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

Computing L- and M-edge spectra using the DFT/CIS method with spin–orbit coupling

Aniket Mandal and John M. Herbert*
Department of Chemistry & Biochemistry, The Ohio State University, Columbus, Ohio 43210, USA. E-mail: herbert@chemistry.ohio-state.edu

Received 2nd May 2025 , Accepted 11th July 2025

First published on 11th July 2025


Abstract

Modeling L-edge spectra at X-ray wavelengths requires consideration of spin–orbit splitting of the 2p orbitals. We introduce a low-cost tool to compute core-level spectra that combines a spin–orbit mean-field description of the Breit–Pauli Hamiltonian with nonrelativistic excited states computed using the semi-empirical density-functional theory configuration-interaction singles (DFT/CIS) method, within the state-interaction approach. Our version of DFT/CIS was introduced recently for K-edge spectra and includes a semi-empirical correction to the core orbital energies, significantly reducing ad hoc shifts that are typically required when time-dependent (TD-)DFT is applied to core-level excitations. In combination with the core/valence separation approximation and spin–orbit couplings, the DFT/CIS method affords semiquantitative L-edge spectra at CIS cost. Spin–orbit coupling has a qualitative effect on the spectra, as demonstrated for a variety of 3d transition metal systems and main-group compounds. The use of different active orbital spaces helps to facilitate spectral assignments. We find that spin–orbit splitting has a negligible effect on M-edge spectra for 3d transition metal species.


1 Introduction

Near-edge X-ray absorption spectroscopy (XAS) is a powerful technique for investigating the electronic structure of molecules and materials, exhibiting atomic and even oxidation-state specificity. K-edge XAS corresponds to excitations originating from elemental 1s orbitals, the localized nature of which means that XAS serves as a reporter on the valence virtual orbitals.1–4 This can be used to probe oxidation states5 and orbital covalency6–8 in transition-metal bonding. L- and M-edge spectra originate with transitions from 2p and 3p orbitals, respectively, that maintain elemental oxidation-state specificity.8–17 This is especially useful for analyzing the electronic structure of transition metal compounds.6,18,19 Probing the valence virtual orbitals enables determination of the local symmetry around a transition metal center,20–23 and can serve as an indicator for strong correlation effects in the 3d orbitals.24

In moving beyond the standard technique of multiplet ligand-field theory,10,25–27 the biggest obstacle to atomistic ab initio simulation of core-level spectra is the high cost of benchmark-quality quantum chemistry methods.3,28–32 Time-dependent density functional theory (TD-DFT) can be applied to much larger systems,33–37 and for X-ray spectroscopy it often affords accurate results for peak splittings.38 However, substantial absolute shifts are typically required in order to match experiment (e.g., ∼20 eV for the L-edge of 3d transition metals),39 even if these shifts amount to only a few percent of the core-to-valence excitation energies. These shifts are primarily artifacts of differential self-interaction error (SIE) between compact core and delocalized valence orbitals,40–42 as evidenced by improved results as the fraction of exact exchange is increased.39,43–45 SIE artifacts are also present in the “real-time” approach to TD-DFT,35 which is also subject to contamination by continuum states.35,37,46–48

To circumvent the need for large empirical shifts, specialized functionals have been developed for core-level spectroscopy,43,49–52 and modified forms of linear response theory have also been suggested.53,54 Alternatively, we have reported a new implementation55 of the so-called DFT/CIS method.56 This approach uses the formalism of configuration interaction with single substitutions (CIS) but incorporates molecular orbitals (MOs) and one-electron energy levels from Kohn–Sham DFT. A few empirical parameters are used to avoid double-counting of electron correlation effects.55,56 Our implementation extends the original DFT/CIS method to core-level spectroscopy using the core/valence separation (CVS) approximation.30,35 An empirical shift of the 1s orbital energies, based on atomic calculations with scalar relativistic corrections, is used to reduce differential SIE effects, with a concomitant reduction in the magnitude of the empirical shifts that are required to reproduce experimental spectra.55

In comparison to K-edge XAS, L-edge spectroscopy has several desirable characteristics including greater intensity because the most interesting transitions are dipole-allowed (e.g., 2p → 3d),57 whereas 1s → 3d transitions that also probe metal–ligand covalency appear as relatively weak pre-edge features58–60 that may be better probed by K-shell fluorescence spectroscopy.58 L-edge linewidths are reduced due to longer core-hole lifetimes, as compared to K-edge transitions,10 making L-edge XAS a formidable experimental tool. M-edge spectroscopy at extreme ultraviolet (XUV) wavelengths shares the same selection rules and is now accessible via tabletop instruments61–63 with ultrafast time resolution,13–16,61–68 providing a sensitive probe of transition metal spin and oxidation states.69

Theoretical simulation of L- and M-edge spectra is challenging due to spin–orbit coupling (SOC) that splits the 2p and 3p orbitals into states with J = 1/2 or J = 3/2.39,70 The present work extends our new DFT/CIS method to include SOC effects, without further parameterization. Results are presented for a variety of 3d transition metal systems and some main-group compounds, indicating favorable agreement with experiment using only modest empirical shifts.

2 Theory

2.1 DFT/CIS method

The DFT/CIS formalism is similar to that of the CIS method,37,71 with a few modifications necessitated by the use of Kohn–Sham orbitals. We use ψi, ψj, … to denote occupied MOs and ψa, ψb, … for virtual MOs, with indices r, s, … referring to arbitrary (occupied or virtual) MOs. We take these to be eigenfunctions of the Fock operator,
 
[F with combining circumflex]ψr = εrψr, (1)
with orbital energy levels εr.

Let |Ψ0〉 denote the reference determinant that solves the ground-state self-consistent field (SCF) problem, with ground-state energy E0. Let |Ψai〉 indicate a singly substituted Slater determinant with respect to that reference state. Using a closed-shell formalism for simplicity, diagonal matrix elements of the Hamiltonian in the singly substituted basis are71

 
Ψai|Ĥ|Ψai〉 = E0 + εaεi + 2(ψiψa|ψiψa) − (ψiψi|ψaψa). (2)
The off-diagonal elements are
 
Ψai|Ĥ|Ψbj〉 = 2(ψiψa|ψjψb) − (ψiψj|ψaψb). (3)
Two-electron integrals in these equations are expressed using Mulliken notation.71 In the CIS method, Brillouin's theorem ensures that
 
Ψ0|Ĥ|Ψai〉 = 0. (4)

Grimme introduced the first version of DFT/CIS,56 based on MOs obtained using the B3LYP functional. For that reason, we refer to Grimme's original method as “B3LYP/CIS”. In attempting to adapt it for core-level spectroscopy, we found that the parameterization was not appropriate for elements beyond the second row of the periodic table (i.e., beyond Ne). As such, we introduced a new parameterization based on the CAM-B3LYP functional,72 a range-separated hybrid that is better suited to TD-DFT applications because it mitigates severe underestimation of Rydberg and other charge-transfer transitions.72–76 We refer to this new parameterization as “CAM-B3LYP/CIS”.55 It was originally parameterized for use with the def2-TZVPD basis set,77 although the results are not strongly sensitive to that choice and smaller basis sets can be used.55

For CAM-B3LYP/CIS we leave eqn (3) and (4) unchanged, following Grimme's prescription,56 even though Brillouin's theorem is not generally satisfied by Kohn–Sham MOs. However, we modify the diagonal matrix elements in eqn (2) to obtain

 
Ψai|Ĥ|Ψai〉 = E0 + εaεi + 2c2Jiac1Kia − Δεi (5)
where c1, c2, and Δεi are empirical parameters (see below) and
 
Jia = (ψiψa|ψiψa) (6a)
 
Kia = (ψiψi|ψaψa). (6b)
These expressions are intended to be used with Kohn–Sham orbitals {ψr} and energy levels {εr}. Unlike the analogous TD-DFT linear-response equations,37 there is no exchange–correlation kernel in the matrix elements defined in eqn (5). Recent work has shown that the effect of this kernel is negligible in core-level TD-DFT calculations.78 Even for valence excitations, the parameterization of DFT/CIS means that it works well without this kernel.55,56

Parameters c1 = 0.525 and c2 = 0.850 in eqn (5) were determined55 using a subset of QuestDB, a benchmark data set of excitation energies for small and medium-size molecules.79 The correction Δεi in eqn (5) is an overall shift, which is different for the first two rows of the periodic table than it is for heavier atoms. Specifically,55

 
image file: d5cp01656h-t1.tif(7)
(Note that εi < 0.) This was determined in order to shift CAM-B3LYP 1s orbital energies {εi} into agreement with those obtained using the short-range corrected (SRC) functional known as SRC1.43 That functional was specifically parameterized to obtain K-edge spectra with experimental accuracy and uses a large fraction of exact exchange on a very short length scale (<1 Å), in order to reduce SIE for the core states. The correction Δεi reduces the inherent errors in core-level eigenvalues obtained from CAM-B3LYP and represents an alternative to shifting spectra to match experiment, as would be required in a XAS calculation using TD-CAM-B3LYP. Absent the Δεi correction, 1s energy levels from CAM-B3LYP are less strongly bound as compared to the corresponding SRC1 values and this discrepancy is adjusted via eqn (7).

For elements H through Ne, corresponding to |εi| ≤ 102Eh, the shift introduced in eqn (7) has essentially the same numerical value as the shift used by Grimme in B3LYP/CIS,56 despite a rather different parameterization and functional form.55 This agreement reflects that fact that core-level eigenvalues from both B3LYP and CAM-B3LYP underestimate excitation energies for XAS by approximately the same amount. However, the CAM-B3LYP/CIS parameterization works for third-row elements (|εi| > 102Eh), whereas Grimme's parameterization does not.

Our choice for the modifications in eqn (5) is motivated by the observation that the majority of the correction in B3LYP/CIS comes from the c1 parameter. Grimme introduced an alternative form for the shift in 〈Ψai|Ĥ|Ψai〉 that involves Coulomb integrals and is more difficult to evaluate.55 This correction was intended to better describe Rydberg and charge-transfer states where the exchange integral Kia is smaller than it is for localized valence excitations. However, the use of a range-separated hybrid functional means that CAM-B3LYP/CIS is inherently better equipped to deal with those types of transitions. Thus, we simplify the form of the empirical shift (Δεi) by modernizing the DFT functional.

2.2 Spin–orbit coupling

Relativistic effects are present in all core-level spectra but in the context of electronic structure calculations they can be split into two categories: scalar and spin–orbit.80 The latter do not play a significant role in K-edge spectroscopy because the 1s orbitals are energetically isolated. So long as peak splittings and other spectral features are accurate, K-edge spectra can simply be shifted to account for relativistic effects, using scalar relativistic corrections for the isolated atoms that can be computed, once and for all, across the periodic table.81–83

This is not the case for L-edge spectra, however, where SOC splits the triply degenerate 2p orbitals into one 2p1/2 orbital and two 2p3/2 orbitals. SOC leads to mixing of spin–orbit-free states with different multiplicities, resulting in two peaks at the L-edge: the L2 peak (originating from the 2p1/2 orbital) and the L3 peak (from the 2p3/2 orbitals). An analogous splitting of the 3p orbitals is observable in M-edge spectroscopy.

SOC effects can be incorporated using either variational or perturbative methods.80 Variational treatments include the zeroth-order regular approximation (ZORA),84 the Douglass–Kroll–Hess (DKH) method,85 and the exact two-component (X2C) approach.86 These methods incorporate spin–orbit operators into the wave function optimization, which is the formally more rigorous approach and is necessary when SOC effects become sufficiently large. However, variational treatment of SOC adds significantly to the complexity of the formalism and increases the computational cost by perhaps an order of magnitude, relative to a nonrelativistic calculation.87 Perturbative treatment of SOC is more affordable and is sufficiently accurate for the first few rows of the periodic table.88 This approach, which is the one adopted here, involves construction of the Breit–Pauli Hamiltonian using nonrelativistic (spin–orbit-free) wave functions as basis states.

2.2.1 Breit–Pauli Hamiltonian. We incorporate SOC as off-diagonal coupling matrix elements between eigenstates of the nonrelativistic (Coulomb electronic) Hamiltonian, in what has been called the “state-interaction” approach.89 In the present work, the nonrelativistic basis states are obtained from DFT/CIS calculations that include a large number of singlet and triplet core-excited states. We denote the DFT/CIS or TD-DFT state energies as E(Sn) for the singlet state Sn and E(T(M)n) for the triplet state Tn, with multiplet components (M ∈ {−1, 0, +1}) that are degenerate in the DFT/CIS calculations:
 
E(T(−1)n) = E(T(0)n) = E(T(+1)n). (8)

Within the state-interaction picture, the Breit–Pauli Hamiltonian incorporates SOC between the aforementioned spin–orbit-free basis states, and can be written in matrix form as

 
image file: d5cp01656h-t2.tif(9)
In principle, one might add scalar relativistic corrections to the diagonal matrix elements in HBP. For the L2,3-edges of silicon and sulfur, such corrections are smaller than 0.05 eV.81 For 3d transition metals considered here, these corrections range from 0.6–2.1 eV,81 which is small compared to shifts that are applied to match experiment.

Off-diagonal matrix elements V(T(M)n,Sm), containing the SOC interaction between spin–orbit-free states T(M)n and Sm, are evaluated here using the spin–orbit mean-field (SOMF) approximation.90,91 These couplings are matrix elements of the Breit–Pauli spin–orbit operator,32,90–95

 
image file: d5cp01656h-t3.tif(10)
Here, α is the fine-structure constant and ŝk is the spin angular momentum for electron k. The one-electron operator ĥssok in eqn (10) is the “spin–same-orbit” coupling, given by
 
image file: d5cp01656h-t4.tif(11)
where rk and RK are the coordinates of electron k and nucleus K, respectively, with the latter having atomic number ZK. One may recognize (rkRK) × pk in eqn (11) as the orbital angular momentum about nucleus K. The two-electron operator ĥsook,l in eqn (10) represents “spin–other-orbit” coupling,
 
image file: d5cp01656h-t5.tif(12)
This form invokes the SOMF approximation,32,90–95 which simplifies calculation of the coupling matrix elements insofar as explicit two-electron integrals are not required. Within this approximation, matrix elements of ĤSO can be evaluated using only one-electron integrals and transition density matrices.95

The formalism introduced above can be used with either DFT/CIS or TD-DFT.95 Matrix elements image file: d5cp01656h-t6.tif in the basis of nonrelativistic states afford the off-diagonal terms in the matrix HBP. In the present work, these states are either singlets (S = 0 = MS) or triplets (S = 1 and MS ∈ {−1, 0, +1}). Diagonalization of HBP in eqn (9) affords the coupled target states.

2.2.2 Oscillator strengths. To compute spectra we need oscillator strengths for the coupled states. We start from the transition dipole moments in the nonrelativistic basis, which have the form
 
image file: d5cp01656h-t7.tif(13)
where α ∈ {x, y, z} and [small mu, Greek, circumflex]α = e[small alpha, Greek, circumflex] is the α component of the dipole moment operator. Suppressing the spin quantum numbers (S, MS) for brevity, the states |Ψ〉 and |Ψ′〉 in eqn (13) range over the SCF ground state |Ψ0〉 and all of the DFT/CIS or TD-DFT excited states. Each of the excited states is a CIS-style linear combination of singly substituted Slater determinants,
 
image file: d5cp01656h-t8.tif(14)
This is reminiscent of the Tamm–Dancoff approximation (TDA) in TD-DFT,37,96 which is implicit in the DFT/CIS approach. The requisite matrix elements in eqn (13) are evaluated using the Slater–Condon rules, in what has been called a “pseudo-wave function” approach to computing transition dipole moments between TD-DFT excited states.97–100 This treatment is consistent with how the Breit–Pauli Hamiltonian has long been used in TD-DFT calculations.101–103 Formally speaking, state-to-state transition dipoles require quadratic rather than linear response theory,100,104 but the pseudo-wave function method is consistent with invocation of the TDA.

The nonrelativistic transition dipole moments in eqn (13) vanish if SS′, and they are identical for all values of MS within a given spin multiplet. However, we retain the indices MS and image file: d5cp01656h-t9.tif in eqn (13) to clarify that the dimension of the matrix μα matches the dimension of HBP. The former can then be transformed into the coupled SOMF basis, using the matrix UBP that diagonalizes HBP:

 
[small mu, Greek, tilde]α = UBPμαUBP. (15)
The matrix [small mu, Greek, tilde]α contains the α components of the transition dipole moments between SOC states, including the ground state. The oscillator strength for |0〉 → |n〉 excitation, from the ground state to coupled state |n〉, is given by37
 
image file: d5cp01656h-t10.tif(16)

3 Computational details

3.1 Implementation

The DFT/CIS method has been implemented in the Q-Chem program,105 and is available starting from v. 6.2.55 SOC matrix elements have been implemented in Q-Chem by Krylov and co-workers;32,92–95 see ref. 95 for the TD-DFT implementation.

Core-level spectra without SOC are computed using either DFT/CIS or TD-DFT within the CVS approximation.35 Subsequently, in a post-processing step, these spin–orbit-free states are transformed to the coupled basis by diagonalizing HBP in eqn (9). For this, we have written a Python code called pySETSOC.106 This program extracts SOC terms from the Q-Chem output, along with state-to-state transition dipole moments that are computed within the pseudo-wave function approach. It then constructs and diagonalizes HBP and computes oscillator strengths for the coupled states. The pySETSOC program can be applied to Q-Chem outputs from TD-DFT or DFT/CIS, using the pseudo-wave function approximation to evaluate the requisite state-to-state transition dipole matrix elements as discussed in Section 2.2.2. The pySETSOC program will also read the output from restricted active-space (RAS)-CI methods,94,107 although that functionality is not used here.

Finally, spectra are plotted using either Lorentzian or Gaussian broadening. The former is probably more appropriate for core-level spectra that are subject to lifetime broadening arising from the core hole. In many cases, we use a Lorentzian function with a full width at half-maximum (FWHM) of 0.3 eV. Although chosen empirically, this is roughly consistent with 2p core-hole states having natural line widths of ∼0.1 eV.108,109

3.2 Core/valence separation

The CVS approximation extends standard excited-state methods to core-level states by eliminating the amplitudes xia [eqn (14)] associated with all but a small number of occupied MOs, ψi.33–35 This dramatically reduces computational cost by eliminating valence excited states, so that core-to-valence excitations appear as the lowest energy states in the spectrum. Furthermore, the CVS approximation decouples these states from the continuum, which is otherwise problematic (e.g., in real-time approaches).35,37 Finally, the CVS approximation enables separation of K-, L-, and M-edge spectra in cases where different elemental edges might overlap with one another. In such situations, peak assignments can be made by comparing DFT/CIS or TD-DFT spectra using different active-space approximations.

All calculations reported here invoke the CVS approximation, which affords negligible errors for K- and L-edge spectra.32,110 For the L-edge spectra presented here, the occupied orbital active space consists of the 2p orbitals for the element in question along with the entire virtual space. In molecules containing more than one atom with the same atomic number, we include all of the 2p orbitals for that element. M-edge spectra are computed using the analogous procedure with 3p orbitals in the active space. For most calculations, we use 200 singlet and 200 triplet (nonrelativistic) basis states to construct HBP.

Conventional TD-DFT calculations reported here also use the CVS approximation as well as the TDA. This is consistent with the CIS-style matrix elements used in DFT/CIS, and with the CIS-style pseudo-wave function method that is used to compute oscillator strengths. Previous work suggests that the TDA has a negligible effect on core-level spectra.35,111

3.3 Functionals and basis sets

DFT/CIS calculations were performed using CAM-B3LYP,72 for which the method was parameterized in previous work.55 The B3LYP-based parameterization introduced by Grimme is not used here, because CAM-B3LYP/CIS is more suitable for computing core-level spectra.

As a point of comparison, conventional TD-DFT calculations are reported using the functionals B3LYP, CAM-B3LYP,72 and SRC1-r2.33,43 The latter is specifically parameterized for K-edge spectra of third-row elements, using a large fraction of exact exchange (87%) on a very short length scale (≲1 Å). Although TD-SRC1 is quite accurate for K-edge spectra,33,35,43,52 it has not yet been tested for L-edge spectra.

CAM-B3LYP/CIS parameters were developed based on calculations using def2-TZVPD.77 That basis set should be sufficient to converge TD-DFT excitation energies,37,112–116 but in fact CAM-B3LYP/CIS provides satisfactory results in smaller basis sets also.55 Except where specified otherwise, the def2-TZVPD basis set is used for all calculations.

4 Results and discussion

4.1 Functional and basis-set effects

The effect of exact exchange on K-edge TD-DFT calculations is well documented,39,43–45 with larger fractions generally affording more accurate results in the sense that smaller absolute shifts are required. This is primarily due to larger SIE associated with the compact 1s orbital, as compared to the relatively delocalized valence virtual orbitals. Core-to-valence excitations can also be considered to manifest a certain type of charge transfer, whose excitation energy is dramatically underestimated by functionals that lack correct asymptotic behavior.37,117–119 Although the latter problem is ameliorated as the fraction of exact exchange is increased, this must be balanced against a semilocal correlation functional that is not designed for fully nonlocal Hartree–Fock exchange. Accuracy may be significantly degraded when the fraction of exact exchange rises above 50%.120 Besley and co-workers solved this problem by introducing SRC functionals with unusually large fractions of exact exchange (>50%) on very short length scales.43 The range-separation scheme and the fractions of exact exchange that are used in these SRC functionals were determined in order to match experimental K-edge excitation energies and these functionals have not previously been examined for L-edge spectra.

Fig. 1 presents conventional TD-DFT results for the Ti L2,3-edge spectrum of TiCl4 using several different functionals, in comparison to CAM-B3LYP/CIS calculations and to experiment.121 All calculations include SOC so that distinct L2 and L3 edges are apparent, separated by ∼12 eV. Conventional TD-B3LYP and TD-CAM-B3LYP calculations underestimate the 2p → valence excitation energies by ∼15 eV, while CAM-B3LYP/CIS and TD-SRC1-r2 underestimate them by only 5.0 eV (TD-SRC1-r2) or 6.8 eV (CAM-B3LYP/CIS). The fact that all four methods underestimate the excitation energies with respect to experiment suggests that differential SIE remains significant even for the 2p orbitals. Note that the core orbital correction Δεi that is used in CAM-B3LYP/CIS was parameterized using 1s orbital energies yet continues to work reasonably well for excitations from the 2p orbitals, as we observed previously.55


image file: d5cp01656h-f1.tif
Fig. 1 Direct relation between the accuracy of TD-DFT calculations and the fraction of exact exchange, for Ti L2,3-edge spectra of TiCl4. Shown are four different TD-DFT approaches along with CAM-B3LYP/CIS, in comparison to an experimental spectrum from ref. 121. All calculations used the def2-TZVPD basis set and include 200 singlet and 200 triplet basis states. Individual transitions were broadened using a Lorentzian function with FWHM = 0.3 eV.

Notably, relative intensities of L2 versus L3 are different between theory and experiment. The same artifact has been noted previously in real-time TD-DFT calculations of the same spectrum.122 Correct intensities are obtained using the ZORA method in conjunction with relativistic basis sets,123,124 pointing to a limitation of the present approach.

Fig. 2 documents basis-set effects on this TiCl4 spectrum. The position of both the L2 and L3 peaks is unchanged amongst the def2-TZVP, def2-TZVPD and aug-cc-pVTZ basis sets. A shoulder on the higher-energy side of the L2 peak becomes progressively more pronounced as one goes from def2-SVPD to def2-TZVPD to aug-cc-pVTZ, but is absent when diffuse functions are omitted. The minimally augmented def2-ma-TZVP basis set,125 which removes higher angular momentum diffuse functions from def2-TZVPD, exaggerates the peak splitting near 455 eV and predicts inconsistent intensities at higher energies, as compared to other basis sets. This suggests a role for the diffuse d functions that are present in def2-TZVPD and aug-cc-pVTZ, at least when it comes to lineshapes.


image file: d5cp01656h-f2.tif
Fig. 2 Basis set effects on the Ti L2,3-edge spectrum of TiCl4, computed at the CAM-B3LYP/CIS + SOC level as described in Fig. 1. Spectra are normalized separately so that the main feature (around 455 eV) has the same intensity in each case. Vertical dashed lines are guides to the eye, aligned with the most intense L2 and L3 features at the top and at the bottom.

Peak excitation energies are a different matter. Using double-ζ rather than triple-ζ basis sets introduces a shift of about 1 eV in both the L2 and L3 features but the L2,3 splitting is essentially identically in all of the basis sets tested, as documented in Table 1. In our view, 1 eV shifts in the absolute excitation energies are of no consequence at X-ray wavelengths, where the empirical shifts required to match experiment are 5–7 eV at a minimum. Thus, smaller basis seem to suffice for excitation energies, even while converged lineshapes require diffuse functions.

Table 1 Basis set effects for CAM-B3LYP/CIS + SOC calculations at the Ti L2,3-edge of TiCl4
Basis Exc. energy (eV) Splitting (eV)
L3 L2
def2-SVP 444.2 455.2 11.0
def2-SVPD 444.2 455.3 11.0
def2-TZVP 443.4 454.4 11.0
def2-ma-TZVP 443.4 454.4 11.0
def2-TZVPD 443.4 454.4 11.0
aug-cc-pVTZ 443.3 454.4 11.1


4.2 Benchmarking L2,3 energies

Fig. 3 shows experimental X(2p) → valence excitation spectra for TiCl4,121 SiCl4,126 and CrO2Cl2,127 at the L2,3-edge of the central atom (X = Ti, Si, or Cr), along with CAM-B3LYP/CIS + SOC calculations. Splitting of the L2,3 features is rather dramatic for TiCl4 (Fig. 3a), highlighting the fact that a spin–orbit-free spectrum contains only one bright feature and is qualitatively incorrect. Additional features in the L2,3 spectrum correspond to splitting of the 3d orbitals into t2 and e sets. Even with SOC, however, intensities of the L2 and L3 peaks are inverted relative to experiment, as noted in Section 4.1.
image file: d5cp01656h-f3.tif
Fig. 3 Experimental L2,3-edge spectra for (a) Ti in TiCl4,121 (b) Si in SiCl4,126 and (c) Cr in CrO2Cl2,127 in comparison to CAM-B3LYP/CIS + SOC calculations. The latter employ the def2-TZVPD basis set with 200 singlet and 200 triplet excited states. Calculated transition energies were broadened using a Lorentzian function (FWHM = 0.3 eV) and shifted by (a) 5.0 eV, (b) 3.2 eV, and (c) 4.8 eV.

Spectra of SiCl4 at the Si L2,3-edge are more complicated (Fig. 3b). One explanation is that the Si(2p) orbitals are less localized as compared to Ti(2p) orbitals, and thus undergo more mixing with the ligand orbitals. Splitting of the Si(2p3/2) and Si(2p1/2) features is very small,126 hence the difference between SOC and nonrelativistic spectra is not as striking in this case as it was for TiCl4. For CrO2Cl2 (Fig. 3c), inclusion of SOC is once again necessary to obtain a spectrum that is even qualitatively correct. The L2 and L3 edges are well separated, as they were in TiCl4, and the low symmetry C2v environment leads to distinctive peak shapes.

Overall, the Breit–Pauli formalism, combined with the SOMF approximation, works rather well for reproducing the L-edge spectra of these third-row elements. Spectral features contain important information about the nature of the valence virtual orbitals that can be connected to features in the experimental spectra by means of the DFT/CIS + SOC calculations.

Table 2 compares the accuracy of CAM-B3LYP/CIS and conventional TD-DFT, with SOC included in either case and compared to experimental transition energies.121,126–131 CIS + SOC results are also tabulated, in order to set a baseline. These L-edge results follow the same trend as the K-edge values that we reported previously.55 Specifically, both the TD-B3LYP and TD-CAM-B3LYP methods underestimate the excitation energies in all cases. This indicates that range separation alone is insufficient to correct the excitation energies, as the errors originate in 2p eigenvalues that are insufficiently bound. The TD-SRC1-r2 method performs better, despite being parameterized for main-group K-edge spectra, because the large fraction of exact exchange at ultrashort range mitigates the eigenvalue problem to some extent. Finally, CAM-B3LYP/CIS + SOC exhibits a mean absolute error (MAE) that is very similar to TD-SRC1-r2 + SOC. This is expected since the Δεi correction in the former was based on SRC1-r2 orbital energies.

Table 2 Simulated L2,3 excitation energies in comparison to experimental valuesa
Molecule Edgeb Expt. (eV) Errors in computed values (eV)
CIS TD-DFT
Standardc DFT/CISd SRC1-r2 CAM-B3LYP B3LYP
a All calculations use def2-TZVPD.b L2 or L3 for the central atom.c Conventional CIS.d CAM-B3LYP/CIS.e Ref. 126.f Ref. 129.g Ref. 128.h Ref. 121.i Ref. 131.j Ref. 127.k Ref. 130.l Mean absolute error.
SiCl4 L3 104.3e 6.8 −3.2 −3.9 −5.3 −5.6
L2 104.9e 6.5 −2.6 −3.3 −4.6 −5.0
H2S L3 164.9f 7.9 −3.4 −1.7 −4.4 −5.0
L2 166.5f 8.7 −2.6 0.6 −2.2 −3.0
Ar L3 244.5g 12.0 −0.7 −1.4 −6.7 −6.7
L2 246.4g 14.2 0.5 1.0 −4.4 −4.5
TiCl4 L3 457.7h 6.7 −6.8 −5.0 −14.3 −14.5
L2 462.8h 13.0 −0.9 0.9 −8.5 −8.6
VO43− L3 518.2i 6.8 −7.9 −5.4 −16.0 −15.9
L2 525.1i 13.8 −1.5 1.1 −9.6 −9.2
CrO2Cl2 L3 577.2j 11.8 −3.4 −2.0 −13.9 −14.0
L2 585.9j 19.7 4.0 5.6 −6.6 −6.5
MnO4 L3 645.6i 9.5 −6.9 −5.3 −18.8 −18.8
L2 656.0i 18.6 2.0 3.9 −9.6 −9.7
Fe(Cp)2 L3 709.1k 1.5 −10.4 −7.9 −22.6 −22.3
L2 722.2k 11.8 −0.3 2.2 −12.5 −12.2
Mean error     10.6 −2.8 −1.3 −10.0 −10.1
MAEl     10.6 3.6 3.2 10.0 10.1


Mean errors in Table 2 provide an additional indication as to the source of the errors. The CIS method, for example, is SIE-free but lacks complete orbital relaxation and as a result the errors are strictly positive, with a mean error of 10.6 eV. On the other hand, the TD-B3LYP and TD-CAM-B3LYP errors are strictly negative, indicating that SIE dominates over orbital relaxation error. The TD-SRC1-r2 approach, which balances these errors (by virtue of its parameterization using experimental data),54 exhibits errors of either sign. The same is true of CAM-B3LYP/CIS.

Table 3 lists the L2,3 splitting [ΔE([thin space (1/6-em)]L2) − ΔE([thin space (1/6-em)]L3)] from Table 2, along with the error in this splitting as computed using each of the methods examined here. In many cases, the errors are of the same order of magnitude as the experimental splitting itself. Errors are also roughly independent of the theoretical method used, averaging about 5 eV but increasing in magnitude as the splitting itself increases. This means that TD-SRC1-r2 is no more accurate than other TD-DFT methods that weren’t parameterized to remove the need for absolute shifts, and furthermore the accuracy of CIS is quite similar. (This is not altogether unexpected, as it speaks to the rather small electron correlation effects on the SOC values, consistent with the accuracy of the SOMF approximation.90,132,133) While the origin of the discrepancies in L2,3 splittings with respect to experiment are unclear, it seems that they have more to do with the treatment of SOC than with the description of electron correlation.

Table 3 Errors in L2,3 splittings in comparison to experimental valuesa
Molecule Expt.b (eV) Error (eV)
CIS DFT/CISc TD-DFT
SRC1-r2 CAM-B3LYP B3LYP
a All calculations use def2-TZVPD.b L2,3 splitting for the central atom, equal to the difference between experimental L2 and L3 values from Table 2.c CAM-B3LYP/CIS.
SiCl4 0.6 0.3 0.6 0.6 0.7 0.6
H2S 1.6 0.8 0.8 2.3 2.2 2.0
Ar 1.9 2.2 1.2 2.4 2.3 2.2
TiCl4 5.1 6.3 5.9 5.9 5.8 5.9
VO43− 6.9 7.0 6.5 6.5 6.4 6.8
CrO2Cl2 8.7 7.9 7.4 7.6 7.3 7.5
MnO4 10.4 9.1 8.9 9.2 9.2 9.1
Fe(Cp)2 13.1 10.3 10.1 10.1 10.1 10.1
Mean   5.5 5.2 5.6 5.5 5.5


4.3 CVS as a diagnostic tool

The CVS approximation is needed to extend CIS-type methods to core-level excitations, because it removes valence excitations that appear at lower transition energies so that core-to-valence transitions can be obtained using reasonable subspace sizes in the iterative eigensolver.35 By performing several calculations with different active occupied orbitals, the CVS approximation also facilitates separation of the K-, L-, and M-edges of different elements. This is a useful feature because overlapping edges can otherwise make it difficult to assign peaks. The vanadate ion (VO43−) affords an example, insofar as the O(1s) K-edge and the V(2p) L-edge are quite close in energy as shown in Fig. 4.
image file: d5cp01656h-f4.tif
Fig. 4 XAS of VO43−, illustrating the important role of the CVS approximation in assigning spectral features. All calculations were performed at the CAM-B3LYP/CIS + SOC level using the def2-TZVPD basis set, including 100 singlet and 100 triplet states broadened with a Lorentzian function (FWHM = 0.3 eV). The calculations labeled V(2p) and O(1s) employ the indicated active occupied orbitals, whereas the “no CVS” calculation uses all MOs. For clarity, the three spectra are normalized separately.

For this small ion, it is possible to perform the calculations without invoking the CVS approximation by computing both valence and core excitations. Upon doing so, subtle dipole-forbidden pre-edge features appear whose origin might at first be unclear. For that matter, given the magnitude of the shifts that are needed to match experiment, it is not clear a priori which features in the “no CVS” spectrum in Fig. 4 even correspond to the oxygen K-edge. Using active-space approximations involving only the O(1s) or V(2p) orbitals, these features are readily assignable and it is clear that the oxygen K-edge appears at 529 eV. Moreover, the close resemblance of the V(2p) spectrum and the exact linear-response result (sans CVS approximation) suggests there is little mixing between O(1s) and V(2p) orbitals in the ground state of VO43−.

The CVS approximation can also be a helpful tool when the same element appears more than once in a molecule, in different coordination environments. An example is the tetrathionate ion, (S4O6)2−, whose terminal sulfur atoms have formal charges of +5 whereas the two interior sulfur atoms have formal charges of zero. Simulated K-edge spectra for this species are shown in Fig. 5 using three different CVS active spaces, consisting of either the S(1s) orbitals on the terminal sulfur atoms, or those on the interior sulfur atoms, or both. The latter spectrum is the true K-edge result but the other two spectra aid in assigning features. These spectra clearly demonstrate how the interior S0 atoms give rise to lower-energy XAS features as compared to the terminal S5+ atoms, which is expected due to the stronger electron–hole interaction when the formal charge is larger.


image file: d5cp01656h-f5.tif
Fig. 5 XAS of (S4O6)2− at the sulfur K-edge, illustrating how calculations with different active spaces can be used to assign spectral features. Spectra were computed using CAM-B3LYP/CIS + SOC with the def2-TZVPD basis set and Lorentzian broadening (FWHM = 0.3 eV). CVS active spaces contain S(1s) orbitals on the atoms indicated in the legend.

Note that this decomposition is not exact, in the sense that the sum of the spectra with limited active spaces need not equal the spectrum that is obtained with the union of these active spaces, which is evident from the feature at 2475 eV in Fig. 5. These discrepancies arise due to coupling between S(1s) orbitals on different atoms, when such coupling is available within the active space. Despite their approximate nature, however, such decompositions can be useful diagnostic tools to understand the origins of various spectral features.

Fig. 6 shows the analogous decomposition of the L-edge spectrum of (S4O6)2−. Here, we use the S(2p) orbitals on either the terminal or the interior sulfur atoms (or both) as the CVS active space, and these calculations once again demonstrate how the lower-energy features originate in transitions localized on the interior S0 atoms. The approximate nature of the decomposition is evident in some of the features above 165 eV in Fig. 6a. When a significantly larger number of nonrelativistic states is used, spectral features fill in at higher energies (Fig. 6b). The larger calculation is significantly more expensive because it requires computing 800 DFT/CIS singlet states, including state-to-state transition moments between them, then another 800 triplet states and SOC couplings between the singlet and triplet states. The total dimension of HBP in eqn (9) is 3200 states. Nevertheless, such a calculation is feasible at the DFT/CIS level using the high-quality def2-TZVPD basis set. At the same time, even an approximate decomposition using atom-specific active spaces can be a useful diagnostic and interpretative tool.


image file: d5cp01656h-f6.tif
Fig. 6 XAS of (S4O6)2− at the sulfur L-edge. All calculations were performed using CAM-B3LYP/CIS + SOC with def2-TZVPD, computing either (a) 200 or (b) 800 singlet and triplet excited states, which were then broadened using a Lorentzian function (FWHM = 0.2 eV). Spectra labeled S(terminal) and S(central) employ only the S(2p) orbitals on the indicated atoms, with a correspondingly smaller number of excited states, whereas S(all) uses all of the S(2p) orbitals.

4.4 Miscellaneous L-edge applications

We next consider a variety of applications of CAM-B3LYP/CIS + SOC to compute L-edge spectra of main-group compounds and transition metal complexes.
4.4.1 L-edge spectra of SF6. XAS of SF6 at the S(2p) L-edge is considered in Fig. 7, juxtaposing the CAM-B3LYP/CIS spectrum with an experimental one.134 Peaks labeled 1 and 2, in the region from 172–175 eV, are assigned as a1g symmetry and correspond to excitation of S(2p3/2) and S(2p1/2) electrons into the lowest unoccupied MO (LUMO). As compared to those features, peaks 3 and 4 (representing transitions into t2g orbitals) are much more intense, in both the computed and the experimental spectrum, although peak 3 has a doublet structure in the CAM-B3LYP/CIS calculation that is not evident in the experiment. Weaker doublet peaks 5 and 6, representing transitions into a second set of t2g orbitals, are not observed experimentally although they have been observed in real-time TD-DFT calculations.46 Finally, peaks 7 and 8 in the simulated spectrum correspond to transitions with eg symmetry. These features are narrow and intense whereas the corresponding region of the experimental spectrum is broad and diffuse.
image file: d5cp01656h-f7.tif
Fig. 7 L-edge XAS of SF6. The CAM-B3LYP/CIS + SOC spectrum (in blue) was obtained using the def2-TZVPD basis set with 200 singlet and 200 triplet states. Individual transition energies were shifted by 5 eV and broadened with a Lorentzian function (FWHM = 0.3 eV). The experimental spectrum (inset) is reproduced from ref. 134; copyright 1993 American Physical Society.

Some Rydberg states between 179–182 eV are highlighted in an inset to Fig. 7. These appear at slightly higher energies as compared to the Rydberg features identified in the experimental spectrum, from 177–180 eV and identified as a1g, eg and t2g transitions.134 Only two peaks can be discerned in the highlighted region of the CAM-B3LYP/CIS spectrum. Ruud and co-workers have identified the same Rydberg features in real-time TD-DFT calculations,46 which are found to be quite sensitive to basis-set effects. Since these features are not as important in the context of using L-edge spectroscopy to probe valence virtual orbitals, they are not discussed any further.

The DFT/CIS spectrum for SF6 in Fig. 7 shows doublet peak structure in all features, with an average splitting of ≈2.5 eV as compared to experimental splittings of ≈1.2 eV. Nominally, the intensity ratio ratio of the 2p3/2 to 2p1/2 features should be about 2[thin space (1/6-em)]:[thin space (1/6-em)]1 for each pair of peaks, corresponding to the 4[thin space (1/6-em)]:[thin space (1/6-em)]2 ratio of microstates. However, both experimental and ligand-field multiplet intensity ratios can deviate from this result in practice,124,135 and in the present example this ratio is inverted. The calculated (CAM-B3LYP/CIS + SOC) intensity ratio is 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 for the a1g features (peaks 1 and 2 in Fig. 7), as compared to an experimental ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1.6. For the t2g features (peaks 3 and 4), the calculations afford an intensity ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2.24 as compared to an experimental ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2.04. Previous work indicates that a significant exchange interaction between the core-excited electron and the valence electrons is responsible for this inversion in the 2p3/2[thin space (1/6-em)]:[thin space (1/6-em)]2p1/2 intensity ratio.134

4.4.2 3d0 complexes. Core-level spectra of tetrahedrally coordinated 3d0 metal oxide anions MO4n (M = Ti, V, Cr, Mn) are examined in Fig. 8. The 2p → 3d transition is dipole-allowed and the peak splitting affords insight into the electronic structure of the metal 3d band and the effect of the ligand field. The d0 nature of the ground state means that each ion has only one reasonable ground-state electron configuration, thus a single-reference method like DFT/CIS may be adequate.
image file: d5cp01656h-f8.tif
Fig. 8 XAS of 3d0 metal oxide ions, MO4n, including (a) spectra at the metal L-edge from CAM-B3LYP/CIS + SOC calculations, (b) the corresponding experimental L-edge spectra, and (c) oxygen K-edge spectra, overlaying theory (in translucent color) and experiment. Spectra are shifted to align the L3- or K-edge feature labeled “B”, as indicated by the dashed vertical lines. This places all spectra within the same energy window, consistent with the presentation in ref. 131. Simulated spectra were computed using def2-TZPVD including 200 singlet and 200 triplet basis states and broadened using a Lorentzian function (FWHM = 0.3 eV). Experimental spectra are reproduced from ref. 131.

Spectra at the metal L2,3 edge are shown in Fig. 8a and b. The most apparent feature is the magnitude of the spin–orbit splitting between the L3 and L2 edges, which decreases in the order MnO4 > CrO42− > VO43− > TiO44− in both CAM-B3LYP/CIS + SOC simulations and experiment. The splitting between experimental features labeled A and B (Fig. 8b) decreases in the same order. A similar trend is observed at the oxygen K-edge; see Fig. 8c. The biggest discrepancy between theory and experiment is an inversion of peak intensities between the L2 and L3 edges, as observed in previous calculations.123,124 This is similar to what we observed for TiCl4 (Fig. 1), where it appears that a variational description of SOC is required in order to obtain the correct relative intensities, as discussed in Section 4.1.

SOC increases with atomic number [eqn (11)] and this is reflected in both theory and experiment. However, these MO4n spectra exhibit ligand-field splitting (into e and t2g orbitals) at both the L3 and L2 edges,131 such that there are two sets of metal d orbitals available to a 3d0 transition metal. Oxygen K-edge spectra of these MO4n complexes show a splitting in their most intense features that correlates well with the L2,3 splitting,131 indicating that the transition from O(1s) either the e or t2g orbitals is dipole-allowed. Brydson et al.131 argue that there is considerable mixing between the O(2p) and M(3d) orbitals, along with some contribution from M(4p) and O(2s) orbitals, resulting in the nominal O(1s) → M(3d) transition becoming dipole-allowed. The increased splitting between the A and B features at the metal L3 edge has been attributed to the increasing X–O bond lengths and decreasing formal charge on the metal ion.131

4.4.3 Ligand effects at the Fe L-edge. Sensitivity of transition metal 3d orbital energies to ligand substitution is especially pronounced for Fe(CO)5 and Fe(Cp)2 (ferrocene),130,136 where “Cp” indicates the cyclopentadienyl anion, (C5H6). L2,3-edge spectra of Fe(CO)5 exhibit a splitting in both the L3 and L2 bands, with one peak that is much stronger than the other. This is characteristic of the metal 2p spectra of all near-covalent complexes.137 In Fig. 9a, we see that this feature is present in both the experimental spectra of Fe(CO)5 and in CAM-B3LYP/CIS + SOC simulations.
image file: d5cp01656h-f9.tif
Fig. 9 L2,3-edge spectra of (a) Fe(CO)5 and (b) Fe(Cp)2, comparing experiment (with an offset baseline) to CAM-B3LYP/CIS + SOC simulations. The calculations employ the def2-TZVPD basis set with 200 singlet and 200 triplet states, broadened using a Lorentzian function (FWHM = 0.7 eV). The computed spectra are were shifted by (a) 8.0 eV and (b) 10.0 eV in order to align with experimental spectra that are reproduced from ref. 130.

The L2,3-edge spectrum of ferrocene show significant differences as compared to the iron pentacarbonyl complex, and these differences are reflected in the calculations. This highlights L-edge sensitivity to metal oxidation state and ligand identity. In Fe(Cp)2, the more intense first peak at 709 eV is a Fe(2p) → e1g transition, as determined in previous calculations.137 The second peak has been attributed to a Fe(2p) → e2u transition where the 3e2u MO contains a significant contribution from Fe(3d) orbitals, giving it metal-to-ligand charge-transfer character.137

CAM-B3LYP/CIS + SOC calculations predict a larger L2,3 splitting for Fe(Cp)2 as compared to what is observed experimentally. This might be due to the presence of large amounts of exact exchange in the CAM-B3LYP/CIS method, which is known to adversely affect the splitting between high- and low-spin configurations of 3d transition metals.138–141 Specifically in the context of L-edge spectroscopy, Kasper et al.122 observe larger d-orbital splittings when the half-and-half functional BH&H-LYP is used, as compared to B3LYP or PBE0. Indeed, ligand-field splittings often vary directly with the fraction of exact exchange,142 and functionals with larger fractions (even 20%) tend to overestimate the d-orbital splitting.138 Overall, the CAM-B3LYP/CIS + SOC approach replicates experimental features and ligand effects with reasonable accuracy, however. This is important for L-edge X-ray spectroscopy where changes in the coordination environment around a metal center are directly observable in experimental spectra.

4.5 M-edge spectra

Finally, we consider M2,3-edge spectra of some 3d transition metal complexes. Fig. 10 presents M-edge spectra of the same two iron complexes whose L-edge spectra were discussed in Section 4.4.3, namely, Fe(CO)5 and Fe(Cp)2.130 The M-edge spectrum of TiO2 is plotted in Fig. 11, from a thin film of the rutile polymorph,143,144 and juxtaposed with a CAM-B3LYP/CIS spectrum of its fundamental octahedral unit, (TiO6)8−.
image file: d5cp01656h-f10.tif
Fig. 10 XUV spectra of (a) Fe(CO)5 and (b) Fe(Cp)2 at the Fe M-edge, comparing CAM-B3LYP/CIS simulations to experimental spectra from ref. 130. Calculations employ the def2-TZVPD basis set with 200 singlet and 200 triplet states. Transition energies were broadened using a Lorentzian function (FWHM = 0.5 eV) but have not been shifted.

image file: d5cp01656h-f11.tif
Fig. 11 XUV spectra at the Ti M-edge, for a rutile TiO2 film (experiment) and its fundamental octahedral unit (TiO6)8− (CAM-B3LYP/CIS calculation). The experimental spectrum is reproduced from ref. 144 but was originally recorded in ref. 143. Calculations use the def2-TZVPD basis set and include 200 singlet and 200 triplet states, with transition energies that are broadened using a Lorentzian function (FWHM = 0.5 eV) but not shifted.

Unlike the L-edge spectra, no shifts are needed to match with experiment in any of these M-edge cases, demonstrating that errors due to SIE are reduced for orbitals with principal quantum number n = 3 (M-edge), as compared to those with n = 2 (L-edge). Furthermore, SOC effects are no longer significant and similar spectral features are obtained using nonrelativistic calculations. This is consistent with the larger electron–nucleus separation for the 3p orbitals and the r−3 distance dependence of SOC [eqn (11) and (12)]. With or without SOC, the CAM-B3LYP/CIS spectra resemble the experimental ones, including the abrupt M-edge in Fe(CO)5, the broad multiplet in Fe(Cp)2, and a low-energy shoulder in TiO2. The spin–orbit splitting is much reduced even as compared to the same Fe complexes at the L-edge (cf. Fig. 9).

5 Conclusions

This work extends a new CAM-B3LYP/CIS variant55 of the DFT/CIS method56 to core-level spectroscopy at elemental L- and M-edges. Specifically, a state-interaction treatment of SOC, using the Breit–Pauli Hamiltonian within a mean-field approximation for the matrix elements, is used to correct the non-relativistic (SOC-free) DFT/CIS states. The same approach works for TD-DFT calculations as well,95 but CAM-B3LYP/CIS was parameterized in such a way that it significantly reduces the ad hoc shifts that are necessary to match experiment for core-to-valence excitation energies. Although these shifts were parameterized at elemental K-edges,55 we find that the same parameterization also yields more accurate results for L- and M-edge spectra, as compared to TD-DFT with general-purpose functionals.

The cost of the DFT/CIS calculations is the same as that of CIS calculations and, when combined with the CVS approximation, is significantly cheaper than the real-time TD-DFT approach.35 SOC corrections implemented in the Q-Chem program105 can be added to nonrelativistic DFT/CIS and TD-DFT calculations a posteriori, using a Python package developed for this work.106

The CAM-B3LYP/CIS + SOC method accurately models L- and M-edge spectra, as assessed herein by comparison to experimental results. The CVS approximation, which reduces the active space to a small number of occupied orbitals appropriate for the elemental edge in question, significantly reduces the cost although a large number of excited states is required to converge spectral lineshapes. The CVS approximation is also vital for disentangling overlapping spectral features and making peak assignments. At present, the SOC corrections are implemented only for closed-shell species and we hope to introduce an open-shell variant in due course, enabling simulations of core-level spectra on a wider range of transition metal systems.

Author contributions

Aniket Mandal: conceptualization (contributing); data curation (lead); formal analysis (lead); investigation (lead); methodology (lead); software (lead); validation (lead); visualization (lead); writing – original draft (lead); writing – review and editing (equal). John M. Herbert: conceptualization (lead); formal analysis (contributing); funding acquisition (lead); methodology (contributing); project administration (lead); supervision (lead); visualization (contributing); writing – review and editing (equal).

Conflicts of interest

J. M. H. is part owner of Q-Chem Inc. and serves on its board of directors.

Data availability

The data that support these findings are available from the corresponding author upon reasonable request.

Acknowledgements

Work by A. M. to develop DFT/CIS was supported by the donors of American Chemical Society Petroleum Research Fund (grant no. 62041-ND6) while work to implement SOC was supported by the National Science Foundation (grant numbers CHE-1955282 and CHE-2402361). Calculations were performed at the Ohio Supercomputer Center.145

References

  1. F. M. F. de Groot, M. Grioni, J. C. Fuggle, J. Ghijsen, G. A. Sawatzky and H. Petersen, Oxygen 1s X-ray-absorption edges of transition-metal oxides, Phys. Rev. B:Condens. Matter Mater. Phys., 1989, 40, 5715–5723 CrossRef CAS PubMed.
  2. F. M. F. de Groot, XANES spectra of transition metal compounds, J. Phys.:Conf. Ser., 2009, 190, 012004 CrossRef.
  3. S. I. Bokarev and O. Kühn, Theoretical X-ray spectroscopy of transition metal compounds, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2020, 10, e1433 CAS.
  4. T. M. Willey, J. R. I. Lee, D. Brehmer, O. A. P. Mellone, L. Landt, P. R. Schreiner, A. A. Fokin, B. A. Tkachenko, A. de Meijere, S. Kozhushkov and A. W. van Buuren, X-ray spectroscopic identification of strain and structure-based resonances in a series of saturated carbon-cage molecules: Adamantane, twistane, octahedrane, and cubane, J. Vac. Sci. Technol., A, 2021, 39, 053208 CrossRef CAS.
  5. S. Adak, M. Hartl, L. Daemen, E. Fohtung and H. Nakotte, Study of oxidation states of the transition metals in a series of Prussian blue analogs using X-ray absorption near edge structure (XANES) spectroscopy, J. Electron Spectrosc. Relat. Phenom., 2017, 214, 8–19 CrossRef CAS.
  6. M. Lundberg and M. G. Delcey, Multiconfigurational approach to X-ray spectroscopy of transition metal compounds, in Transition Metals in Coordination Environments, ed. E. Broclawik, T. Borowski, and M. Radoń, Springer Nature, Switzerland, 2019, vol. 29 of Challenges and Advances in Computational Chemistry and Physics, pp. 185–217 Search PubMed.
  7. K. Stanistreet-Welsh and A. Kerridge, Bounding [AnO2]2+ (An = U, Np) covalency by simulated O K-edge and An M-edge X-ray absorption near-edge spectroscopy, Phys. Chem. Chem. Phys., 2023, 25, 23753–23760 Search PubMed.
  8. M. L. Baker, M. W. Mara, J. J. Yan, K. O. Hodgson, B. Hedman and E. I. Solomon, K- and L-edge X-ray absorption spectroscopy (XAS) and resonant inelastic X-ray scattering (RIXS) determination of differential orbital covalency (DOC) of transition metal sites, Coord. Chem. Rev., 2017, 345, 182–208 CrossRef CAS PubMed.
  9. G. Kaindl, G. Kalkowski, W. D. Brewer, B. Perscheid and F. Holtzberg, M-edge X-ray absorption spectroscopy of 4f instabilities in rare-earth systems (invited), J. Appl. Phys., 1984, 55, 1910–1915 CrossRef CAS.
  10. F. de Groot and A. Kotani, Core Level Spectroscopy of Solids, CRC Press, Boca Raton, 2008 Search PubMed.
  11. H. Wang, A. T. Young, J. Guo, S. P. Cramer, S. Friedrich, A. Braun and W. Gu, Soft X-ray absorption spectroscopy and resonant inelastic X-ray scattering spectroscopy below 100 eV: Probing first-row transition-metal M-edges in chemical complexes, J. Synchrotron Radiat., 2013, 20, 614–619 CrossRef CAS PubMed.
  12. W. Gu, H. Wang and K. Wang, Nickel L-edge and K-edge X-ray absorption spectroscopy of non-innocent Ni[S2C2(CF3)2]2n series (n = −2, −1, 0): Direct probe of nickel fractional oxidation state changes, Dalton Trans., 2014, 43, 6406–6413 RSC.
  13. J. Vura-Weis, C.-M. Jiang, C. Liu, H. Gao, J. M. Lucas, F. M. F. de Groot, P. Yang, A. P. Alivasatos and S. R. Leone, Femtosecond M2,3-edge spectroscopy of transition metal oxides: Photoinduced oxidation state change in α-Fe2 O3, J. Phys. Chem. Lett., 2013, 4, 3667–3671 CrossRef CAS.
  14. C.-M. Jiang, L. R. Baker, J. M. Lucas, J. Vura-Weis, A. P. Alivisatos and S. R. Leone, Characterization of photo-induced charge transfer and hot carrier relaxation pathways in spinel cobalt oxide (Co3O4), J. Phys. Chem. C, 2014, 118, 22774–22784 CrossRef CAS.
  15. C. A. Leahy and J. Vura-Weis, Femtosecond extreme ultraviolet spectroscopy of an iridium photocatalyst reveals oxidation state and ligand field specific dynamics, J. Phys. Chem. A, 2022, 126, 9510–9518 CrossRef CAS PubMed.
  16. L. R. Baker, C.-M. Jiang, S. T. Kelly, J. M. Lucas, J. Vura-Weis, M. K. Gilles, A. P. Alivisatos and S. R. Leone, Charge carrier dynamics of photoexcited Co3O4 in methanol: Extending high harmonic transient absorption spectroscopy to liquid environments, Nano Lett., 2014, 14, 5883–5890 CrossRef CAS PubMed.
  17. M. Kubin, M. Guo, T. Kroll, H. Löchel, E. Källman, M. L. Baker, R. Mitzner, S. Gul, J. Kern, A. Föhlisch, A. Erko, U. Bergmann, V. Yachandra, J. Yano, M. Lundberg and P. Wernet, Probing the oxidation state of transition metal complexes: A case study on how charge and spin densities determine Mn L-edge X-ray absorption energies, Chem. Sci., 2018, 9, 6813–6829 RSC.
  18. F. M. F. de Groot, Z. W. Hu, M. F. Lopez, G. Kaindl, F. Guillot and M. Tronc, Differences between L3 and L2 X-ray absorption spectra of transition metal compounds, J. Chem. Phys., 1994, 101, 6570–6576 CrossRef.
  19. J. N. Ehrman, K. Shumilov, A. J. Jenkins, J. M. Kasper, T. Vitova, E. R. Batista, P. Yang and X. Li, Unveiling hidden shake-up features in the uranyl M4-edge spectrum, J. Am. Chem. Soc. Au, 2024, 4, 1134–1141 CAS.
  20. G. Cressey, C. M. B. Henderson and G. van der Laan, Use of L-edge X-ray absorption spectroscopy to characterize multiple valence states of 3d transition metals; a new probe for mineralogical and geochemical research, Phys. Chem. Miner., 1993, 20, 111–119 CrossRef CAS.
  21. J. T. Lau, K. Hirsch, P. Klar, A. Langenberg, F. Lofink, R. Richter, J. Rittmann, M. Vogel, V. Zamudio-Bayer, T. Möller and B. V. Issendorff, X-ray spectroscopy reveals high symmetry and electronic shell structure of transition-metal-doped silicon clusters, Phys. Rev. A:At., Mol., Opt. Phys., 2009, 79, 053201 CrossRef.
  22. G. Lucovsky, L. Miotti, D. Zeller, C. Adamo and D. Scholm, X-ray absorption studies of elemental and complex transition metal (TM) oxides: Differences between: (i) chemical, and (ii) local site symmetry multivalency, in Ulis 2011 Ultimate Integration on Silicon, 2011, pp. 1–4 Search PubMed.
  23. T. Hayakawa, K. Egashira, M. Arakawa, T. Ito, S. Sarugaku, K. Ando and A. Terasaki, X-ray absorption spectroscopy of Ce2O3+ and Ce2O5+ near Ce M-edge, J. Phys. B:At., Mol. Opt. Phys., 2016, 49, 075101 CrossRef.
  24. J. Lüder, J. Schött, B. Brena, M. W. Haverkort, P. Thunström, O. Eriksson, B. Sanyal, I. Di Marco and Y. O. Kvashnin, Theory of L-edge spectroscopy of strongly correlated systems, Phys. Rev. B, 2017, 96, 245131 CrossRef.
  25. F. de Groot, Multiplet effects in X-ray spectroscopy, Coord. Chem. Rev., 2005, 249, 31–63 CrossRef CAS.
  26. M. W. Haverkort, M. Zwierzycki and O. K. Andersen, Multiplet ligand-field theory using Wannier orbitals, Phys. Rev. B:Condens. Matter Mater. Phys., 2012, 85, 165113 CrossRef.
  27. F. M. F. de Groot, H. Elnaggar, F. Frati, R. Wang, M. U. D. M. van Veenendaal, J. Fernandez-Rodriguez, M. W. Haverkort, R. J. Green, G. van der Laan, Y. Kvashnin, A. Hariki, H. Ikeno, H. Ramanantoanina, C. Daul, B. Delley, M. Odelius, M. Lundberg, O. Kuhn, S. I. Bokarev, E. Shirley, J. Vinson, K. Gilmore, M. Stener, G. Fronzoni, P. Decleva, P. Kruger, M. Retegan, Y. Joly, C. Vorwerk, C. Draxl, J. Rehr and A. Tanaka, 2p X-ray absorption spectroscopy of 3d transition metal systems, J. Electron Spectrosc. Relat. Phenom., 2021, 249, 147061 CrossRef CAS.
  28. I. Josefsson, K. Kunnus, S. Schreck, A. Föhlisch, F. de Groot, P. Wernet and M. Odelius, Ab initio calculations of X-ray spectra: Atomic multiplet and molecular orbital effects in a multiconfigurational SCF approach to the L-edge spectra of transition metal complexes, J. Phys. Chem. Lett., 2012, 3, 3565–3570 CrossRef CAS PubMed.
  29. A. Chantzis, J. K. Kowalska, D. Maganas, S. DeBeer and F. Neese, Ab initio wave function-based determination of element specific shifts for the efficient calculation of X-ray absorption spectra of main group elements and first row transition metals, J. Chem. Theory Comput., 2018, 14, 3686–3702 CrossRef CAS PubMed.
  30. P. Norman and A. Dreuw, Simulating X-ray spectroscopies and calculating core-excited states of molecules, Chem. Rev., 2018, 118, 7208–7248 CrossRef CAS PubMed.
  31. I. Seidu, S. P. Neville, M. Kleinschmidt, A. Heil, C. M. Marian and M. S. Schuurman, The simulation of X-ray absorption spectra from ground and excited electronic states using core-valence separated DFT/MRCI, J. Chem. Phys., 2019, 151, 144104 CrossRef PubMed.
  32. M. L. Vidal, P. Pokhilko, A. I. Krylov and S. Coriani, Equation-of-motion coupled-cluster theory to model L-edge X-ray absorption and photoelectron spectroscopy, J. Phys. Chem. Lett., 2020, 11, 8314–8321 CrossRef CAS PubMed.
  33. N. A. Besley, Density functional theory based methods for the calculation of X-ray spectroscopy, Acc. Chem. Res., 2020, 53, 1306–1315 CrossRef CAS PubMed.
  34. N. A. Besley, Modeling of the spectroscopy of core electrons with density functional theory, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2021, 12, e1527 Search PubMed.
  35. J. M. Herbert, Y. Zhu, B. Alam and A. K. Ojha, Time-dependent density functional theory for X-ray absorption spectra: Comparing the real-time approach to linear response, J. Chem. Theory Comput., 2023, 19, 6745–6760 CrossRef CAS PubMed.
  36. D. R. Nascimento and N. Govind, Computational approaches for XANES, VtC-XES, and RIXS using linear-response time-dependent density functional theory based methods, Phys. Chem. Chem. Phys., 2022, 24, 14680–14691 RSC.
  37. J. M. Herbert, Density-functional theory for electronic excited states, in Theoretical and Computational Photochemistry: Fundamentals, Methods, Applications and Synergy with Experimental Approaches, ed. C. García-Iriepa and M. Marazzi, Elsevier, Amsterdam, 2023, ch. 3, pp. 69–118 Search PubMed.
  38. T. Fransson, I. E. Brumboiu, M. L. Vidal, P. Norman, S. Coriani and A. Dreuw, XABOOM: An X-ray absorption benchmark of organic molecules based on carbon, nitrogen, and oxygen 1s → π* transitions, J. Chem. Theory Comput., 2021, 17, 1618–1637 CrossRef CAS PubMed.
  39. L. Konecny, J. Vicha, S. Komorovsky, K. Ruud and M. Repisky, Accurate X-ray absorption spectra near L- and M-edges from relativistic four-component damped response time-dependent density functional theory, Inorg. Chem., 2021, 61, 830–846 CrossRef PubMed.
  40. Y. Imamura and H. Nakai, Analysis of self-interaction correction for describing core excited states, Int. J. Quantum Chem., 2007, 107, 23–29 CrossRef CAS.
  41. Y. Imamura, T. Otsuka and H. Nakai, Description of core excitations by time-dependent density functional theory with local density approximation, generalized gradient approximation, meta-generalized gradient approximation, and hybrid functionals, J. Comput. Chem., 2007, 28, 2067–2074 CrossRef CAS PubMed.
  42. Y. Imamura and H. Nakai, Description of core-ionized and core-excited states by density functional theory and time-dependent density functional theory, in Quantum Systems in Chemistry and Physics, ed. K. Nishikawa, J. Maruani, E. J. Brändas, G. Delgado-Barrio and P. Piecuch, Spring Science + Business Media, Dordrecht, 2012, vol. 26 of Progress in Theoretical Chemistry and Physics, ch. 14, pp. 275–308 Search PubMed.
  43. N. A. Besley, M. J. G. Peach and D. J. Tozer, Time-dependent density functional theory calculations of near-edge X-ray absorption fine structure with short-range corrected functionals, Phys. Chem. Chem. Phys., 2009, 11, 10350–10358 RSC.
  44. J. D. Wadey and N. A. Besley, Quantum chemical calculations of X-ray emission spectroscopy, J. Chem. Theory Comput., 2014, 10, 4557–4564 CrossRef CAS PubMed.
  45. I. P. E. Roper and N. A. Besley, The effect of basis set and exchange-correlation functional on the time-dependent density functional theory calculations within the Tamm-Dancoff approximation of the X-ray emission spectroscopy of transition metal complexes, J. Chem. Phys., 2016, 144, 114104 CrossRef PubMed.
  46. M. Kadek, L. Konecny, B. Gao, M. Repisky and K. Ruud, X-ray absorption resonances near L2,3-edges from real-time propagation of the Dirac-Kohn-Sham density matrix, Phys. Chem. Chem. Phys., 2015, 17, 22566–22570 RSC.
  47. R. G. Fernando, M. C. Balhoff and K. Lopata, X-ray absorption in insulators with non-Hermitian real-time time-dependent density functional theory, J. Chem. Theory Comput., 2015, 11, 646–654 CrossRef CAS PubMed.
  48. M. Yang, A. Sissay, M. Chen and K. Lopata, Intruder peak-free transient inner-shell spectra using real-time simulations, J. Chem. Theory Comput., 2022, 18, 992–1002 CrossRef CAS PubMed.
  49. A. Nakata, Y. Imamura, T. Otsuka and H. Nakai, Time-dependent density functional theory calculations for core-excited states: Assessment of standard exchange-correlation functionals and development of a novel hybrid functional, J. Chem. Phys., 2006, 124, 094105 CrossRef PubMed.
  50. A. Nakata, Y. Imamura and H. Nakai, Hybrid exchange-correlation functional for core, valence, and Rydberg excitations: Core-valence-Rydberg B3LYP, J. Chem. Phys., 2006, 125, 064109 CrossRef PubMed.
  51. A. Nakata, Y. Imamura and H. Nakai, Extension of the core-valence-Rydberg B3LYP functional to core-excited-state calculations of third-row atoms, J. Chem. Theory Comput., 2007, 3, 1295–1305 CrossRef PubMed.
  52. N. A. Besley and F. A. Asmuruf, Time-dependent density functional theory calculations of the spectroscopy of core electrons, Phys. Chem. Chem. Phys., 2010, 12, 12024–12039 RSC.
  53. D. Hait, K. J. Oosterbaan, K. Carter-Fenk and M. Head-Gordon, Computing X-ray absorption spectra from linear-response particles atop optimized holes, J. Chem. Phys., 2022, 156, 201104 CrossRef CAS PubMed.
  54. K. Carter-Fenk, L. A. Cunha, J. E. Arias-Martinez and M. Head-Gordon, Electron-affinity time-dependent density functional theory: Formalism and applications to core-excited states, J. Phys. Chem. Lett., 2022, 13, 9664–9672 CrossRef CAS PubMed.
  55. A. Mandal, E. J. Berquist and J. M. Herbert, A new parameterization of the DFT/CIS method with applications to core-level spectroscopy, J. Chem. Phys., 2024, 161, 044114 CrossRef CAS PubMed.
  56. S. Grimme, Density functional calculations with configuration interaction for the excited states of molecules, Chem. Phys. Lett., 1996, 259, 128–137 CrossRef CAS.
  57. A. Stolow, Femtosecond time-resolved photoelectron spectroscopy of polyatomic molecules, Annu. Rev. Phys. Chem., 2003, 54, 89–119 CrossRef CAS PubMed.
  58. P. Glatzel and U. Bergmann, High resolution 1s core hole X-ray spectroscopy in 3d transition metal complexes--electronic and structural information, Coord. Chem. Rev., 2005, 249, 65–95 CrossRef CAS.
  59. T. Yamamoto, Assignment of pre-edge peaks in K-edge X-ray absorption spectra of 3d transition metal compounds: Electric dipole or quadrupole?, X-Ray Spectrom., 2008, 37, 572–584 CrossRef CAS.
  60. F. de Groot, G. Vankó and P. Glatzel, The 1s X-ray absorption pre-edge structures in transition metal oxides, J. Phys.:Condens. Matter, 2009, 21, 104207 CrossRef PubMed.
  61. K. Zhang, M.-F. Lin, E. S. Ryland, M. A. Verkamp, K. Benke, F. M. F. de Groot, G. S. Girolami and J. Vura-Weis, Shrinking the synchrotron: Tabletop extreme ultraviolet absorption of transition-metal complexes, J. Phys. Chem. Lett., 2016, 7, 3383–3387 CrossRef CAS PubMed.
  62. E. S. Ryland, M.-F. Lin, M. A. Verkamp, K. Zhang, K. Benke, M. Carlson and J. Vura-Weis, Tabletop femtosecond M-edge X-ray absorption near-edge structure of FeTPPCI: Metalloporphyrin photophysics from the perspective of the metal, J. Am. Chem. Soc., 2018, 140, 4691–4696 CrossRef CAS PubMed.
  63. K. Zhang, R. Ash, G. S. Girolami and J. Vura-Weis, Tracking the metal-centered triplet in photoinduced spin crossover of Fe(phen)32+ with tabletop femtosecond M-edge X-ray absorption near-edge structure spectroscopy, J. Am. Chem. Soc., 2019, 141, 17180–17188 CrossRef CAS PubMed.
  64. A. S. Chatterley, F. Lackner, C. D. Pemmaraju, D. M. Neumark, S. R. Leone and O. Gessner, Dissociation dynamics and electronic structures of highly excited ferrocenium ions studied by femtosecond XUV absorption spectroscopy, J. Phys. Chem. A, 2016, 120, 9509–9518 CrossRef CAS PubMed.
  65. A. Cirri, J. Husek, S. Biswas and L. R. Baker, Achieving surface sensitivity in ultrafast XUV spectroscopy: M2,3-edge reflection-absorption of transition metal oxides, J. Phys. Chem. C, 2017, 121, 15861–15869 CrossRef CAS.
  66. J. Husek, A. Cirri, S. Biswas and L. R. Baker, Surface electron dynamics in hematite (α-Fe2O3): Correlation between ultrafast surface electron trapping and small polaron formation, Chem. Sci., 2017, 8, 8170–8178 RSC.
  67. S. Biswas, J. Husek and L. R. Baker, Elucidating ultrafast electron dynamics at surfaces using extreme ultraviolet (XUV) reflection-absorption spectroscopy, Chem. Commun., 2018, 54, 4216–4230 RSC.
  68. S. Biswas, J. Husek, S. Londo and L. R. Baker, Ultrafast electron trapping and defect-mediated recombination in NiO probed by femtosecond extreme ultraviolet reflection-absorption spectroscopy, J. Phys. Chem. Lett., 2018, 9, 5047–5054 CrossRef CAS PubMed.
  69. J. Vura-Weis, Femtosecond extreme ultraviolet absorption spectroscopy of transition metal complexes, Annu. Rev. Phys. Chem., 2025, 76, 455–470 CrossRef CAS PubMed.
  70. J. M. Kasper, T. F. Stetina, A. J. Jenkins and X. Li, Ab initio methods for L-edge X-ray absorption spectroscopy, Chem. Phys. Rev., 2020, 1, 011304 CrossRef.
  71. A. Dreuw and M. Head-Gordon, Single-reference ab initio methods for the calculation of excited states of large molecules, Chem. Rev., 2005, 105, 4009–4037 CrossRef CAS PubMed.
  72. T. Yanai, D. P. Tew and N. C. Handy, A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP), Chem. Phys. Lett., 2004, 393, 51–57 CrossRef CAS.
  73. R. Kobayashi and R. D. Amos, The application of CAM-B3LYP to the charge-transfer band problem of the zincbateriochlorin-bacteriochlorin complex, Chem. Phys. Lett., 2006, 420, 106–109 CrossRef CAS.
  74. Z.-L. Cai, M. J. Crossley, J. R. Reimers, R. Kobayashi and R. D. Amos, Density functional theory for charge transfer: The nature of the N-bands of porphyrins and chlorophylls revealed through CAM-B3LYP, CASPT2, and SAC-CI calculations, J. Phys. Chem. B, 2006, 110, 15624–15632 CrossRef CAS PubMed.
  75. M. J. G. Peach, P. Benfield, T. Helgaker and D. J. Tozer, Excitation energies in density functional theory: An evaluation and a diagnostic test, J. Chem. Phys., 2008, 128, 044118 CrossRef PubMed.
  76. B. Alam, A. F. Morrison and J. M. Herbert, Charge separation and charge transfer in the low-lying excited states of pentacene, J. Phys. Chem. C, 2020, 124, 24653–24666 CrossRef CAS.
  77. D. Rappoport and F. Furche, Property-optimized Gaussian basis sets for molecular response calculations, J. Chem. Phys., 2010, 133, 134105 CrossRef PubMed.
  78. S. Pak and D. R. Nascimento, The role of the coupling matrix elements in time-dependent density functional theory on the simulation of core-level spectra of transition metal complexes, Electron. Struct., 2024, 6, 015014 CrossRef CAS.
  79. P.-F. Loos, A. Scemama, A. Blondel, Y. Garniron, M. Caffarel and D. Jacquemin, A mountaineering strategy to excited states: Highly accurate reference energies and benchmarks, J. Chem. Theory Comput., 2018, 14, 4360–4379 CrossRef CAS PubMed.
  80. T. Saue, Relativistic Hamiltonians for chemistry: A primer, ChemPhysChem, 2011, 12, 3077–3094 CrossRef CAS PubMed.
  81. O. Takahashi, Relativistic corrections for single- and double-core excitations at the K- and L-edges from Li to Kr, Comput. Theor. Chem., 2017, 1102, 80–86 CrossRef CAS.
  82. S. Jana and J. M. Herbert, Slater transition methods for core-level electron binding energies, J. Chem. Phys., 2023, 158, 094111 CrossRef CAS PubMed.
  83. S. Jana and J. M. Herbert, Fractional-electron and transition-potential methods for core-to-valence excitation energies using density functional theory, J. Chem. Theory Comput., 2023, 19, 4100–4113 CrossRef CAS PubMed.
  84. E. van Lenthe, J. G. Snijders and E. J. Baerends, The zero-order regular approximation for relativistic effects: The effect of spin-orbit coupling in closed shell molecules, J. Chem. Phys., 1996, 105, 6506–6516 Search PubMed.
  85. M. Douglas and N. M. Kroll, Quantum electrodynamical corrections to the fine structure of helium, Ann. Phys., 1974, 82, 89–155 CAS.
  86. I. Miroslav and T. Saue, An infinite-order two-component relativistic Hamiltonian by a simple one-step transformation, J. Chem. Phys., 2007, 126, 064102 CrossRef PubMed.
  87. L. Cheng, F. Wang, J. F. Stanton and J. Gauss, Perturbative treatment of spin-orbit-coupling within spin-free exact two-component theory using equation-of-motion coupled-cluster methods, J. Chem. Phys., 2018, 148, 044108 CrossRef PubMed.
  88. U. Wahlgren, M. Sjøvoll, H. Fagerli, O. Gropen and B. Schimmelpfennig, Ab initio calculation of the 2P1/22P3/2 splitting in the thallium atom, Theor. Chem. Acc., 1997, 97, 324–330 Search PubMed.
  89. Z. Lin, C. Zhang and L. Cheng, Comparison of state-interaction and spinor-representation calculations of spin-orbit coupling within exact two-component coupled-cluster theories, Mol. Phys., 2025, 123, e2256423 CrossRef.
  90. B. A. Heß, C. M. Marian, U. Wahlgren and O. Gropen, A mean-field spin-orbit method applicable to correlated wavefunctions, Chem. Phys. Lett., 1996, 251, 365–371 CrossRef.
  91. C. M. Marian, Spin-orbit coupling and intersystem crossing in molecules, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2012, 2, 187–203 CAS.
  92. E. Epifanovsky, K. Klein, S. Stopkowicz, J. Gauss and A. I. Krylov, Spin-orbit couplings within the equation-of-motion coupled-cluster framework: Theory, implementation, and benchmark calculations, J. Chem. Phys., 2015, 143, 064102 CrossRef PubMed.
  93. P. Pokhilko, E. Epifanovsky and A. I. Krylov, General framework for calculating spin-orbit couplings using spinless one-particle density matrices: Theory and application to the equation-of-motion coupled-cluster wave functions, J. Chem. Phys., 2019, 151, 034106 CrossRef PubMed.
  94. A. Carreras, H. Jiang, P. Pokhilko, A. I. Krylov, P. M. Zimmerman and D. Casanova, Calculation of spin-orbit couplings using RASCI spinless one-particle density matrices: Theory and applications, J. Chem. Phys., 2020, 153, 214107 CrossRef CAS PubMed.
  95. S. Kotaru, P. Pokhilko and A. I. Krylov, Spin-orbit couplings within spin-conserving and spin-flipping time-dependent density functional theory: Implementation and benchmark calculations, J. Chem. Phys., 2022, 157, 224110 CrossRef CAS PubMed.
  96. J. M. Herbert, Visualizing and characterizing excited states from time-dependent density functional theory, Phys. Chem. Chem. Phys., 2024, 26, 3755–3794 RSC.
  97. X. Zhang and J. M. Herbert, Analytic derivative couplings in time-dependent density functional theory: Quadratic response theory versus pseudo-wavefunction approach, J. Chem. Phys., 2015, 142, 064109 CrossRef PubMed.
  98. E. C. Alguire, Q. Ou and J. E. Subotnik, Calculating derivative couplings between time-dependent Hartree-Fock excited states with pseudo-wavefunctions, J. Phys. Chem. B, 2015, 119, 7140–7149 CrossRef CAS PubMed.
  99. D. R. Nascimento, E. Biasin, B. I. Poulter, M. Khalil, D. Sokaras and N. Govind, Resonant inelastic X-ray scattering calculations of transition metal complexes within a simplified time-dependent density functional theory framework, J. Chem. Theory Comput., 2021, 17, 3031–3038 CrossRef CAS PubMed.
  100. X. Zhang and J. M. Herbert, Spin-flip, tensor equation-of-motion configuration interaction with a density-functional correction: A spin-complete method for exploring excited-state potential energy surfaces, J. Chem. Phys., 2015, 143, 234107 CrossRef PubMed.
  101. S. G. Chiodo and N. Russo, DFT spin-orbit coupling between singlet and triplet excited states: A case of psoralen compounds, Chem. Phys. Lett., 2010, 490, 90–96 CrossRef CAS.
  102. S. G. Chiodo and M. Leopoldini, MolSOC: A spin-orbit coupling code, Comput. Phys. Commun., 2014, 185, 676–683 CrossRef CAS.
  103. M. Roemelt, D. Maganas, S. DeBeer and F. Neese, A combined DFT and restricted open-shell configuration interaction method including spin-orbit coupling: Application to transition metal L-edge X-ray absorption spectroscopy, J. Chem. Phys., 2013, 138, 204101 CrossRef PubMed.
  104. F. Furche, On the density matrix based approach to time-dependent density functional response theory, J. Chem. Phys., 2001, 114, 5982–5992 CrossRef CAS.
  105. E. Epifanovsky, A. T. B. Gilbert, X. Feng, J. Lee, Y. Mao, N. Mardirossian, P. Pokhilko, A. F. White, M. P. Coons, A. L. Dempwolff, Z. Gan, D. Hait, P. R. Horn, L. D. Jacobson, I. Kaliman, J. Kussmann, A. W. Lange, K. U. Lao, D. S. Levine, J. Liu, S. C. McKenzie, A. F. Morrison, K. D. Nanda, F. Plasser, D. R. Rehn, M. L. Vidal, Z.-Q. You, Y. Zhu, B. Alam, B. J. Albrecht, A. Aldossary, E. Alguire, J. H. Andersen, V. Athavale, D. Barton, K. Begam, A. Behn, N. Bellonzi, Y. A. Bernard, E. J. Berquist, H. G. A. Burton, A. Carreras, K. Carter-Fenk, R. Chakraborty, A. D. Chien, K. D. Closser, V. Cofer-Shabica, S. Dasgupta, M. de Wergifosse, J. Deng, M. Diedenhofen, H. Do, S. Ehlert, P.-T. Fang, S. Fatehi, Q. Feng, T. Friedhoff, J. Gayvert, Q. Ge, G. Gidofalvi, M. Goldey, J. Gomes, C. E. González-Espinoza, S. Gulania, A. O. Gunina, M. W. D. Hanson-Heine, P. H. P. Harbach, A. Hauser, M. F. Herbst, M. Hernández Vera, M. Hodecker, Z. C. Holden, S. Houck, X. Huang, K. Hui, B. C. Huynh, M. Ivanov, A. Jász, H. Ji, H. Jiang, B. Kaduk, S. Kähler, K. Khistyaev, J. Kim, G. Kis, P. Klunzinger, Z. Koczor-Benda, J. H. Koh, D. Kosenkov, L. Koulias, T. Kowalczyk, C. M. Krauter, K. Kue, A. Kunitsa, T. Kus, I. Ladjánszki, A. Landau, K. V. Lawler, D. Lefrancois, S. Lehtola, R. R. Li, Y.-P. Li, J. Liang, M. Liebenthal, H.-H. Lin, Y.-S. Lin, F. Liu, K.-Y. Liu, M. Loipersberger, A. Luenser, A. Manjanath, P. Manohar, E. Mansoor, S. F. Manzer, S.-P. Mao, A. V. Marenich, T. Markovich, S. Mason, S. A. Maurer, P. F. McLaughlin, M. F. S. J. Menger, J.-M. Mewes, S. A. Mewes, P. Morgante, J. W. Mullinax, K. J. Oosterbaan, G. Paran, A. C. Paul, S. K. Paul, F. Pavošević, Z. Pei, S. Prager, E. I. Proynov, A. Rák, E. Ramos-Cordoba, B. Rana, A. E. Rask, A. Rettig, R. M. Richard, F. Rob, E. Rossomme, T. Scheele, M. Scheurer, M. Schneider, N. Sergueev, S. M. Sharada, W. Skomorowski, D. W. Small, C. J. Stein, Y.-C. Su, E. J. Sundstrom, Z. Tao, J. Thirman, G. J. Tornai, T. Tsuchimochi, N. M. Tubman, S. P. Veccham, O. Vydrov, J. Wenzel, J. Witte, A. Yamada, K. Yao, S. Yeganeh, S. R. Yost, A. Zech, I. Y. Zhang, X. Zhang, Y. Zhang, D. Zuev, A. Aspuru-Guzik, A. T. Bell, N. A. Besley, K. B. Bravaya, B. R. Brooks, D. Casanova, J.-D. Chai, S. Coriani, C. J. Cramer, G. Cserey, A. E. DePrince III, R. A. DiStasio Jr., A. Dreuw, B. D. Dunietz, T. R. Furlani, W. A. Goddard III, S. Hammes-Schiffer, T. Head-Gordon, W. J. Hehre, C.-P. Hsu, T.-C. Jagau, Y. Jung, A. Klamt, J. Kong, D. S. Lambrecht, W. Liang, N. J. Mayhall, C. W. McCurdy, J. B. Neaton, C. Ochsenfeld, J. A. Parkhill, R. Peverati, V. A. Rassolov, Y. Shao, L. V. Slipchenko, T. Stauch, R. P. Steele, J. E. Subotnik, A. J. W. Thom, A. Tkatchenko, D. G. Truhlar, T. Van Voorhis, T. A. Wesolowski, K. B. Whaley, H. L. Woodcock III, P. M. Zimmerman, S. Faraji, P. M. W. Gill, M. Head-Gordon, J. M. Herbert and A. I. Krylov, Software for the frontiers of quantum chemistry: An overview of developments in the Q-Chem 5 package, J. Chem. Phys., 2021, 155, 084801 CrossRef CAS PubMed.
  106. pySETSOC, a Q-Chem post-processing program for single-excitation theories with spin–orbit coupling, https://gitlab.com/john-herbert-group/pysetsoc.
  107. D. Casanova, Efficient implementation of restricted active space configuration interaction with the hole and particle approximation, J. Comput. Chem., 2012, 34, 720–730 CrossRef PubMed.
  108. K. C. Prince, M. Vondráček, J. Karvonen, M. Coreno, R. Camilloni, L. Avaldi and M. de Simone, A critical comparison of selected 1s and 2p core hole widths, J. Electron Spectrosc. Relat. Phenom., 1999, 101–103, 141–147 CrossRef CAS.
  109. C. Nicolas and C. Miron, Lifetime broadening of core-excited and -ionized states, J. Electron Spectrosc. Relat. Phenom., 2012, 185, 267–272 CrossRef CAS.
  110. M. F. Herbst and T. Fransson, Quantifying the error of the core-valence separation approximation, J. Chem. Phys., 2020, 153, 054114 CrossRef CAS PubMed.
  111. T. Fransson and L. G. M. Pettersson, Evaluating the impact of the Tamm-Dancoff approximation on X-ray spectrum calculations, J. Chem. Theory Comput., 2024, 20, 2181–2191 CrossRef CAS PubMed.
  112. P. Elliott, F. Furche and K. Burke, Excited states from time-dependent density functional theory, in Reviews in Computational Chemistry, ed. K. B. Lipkowitz and T. R. Cundari, Wiley-VCH, New York, 2009, vol. 26, ch. 3, pp. 91–165 Search PubMed.
  113. M. Gronowski, TD-DFT benchmark: Excited states of atoms and atomic ions, Comput. Theor. Chem., 2017, 1108, 50–56 CrossRef CAS.
  114. D. Escudero, A. D. Laurent and D. Jacquemin, Time-dependent density functional theory: A tool to explore excited states, in Handbook of Computational Chemistry, ed. J. Leszczynski, A. Kaczmarek-Kedziera, T. Puzyn, M. G. Papadopoulos, H. Reis and M. K. Shukla, Springer International Publishing, Switzerland, 2nd edn, 2017, ch. 21, pp. 927–961 Search PubMed.
  115. J. Liang, X. Feng, D. Hait and M. Head-Gordon, Revisiting the performance of time-dependent density functional theory for electronic excitations: Assessment of 43 popular and recently developed functionals from rungs one to four, J. Chem. Theory Comput., 2022, 18, 3460–3473 CrossRef CAS PubMed.
  116. V. Mahamiya, P. Bhattacharyya and A. Shukla, Benchmarking Gaussian basis sets in quantum-chemical calculations of photoabsorption spectra of light atomic clusters, ACS Omega, 2022, 7, 48261–48271 CrossRef CAS PubMed.
  117. A. Dreuw, J. L. Weisman and M. Head-Gordon, Long-range charge-transfer excited states in time-dependent density functional theory require non-local exchange, J. Chem. Phys., 2003, 119, 2943–2946 CrossRef CAS.
  118. A. Lange and J. M. Herbert, Simple methods to reduce charge-transfer contamination in time-dependent density-functional calculations of clusters and liquids, J. Chem. Theory Comput., 2007, 3, 1680–1690 CrossRef CAS PubMed.
  119. A. W. Lange, M. A. Rohrdanz and J. M. Herbert, Charge-transfer excited states in a π-stacked adenine dimer, as predicted using long-range-corrected time-dependent density functional theory, J. Phys. Chem. B, 2008, 112, 6304–6308 CrossRef CAS PubMed.
  120. M. Gray, A. Mandal and J. M. Herbert, Revisiting the half-and-half functional, J. Phys. Chem. A, 2025, 129, 3969–3982 CrossRef CAS PubMed.
  121. A. T. Wen and A. P. Hitchcock, Inner shell spectroscopy of (η5-C5H5)2TiCl2, (η5-C5H5)2TiCl3, and TiCl4, Can. J. Chem., 1993, 71, 1632–1644 CrossRef CAS.
  122. J. M. Kasper, P. J. Lestrange, T. F. Stetina and X. Li, Modeling L2,3-edge X-ray absorption spectroscopy with real-time exact two-component relativistic time-dependent density functional theory, J. Chem. Theory Comput., 2018, 14, 1998–2006 CrossRef CAS PubMed.
  123. G. Fronzoni, M. Stener, P. Decleva, F. Wang, T. Ziegler, E. van Lenthe and E. Baerends, Spin-orbit relativistic time dependent density functional theory calculations for the description of core electron excitations: TiCl4 case study, Chem. Phys. Lett., 2005, 416, 56–63 CrossRef CAS.
  124. M. Casarin, P. Finetti, A. Vittadini, F. Wang and T. Ziegler, Spin-orbit relativistic time-dependent density functional calculations of the metal and ligand pre-edge XAS intensities of organotitanium complexes: TiCl4, Ti(η5-C5H5)Cl3, and Ti(η5-C5H5)2Cl2, J. Phys. Chem. A, 2007, 111, 5270–5279 CrossRef CAS PubMed.
  125. M. Gray and J. M. Herbert, Comprehensive basis-set testing of extended symmetry-adapted perturbation theory and assessment of mixed-basis combinations to reduce cost, J. Chem. Theory Comput., 2022, 18, 2308–2330 CrossRef CAS PubMed.
  126. J. D. Bozek, K. H. Tan and G. M. Bancroft, High resolution gas phase photoabsorption spectra of SiCl4 and Si(CH3)4 at the silicon L edges: Characterization and assignment of resonances, Chem. Phys. Lett., 1987, 138, 33–42 CrossRef CAS.
  127. G. Fronzoni, M. Stener, P. Decleva, M. de Simone, M. Coreno, P. Franceschi, C. Furlani and K. C. Prince, X-ray absorption spectroscopy of VOCl3, CrO2Cl2, and MnO3Cl: An experimental and theoretical study, J. Phys. Chem. A, 2009, 113, 2914–2925 CrossRef CAS PubMed.
  128. M. Nakamura, M. Sasanuma, S. Sato, M. Watanabe, H. Yamashita, Y. Iguchi and A. Ejiri, Absorption structure near the LII,III edge of argon gas, Phys. Rev. Lett., 1968, 21, 1303–1305 CrossRef CAS.
  129. S. Bodeur and J. M. Esteva, Photoabsorption spectra of H2S, CH3SH and SO2 near the sulfur K edge, Chem. Phys., 1985, 100, 415–427 CrossRef CAS.
  130. A. T. Wen, E. Ruehl and A. P. Hitchcock, Inner-shell excitation of organoiron compounds by electron impact, Organometallics, 1992, 11, 2559–2569 CrossRef CAS.
  131. R. Brydson, L. A. J. Garvie, A. J. Craven, H. Sauer, F. Hofer and G. Cressey, L2,3 edges of tetrahedrally coordinated d0 transition-metal oxyanions XO4n, J. Phys.:Condens. Matter, 1993, 5, 9379–9392 CrossRef CAS.
  132. D. Danovich, C. M. Marian, T. Neuheuser, S. D. Peyerimhoff and S. Shaik, Spin-orbit coupling patterns induced by twist and pyramidalization modes in C2H4: A quantitative study and a qualitative analysis, J. Phys. Chem. A, 1998, 102, 5923–5936 CrossRef CAS.
  133. J. Tatchen and C. M. Marian, On the performance of approximate spin-orbit Hamiltonians in light conjugated molecules: The fine-structure splitting of HC6H+, NC5H+, and NC4N+, Chem. Phys. Lett., 1999, 313, 351–357 CrossRef CAS.
  134. E. Hudson, D. A. Shirley, M. Domke, G. Remmers, A. Puschmann, T. Mandel, C. Xue and G. Kaindl, High-resolution measurements of near-edge resonances in the core-level photoionization spectra of SF6, Phys. Rev. A:At., Mol., Opt. Phys., 1993, 47, 361–373 CrossRef CAS PubMed.
  135. F. M. F. de Groot, Differences between L3 and L2 X-ray absorption spectra, Phys. B, 1995, 208–209, 15–18 CrossRef.
  136. K. Godehusen, T. Richter, P. Zimmermann and P. Wernet, Iron L-edge absorption spectroscopy of iron pentacarbonyl and ferrocene in the gas phase, J. Phys. Chem. A, 2017, 121, 66–72 CrossRef CAS PubMed.
  137. A. P. Hitchcock, A. T. Wen and E. Rühl, Transition metal 2p excitation of organometallic compounds studied by electron energy loss spectroscopy, Chem. Phys., 1990, 147, 51–63 CrossRef CAS.
  138. M. Reiher, O. Salomon and B. A. Hess, Reparameterization of hybrid functionals based on energy differences of states of different multiplicity, Theor. Chem. Acc., 2001, 107, 48–55 Search PubMed.
  139. J. N. Harvey, DFT computation of relative spin-state energetics of transition metal compounds, Struct. Bonding, 2004, 112, 151–183 CrossRef CAS.
  140. J. N. Harvey, On the accuracy of density functional theory in transition metal chemistry, Annu. Rep. Prog. Chem., Sect. C:Phys. Chem., 2006, 102, 203–226 RSC.
  141. M. Radoń, Revisiting the role of exact exchange in DFT spin-state energetics of transition metal complexes, Phys. Chem. Chem. Phys., 2014, 16, 14479–14488 RSC.
  142. G. Capano, T. J. Penfold, N. A. Besley, C. J. Milne, M. Reinhard, H. Rittmann-Frank, P. Glatzel, R. Abela, U. Rothlisberger, M. Chergui and I. Tavernelli, The role of Hartree-Fock exchange in the simulation of X-ray absorption spectra: A study of photoexcited [Fe(bpy)3]2+, Chem. Phys. Lett., 2013, 580, 179–184 CrossRef CAS.
  143. C.-M. Jiang, Charge Carrier Dynamics in Transition Metal Oxides Studied by Femtosecond Transient Extreme Ultraviolet Absorption Spectroscopy, PhD thesis, University of California, Berkeley, 2015 Search PubMed.
  144. A. Kubas, M. Verkamp, J. Vura-Weis, F. Neese and D. Maganas, Restricted open-shell configuration interaction singles study on M- and L-edge X-ray absorption spectroscopy of solid chemical systems, J. Chem. Theory Comput., 2018, 14, 4320–4334 CrossRef CAS PubMed.
  145. Ohio Supercomputer Center, https://osc.edu/ark:/19495/f5s1ph73 (accessed 2025-07-08).

Footnotes

Electronic supplementary information (ESI) available: Coordinates of all systems examined. See DOI: https://doi.org/10.1039/d5cp01656h
Present address: Dept. of Physics, Rutgers University, Newark, New Jersey USA.

This journal is © the Owner Societies 2025
Click here to see how this site uses Cookies. View our privacy policy here.