Open Access Article
Ilya Popov
a,
Petros-Panagis Filippatos
a,
Shayantan Chaudhuri
a,
Andrei L. Tchougréeff
b,
Katherine Inzani
a and
Elena Besley
*a
aSchool of Chemistry, University of Nottingham, Nottingham NG7 2RD, UK. E-mail: Elena.Besley@nottingham.ac.uk
bFrumkin Institute of Physical Chemistry and Electrochemistry of the Russian Academy of Sciences, Moscow 119071, Russia
First published on 24th December 2025
Transition metal doping is commonly used for altering the properties of solid-state materials to suit applications in science and technology. Partially filled d-shells of transition metal atoms lead to electronic states with diverse spatial and spin symmetries. Chromium(III) cations have shown great potential for designing laser materials and, more recently, for developing spin qubits in quantum applications. They also represent an intriguing class of chemical systems with strongly correlated multi-reference excited states, due to the d3 electron configuration. These states are difficult to describe accurately using single-reference quantum chemical methods such as density functional theory (DFT), the most commonly used method to study the electronic structures of solid-state systems. Recently, the periodic effective Hamiltonian of crystal field (pEHCF) method has been shown to overcome some limitations arising in the calculations of excited d-states. In this work, we assess the suitability of DFT and pEHCF to calculate the electronic structure and d–d excitations of chromium(III) dopants in wide band gap host materials. The results will aid computational development of novel transition metal-doped materials and provide a deeper understanding of the complex nature of transition metal dopants in solids.
Dopants with open d-shells exhibit electronic correlation effects17 of both static and dynamical nature, which occur due to the strong electron–electron interactions characteristic of d-electrons. Static correlations appear when the ground state (or an excited state of interest) exhibits a multi-reference character, i.e. it cannot be represented quantitatively by a single Slater determinant (see Benavides-Riveros et al.18 and references therein for further details). Such correlation effects are present in systems with electron (quasi-)degeneracy, which are typical for the d-states of transition metals in highly symmetrical crystal fields,19 either ideal or slightly distorted. These strongly correlated materials pose a significant challenge to computational research20,21 as the electronic structure of d-multiplets cannot be adequately captured by standard density functional theory22,23 (DFT) methods, which are commonly employed to investigate the electronic structures of solid-state systems. Limitations of DFT are most pronounced when calculating the d–d excitations and band gaps of strongly-correlated systems. Advanced methods based on hybrid density-functional approximations can yield better results than standard semi-local functionals,24 however these solutions are not universal and have limitations on the types of excitations that can be accurately predicted.25
This highlights the pressing requirement to search beyond DFT for more reliable methods for electronic structure calculations. One of the most direct ways for addressing strong correlations involves the use of post-Hartree–Fock approaches, such as complete active space self-consistent field (CASSCF),26 second-order Møller–Plesset perturbation theory27 and coupled cluster methods.28 While examples of applying post-Hartree–Fock treatment to solid-state materials exist in the literature,29–31 the range of applications is limited due to a strong scaling with system size. CASSCF has occasionally been used to calculate the energies of local d–d excitations in small (finite) clusters that represent the first coordination sphere of a transition metal atom in a crystalline system.14 This approach is, however, associated with significant underestimation of the delocalization effects taking place in extended systems. An alternative to post-Hartree–Fock approaches is to combine multiple electronic structure methods within a hybrid embedding approach, whereby the electronic structure of localized d-shells is treated using a correlated method and the host is described within a weakly correlated (e.g. one-electron self-consistent field) treatment. Such a hybrid approach allows one to reduce computational costs while providing a multi-reference description of strongly correlated d-electrons. One of the examples of hybrid electronic structure methods used in solid state theory is dynamical mean-field theory.32
In our previous work33 we extended a hybrid embedding method, called the effective Hamiltonian of crystal field (EHCF),34,35 to periodic systems (pEHCF). pEHCF has been shown to be successful in describing the d-multiplet structure of various solid materials including oxides,33 carbodiimides and metal–organic frameworks.36,37 In this work, we evaluate and compare the suitability of DFT and pEHCF to investigate the electronic structure and d–d excitations of Cr3+ dopants in three wide band gap host materials: corundum (α-Al2O3), aluminium oxonitridoborate (AlB4O6N)38 and chrysoberyl (BeAl2O4).
α-Al2O3 with Cr3+ dopants is a well-known laser material15,39 with excellent optical properties and distinctive fluorescence which has been also explored for quantum applications.13 AlB4O6N recently synthesized by Widmann et al.38 possesses interesting fluorescence and luminescence properties as well as high thermal stability. Finally, BeAl2O4 has been widely utilized in solid-state laser technologies due to its exceptional emission properties in the 700–800 nm range.40 All three host materials show characteristics that are promising for quantum technologies which will benefit from deeper understanding of the complex electronic structure of transition metal dopants in solids provided in this study.
| Ψ = Ψd(nd) ∧ Ψl(N − nd) | (1) |
![]() | (2) |
The wavefunction in eqn (1) assumes the number of electrons in both d- and l-subsystems to be fixed and therefore excludes charge transfer states. The presence of such states is taken into account in pEHCF by the Löwdin partitioning technique,42 which provides effective corrections to the Hamiltonian operators of the subsystems arising due to the electron hopping between them.33,34 This results in the effective Hamiltonian for the d-subsystem,
, having the following form:
![]() | (3) |
includes the bare Hamiltonian of the d-subsystem
, and both Coulomb
and resonance
interactions of d-electrons with electrons and nuclei in the l-subsystem. Contributions from
and
determine the ‘splitting parameter’ of the d-orbitals, as referred to within the crystal field theory. In pEHCF, unlike in the crystal field theory, the main contribution to the splitting of d-orbitals is the resonance interactions33,34 corresponding to the one-electron transfers between the d- and l-subsystems. Therefore,
is the most important factor when analyzing variations in the splitting parameters during spin-crossover processes. The matrix elements of
have the following form:33,43
![]() | (4) |
![]() | (5) |
![]() | (6) |
In eqn (5) and (6), k is a vector in the first Brillouin zone, n enumerates bands of the l-subsystem, and εnk and fnk are energies and occupation numbers, respectively, of the l-bands. Spin variables are omitted for clarity. As can be seen, the resonance term (and therefore the splitting of the d-orbitals) depends on three main factors: the geometry of the first coordination sphere through the resonance integrals between local atomic orbitals (βμa), the occupations of local atomic orbitals forming the first coordination sphere, and the energy difference between d-states and the valence/conduction bands of the l-subsystem.
Solving the linear Schrödinger equation for the wavefunction of the d-system, as shown in eqn (2), with the effective Hamiltonian in eqn (3) produces the whole spectrum of energies for the d-multiplets with all allowed spins and symmetries, among them the ground state. pEHCF electronic structure calculations were performed for geometries obtained by DFT with use of the r2SCAN44 meta-generalized gradient approximation as described below.
c,54 P63mc38 and Pbnm,55 respectively. Their experimental crystal parameters along with the values calculated by r2SCAN and HSE06 are collected in Table 1. The calculated values are within 1% of the reported experimental data for all three structures. In all materials, the Cr3+ dopant substitutes an aluminium cation (Al3+), leading to the formation of six-coordinate dopant sites of various symmetries, as shown in Fig. 1. The geometries of the Cr3+ dopant in each host, obtained with use of r2SCAN and HSE06, are illustrated in Fig. 2 (all structures are given in SI). The dopant site in AlB4O6N is of high Oh symmetry with minimal distortion from the perfect octahedral coordination characterized by bond length deviation not exceeding 0.01 Å and bond angles deviating from the perfect 90° by 0.8°. α-Al2O3 also has a single dopant site that exhibits C3 symmetry and BeAl2O4 has two dopant sites characterized by Cs (Wyckoff position 4c) and Ci (Wyckoff position 4a) local symmetry point groups. In all cases, the geometries of the coordination spheres are close to that of a regular octahedron, with fairly small distortions resulting in symmetry lowering. In CrAl3+:α-Al2O3, the distortion is characterized by a maximum deviation of 0.05 Å in Cr–O bond lengths and a maximum deviation of 10° in bond angles, whereas in the case of CrAl3+:BeAl2O4 these values are 0.09 Å and 10°, respectively. Due to this, we use the notations of irreducible representations of the Oh point group to label electronic states in all three systems and separately discuss splittings of the high-symmetry multiplets caused by the imperfections of dopant sites. As follows from the Tanabe–Sugano diagram of the d3 configuration in an octahedral field, Cr3+ cations must always have a high-spin quartet (S = 3/2) ground state (4A2), and a set of low-spin quartet and doublet (S = 1/2) excited states, the order and energies of which depend on the interactions of the Cr3+ dopant with its host. Transitions between these states correspond to the class of crystal field d–d excitations and can be experimentally probed via ultraviolet-visible and photoluminescence spectroscopy. We further test the capabilities of DFT and pEHCF in reproducing the energies and spin-symmetries of excited d-multiplets of Cr3+ dopants.
![]() | ||
| Fig. 2 Ground-state geometries of the Cr coordination sphere and the splitting diagrams of the one-electron d-states for (a) α-Al2O3, (b) AlB4O6N, and BeAl2O4 at (c) Cs- and (d) Ci-sites. The corresponding splitting parameters of the one-electron d-states calculated with r2SCAN, HSE06, and pEHCF are presented in Table 2. | ||
First, we analyse the one-electron states of the three systems, as calculated using DFT and pEHCF methods. The atomic orbital-projected density of states (DOS) for the ground states of Cr3+-doped α-Al2O3, AlB4O6N, and BeAl2O4 are shown in Fig. 3. For all systems, as can be seen from the r2SCAN-calculated DOS, the narrow d-bands of Cr3+ lie within the gap between the wide sp valence and conduction bands of the host. The small width of the d-bands indicates a small degree of coupling between the d-shell and sp-states, supporting the assumption regarding the locality of d-shells that is employed within pEHCF. In the case of DFT, there is a minor contribution from the oxygen 2p states in the gap states reflecting a small degree of hybridization with the chromium 3d states. The pEHCF-calculated DOS plots qualitatively agree with DFT regarding the position of the one-electron d-states and the structure of the frontier sp-bands. The valence band of the sp-subsystem mostly comprises oxygen 2p orbitals for each host material, whereas the conduction band has significant contributions from aluminium 3p orbitals. If the concentration of Cr3+ dopants is sufficiently low, the gap between the valence and conduction sp-bands should be close to the band gap of the host material.
Our calculations show that for the given unit cells, the sp-band gaps of the doped materials are smaller than the band gaps of the pure hosts by 0.1–0.2 eV, which can be compared against corresponding experimental values. Experimental values of the band gap of α-Al2O3 have been reported to range from 8.15–9.40 eV depending on conditions,57 while the reported experimental value for BeAl2O4 is 9.00 eV.58 The calculated values of the band gap are compared to the available experimental data in Table 2. It shows that pEHCF systematically overestimates the gap by 0.50–1.50 eV for CrAl3+:α-Al2O3 and by 2.00 eV for CrAl3+:BeAl2O4 as the Hartree–Fock method is used to calculate the electronic structure of the sp-subsystem. In contrast, r2SCAN underestimates the band gap by about 1.5 eV for both materials whilst HSE06 gives the improved values of 8.15 eV for α-Al2O3 and 8.38 eV for BeAl2O4. No experimental band gap value has been reported for AlB4O6N, but we calculate a band gap of 9.78 eV using HSE06. Other reported theoretical band gap values range from 7.32–9.31 eV, depending on the computational method.38
| Host → | α-Al2O3 | AlB4O6N | BeAl2O4 (Cs) | BeAl2O4 (Ci) | ||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| pEHCF | r2SCAN | HSE06 | pEHCF | r2SCAN | HSE06 | pEHCF | r2SCAN | HSE06 | pEHCF | r2SCAN | HSE06 | |
| t2g [eV] | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 |
| 0.00 | 0.00 | 0.00 | 0.06 | 0.05 | 0.11 | 0.19 | 0.05 | 0.05 | 0.14 | 0.09 | 0.08 | |
| 0.03 | 0.10 | 0.10 | 0.10 | 0.05 | 0.11 | 0.30 | 0.07 | 0.09 | 0.37 | 0.19 | 0.15 | |
| eg [eV] | 2.08 | 4.04 | 6.21 | 2.40 | 4.15 | 6.32 | 2.50 | 3.92 | 6.11 | 2.95 | 4.29 | 6.47 |
| 2.15 | 4.04 | 6.21 | 2.55 | 4.15 | 6.32 | 2.69 | 4.06 | 6.26 | 3.17 | 4.40 | 6.59 | |
| Espgap [eV] | 9.86 | 7.22 | 8.21 | 13.45 | 7.99 | 9.72 | 11.71 | 7.45 | 8.45 | 11.77 | 7.45 | 8.42 |
| Exp. Espgap [eV] | 8.15–9.4057 | — | 9.0058 | 9.0058 | ||||||||
Splitting diagrams for the one-electron d-states and the ground state, as calculated with pEHCF, r2SCAN and HSE06, are shown in Fig. 2. All methods qualitatively follow the same symmetry considerations, exhibiting a significant t2g–eg splitting characteristic of an octahedral crystal field. An additional minor splitting within the threefold degenerate t2g manifold is present in CrAl3+:α-Al2O3 (C3 symmetry), while full degeneracy lifting occurs for CrAl3+:BeAl2O4, which possesses lower-symmetry (Cs and Ci) dopant sites. Quantitatively, the absolute magnitudes of crystal-field splittings differ significantly between pEHCF and the DFT methods (Table 2). For instance, the r2SCAN splitting of t2g–eg has values of around 4.00–4.40 eV; HSE06 provides splittings of 6.10–6.60 eV, while pEHCF shows splittings of 2.08–3.47 eV. This disparity is to be expected due to the nature of the one-electron energies in both methods. pEHCF splitting diagrams correspond to the eigenvalues of the one-electron part of the effective Hamiltonian, whereas energy levels produced by DFT already include an effect of d–d electron interactions. From this point of view, pEHCF produces splitting parameters that are usually discussed in the literature related to spectroscopy of transition metal ions, such as 10Dq in the Oh crystal field.
Many-electron multiplet energies calculated using r2SCAN, HSE06, and pEHCF are detailed in Tables 3–5 for the three materials, where they are compared against experimental values. Theoretically, the simplest transition is 4A2 → 4T2, which corresponds to the promotion of one electron from the t2g-orbital to the eg-orbital. Both multiplets can be accurately described using a single determinant wavefunction;19 the effect of static correlations should therefore be minor. r2SCAN and HSE06 accurately reproduce the energy of this transition for all materials with absolute errors compared to experiment ranging between 0.10–0.15 eV, and pEHCF also provides the same level of accuracy. As shown in Table 3, CASSCF calculations previously performed14 for a finite [CrO6]9− cluster cut out from Cr3+:Al2O3 give slightly larger errors for the 4A2 → 4T2 transition as compared to DFT and pEHCF.
| Transition | Experiment | pEHCF | r2SCAN | HSE06 | CASSCF |
|---|---|---|---|---|---|
| 4A2 → 2E | 1.78;60 1.8061 | 1.90 | 1.21 | 1.14 | 1.7814 |
| 4A2 → 2T1 | Not resolved | 1.97; 2.00 | — | — | — |
| 4A2 → 4T2 | 2.22;60,61 2.25;62 2.2860 | 2.08; 2.14 | 2.37 | 2.36 | 2.4914 |
| 4A2 → 2T2 | Not resolved | 2.82; 2.83; 2.90 | — | — | — |
| 4A2 → 4T1 | 3.01–3.03;60–62 3.1260 | 2.94; 2.97; 3.11 | 3.31 | 3.05 | — |
| Transition | Experiment | pEHCF | r2SCAN | HSE06 |
|---|---|---|---|---|
| 4A2 → 2E | 1.81 | 1.91 | 1.25 | 1.12 |
| 4A2 → 2T1 | 1.89 | 1.99; 2.01 | — | — |
| 4A2 → 4T2 | 2.43 | 2.40; 2.42; 2.44 | 2.46 | 2.49 |
| 4A2 → 2T2 | 2.70 | 2.89; 2.92 | — | — |
| 4A2 → 4T1 | 3.26 | 3.30; 3.32 | 3.49 | 3.46 |
| Transition | Experiment | Cs site | Ci site | ||||
|---|---|---|---|---|---|---|---|
| pEHCF | r2SCAN | HSE06 | pEHCF | r2SCAN | HSE06 | ||
| 4A2 → 2E | 1.82 | 1.89; 1.90 | 1.27 | 1.17 | 1.80; 1.89 | 1.22 | 1.15 |
| 4A2 → 2T1 | 1.91 | 1.95; 1.99; 2.00 | — | — | 1.91; 2.01; 2.04 | — | — |
| 4A2 → 4T2 | 2.10 | 2.34; 2.41; 2.52 | 2.18 | 2.16 | 2.72; 2.91; 2.95 | 2.47 | 2.49 |
| 4A2 → 2T2 | — | 2.91; 2.95; 2.96 | — | — | 3.00; 3.04; 3.09 | — | — |
| 4A2 → 4T1 | 3.02 | 3.23; 3.37 | 2.98 | 2.72 | 3.64 | 3.16 | 3.02 |
Other excited states, such as 2E, have significant multi-reference character, and correct description of their electronic structure is therefore much more challenging for DFT. As has been previously shown with ΔSCF, the low-spin excited states may not be achievable or could be significantly underestimated with respect to experimental values.59 This is clearly illustrated by the results for the 2E multiplet that presents an interest for Cr3+-based spin qubits as it plays an important role in the inter-system crossing pathway.59 Both r2SCAN and HSE06 consistently underestimate the energy of the 2E state as compared to experiments, with absolute errors ranging between 0.50–0.60 eV. At the same time, pEHCF provides good accuracy for the 4A2 → 2E lines for all materials, with absolute errors ranging between 0.07–0.11 eV. The 2T1 and 2T2 states are even more complicated, as their energies cannot be calculated using ΔSCF at all due to their multi-reference nature. This is because their many-determinant wavefunctions cannot be approximated by the one-electron population matrix that is used to set up a trial Kohn–Sham wavefunction in ΔSCF. Full configuration interaction treatment of the d-shell, as implemented in pEHCF, permits the 2T1 and 2T2 wavefunctions to be determined and results in energies that are in good agreement with experimental data for all systems.
Regarding the second excited quartet state, 4T1, all methods agree well with experimental data. For CrAl3+:α-Al2O3, pEHCF successfully captures the reported experimental values ranging from 3.01–3.12 eV.60–62 r2SCAN and HSE06 give excitation energies of 3.31 eV and 3.05 eV, respectively. For CrAl3+:AlB4O6N, the second excited quartet is calculated to be 3.30 eV in pEHCF and 3.46–3.49 eV by DFT methods, which can be compared to the experimental value of 3.26 eV.38 Finally, for CrAl3+:BeAl2O4, the experimentally reported value of 3.02 eV63 agrees well with our r2SCAN-calculated energy of 2.98 eV for the Cs-symmetrized site and our HSE06-calculated energy of 3.02 eV for the Ci-symmetrized site. In contrast, pEHCF overestimates the energy of 4A2 → 4T1 transition by 0.20 eV and 0.10 eV for the Cs- and Ci-symmetrized sites, respectively.
We also note an interesting discrepancy with experimental data for CrAl3+:BeAl2O4 at the Ci dopant site. As shown in Table 5, the pEHCF-, r2SCAN-, and HSE06-calculated energies for the 4A2 → 4T2 transition are in good agreement with the experimentally observed line for the Cs dopant site. However, for the Ci dopant site, the 4A2 → 4T2 transition is significantly overestimated by all computational methods: by around 0.6–0.9 eV with pEHCF and 0.4 eV with DFT. This indicates that the Cs dopant site might be largely responsible for the experimental emission, whereas the Ci dopant site plays a minor role. This conclusion is further supported by our DFT calculations showing that Cr3+ in the Cs-symmetrized site lies lower in energy than the Ci-symmetrized site by 0.17 eV and 0.19 eV for r2SCAN and HSE06, respectively. A wide range of experimental studies confirm the preference of the Cs site. For instance, electron paramagnetic resonance and optical absorption spectroscopic analyses of BeAl2O4 indicate that 75% of Cr3+ ions substitute the Cs-symmetrized site.64 Using X-ray absorption spectroscopy, Bordage et al.65 also confirmed Cr3+ substitution at the Cs-symmetrized site to be 70%. These experimental findings are consistent with our computational results suggesting that the optical features observed experimentally predominantly correspond to Cr3+ substitution at the energetically and structurally favoured Cs-symmetrized site.
Our results show that, for the excited multiplets of Cr3+, r2SCAN and HSE06 functionals consistently give close energy values and perform particularly well for the high-spin states. Previous work on NV-like defects has shown that both hybrid density-functional approximation and meta-GGA can yield reliable predictions for the formation energy and charge transition levels.66 However, as shown in ref. 67, HSE06 may fail for transition metal containing systems due to the Coulombic interactions between localised d-electrons not being sufficiently screened. This results in an incomplete removal of self-interaction error and violation of the generalized Koopmans' condition. This makes r2SCAN a competitive, lower-cost alternative to hybrid density-functional approximations for systems containing transition metal atoms, particularly when combined with the pEHCF multi-reference treatment of the d-shell.
| This journal is © The Royal Society of Chemistry 2026 |