 23 24
acac3-phen KD1 57 12 13 24 2 4
KD2 31 20 20 30 15 15
pc2 KD1 17 5 9 3 0 2
KD2 17 5 4 7 0 9
paah2-2NO3-meOH KD1 10 9 9 7 5 6
KD2 86 11 12 1 7 9
tta3-bipy KD1 81 8 8 72 1 1
KD2 38 35 41 61 32 39
tta3-phen KD1 42 7 9 19 2 3
KD2 58 16 11 50 15 9
tfpb3-dppz KD1 62 8 10 12 3 3
KD2 78 3 9 53 5 5
hfac3-glyme KD1 89 6 8 46 2 3
KD2 90 21 20 41 24 21


Table 4 Main magnetic axes for the two lowest lying Kramers doublets (KD) within the ground spin–orbit multiplet 6H15/2 of different Dy(III) complexes: angle (in degrees) between the axes estimated by the different methods proposed in this paper and the ab initio computation on the whole systems at the CAHF/CASCI-SO level
FAIMP1 FAIMP2 FAIMP3 NOGF1 NOGF2 NOGF3
acac3-2H2O KD1 75 11 5 28 2 2
KD2 73 8 2 49 2 5
acac3-dpq KD1 51 15 15 18 6 6
KD2 43 12 11 13 10 11
acac3-dppz KD1 86 10 11 11 2 3
KD2 83 7 9 16 4 8
acac3-phen KD1 70 14 13 19 6 5
KD2 23 15 18 9 8 13
pc2 KD1 12 9 9 21 2 4
KD2 4 9 10 16 5 6
paah2-2NO3-meOH KD1 37 4 3 2 2 2
KD2 28 1 1 14 0 0
tta3-bipy KD1 85 15 19 9 4 7
KD2 79 36 32 62 24 25
tta3-phen KD1 78 19 16 20 6 6
KD2 45 23 23 24 13 15
tfpb3-dppz KD1 89 16 18 10 4 5
KD2 50 22 21 30 9 12
hfac3-glyme KD1 73 14 16 19 6 8
KD2 39 16 22 14 10 16


All results presented here clearly show that our proposed NOGF methodologies for the description of non-covalent crystal field interactions in lanthanide complexes perform significantly better than the FAIMP methods of equivalent quality, and in fact, quite remarkably, they often manage to reproduce the fully interacting results obtained via CAHF/CASCI-SO to an unexpected degree of accuracy. To appreciate the meaning of these results, it is important to remember that the CAHF/CASCI-SO and CASSCF/SI-SO methods, while neglecting the dynamical correlation between metal and ligand electrons, do provide a mean-field description of metal–ligand charge-transfer excitations, and hence of covalency effects, via the refinement of non-redundant active-to-virtual and inactive-to-active orbital rotations carried out during the orbital optimisation process (in other words, via the optimisation of orbitals containing a small degree of metal–ligand hybridisation). On the other hand, FAIMP and NOGF by definition do not formally account for charge transfer processes between different groups; hence they are formally incapable of describing any metal–ligand interactions other than non-covalent interactions. Thus two points of discussion immediately arise from these results.

First, the observation that NOGF performs consistently better than the FAIMP version of equivalent quality would seem to suggest that in NOGF, the process of building antisymmetrised product functions from non-orthogonal wave functions associated with the single groups, and of producing linear combinations of such product functions to form the eigenstates of the first-order degenerate perturbation theory interaction Hamiltonian, can lead to a recovery of some degree of effective metal–ligand hybridisation, as much as it is necessary to ensure orthogonality between different orbitals in an effective orbital-delocalised description leading to equivalent results. The minimal hybridisation required between localised non-orthogonal orbitals within an orthogonalisation process can be considered as a component of covalency as described by an orthogonal and hybridised basis, accounting for some charge transfer processes between metal-centered and ligand-centered orbitals. One should also consider that a rather significant difference between NOGF and FAIMP methods is not fully unexpected, given that non-orthogonality makes the matrix elements of the electrostatic Hamiltonian between group product functions significantly more complex and rich in information about group–group interactions with respect to the effective matrix elements of the FAIMP method, where the strong orthogonality constraint makes the Hamiltonian describing the interaction between metal and ligands eqn (23) significantly simpler than the full non-orthogonal results of the NOGF method, eqn (48) and (49). After all, FAIMP can be considered as an approximation to the full NOGF method, hence bound to lead to a poorer performance.

Second, the observation that, generally speaking, the NOGF3 approach appears to reproduce to a high degree of accuracy the fully interacting CAHF/CASCI-SO results suggests that fully-interacting multiconfigurational ab initio methods ignoring the dynamical correlation, such as CAHF/CASCI-SO or CASSCF/SI-SO, lead to a description of the Ln(III)–ligand interactions that is strongly dominated by intermolecular interactions, and thus probably tend to underestimate covalency effects, so much so that for most systems explored here the mean field covalency effects captured by the molecular orbital optimisation process are captured also by our non-covalent NOGF methods.

However, we have also exposed here a few cases where even our best NOGF approach displays some difficulties in recovering the fully interacting CAHF/CASCI-SO results. We can identify here a major systematic trend in these more difficult cases. We note in fact that, generally speaking, the NOGF methodology tends to perform much better for complexes of Dy(III) and Tb(III) ions, rather than for equivalent complexes of Er(III) and Ho(III) ions. We provide here a rationalisation of this behaviour in terms of the multipolar expansion of the 4f-charge distributions associated with the MJ states of a spin–orbit multiplet with a total angular momentum quantum number J. As shown in fact first by Sievers,10 the asphericity of the 4f charge density distribution, ρJ,MJ(θ, ϕ), can be expanded in terms of just three spherical harmonics Ckq(θ, ϕ), corresponding to the 4f multipolar expansion of the charge density associated with a given state MJ as in eqn (70)

 
ρJ,MJ(θ,ϕ) = A2,MJC20(θ, ϕ) + A4,MJC40(θ,ϕ) + A6,MJC60(θ, ϕ),(70)
where A2,MJ is the quadrupole coefficient for the state MJ, while A4,MJ and A6,MJ are the hexadecapole and hexacontadecapole coefficients, respectively, for the same state. In the approximation that the J-multiplet is dominated by a single Russell–Saunders term with specific values of the L and S total angular momentum and spin quantum numbers (known to be a very good approximation for complexes of trivalent lanthanide ions), the three multipolar coefficients can be exactly calculated as
 
image file: d1cp05488k-t56.tif(71)
Eqn (71) shows that the magnitude of a given multipolar term describing the charge density of different crystal field states for a given lanthanide, aside from a simple Wigner 3j symbol differentiating different MJ states within the same multiplet, is essentially determined by the reduced matrix element 〈LSJ||Ck||LSJ〉, for which an expression can be found e.g. in ref. 10. The direct calculation of this reduced matrix element determining the magnitude of the quadrupolar component of the charge density (k = 2), for ions Tb (L = 3, S = 3, J = 6), Dy (L = 5, S = 5/2, J = 15/2), Ho (L = 6, S = 2, J = 8), and Er (L = 6, S = 3/2), gives
 
LSJC2LSJ〉 = −0.42(Tb); −0.40(Dy); −0.16(Ho); 0.16(Er).(72)

Furthermore, within a simple point charge model, we can easily show that the strength of the electrostatic interactions between metal and ligands in lanthanides is strongly dominated by the quadrupolar term. This can be shown in the following manner. If a ligand charge is at a distance R from the lanthanide ion, the strength of the electrostatic potential generated at distance R by the 4f-multipole of order k associated with the 4f charge density distribution can be estimated as image file: d1cp05488k-t57.tif. For a typical metal–ligand distance of R = 5a0(2.65 Å), and using the radial multipoles tabulated for all trivalent lanthanide ions by Edvardsson and Klintenberg,69 this leads to rather small ratios between the hexadecapolar (k = 4) and quadrupolar (k = 2) crystal field strengths of image file: d1cp05488k-t58.tif, and to even smaller ratios between hexacontadecapolar (k = 6) and quadrupolar terms for all four ions: image file: d1cp05488k-t59.tif, which is evidence that metal–ligand electrostatics are dominated by the k = 2 quadrupolar interactions (quadrupole–charge interactions in this simple example).

This simple argument thus shows that intermolecular electrostatic interactions between the ligands and the lanthanide ions in a complex are dominated by quadrupolar terms, and thus are much stronger in Dy and Tb than in Ho and Er, where the strength of the quadrupolar terms of the 4f charge density expansion calculated in eqn (72) is found to be less than half the magnitude of those found for Dy and Tb. The fact that in Er and Ho the interactions between metal and ligands will be less strongly dominated by pure electrostatics provides a rationalisation of our finding that ab initio non-covalent approaches like NOGF have more difficulties in providing a satisfactory description of metal–ligand interactions with respect to methods like CAHF/CASCI-SO, in which covalent ligand–metal interactions can be partially captured via optimisation of ligand–metal orbital hybridisation.

At this time, we were not able to rationalise the relatively poor performance of the NOGF methods for Dy(III) and Tb(III) for the single case of the paah2-2NO3-meOH ligands.

Conclusions

We presented a novel multiconfigurational non-orthogonal group function (MC-NOGF) approach for the calculation of the crystal field energies and magnetic properties of trivalent lanthanide complexes. The novel MC-NOGF approach has been implemented in the ab initio program Computational Emulator of Rare Earth Systems (CERES).35,45 By direct comparison of two different fragment approaches of the crystal field energies of 44 complexes, namely the FAIMP and the here proposed MC-NOGF, we showed that the MC-NOGF model offers a better performing approach to an ab initio non-covalent description of metal–ligand interactions in trivalent lanthanide complexes. This suggests that many bonding situations in lanthanide complexes, especially involving lanthanide ions whose 4f charge density has a strong quadrupolar component such as Tb(III) and Dy(III), can be satisfactorily described in terms of “intermolecular interactions” between ligands and 4f electrons, a fact that could be used not only to better understand the chemistry of 4f coordination related to different ligands, but also to design more efficient and localised ab initio methods for the inclusion of dynamical correlation effects, which could focus on the description of those excitations that are by definition absent from NOGF strategies.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

A. S. thankfully acknowledges support from the Australian Research Council, Future Fellowship FT180100519, and Discovery Project Grant DP210103208. This research was supported by the Nectar Research Cloud (https://nectar.org.au), a collaborative Australian research platform supported by the National Collaborative Research Infrastructure Strategy (NCRIS).

References

  1. R. M. Pallares and R. J. Abergel, Nanoscale, 2020, 12, 1339–1348 RSC.
  2. I. Georgieva, T. Zahariev, A. J. A. Aquino, N. Trendafilova and H. Lischa, Spectrochim. Acta, Part A, 2020, 240, 118591 CrossRef CAS PubMed.
  3. A. G. Cosby, J. J. Woods, P. Nawrocki, T. J. Sørensen, J. J. Wilson and E. Boros, Chem. Sci., 2021, 12, 9442–9451 RSC.
  4. S. E. Bodman, C. Breen, S. Kirkland, S. Wheeler, E. Robertson, F. Plasser and S. J. Butler, Chem. Sci., 2022, 13, 3386–3394 RSC.
  5. J. Tang and P. Zhang, Lanthanide Single Molecule Magnets, Springer-Verlag Berlin Heidelberg, 1st edn, 2015 Search PubMed.
  6. C. A. P. Goodwin, F. Ortu, D. Reta, N. F. Chilton and D. P. Mills, Nature, 2017, 548, 439–442 CrossRef CAS PubMed.
  7. F. S. Guo, B. M. Day, Y. C. Chen, M. L. Tong, A. Mansikkamaki and R. A. Layfield, Science, 2018, 362, 1400 CrossRef CAS PubMed.
  8. H. A. Bethe, Splitting of terms in crystals: Complete English translation, Consultants Bureau, 1929 Search PubMed.
  9. J. D. Rinehart and J. R. Long, Chem. Sci., 2011, 2, 2078–2085 RSC.
  10. J. Sievers, Z. Phys. B, 1982, 45, 289 CrossRef CAS.
  11. N. F. Chilton, D. Collison, E. J. L. McInnes, R. E. P. Winpenny and A. Soncini, Nat. Commun., 2013, 4, 2551 CrossRef PubMed.
  12. C. Gao, A. Genoni, S. Gao, S. Jiang, A. Soncini and J. Overgaard, Nat. Chem., 2020, 12, 213–219 CrossRef CAS PubMed.
  13. M. T. Hutchings and D. K. Ray, Proc. Phys. Soc., 1963, 81, 663–676 CrossRef CAS.
  14. D. K. Ray, Proc. Phys. Soc., 1963, 82, 47–57 CrossRef CAS.
  15. C. J. Lenander and E. Y. Wong, J. Chem. Phys., 1963, 38, 2750–2752 CrossRef CAS.
  16. R. E. Watson and A. J. Freeman, Phys. Rev., 1964, 134, A1526–A1546 CrossRef CAS.
  17. K. Rajnak and B. G. Wybourne, J. Chem. Phys., 1964, 41, 565–569 CrossRef CAS.
  18. D. Newman, Adv. Phys., 1971, 20, 197–256 CrossRef CAS.
  19. C. A. Morrison and R. P. Leavitt, Handbook on the Physics and Chemistry of Rare Earths. vol. 5, North Holland, 1982, vol. 5, pp. 461–692 Search PubMed.
  20. O. Malta, S. Ribeiro, M. Faucher and P. Porcher, J. Phys. Chem. Solids, 1991, 52, 587–593 CrossRef CAS.
  21. L. Ungur and L. F. Chibotaru, Chem. – Eur. J., 2017, 23, 3708–3718 CrossRef CAS PubMed.
  22. R. Alessandri, H. Zulfikri, J. Autschbach and H. Bolvin, Chem. – Eur. J., 2018, 24, 5538–5550 CrossRef CAS PubMed.
  23. L. Ungur, Lanthanide-Based Multifunctional Materials, Elsevier, 2018, pp. 1–58 Search PubMed.
  24. J. D. Weeks and S. A. Rice, J. Chem. Phys., 1968, 49, 2741–2755 CrossRef CAS.
  25. S. Huzinaga and A. A. Cantu, J. Chem. Phys., 1971, 55, 5543–5549 CrossRef CAS.
  26. S. Huzinaga, D. Mcwilliams and A. A. Cantu, Adv. Quantum Chem., 1973, vol. 7, 187–220 CrossRef.
  27. B. Swerts, L. F. Chibotaru, R. Lindh, L. Seijo, Z. Barandiaran, S. Clima, K. Pierloot and M. F. A. Hendrickx, J. Chem. Theory Comput., 2008, 4, 586–594 CrossRef CAS PubMed.
  28. B. Jeziorski, R. Moszynski and K. Szalewicz, Chem. Rev., 1994, 94, 1887–1930 CrossRef CAS.
  29. K. Szalewicz, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 254–272 CAS.
  30. K. Patkowski, P. S. Zuchowski and D. G. A. Smith, J. Chem. Phys., 2018, 148, 164110 CrossRef PubMed.
  31. P. W. Payne, J. Chem. Phys., 1982, 77, 5630–5638 CrossRef CAS.
  32. P. A. Malmqvist, Int. J. Quantum Chem., 1986, 30, 479–494 CrossRef CAS.
  33. P.-Å. Malmqvist and B. O. Roos, Chem. Phys. Lett., 1989, 155, 189–194 CrossRef CAS.
  34. A. Mitrushchenkov and H.-J. Werner, Mol. Phys., 2007, 105, 1239–1249 CrossRef CAS.
  35. A. Soncini, S. Calvello, M. Piccardo and S. V. Rao, Ceres, an ab initio quantum chemistry package for the electronic structure and magnetic properties of lanthanide complexes, 2018 Search PubMed.
  36. A. L. Tchougréeff, Hybrid Methods of Molecular Modeling, Springer, Netherlands, Dordrecht, 2008, vol. 17 Search PubMed.
  37. J. G. Ángyán and G. Náray-Szabó, in Theoretical Treatment of Large Molecules and Their Interactions, ed. Z. B. Maksic, Springer Berlin Heidelberg, Berlin, Heidelberg, 1991, pp. 1–49 Search PubMed.
  38. R. McWeeny and B. T. Sutcliffe, Methods of molecular quantum mechanics, Academic Press, San Diego, 1969 Search PubMed.
  39. E. L. Mehler, J. Math. Chem., 1992, 10, 57–91 CrossRef CAS.
  40. E. L. Mehler, J. Chem. Phys., 1981, 74, 6298 CrossRef CAS.
  41. E. L. Mehler, J. Chem. Phys., 1977, 67, 2728 CrossRef CAS.
  42. C. Amovilli and R. McWeeny, Chem. Phys., 1990, 140, 343–361 CrossRef CAS.
  43. R. McWeeny, Proc. R. Soc. A, 1959, 253, 242–259 CAS.
  44. T. Shimazaki, K. Kitaura, D. G. Fedorov and T. Nakajima, J. Chem. Phys., 2017, 146, 084109 CrossRef PubMed.
  45. S. Calvello, M. Piccardo, S. V. Rao and A. Soncini, J. Comput. Chem., 2018, 39, 328–337 CrossRef CAS PubMed.
  46. P.-O. Löwdin, J. Chem. Phys., 1950, 18, 365–375 CrossRef.
  47. W. J. Carr, Phys. Rev., 1953, 92, 28–35 CrossRef.
  48. P.-O. Löwdin, Phys. Rev., 1955, 97, 1490–1508 CrossRef.
  49. P.-O. Löwdin, Phys. Rev., 1955, 97, 1474–1489 CrossRef.
  50. F. Prosser and S. Hagstrom, Int. J. Quantum Chem., 1968, 2, 89–99 CrossRef CAS.
  51. S. C. Leasure and G. Balint-Kurti, Phys. Rev. A: At., Mol., Opt. Phys., 1985, 31, 2107–2113 CrossRef CAS PubMed.
  52. J. Verbeek and J. H. Van Lenthe, Theochem, 1991, 229, 115–137 CrossRef.
  53. I. Hayes and A. Stone, Mol. Phys., 1984, 53, 69–82 CrossRef CAS.
  54. G. Figari and V. Magnasco, Mol. Phys., 1985, 55, 319–330 CrossRef CAS.
  55. W. W. Hager, SIAM Rev., 1989, 31, 221–239 CrossRef.
  56. M. Piccardo and A. Soncini, J. Comput. Chem., 2017, 38, 2775–2783 CrossRef CAS PubMed.
  57. W. Van den Heuvel, S. Calvello and A. Soncini, Phys. Chem. Chem. Phys., 2016, 18, 15807–15814 RSC.
  58. S.-D. Jiang, B.-W. Wang, G. Su, Z.-M. Wang and S. Gao, Angew. Chem., 2010, 122, 7610–7613 CrossRef.
  59. G.-J. Chen, C.-Y. Gao, J.-L. Tian, J. Tang, W. Gu, X. Liu, S.-P. Yan, D.-Z. Liao and P. Cheng, Dalton Trans., 2011, 40, 5579 RSC.
  60. G.-J. Chen, Y.-N. Guo, J.-L. Tian, J. Tang, W. Gu, X. Liu, S.-P. Yan, P. Cheng and D.-Z. Liao, Chem. – Eur. J., 2012, 18, 2484–2487 CrossRef CAS PubMed.
  61. E. M. Fatila, E. E. Hetherington, M. Jennings, A. J. Lough and K. E. Preuss, Dalton Trans., 2012, 41, 1352–1362 RSC.
  62. N. F. Chilton, S. K. Langley, B. Moubaraki, A. Soncini, S. R. Batten and K. S. Murray, Chem. Sci., 2013, 4, 1719 Search PubMed.
  63. Z.-G. Wang, J. Lu, C.-Y. Gao, C. Wang, J.-L. Tian, W. Gu, X. Liu and S.-P. Yan, Inorg. Chem. Commun., 2013, 27, 127–130 CrossRef CAS.
  64. Y. Bi, Y.-N. Guo, L. Zhao, Y. Guo, S.-Y. Lin, S.-D. Jiang, J. Tang, B.-W. Wang and S. Gao, Chem. – Eur. J., 2011, 17, 12476–12481 CrossRef CAS PubMed.
  65. M. Habib, S. Sain, B. Das and S. Chandra, J. Indian Chem. Soc., 2011, 88, 1501–1508 CAS.
  66. Q. Sun, P. Yan, W. Niu, W. Chu, X. Yao, G. An and G. Li, RSC Adv., 2015, 5, 65856–65861 RSC.
  67. B. O. Roos, R. Lindh, P.-Å. Malmqvist, V. Veryazov, P.-O. Widmark and A. C. Borin, J. Phys. Chem. A, 2008, 112, 11431–11435 CrossRef CAS PubMed.
  68. M. Reiher and A. Wolf, Relativistic quantum chemistry: the fundamental theory of molecular science, Wiley-VCH, 2015 Search PubMed.
  69. S. Edvardsson and M. Klintenberg, J. Alloys Compd., 2013, 275–277, 230–233 Search PubMed.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d1cp05488k
These authors contributed equally to this work.

This journal is © the Owner Societies 2022
Click here to see how this site uses Cookies. View our privacy policy here.