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

Electronic spectra of ytterbium fluoride from relativistic electronic structure calculations

Johann V. Pototschnig *ab, Kenneth G. Dyall c, Lucas Visscher b and André Severo Pereira Gomes d
aInstitut für Experimentalphysik, Technische Universität Graz, Petersgasse 16, 8010 Graz, Austria. E-mail: johann.pototschnig@tugraz.at
bDepartment of Chemistry and Pharmaceutical Sciences, Faculty of Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1083, NL-1081 HV Amsterdam, The Netherlands. E-mail: visscher@chem.vu.nl
cDirac Solutions, 10527 NW Lost Park Drive, Portland, OR 97229, USA. E-mail: diracsolutions@gmail.com
dUniversité de Lille, CNRS, UMR 8523 - PhLAM - Physique des Lasers, Atomes et Molécules, F-59000 Lille, France. E-mail: andre.gomes@univ-lille.fr

Received 11th August 2021 , Accepted 23rd September 2021

First published on 1st October 2021


Abstract

We report an investigation of the low-lying excited states of the YbF molecule-a candidate molecule for experimental measurements of the electron electric dipole moment-with 2-component based multi-reference configuration interaction (MRCI), equation of motion coupled cluster (EOM-CCSD) and the extrapolated intermediate Hamiltonian Fock-space coupled cluster (XIHFS-CCSD). Specifically, we address the question of the nature of these low-lying states in terms of configurations containing filled or partially-filled Yb 4f shells. We show that while it does not appear possible to carry out calculations with both kinds of configurations contained in the same active space, reliable information can be extracted from different sectors of Fock space-that is, by performing electron attachment and detachment IHFS-CCSD and EOM-CCSD calculation on the closed-shell YbF+ and YbF species, respectively. From these calculations we predict Ω = 1/2, 3/2 states, arising from the 4f13σ26s, 4f145d1/6p1, and 4f135d1σ16s configurations to be able to interact as they appear in the same energy range around the ground-state equilibrium geometry. As these states are generated from different sectors of Fock space, they are almost orthogonal and provide complementary descriptions of parts of the excited state manifold. To obtain a comprehensive picture, we introduce a simple adiabatization model to extract energies of interacting Ω = 1/2, 3/2 states that can be compared to experimental observations.


1 Introduction

In a previous paper,1 we introduced all-electron relativistic basis sets for the lanthanides (La–Lu) and discussed their performance for the determination of spectroscopic constants for the ground state of ytterbium fluoride (YbF), an open-shell molecule with a 2Σ+ ground state. This molecule has received a fair amount of experimental and theoretical attention because of its potential application in the observation of parity-violating interactions2–4via determination of the electric dipole moment of the electron (eEDM)-see, for instance ref. 4–11 and references therein). There is also some interest in the Yb atom, cation and dimer in connection to ultracold physics.12–14 An example15 is the parity violation observed in the Yb atom.

A noteworthy finding in our previous work was the sensitivity of coupled cluster calculations to the basis set in use and, indirectly, to the amount of electron correlation recovered in the calculations. We observed a spike in the values of the T1 diagnostic around the ground state equilibrium geometry, so it appears that the perturbative treatment of triple excitations in the CCSD(T) calculations breaks down in this region of the potential energy curves unless there is enough flexibility in the correlation treatment. The same was recently observed by Pasteka et al.16 for the nuclear quadruple coupling constant. This suggests the existence of a low lying perturbing state, which we want to investigate further in the current work.

Experimental17,18 and previous theoretical19–21 investigations suggest that in the ground state the unpaired electron is located in a σs orbital with dominant contributions from the 6s orbital of Yb, corresponding to a Yb(4f14σ16s)F configuration. This 2Σ1/2 state ground state was studied in greater detail by combining microwave and optical spectroscopy for the odd 171Yb isotope.22

Experimentally,18 the lowest excited state observed is assigned as 2Π1/2, with an energy of 18[thin space (1/6-em)]106 cm−1, while the 2Π3/2 component is found at 19[thin space (1/6-em)]471 cm−1, yielding a spin–orbit splitting of 1365 cm−1 of this spin–orbit split A 2Π state. The lower component will be denoted 31/2 in the current work. Experiments indicate a perturbation of its vibrational levels,18,23–25 which was attributed to the presence of a perturbing state (denoted by 41/2 here) found at 18[thin space (1/6-em)]705 cm−1.18 This perturbing 41/2 state is sometimes referred to as [18.6]1/2 by experimentalists18,23 (energy in cm−1 divided by 103 in the square brackets, and Ω-value as subscript). The mixing of these two Ω = 1/2 states gives rise to states designated as [557] and [561] (the values in square brackets referring to transition energies in Thz from the vibronic ground state) with transition energies of 18[thin space (1/6-em)]574 and 18[thin space (1/6-em)]699 cm−1, respectively.24,25 These two states are of importance for laser cooling schemes that have been investigated26 and tested25 with the purpose of realizing high-accuracy measurements of YbF at very low temperatures. Besides these first excited states, Smallman26 investigated also two not yet fully characterized mixed states, [574](≈19[thin space (1/6-em)]150 cm−1) and [578](≈19[thin space (1/6-em)]280 cm−1) at higher energies. These can be compared with the Ω = 3/2 state at 19[thin space (1/6-em)]471 cm−1 found earlier by Dunfield,18 which will be denoted 23/2 in the current work. Uttam et al.27,28 furthermore measured additional unidentified higher bands at about 23[thin space (1/6-em)]035, 23[thin space (1/6-em)]256 and 26[thin space (1/6-em)]015 cm−1, which they denoted as C1, C2 and D, respectively.

Theoretically, excited states arising from the Yb(4f146p1)F and Yb(4f145d1)F configurations were considered by Nayak and Chaudhuri6 with RAS-CI based on 4-component spinors, yielding the A2Π1/2 (31/2), A2Π3/2 (23/2), and a 2Σ1/2 state. Earlier multireference CI calculations by Dolg et al.19 furthermore indicate the possibility of low-lying Ω = 1/2 states arising from the image file: d1cp03701c-t1.tif or image file: d1cp03701c-t2.tif configurations, to be lying below or close to the Yb(4f146p1)F states. This was also found in the DFT calculations of Liu et al.20 who place excited states arising from the image file: d1cp03701c-t3.tif configuration in the range from 9000 to 15[thin space (1/6-em)]000 cm−1 relative to the Yb(4f14σ16s)F ground state. These findings make it of interest to explicitly consider the configuration interaction between the f13 and f14 configurations in the Yb atom.29

The vibronic states are additionally split due to hyperfine interactions. In atomic experiments they were measured for the ground and excited states9,22,30–33 also using Zeeman spectroscopy.34 The hyperfine interaction of the atom35–37 and molecule38 were studied theoretically, and should have similar uncertainties to the contribution of the eEDM to the spectrum due to the similarity of the matrix elements. Recently, uncertainties of the hyperfine constants arising in relativistic coupled cluster computations have been studied.39

It is clear from the above that a proper description of the Yb atom and the YbF molecule requires an accurate treatment of both spin–orbit coupling and electron correlation, for ground as well as excited states. A popular approach is the so-called two-step approach to spin–orbit coupling (SOC), in which electron correlation methods based on non-relativistic or scalar relativistic Hamiltonians are used to obtain excited state energies, that are in turn used to dress a spin–orbit configuration interaction (SOCI) matrix. This approach can yield quite accurate spin–orbit coupled states, but results are particularly sensitive to the number of spin-free states serving as a basis for the SOCI step.40–46 An alternative is to include SOC already at the mean-field level, and use fully SO-coupled molecular spinors to construct the correlated wave functions.47 This can be done with four-component Hamiltonians, as done for the ground1 and excited states6,37,48,49 of YbF, or with more computationally efficient two-component Hamiltonians based on the eXact 2-Component (X2C) approach,50–57 in which a transformation to decouple the positive and negative energy states of the Dirac Hamiltonian can be carried out in matrix form, yielding the same positive energy spectrum as the original 4-component Hamiltonian. More details can be found in the recent review by Liu.47 Among the different X2C flavors, we can distinguish two main strategies for the decoupling, which is performed based on: (i) the one-electron Dirac Hamiltonian prior to the mean-field step,54,57 and for which two-electron spin–orbit contributions due to the untransformed two-electron potential are included via atomic mean-field contributions calculated with the AMFI code41,58,59 (X2C-AMFI); (ii) after a converged 4-component mean-field calculation on atoms52,53,56 or molecules55 (2DCM). Recent benchmarks show that 2DCM calculations closely reproduce equivalent 4-component ones for valence60 or core61 states.

Moreover, the aforementioned calculations for the excited states of YbF have mostly employed multireference CI (MRCI) approaches. While these can provide great flexibility in capturing static correlation, it remains the case that dynamical correlation is better accounted for with coupled cluster approaches. Among the coupled cluster singles and doubles (CCSD) approaches for excited states, we have the equation of motion (EOM-CCSD) method as well as Fock-space (IHFS-CCSD) methods,62 of which the single electron attachment, detachment, and singly excited states variants are the most commonly used. The two approaches have been found to yield very accurate results in general and in particular for calculations with relativistic Hamiltonians as discussed elsewhere (see ref. 60 and references therein).

The first goal of this work is therefore to go beyond the investigations performed to date in the literature, and apply the relativistic EOM-CCSD and IHFS-CCSD approaches to describe the low-lying excited states of YbF. For such states, where the most important excited state configurations appear have a single open-shell character (4f14σ16s, image file: d1cp03701c-t4.tif 4f145d1, image file: d1cp03701c-t5.tif 4f146p1), these coupled cluster approaches are in principle applicable, provided one starts from closed shell configurations such as Yb(4f14σ26s)F or Yb(4f14)F+. Additionally, we assess the performance of relativistic MRCI with respect to the coupled cluster methods. Our second goal is to confirm whether any low-lying state is close enough to the ground state to perturb the latter, and explain the anomalous behavior observed in the open-shell ground-state calculations in the literature.

2 Computational details

All relativistic electronic structure calculations were performed with a development version of the Dirac program suite63 (revision 6e10c5d3), employing for Yb the valence double-zeta (24s19p13d9f2g), triple-zeta (30s24p18d14f4g2h) and quadruple-zeta (35s30p19d16f6g4h2i) basis sets from the previous work,1 along with the matching augmented correlation-consistent (aug-cc-pVnZ, n = 2, 3, 4) basis sets of Dunning64 for F. All basis sets were kept uncontracted, with the small component basis generated by restricted kinetic balance. In addition to these individual basis sets, we have used the calculations with triple- and quadruple-zeta sets to construct extrapolations to the complete basis set limit (E) for the underlying potential energy curves, using the relation65
 
image file: d1cp03701c-t6.tif(1)
where the subscripts denote the cardinal numbers for the basis sets and En(R) the energy for a given geometry and electronic structure method for a basis of cardinal number n(= 2, 3, 4).

In the coupled cluster computations the 2DCM Hamiltonian54,55 was applied, all two-electron integrals over small component (S) basis sets (i.e. the so-called (SS|SS)-type integrals) appearing in the SCF step have been replaced by a simple correction.66 In order to account for spin–orbit coupling and other relativistic effects the X2C-AMFI Hamiltonian was employed for the Kramers-restricted configuration interaction (KRCI) method.

Spectroscopic constants (re, De, ωe and ωeχe) were determined from a Morse potential fit in the vicinity of the potential energy minima. The potential energy curves were determined for bond lengths between 1.6 Å and 2.3 Å spaced by 0.02 Å and additional points with larger spacing up to 3.5 Å. In the calculation of De the asymptotic dissociation limit is calculated from the energies of the isolated neutral atoms, F in the 2P3/2 state and Yb in the 1S0 state.

The dataset associated with this manuscript (outputs from calculations, codes to extract and process information from these, and code to obtain the spectroscopic constants) is provided in ref. 67.

2.1 Kramers-restricted configuration interaction

For YbF we first consider Kramers-restricted configuration interaction (KRCI) based on an average-of-configuration Hartree–Fock approach (AOC-SCF).68 This method was employed in order to treat the open shells, where one or two valence electrons were distributed over the s- and d-orbitals and the f-shell was either completely filled or contained one hole, depending on the states of interest. The AOC-SCF reference wave function in the KRCI computation is occupied according to a definition given by a generalized active space (GAS).69 In this approach the Hamiltonian is computed for all allowed configurations and then diagonalized. The GAS space was defined by a f-shell which was completely filled or contained one hole and one or two electrons distributed over 29 orbitals.

2.2 Equation-of-motion coupled cluster

The first approach we use to describe the dynamical correlation that is largely missing in KRCI is EOM-CCSD, which can give access to electronic states of different kinds, depending on the single determinant wave function that is chosen as the starting point. In it, the CCSD amplitudes are determined for the chosen ground state in the first step, and subsequently the similarity transformed Hamiltonian is constructed using these amplitudes and the desired states are generated by an operator that either removes or adds an electron.

The first set of states was obtained by electron attachment on Yb(4f14F+ ion, where the HOMO (σ6s,1/2) of YbF was initially empty. This computation on the (0h,1p) sector of Fock space yielded states with 4f14 and a valence electron in the σ6s, d or p orbital. This means that, in the process of obtaining the potentials for the ground and excited states of YbF, we immediately obtain energies of CCSD quality for YbF+, and therefore vertical ionization potentials (IP) at each geometry.

Another set of states was obtained by ionizing the Yb(4f14σ26s)F anion, where the HOMO (σ6s,1/2) of YbF was initially doubly occupied. States arising from the Yb(4f14σ16s)F, image file: d1cp03701c-t7.tifimage file: d1cp03701c-t8.tif and Yb(4f14σ26s)F(2p5) configurations were obtained by considering the (1h,0p) sector of Fock space. This means that, in the process of obtaining the potentials for the ground and excited states of YbF, we immediately obtain energies of CCSD quality for YbF, and therefore vertical electron affinities (EA) at each geometry. We note that states arising from the (2h,1p) and (1h,2p) manifolds are also accessible from EOM-IP and EOM-EA calculations, though the energy of electronic states determined by such configurations will be less accurate than states dominated by single detachment or attachment configurations.

The EOM-CCSD electronic states are obtained by an iterative diagonalization (Davidson) procedure in which only the energies of a certain number of the lowest states are determined. For the IP-EOM-CCSD we obtained 16 Ω = 1/2, 8 Ω = 3/2, 6 Ω = 5/2 and 2 Ω = 7/2 states, whereas for EA-EOM-CCSD we obtained 8 Ω = 1/2, 6 Ω = 3/2, 4 Ω = 5/2 and 2 Ω = 7/2 states.

As transition moments are not yet available for the EOM-CCSD implementation in Dirac, we have only obtained the potential energy curves. These are nevertheless useful since, by not requiring the definition of model spaces or the use of an extrapolation procedure, they serve as a cross validation of the IHFS-CCSD calculations below.

2.3 Fock-space coupled cluster

Fock-space coupled cluster70 (FS-CCSD) is our second approach to include dynamical correlation in the electronically excited states. Here it was employed in a similar fashion to EOM-CCSD, starting from YbF+ or YbF and proceeding to the (0h,1p) and (1h,0p) sectors of Fock space, respectively. For FS-CCSD a model space is defined by selecting a number of occupied and virtual orbitals and how many electrons are added and removed. The matrix for this subspace is constructed and subsequently diagonalized, thus yielding all states within the chosen model space, in this case states arising from single electron attachment (EA) or single electron detachment (IP). This method requires solving first for the underlying sectors, starting with (0h,0p), which corresponds to CCSD. Due to computational constraints, we have truncated the virtual space so that 117, 230 and 296 orbitals were used in the double-, triple- and quadruple-zeta CCSD calculations, respectively.

The separation into a model and external space leads to the appearance of the so-called intruder states, a well-known difficulty with Fock-space coupled cluster and other effective Hamiltonian approaches, that can be dealt with in many cases by the intermediate Hamiltonian (IH) Fock-space coupled cluster (IHFS-CCSD) method.71,72

The IH approach was employed to compute Yb(4f146s,6p,5d…}1)F states starting from YbF+. The active P space in such calculations contained about 50 spinors varying slightly with bond distance and basis set. Of these 26 spinors are always present in the model (Pm) space, whereas the remaining active spinors are placed in the intermediate (PI) space. Due to using the (0h,1p) sector for the cation, states arising from configurations where the Yb 4f shell is partially filled (such as image file: d1cp03701c-t9.tifimage file: d1cp03701c-t10.tif or image file: d1cp03701c-t11.tif) are not accessible in this calculation.

The approach outlined above was, however, not enough to avoid divergence in the computation of the (1h,0p) sector using the anion as a reference. Therefore, the extrapolated intermediate Hamiltonian (XIH) Fock-space coupled-cluster approach73 (XIHFS-CCSD) was applied using the same shifts as in ref. 73. Values of 0.1 and 0.2 Hartree were selected if one of the holes is not in the model space. These shifts were doubled for two holes outside the model space. Using the determined energies an extrapolation to the system without shifts was performed. The model (Pm) space in these computations contained 22 spinors, the intermediate (PI) space about 24 spinors depending on the bond distance and basis set. Since we start out from the anion and only allow holes, only Yb(4f14σ16s)F, image file: d1cp03701c-t12.tifimage file: d1cp03701c-t13.tif and Yb(4f14σ26s)F(2p5) configurations are accessible in this computation.

Combining the two sectors allows us to get different excited states of YbF, although there are limitations. Firstly, the interactions between configurations with open f-shell and the ones with an electron in the p- or d-shell are not included, since they will be obtained for different sectors of Fock space. This interaction will nevertheless be treated with a simple adiabatization approach, described in Sections 2.4 and 3.6. Secondly, configurations such as image file: d1cp03701c-t14.tif or image file: d1cp03701c-t15.tif are not included in the current treatment. This limitations is not as significant because these states have higher energies than the ones we are interested in. Both of these problems could be dealt with by using the (1h,1p) sector, but this goes beyond the current work as convergence is very unstable for this sector and it requires the use of an open-shell reference.

2.4 Adiabatization of electronic states

As we separated the computations of states with 4f13 and 4f14 character, these states cannot interact with each other, and states with the same Ω value cross although they should have an avoided crossing. In order to correct this deficiency we considered a simple adiabatization model, in which we set up and diagonalize the following matrix for each Ω value:
 
image file: d1cp03701c-t16.tif(2)
where C is a matrix where every entry is a coupling constant (whose value is kept constant for all states and geometries considered; we have investigated values of 0.01, 0.001 and 0.0001 a.u.), E are matrices with the eigenvalues of the different electronic states on the diagonal. The potential curves were computed for different coupling constants and the results are shown in Section 3.6. We note that since the ground state energy, associated with the Yb(4f14σ16s)F configuration, appears in both coupled cluster approaches, we have only considered one such energy. As we shall see in the discussion, this is valid in the region between 1.8 Å and 2.5 Å, since for these distances the ground-state energies from IP and EA calculations are nearly identical.

3 Results and discussion

We start our discussion with the electronic transitions of the atomic Yb+ cation, before moving on to the YbF molecule. This is because the cation's electronic structure is similar to the Yb in YbF since, due to the large electronegativity of fluorine, one electron is almost completely removed from the Yb atom.

Subsequently, the potential energy curves for Kramers-restricted configuration interaction are presented. A discussion of the coupled cluster approaches follows, with a focus on the comparison of the coupled cluster results for the Fock space and equation-of-motion approach. This section is followed by a presentation of the spectroscopic data for the ground and excited states. In the last part we take a closer look at the mixing of states at around 18[thin space (1/6-em)]000 cm−1 and apply the adiabatization procedure.

3.1 Ytterbium cation

As discussed in the introduction, states from both the 4f13 and the 4f14 configurations are of importance. This is difficult to realize in a balanced manner when using one set of orbitals to describe all states. Any change in the occupation of the 4f-shell will alter the screening of the 5s and 5p orbitals of Yb, resulting in differences between orbital sets optimized for a 4f13 or a 4f14 configuration. Additionally, the 4f orbitals are very compact and since they are the first f-shell there are no orthogonality conditions limiting the radial expansion or contraction of the orbital. Depictions of the orbitals for both configurations can be found in Table S1 in the ESI.

These observations help to understand why it turned out to be very difficult to treat both sets of states in the same calculation, which we attempted to do from AOC-SCF on the Yb+. We started out by performing AOC-SCF computations on the atom, based on the 4f closed shell configuration. While we obtained the correct ground state configuration, the image file: d1cp03701c-t17.tif has an energy of about 46[thin space (1/6-em)]000 cm−1 (over two times higher than the experimental value), and the wrong order for the hole states is observed. If the wave function is optimized for a 4f13 configuration, one obtains the image file: d1cp03701c-t18.tif as the lowest state and finds the true ground state more than 20[thin space (1/6-em)]000 cm−1 higher. Because of these difficulties, the KRCI calculations discussed below all follow the strategy of different orbital sets that is also employed in the subsequent coupled cluster calculations.

Table 1 contains KRCI values of electronic transitions for the cation. The transition energies show deviations of about 10% and the spin–orbit splitting is underestimated for states with a 4f14 configuration. The squared transition dipole moment (TDM2) of the image file: d1cp03701c-t19.tif state is underestimated by about 13%, the one for the image file: d1cp03701c-t20.tif state overestimated by about 47%. The second set of states with a hole in the f-shell and different distributions of the 2 valence electrons are given in the lower part of Table 1, the energies are relative to the image file: d1cp03701c-t21.tif state. In this case the two valence electrons are distributed over the s- and d-shell. The lowest state with a image file: d1cp03701c-t22.tif configuration has a transition energy of 47912.31 cm−1 and was not included in the current treatment.

Table 1 Kramers-restricted configuration interaction transition energies (in cm−1), squared transition dipole moments (TDM2), and line strength (S) for the Yb+ cation, the latter two are in atomic units(e2a02). Reference values and notation have been taken from the NIST database.74 For the 4f13 configurations, energies relative to the image file: d1cp03701c-t23.tif state are also given. 2z, 3z, 4z, and extr. indicate double, triple, quadruple zeta and extrapolated results, respectively
State Conf. NIST74 2z 3z 4z extr.
E S E TDM2 E TDM2 E TDM2 E
2S1/2 4f146s1 0 0 0 0 0
2D3/2 4f145d1 22[thin space (1/6-em)]961 23[thin space (1/6-em)]322 0.0 22[thin space (1/6-em)]802 0.0 23[thin space (1/6-em)]606 0.0 24[thin space (1/6-em)]192
2D5/2 4f145d1 24333 23[thin space (1/6-em)]882 0.0 23[thin space (1/6-em)]321 0.0 24[thin space (1/6-em)]117 0.0 24[thin space (1/6-em)]698
image file: d1cp03701c-t24.tif 4f146p1 27062 6.1 25[thin space (1/6-em)]210 3.5 24[thin space (1/6-em)]533 3.8 25[thin space (1/6-em)]331 3.6 25[thin space (1/6-em)]914
image file: d1cp03701c-t25.tif 4f146p1 30392 11.4 28[thin space (1/6-em)]104 16.9 27[thin space (1/6-em)]385 18.9 28[thin space (1/6-em)]153 17.4 28[thin space (1/6-em)]712

State Conf. E ΔE ΔE ΔE ΔE ΔE
image file: d1cp03701c-t26.tif 4f136s2 21[thin space (1/6-em)]419 0 0 0 0 0
image file: d1cp03701c-t27.tif 4f135d16s1 26[thin space (1/6-em)]759 5340 4260 5538 4618 3946
image file: d1cp03701c-t28.tif 4f135d16s1 28[thin space (1/6-em)]758 7339 6387 7822 7123 6613
image file: d1cp03701c-t29.tif 4f135d16s1 30[thin space (1/6-em)]224 8806 8214 9325 8314 7576
image file: d1cp03701c-t30.tif 4f135d16s1 30[thin space (1/6-em)]563 9144 8320 9431 8447 7729


The excited states of the Yb atom for the 4f14 configuration have already been investigated by relativistic Fock-space coupled cluster36,37,75,76 as well as for the cation,75 including the transition moment of magnetic transitions.76 With our current calculations we can go beyond these studies and investigate the f13 configurations as well. Before discussing our IHFS-CCSD calculations for Yb+, we focus on the EOM-CCSD excitation energies, shown in Table 2. The EOM-IP-CCSD energies of 4f13 states obtained from the extrapolation to the complete basis set limit underestimate the experimental transition energies by around 3000 cm−1, whereas the values for 4f14 states, obtained with EOM-EA-CCSD are within 1000 cm−1 of the experimental values, which yields a quantitative improvement over the KRCI ones for both configurations, even though qualitatively the two methods provide a similar picture. From that and the preceeding discussion, we attribute the relatively lower accuracy for the 4f13 to arise from the incomplete account of the relaxation of the wave function upon the creation of the hole in the f shell. Beyond the states presented in Table 2, which are dominated by single electron attachment and detachment, we are able to access states with significant (1h,2p) and (2h,1p) character with EOM-CCSD. These states, available in Table S14 in the ESI, are about 10[thin space (1/6-em)]000 cm−1 higher in energy than the experimental ones.

Table 2 Transition energies (in cm−1) for the Yb+ cation, obtained for different basis set with EOM-IP-CCSD (4f13) and EOM-EA-CCSD (4f14), except for the ground state, for which both methods yield the same configuration and total energy. 2z, 3z, 4z, and extr. indicate double, triple, quadruple zeta and extrapolated results, respectively. Reference values were obtained from the NIST database74
State Conf. NIST74 2z 3z 4z Extr.
2S1/2 4f146s1 0 0 0 0 0
image file: d1cp03701c-t31.tif 4f136s2 21[thin space (1/6-em)]419 12[thin space (1/6-em)]054 13[thin space (1/6-em)]524 16[thin space (1/6-em)]092 17[thin space (1/6-em)]966
image file: d1cp03701c-t32.tif 4f136s2 31[thin space (1/6-em)]568 22[thin space (1/6-em)]629 24[thin space (1/6-em)]139 26[thin space (1/6-em)]655 28[thin space (1/6-em)]491
2D3/2 4f145d1 22[thin space (1/6-em)]961 24[thin space (1/6-em)]073 24[thin space (1/6-em)]209 24[thin space (1/6-em)]060 23[thin space (1/6-em)]951
2D5/2 4f145d1 24[thin space (1/6-em)]333 25[thin space (1/6-em)]351 25[thin space (1/6-em)]457 25[thin space (1/6-em)]341 25[thin space (1/6-em)]257
image file: d1cp03701c-t33.tif 4f146p1 27[thin space (1/6-em)]062 27[thin space (1/6-em)]539 27[thin space (1/6-em)]780 27[thin space (1/6-em)]857 27[thin space (1/6-em)]913
image file: d1cp03701c-t34.tif 4f146p1 30[thin space (1/6-em)]392 30[thin space (1/6-em)]954 31[thin space (1/6-em)]246 31[thin space (1/6-em)]323 31[thin space (1/6-em)]380


Finally, our IHFS-CCSD results are presented in Table 3. The transition energies for 4f14 configuration reproduce well the experimental ones, with errors below 6%, and only show a small dependence on the basis set. The states arising from the 4f13 configuration (image file: d1cp03701c-t35.tifetc.), in contrast, show a significant dependence on the basis and a rather slow convergence and underestimate the value by about 30%, which makes them less accurate than the EOM-CCSD ones. This lower accuracy is a consequence of the reduced flexibility in the model spaces, due to the need of adding the 5p-shell just below the 4f-shell to the intermediate space, in order to achieve convergence. These results are in line with the observations of Shee et al.,60 in that the formal equivalence between EOM-CCSD and IHFS-CCSD for the sectors of Fock space considered depends, in fact, on the flexibility of the main model space.

Table 3 Transition energies for the Yb+ cation. Reference values have been obtained from the NIST database,74 the computed values were obtained for different basis set sizes with Fock-space coupled cluster
State Conf. NIST74 2z 3z 4z Extr. DCB75
2S1/2 4f146s1 0 0 0 0 0 0
image file: d1cp03701c-t36.tif 4f136s2 21[thin space (1/6-em)]419 11[thin space (1/6-em)]087 12[thin space (1/6-em)]390 13[thin space (1/6-em)]618 14[thin space (1/6-em)]514
image file: d1cp03701c-t37.tif 4f136s2 31[thin space (1/6-em)]568 21[thin space (1/6-em)]631 22[thin space (1/6-em)]976 24[thin space (1/6-em)]170 25[thin space (1/6-em)]042
2D3/2 4f145d1 22[thin space (1/6-em)]961 24[thin space (1/6-em)]058 24[thin space (1/6-em)]223 24[thin space (1/6-em)]059 23[thin space (1/6-em)]938 23[thin space (1/6-em)]720
2D5/2 4f145d1 24[thin space (1/6-em)]333 25[thin space (1/6-em)]336 25[thin space (1/6-em)]469 25[thin space (1/6-em)]340 25[thin space (1/6-em)]246 24[thin space (1/6-em)]998
image file: d1cp03701c-t38.tif 4f146p1 27[thin space (1/6-em)]062 27[thin space (1/6-em)]518 27[thin space (1/6-em)]774 27[thin space (1/6-em)]851 27[thin space (1/6-em)]907 27[thin space (1/6-em)]870
image file: d1cp03701c-t39.tif 4f146p1 30[thin space (1/6-em)]392 30[thin space (1/6-em)]934 31[thin space (1/6-em)]241 31[thin space (1/6-em)]316 31[thin space (1/6-em)]371 31[thin space (1/6-em)]312


Furthermore, the removal of the 5p spinors from the main model space underscores the importance of the 5p for the energetics of the states with a hole in the 4f shell, since by doing so, we undress the contributions from the 5p configurations, and thus prevent them from interacting effectively with 4f13 determinants.

3.2 Kramers-restricted configuration interaction potential energy curves

In Fig. 1 the potential energy curves obtained by an approach corresponding to the one used for Yb+ are shown for f14 and f13 configurations. For separate potential energy curves and transition dipole moments we refer the interested reader to the ESI (Fig. S2–S7).
image file: d1cp03701c-f1.tif
Fig. 1 Combination of the sets of KRCI potentials obtained by extrapolating triple and quadruple zeta basis sets. The lowest Ω = 1/2 states are denoted by their dominant configuration.

The potential energy curves have been determined up to 15 Å for the closed and open f-shells. The energy difference in the atom between these states is 21418.75 cm−1. Accordingly, the PECs for the hole states were shifted to obtain this separation at this distance. There is still some interaction between ytterbium and fluorine at 15 Å, but the long range behaviour can be expected to be similar for the two configurations (this assumption was checked, see Fig. S8 in the ESI, for further details). Taking into account the position of the minima, the curvatures, spin–orbit splitting, the avoided crossings and asymptotes the states can be assigned to a dominant configuration, shown in Fig. 1.

Regarding the 4f14 manifold, the lowest two excited states in the figure belong to the Yb(4f146p1)F configuration, but approach asymptotically the 2D3/2 state. The asymptote of the next three states is 2D5/2 corresponding to the Yb(4f145d1)F configuration for smaller internuclear separations. For Ω = 5/2 the transition dipole moment with the ground state is zero, for the other four the values are shown in Fig. S3 in the ESI. The first Ω = 3/2 and the third Ω = 1/2 state have a larger transition dipole moment close to equilibrium, but get close to each other at the largest internuclear separations.

Regarding states of the 4f13 manifold, the lowest four states belonging to the image file: d1cp03701c-t40.tif configuration are well separated from a dense region with a lot of states about 12[thin space (1/6-em)]000 cm−1 higher. Most of these states are of the image file: d1cp03701c-t41.tif configuration, with the image file: d1cp03701c-t42.tif state slightly higher in energy asymptotically and more strongly bound, resulting in several avoided crossings. For each of the four image file: d1cp03701c-t43.tif states the transition dipole moments with higher excited states of varying Ω are plotted in Fig. S6 in the ESI. The transition dipole moments are substantially smaller than the ones for the closed 4f-shells but some of them are non-zero.

An alternative to AOC-SCF for obtaining orbitals for several configurations is multiconfigurational SCF, but similar difficulties as for AOC-SCF in obtaining a balanced description of the 4f14 and 4f13 states are observed: either the wrong ground state is obtained (if only the hole states are optimized in MCSCF), or the hole states are too high in energy by about 20[thin space (1/6-em)]000 cm−1 (if the ground state is optimized). We also made attempts using state-averaged MCSCF in a non-relativistic quantum chemistry code and observed the same difficulties (see dataset67). If the 4f13 configurations are excluded one obtains meaningful results, but at the expense of obtaining a image file: d1cp03701c-t44.tif states too high in energy. If all the states are included, the wrong ground state is obtained.

3.3 Coupled cluster potential energy curves

The potential energy curves of excited states obtained by the equation-of-motion and Fock space methods are displayed in Fig. 2, the values for the complete basis set limit are shown. The basis set dependence in the molecule is similar to the one observed for Yb+: energies for 4f14 states depend only weakly on the basis set, while the gap between the ground state and the excited states corresponding to the image file: d1cp03701c-t45.tif configuration increases upon improving the basis sets.
image file: d1cp03701c-f2.tif
Fig. 2 Potential energy curves obtained by extrapolating triple and quadruple zeta basis sets. EOM-CCSD results in the upper part, IHFS-CCSD in the lower part. The Ω values of 1/2, 3/2, 5/2 and 7/2 are indicated by the colors red, blue, green and orange. The light colors are used for the (0h,1p) sector, dark ones for (1h,0p). For the states below 30[thin space (1/6-em)]000 cm−1 we employ the same color coding and state notation as in Fig. 1.

While the EOM-CCSD excitations energies of Yb+ are closer to the experimental ones, the image file: d1cp03701c-t46.tif states are too high to perturb the Yb(4f146p1/5d1)F PECs. From the extended potential energy curves provided in the ESI (Fig. S15 and S16), we can observe that the ground state of the non-interacting system (Yb(4f14σ26s)F(2p5)) is repulsive and has a high energy at the equilibrium distance. This results in several avoided crossings being observed at 3, 3.5, and 5 Å.

Since the Yb(4f14σ16s)F ground state is accessible for both sectors employed in the coupled cluster calculations (Fock space as well as EOM), we can assess the compatibility of the two separate sets of calculations (in the sense of having comparable accuracies) by looking more closely at the differences between the ground states in Fig. 2. From that, we can see that the EOM-CCSD and IHFS-CCSD approaches the curves are on top of each other from the smallest considered internuclear separation up to about 2.8 Å. This assures us that there should not be artifacts in putting together and comparing the calculations on the two sectors.

3.4 Dissociation and ionization energies

Since the (1h,0p) and (0h,1p) sectors have been considered in our EOM-CCSD and IHFS-CCSD calculations, we have as a by-product of our calculations the ionization potentials (IP) and electron affinities (EA) for YbF for all computed distances. Therefore these quantities are presented first in Table 4, before proceeding to the spectroscopic constants.
Table 4 Ionization potential (IP), electron affinity (EA) and dissociation energy (De) of Yb, F, and YbF. All values in cm−1. They are listed for a quadrupole zeta basis set and a basis set extrapolation. The minimum of the potentials were determined using a Morse fit and used to compute the adiabatic values listed here
Quant. System KRCI EOM-CCSD IHFS-CCSD Experiment
4z Extr. 4z Extr. 4z Extr.
IP Yb 38[thin space (1/6-em)]406 39[thin space (1/6-em)]128 50[thin space (1/6-em)]735 50[thin space (1/6-em)]822 50[thin space (1/6-em)]740 50[thin space (1/6-em)]837 50[thin space (1/6-em)]44374
IP2 Yb 90[thin space (1/6-em)]581 90[thin space (1/6-em)]926 97[thin space (1/6-em)]919 98[thin space (1/6-em)]035 97[thin space (1/6-em)]918 98[thin space (1/6-em)]040 98[thin space (1/6-em)]23274
IP F 127[thin space (1/6-em)]131 126[thin space (1/6-em)]617 144[thin space (1/6-em)]153 143[thin space (1/6-em)]321 144[thin space (1/6-em)]076 144[thin space (1/6-em)]703 140[thin space (1/6-em)]52574
EA F 13[thin space (1/6-em)]326 11[thin space (1/6-em)]978 27[thin space (1/6-em)]279 27[thin space (1/6-em)]740 27[thin space (1/6-em)]246 27[thin space (1/6-em)]759 27[thin space (1/6-em)]43277
IP YbF 58[thin space (1/6-em)]478 56[thin space (1/6-em)]884 48[thin space (1/6-em)]471 48[thin space (1/6-em)]578 48[thin space (1/6-em)]426 49[thin space (1/6-em)]901 47[thin space (1/6-em)]70078
EA YbF 7423 7326 9713 9876 9579 8197
D e(IP) YbF 26[thin space (1/6-em)]059 25[thin space (1/6-em)]887 43[thin space (1/6-em)]824 47[thin space (1/6-em)]782 40[thin space (1/6-em)]591 40[thin space (1/6-em)]931 43[thin space (1/6-em)]26079
D e(EA) YbF 40[thin space (1/6-em)]394 39[thin space (1/6-em)]660 45[thin space (1/6-em)]534 49[thin space (1/6-em)]629 40[thin space (1/6-em)]430 49[thin space (1/6-em)]053 43[thin space (1/6-em)]26079


Unlike coupled cluster calculations, for KRCI a consistent definition of active spaces is difficult, and its lack of size-consistency results in large deviations from experiment and from the coupled cluster values. For adiabatic electron affinities, for which to the best of our knowledge there are no experimental values, the extrapolated values are 8393 and 8197 cm−1 for EOM-CCSD and IHFS-CCSD, respectively. For a distance of 6.5 Å a value of 28[thin space (1/6-em)]651 cm−1 was obtained, which is reasonably close to the electron affinity of fluorine (27[thin space (1/6-em)]432 cm−1).77 Corresponding results for the atoms are listed in the table, which allow to calculate the dissociation energies (De). They deviate from the experimental values of 43[thin space (1/6-em)]600 ± 800 cm−1 by Kaledin et al.78 and 43[thin space (1/6-em)]260 ± 800 cm−1 by Yokozeki and Menzinger.79 The ionization potentials in Table 4 show acceptable agreement with experimental values.

3.5 Spectroscopic constants

The spectroscopic constants for the ground state are now considered. In Table 5 our results are summarized, along those from the literature.
Table 5 Spectroscopic constants for ground state parameters for different approaches. Dissociation energies (De), harmonic frequencies (ωe) and anharmonicity constants (ωeχe) are given in cm−1, the equilibrium bond distances (re) in Å. For the theoretical results we listed the values obtained by extrapolating triple and quadruple zeta basis sets (CBS)
Method Ref. r e ω e ω e χ e D e
KRCI YbF 2.0829 465 2.40 39 660
EOM-CCSD YbF+ 2.0230 511 2.80 49[thin space (1/6-em)]629
YbF 2.0250 508 2.53 47[thin space (1/6-em)]782
IHFS-CCSD YbF+ 2.0176 515 2.82 49[thin space (1/6-em)]053
YbF 2.0159 513 2.42 40[thin space (1/6-em)]931
CCSD1 YbF 2.0174 507.6 2.357 40[thin space (1/6-em)]904
CCSD(T)1 YbF 2.0289 528.2 1.939 41[thin space (1/6-em)]156
CCSD21 2.0127 566.8 3.7885 55[thin space (1/6-em)]650
RASCI6 2.051 529
CCSD(T)80 2.03 38[thin space (1/6-em)]900
CISD81 2.034 502 42[thin space (1/6-em)]100
DFT20 1.987 532 45[thin space (1/6-em)]000
Exp.79 43[thin space (1/6-em)]260
Exp.18 2.0158 506.6674 2.2452
Exp.27 505.5 1.9
Exp.78 43[thin space (1/6-em)]600
Exp.30 2.016514
Exp.24 2.0195 506.616 2.235


We observe that the extrapolated KRCI bond distances, at about 2.058 Å, are significantly longer (by around 0.04 Å) than experiment,30 whereas the coupled cluster calculations show differences from experiment smaller than 0.01 Å, with EOM-CCSD showing slightly larger discrepancies than IHFS-CCSD. Between the extrapolated EOM-CCSD and IHFS-CCSD, we also see small differences between the 4f14 and 4f13 for EOM-CCSD these differ by around 0.001 Å whereas for IHFS-CCSD the difference is slightly under 0.002 Å, with the 4f13 configuration yielding a slightly underestimated value, compared to experiment, something that can be traced back to the differences in model spaces for this configuration.

Our results for harmonic frequencies further indicate that KRCI seems to underestimate the bonding strength in YbF, as the harmonic vibrational frequency is smaller (491 cm−1) than experiment (between 505.5 and 506.7 cm−1 depending on the experiment). The coupled cluster results, on the other hand, show the typical 5–6 cm−1 overestimation of the harmonic frequencies with respect to experiment (something also seen for the anharmonic constants), which can be attributed to lack of triples in the EOM or FS treatment, that would introduce further orbital relaxation. This can be seen in comparison to the unrestricted coupled cluster calculations of Gomes et al.,1 which in spite of the large value of the T1 diagnostic, reproduce well the experimental bond lengths, harmonic frequencies and anharmonic constants.

Taken together, our 2-component CCSD-based calculations and the 4-component ones of Gomes et al.1 compare consistently better to experiment than the other theoretical works for bond lengths, vibrational frequencies and anharmonic constants. For the dissociation energies, on the other hand, the extrapolated calculations presented here do not provide a significant improvement over the results of prior theoretical investigations (quadruple zeta values are closer to the experimental ones for this quantity, see Table 4). Especially, electron attachment values are off, which might be related to the absence of the configuration with a hole in the p orbitals of fluorine.

Moving now to excited states, we start by considering the four lowest excited states, which belong to the image file: d1cp03701c-t47.tif configuration. These states are well separated from the ground state (the lowest excited state is about 10[thin space (1/6-em)]000 cm−1 above the ground state) and higher excited states. That such states are quite well separated from the ground state would, in our view, tend to exclude the interaction with a low-lying excited state as an explanation for the appearance of the large T1 diagnostic values observed by Gomes et al.1 From their spectroscopic constants, presented in Table 6, we see that with the exception of DFT all methods yield similar level splittings of about 500, 1200, and 2000 cm−1. To the best of the authors knowledge there is no experimental data available for these states, due to their negligible transition dipole moments for dipole excitations (see for instance Fig. S6 in the ESI) and small Franck–Condon factors due to the difference in bond distances between these states and the ground state, see Section 3.6.

Table 6 Spectroscopic constants for the lowest excited states image file: d1cp03701c-t48.tif for different wave function methods using the potential energy curves extrapolated to the basis set limit. In the case of KRCI and MRCI19 the ground state is not included in the computation and absolute transition energies are not available. The transition energy (Te), level splitting (Trel, energy relative to 21/2), harmonic frequencies (ωe) and anharmonicity constants (ωeχe) are given in cm−1, the equilibrium bond distances (re) in Å
State Method T e r e ω e ω e χ e T rel
21/2 KRCI 1.9200 631 2.51
EOM-CCSD 12[thin space (1/6-em)]568 1.9432 591 2.59
IHFS-CCSD 9627 1.9396 599 2.79
DFT20 3790 1.9570 561
MRCI19 1.9480 600
13/2 KRCI 1.9253 628 2.50 540
EOM-CCSD 13[thin space (1/6-em)]211 1.9494 588 2.61 643
IHFS-CCSD 10[thin space (1/6-em)]180 1.9438 595 2.79 553
DFT20 9520 1.9440 597 5730
MRCI19 1.951 0 598 428
15/2 KRCI 1.9296 622 2.45 1223
EOM-CCSD 13[thin space (1/6-em)]703 1.9553 582 2.61 1135
IHFS-CCSD 10[thin space (1/6-em)]968 1.9493 589 2.78 1341
DFT20 10[thin space (1/6-em)]970 1.9360 598 7180
MRCI19 1.9540 594 1021
17/2 KRCI 1.9315 616 2.43 1933
EOM-CCSD 14[thin space (1/6-em)]685 1.9556 577 2.62 2117
IHFS-CCSD 11[thin space (1/6-em)]645 1.9496 583 2.77 2018
DFT20 16[thin space (1/6-em)]530 1.936 592 12[thin space (1/6-em)]740
MRCI19 1.954 589 1709


The smallest equilibrium distance was obtained for the 21/2 state with 1.94 Å for the coupled cluster methods and 0.02 Å less for KRCI. The vibrational frequencies are between 570 and 600 cm−1 for the coupled cluster methods and about 30 cm−1 higher for KRCI.

For higher excited states, as apparent from the figures in the previous section, the identification and assignment of states gets more difficult and there are differences between the methods. We have nevertheless provided in Table 7 the spectroscopic constants for excited with Ω values of 1/2, 3/2, and 5/2, respectively.

Table 7 Spectroscopic constants for selected excited states (complete list presented in Table S17 in the ESI) with Ω = 1/2, 3/2, 5/2, starting from 18[thin space (1/6-em)]000 cm−1 for different methods using the values after extrapolation to the basis set limit. Transition energy (Te), vibrational constant (ωe), and anharmonicty constant (ωeχe) are given in cm−1, the equilibrium bond distance (re) in Å. Experimental transitions that were not assigned (n.a.) are also listed. Labels of experimental results are defined in the introduction
Ω Method State Configuration T e r e ω e ω e χ e
a KRCI transition energies for the 4f13 sector were obtained by adding 4144 cm−1, an estimate for the energy of the lowest state in this manifold.
1/2 KRCIa 3 4f13σ26s 15[thin space (1/6-em)]572 1.9038 655 13.57
5 4f146p1 16[thin space (1/6-em)]189 2.0504 496 2.38
10 4f145d1 19[thin space (1/6-em)]631 2.0552 490 2.49
EOM-CCSD 3 4f146p1 18[thin space (1/6-em)]373 2.0004 536 2.72
4 4f145d1 21[thin space (1/6-em)]448 2.0079 532 2.78
6 4f13σ26s 23[thin space (1/6-em)]241 1.9432 582 4.06
IHFS-CCSD 3 4f146p1 18[thin space (1/6-em)]249 1.9953 539 2.63
4 4f13σ26s 20[thin space (1/6-em)]267 1.9397 597 2.78
5 4f145d1 21[thin space (1/6-em)]375 2.0032 533 2.73
MRCI19 4f13σ26s 1.948 600
Exp.18[thin space (1/6-em)]a 3 18106.20 537 3
Exp.18 4 [18.6]1/2 18705.06
Exp.24 [557] 18[thin space (1/6-em)]574 1.9656 502.15
Exp.24 [561] 18[thin space (1/6-em)]699 1.9571
3/2 KRCIa 2 4f13σ26s 16[thin space (1/6-em)]206 1.9331 711 8.45
4 4f146p1 17[thin space (1/6-em)]123 2.0473 499 2.37
>7 4f145d1 24[thin space (1/6-em)]583 2.0669 470 2.50
EOM-CCSD 2 4f146p1 19[thin space (1/6-em)]672 1.9971 540 2.72
4 4f13σ26s 24[thin space (1/6-em)]251 1.9537 584 2.64
5 4f145d1 24[thin space (1/6-em)]468 2.0177 509 2.78
IHFS-CCSD 2 4f146p1 19[thin space (1/6-em)]543 1.9920 542 2.63
3 4f13σ26s 21[thin space (1/6-em)]222 1.9480 591 2.80
4 4f145d1 24[thin space (1/6-em)]363 2.0120 512 2.73
MRCI19 4f13σ26s 1.953 596
Exp.18 2 19471.49
5/2 KRCIa 3 4f13σ26s 17[thin space (1/6-em)]063 1.9302 635 1.41
>6 4f145d1 24[thin space (1/6-em)]744 2.0639 474 2.48
EOM-CCSD 2 4f13σ26s 24[thin space (1/6-em)]957 1.9536 577 2.62
3 4f145d1 25[thin space (1/6-em)]023 2.0146 513 2.80
IHFS-CCSD 2 4f13σ26s 22[thin space (1/6-em)]127 1.9499 584 2.77
3 4f145d1 24[thin space (1/6-em)]919 2.0089 515 2.77
MRCI19 4f13σ26s 1.954 590
n.a. Exp.26 [574] 19[thin space (1/6-em)]150
Exp.26 [578] 19[thin space (1/6-em)]280
Exp.27 C1 23035.3 523 2
Exp.27 C2 23256.0 507 2
Exp.27 D 26014.8 574.6 2.8


The comparison with experimental results allows assignment of the lowest excited state reported in experiments and give some indications for higher states. The lowest Ω = 1/2 state observed in experiment can be identified as the 31/2 state (51/2 for KRCI). Spectroscopic parameters agree well with the ones obtained by fitting to the A2Π1/2 in experiments. A bond distance of 1.9935 Å obtained by fitting to the same states in ref. 34 agrees well with the coupled cluster values for the 31/2 state, the vibrational constant of about 540 cm−1 is close to the experimental value of ref. 18. Similarly, the lowest Ω = 3/2 state reported by Dunfield et al.18 can be identified as the 23/2 state (43/2 for KRCI), see Table 7.

The lowest states with Ω = 1/2 and Ω = 3/2 for this energy range approach asymptotically a state with a Yb(4f145d1)F configuration, but if one analyses the EOM-CCSD and IHFS-CCSD orbital composition, significant contributions of the atomic 6p are identified. The Ω = 3/2 state is dominated (97%) by a single configuration, corresponding to a HOMO(σ6s,1/2) → LUMO+1 where the latter is made up of a mixture of 6pπ and 5dπ orbitals (the 6pπ contributions being the dominant − ≃80% – in the reference YbF+ orbitals). The few other significant configurations arise from excitations to higher-lying orbitals with increasingly large (≥50% 5dπ) contributions. The Π1/2 state is also dominated by a single configuration, now corresponding to a HOMO(σ1/2) → LUMO transition, and shows a rather similar picture in terms of the relative weights of the 6pπ and 5dπ orbitals, with very small contributions from the ground-state mixing due to spin–orbit coupling. The splitting in Yb+ of 2D3/2 and 2D5/2 is 1372 cm−1, for 2P1/2 and 2P3/2 it is 3330 cm−1. The separation between the lowest excited Ω = 1/2 states in the closed shell computation is 3779 and 3126 cm−1 for EOM-CCSD and IHFS-CCSD, respectively. This is an indication that the two states must be regarded as a fairly strong admixture of 6p and 5d orbitals of j = 1/2 or 3/2, as one can expect a much smaller spin–orbit splitting in the axial field of the molecule (about 1/3 of the atomic spin–orbit splitting for the P state). This picture also finds experimental support in recent measurements of hyperfine constants (d and eq0Q) for the ground and Π1/2 excited state of YbF9, where a simple ligand-field model disregarding the contributions from 5dπ orbitals predicted values of d a factor of 2 larger than the measurements. For bond distances much larger that the equilibrium one the system gets closer to the configurations in Yb+ with a dominating 5d contribution.

As already mentioned this energy range above 18[thin space (1/6-em)]000 cm−1 is dense with a large number of excited states that can mix with each other and result in new mixed states, like the [557] and [561] ones.24 These will be addressed in Section 3.6.

Uttam et al.27 reported three unidentified states with energies above 22[thin space (1/6-em)]000 cm−1 which are listed in Table 7 and cannot uniquely be identified with the current results. The one at 26014.8 cm−1 has a larger vibrational constant indicating a more strongly bound state, possibly of the 4f13σ26s configuration. The vibrational spacing of the two states at 23[thin space (1/6-em)]000 cm−1 rather points to states with a closed f shell.

3.6 Perturbation of the 31/2 excited state

Due to the use of different sectors of Fock space to obtain the 4f14 and 4f13 configurations, the excited states with the same Ω values cannot interact among themselves, as is the case within each sector. However, from the discussion above, it is clear that dealing with states which are artificially prevented from interaction makes it difficult to establish a comparison to experiment, for states from about 18[thin space (1/6-em)]000 cm−1 to about 26[thin space (1/6-em)]000 cm−1, which is where these configurations should be the most entangled. In order to remedy that, in the following we introduce a simple adiabatization model (eqn (2)) that allows us to investigate how coupling such states would affect the overall spectra in the aforementioned energy region.

In the following we only consider the IHFS-CCSD potential energy curves, as the spectroscopic parameters are more reliable for CCSD than for KRCI. The coupled cluster results for the two methods are quite similar, and FS-CCSD was selected (because it does not include the (2h,1p) and (1h,2p) transitions with rather large uncertainties). Fig. 3 contains the original FS-CCSD curves as well as the ones obtained after adiabatzation with three different coupling constants. Looking at the potential energy curves for this energetic region, there are two Ω = 1/2 and two Ω = 3/2 states of Yb(4f146p1)F and Yb(4f145d1)F configurations originating from the (0h,1p) sector. For both Ω values there is an additional state with a image file: d1cp03701c-t49.tif configuration stemming from the (1h,0p) sector. By looking at the KRCI results one expects additional states belonging to the image file: d1cp03701c-t50.tif configuration for this energy range, which will not be included in the current considerations.


image file: d1cp03701c-f3.tif
Fig. 3 Frank-Condon factors before and after adiabatization for the IHFS-CCSD potential energy curves. C is the coupling strength in Hartree. The lowest 10 vibrational levels of the ground state as well as the lowest 60 vibrational levels of the excited state were computed using the LEVEL program.82 The experimental values18,24,26 have been added as straight lines, the labels are defined in the introduction and Table 7.

As already mentioned earlier the lowest Ω = 1/2 and Ω = 3/2 states can be identified clearly and assigned to experimental observations. There are several experimental states in this energy region attributed to the mixing of states. The [557] and [561] ones24 are assumed to arise from a mixing of the 31/2 and 41/2. The vibrational constant of the perturbing state (41/2) was estimated to be 605 cm−1 in ref. 18. This agrees with the 41/2 state in Fig. 3 with a image file: d1cp03701c-t51.tif configuration, see also Table 7. [574] and [578]26 have not been identified and since their Ω value is unknown, we were not able to assign them to a configuration.

Next we take a look at the changes introduced by adiabatization. For small and intermediate coupling strengths there are no major differences in the potential energy curves, although close to the crossing points the potentials are deformed. Intermediate coupling strengths with slightly deformed potentials close to the avoided crossings will be the most realistic description. For very large coupling strengths one obtains parallel potential energy curves due to the strong repulsion. This also results in a major change of the spectra above 19[thin space (1/6-em)]000 cm−1. One of the differences between the adiabatic spectrum and the upermost one in Fig. 3 is that the Frank-Condon factors of the 41/2 state, which is of the Yb(4f13σ26s)F configuration, are now noticeable and the spacing of the energy levels of the 51/2 is changed. Similarly, transitions belonging to the 51/2 appear.

The influence of adiabatization on spectroscopic parameters can be investigated by comparing spectroscopic constants calculated for the IHFS-CCSD curves without and with a coupling of 100 cm−1 (Table 8). We observe that for this coupling strength, there are small but non-negligible changes for the excitation energies, harmonic frequencies and anharmonicity constants, for all but the fourth and fifth Ω = 1/2 states; there, the coupling does seem to significantly change the anharmonicity constants. Equilibrium distances, on the other hand, are largely unperturbed in all cases. Furthermore, as expected from the preceding discussion, no changes are observed for the ground-state, since it is too separated in energy from the other electronic states.

Table 8 Spectroscopic data obtained by fitting Morse potentials to the lowest points of the potential energy curves obtained with FSCC for the extrapolated basis set (CBS). This table combines results from both sectors starting either with a closed (f14) or open (f13) f-shell. Additionally, the table contains spectroscopic parameters after adiabatization with a specific coupling constant (C). The transition energy (Te), vibrational constant (ωe), and anharmonicty constant (ωeχe) are given in cm−1, the equilibrium bond distance (re) in Å
Ω CBS C = 100 cm−1
State r e ω e ω e χ e T e State r e ω e ω e χ e T e
1/2 f14 – 1 2.018 515 2.9 0 1 2.018 515 2.8 0
f13 – 2 1.940 599 2.8 9627 2 1.940 599 2.8 9617
f14 – 2 1.995 539 2.6 18[thin space (1/6-em)]249 3 1.995 538 2.6 18[thin space (1/6-em)]247
f13 – 3 1.940 597 2.8 20[thin space (1/6-em)]267 4 1.935 603 8.6 20[thin space (1/6-em)]258
f14 – 3 2.003 533 2.7 21[thin space (1/6-em)]375 5 2.002 586 0.4 21[thin space (1/6-em)]359
f14 – 4 1.964 581 1.8 31[thin space (1/6-em)]416 6 1.964 581 1.8 31[thin space (1/6-em)]419
3/2 f13 – 1 1.944 595 2.8 10[thin space (1/6-em)]180 1 1.944 594 2.8 10[thin space (1/6-em)]170
f14 – 1 1.992 542 2.6 19[thin space (1/6-em)]543 2 1.992 542 2.6 19[thin space (1/6-em)]540
f13 – 2 1.948 591 2.8 21[thin space (1/6-em)]222 3 1.948 591 2.8 21[thin space (1/6-em)]217
f14 – 2 2.012 512 2.7 24[thin space (1/6-em)]363 4 2.012 512 2.7 24[thin space (1/6-em)]369
5/2 f13 – 1 1.949 589 2.8 10[thin space (1/6-em)]967 1 1.949 589 2.8 10[thin space (1/6-em)]960
f13 – 2 1.950 584 2.8 22[thin space (1/6-em)]127 2 1.950 583 2.8 22[thin space (1/6-em)]117
f14 – 1 2.009 515 2.8 24[thin space (1/6-em)]919 3 2.009 516 2.7 24[thin space (1/6-em)]926


4 Conclusion

In this manuscript we have presented a study of the ground and excited states of the YbF molecule, with 2-component multireference CI, equation-of-motion and Fock space coupled cluster approaches (in all cases, performing extrapolations to the complete basis set limit). In particular, we have focused on obtaining electronic states up to around 24[thin space (1/6-em)]000 cm−1 arising from configurations which differ in the occupation of the 4f shell (4f14 and 4f13), which are very difficult to treat on the same footing due to a number of subtle correlation and relaxation effects.

In order to achieve such a balanced description, our strategy consisted of starting from YbF+ and YbF, in order to arrive at the wavefunctions for YbF through the (1h,0p) and (0h,1p) sectors of Fock space. Once obtained, electronic states with same Ω values coming from these different sectors are further coupled through a simple adiabatization model in which the coupling strength is taken as a constant.

As a general rule we find that the CI calculations do capture the essential physics of the system, though they are not as reliable as the coupled cluster approaches for excitation energies, bond lengths, harmonic vibrational frequencies and anharmonic constants. In effect, the coupled cluster calculations for the (1h,0p) and (0h,1p) sectors yield the same potential energy curves for the ground state, for internuclear distances up to around 2.8 Å, which is sufficient to capture the bound regions of all states under consideration,

We have determined that the lowest lying excited states arise from the image file: d1cp03701c-t52.tif configuration, with transition energies of around 10[thin space (1/6-em)]000 cm−1, and a splitting about 2000 cm−1. These states are, however, not generally accessible in experiment due to their low dipolar intensity and significantly shifted minima of the potential energy curve resulting in small Frank-Condon factors.

The next set of states, coming above 18[thin space (1/6-em)]000 cm−1, arise from the Yb(4f146p1)F, Yb(4f145d1)F, image file: d1cp03701c-t53.tif and image file: d1cp03701c-t54.tif configurations. Among these, the image file: d1cp03701c-t55.tif configurations generally display the shortest equilibrium distances and deepest potential well, while the Yb(4f145d1)F and Yb(4f146p1)F configurations exhibits the largest bond distances and smallest harmonic frequencies, with the other configurations falling somewhere in between. The lowest Ω = 1/2 and Ω = 3/2 states of this group show a Yb(4f146p1)F orbital composition around the ground-state equilibrium structure, though for longer bond lengths they asymptotically approach the Yb(4f145d1)F configuration.

We note that configurations with three unpaired electrons, such as image file: d1cp03701c-t56.tif were only considered with the KRCI method, which has larger uncertainties. This only allows us to make some qualitative statements, e.g. that their bond distances and vibrational constant should be between the values for the other configurations and that they should be higher in energy than the lowes excited Yb(4f146p1)F and image file: d1cp03701c-t57.tif states.

A simple method was applied in order to adiabatize the curves obtained for different sectors and reference wave functions. It was applied to potential energy curves between 18[thin space (1/6-em)]000 and 26[thin space (1/6-em)]000 cm−1 and small changes of the Franck–Condon factors were observed. The influence on spectroscopic constant was minor, with the exception of the asymmetry constant for two states. However, the approximation introduced (same coupling strength for all states and all geometries) is perhaps not flexible enough, and more sophisticated models should be investigated.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

L. V. and J. V. P. wish to thank The Netherlands Organization for Scientific Research (NWO) for financial support via the ECHO and computer time. J. V. P. acknowledges funding by the Austrian Science Fund (FWF):J 4177-N36. ASPG acknowledges support from PIA ANR project CaPPA (ANR-11-LABX-0005-01), the Franco-German project CompRIXS (Agence nationale de la recherche ANR-19-CE29-0019, Deutsche Forschungsgemeinschaft JA 2329/6-1), I-SITE ULNE projects OVERSEE, the French Ministry of Higher Education and Research, region Hauts de France council and European Regional Development Fund (ERDF) project CPER CLIMIBIO, and the French national supercomputing facilities (grants DARI A0070801859 and A0090801859). A. S. P. G., L. H., J. P. and L. V. acknowledge support from MESONM International Associated Laboratory (LAI) (ANR-16-IDEX-0004).

References

  1. A. S. P. Gomes, K. G. Dyall and L. Visscher, Theor. Chem. Acc., 2010, 127, 369–381 Search PubMed.
  2. J. Sapirstein, in Theoretical and Computational Chemistry, ed. P. Schwerdtfeger, Elsevier, 2002, vol. 11, pp. 468–522 Search PubMed.
  3. R. Berger, in Theoretical and Computational Chemistry, ed. P. Schwerdtfeger, Elsevier, 2004, vol. 14, pp. 188–288 Search PubMed.
  4. A. Sunaga, M. Abe, M. Hada and B. P. Das, Phys. Rev. A, 2016, 93, 042507 CrossRef.
  5. F. A. Parpia, J. Phys. B: At., Mol. Opt. Phys., 1998, 31, 1409–1430 CrossRef CAS.
  6. M. K. Nayak and R. K. Chaudhuri, Chem. Phys. Lett., 2006, 419, 191–194 CrossRef CAS.
  7. A. V. Titov, N. S. Mosyagin, A. N. Petrov and T. A. Isaev, Int. J. Quantum Chem., 2005, 104, 223–239 CrossRef CAS.
  8. J. J. Hudson, B. E. Sauer, M. R. Tarbutt and E. A. Hinds, Phys. Rev. Lett., 2002, 89, 023003 CrossRef CAS PubMed.
  9. T. C. Steimle, T. M. Ma and C. Linton, J. Chem. Phys., 2007, 127, 234316 CrossRef PubMed.
  10. M. Fukuda, K. Soga, M. Senami and A. Tachibana, Phys. Rev. A, 2016, 93, 012518 CrossRef.
  11. N. J. Fitch, J. Lim, E. A. Hinds, B. E. Sauer and M. R. Tarbutt, Quantum Sci. Technol., 2020, 6, 014006 CrossRef.
  12. Y. Takasu, K. Maki, K. Komori, T. Takano, K. Honda, M. Kumakura, T. Yabuzaki and Y. Takahashi, PRL, 2003, 91, 040404 CrossRef PubMed.
  13. T. Fukuhara, S. Sugawa, Y. Takasu and Y. Takahashi, PRA, 2009, 79, 021601 CrossRef.
  14. P. Tecmer, K. Boguslawski, M. Borkowski, P. S. Żuchowski and D. Kedziera, Int. J. Quantum Chem., 2019, 119, e25983 CrossRef.
  15. K. Tsigutkin, D. Dounas-Frazer, A. Family, J. E. Stalnaker, V. V. Yashchuk and D. Budker, PRL, 2009, 103, 071601 CrossRef CAS PubMed.
  16. L. F. Pasteka, R. J. Mawhorter and P. Schwerdtfeger, Mol. Phys., 2016, 114, 1110–1117 CrossRef CAS.
  17. K. Huber and G. Herzberg, in Constants of Diatomic Molecules, National Institute of Standards and Technology, ed. P. J. Linstrom and W. G. Mallard, Gaithersburg MD, 20899, 2005, (data prepared by J.W. Gallagher and R.D. Johnson, III) in NIST Chemistry WebBook, NIST Standard Reference Database, Number 69 Search PubMed.
  18. K. L. Dunfield, C. Linton, T. E. Clarke, J. McBride, A. G. Adam and J. R. D. Peers, J. Mol. Spectrosc., 1995, 174, 433–445 CrossRef CAS.
  19. M. Dolg, H. Stoll and H. Preuss, Chem. Phys., 1992, 165, 21–30 CrossRef CAS.
  20. W. Liu, M. Dolg and L. Li, J. Chem. Phys., 1998, 108, 2886–2895 CrossRef CAS.
  21. T. Su, C. L. Yang, X. Q. Wang, F. J. Bai and M. S. Wang, Chem. Phys. Lett., 2009, 467, 265–269 CrossRef CAS.
  22. Z. Glassman, R. Mawhorter, J. U. Grabow, A. Le and T. C. Steimle, J. Mol. Spectrosc., 2014, 300, 7–11 CrossRef CAS.
  23. B. E. Sauer, S. B. Cahn, M. G. Kozlov, G. D. Redgrave and E. A. Hinds, J. Chem. Phys., 1999, 110, 8424–8428 CrossRef CAS.
  24. J. Lim, J. R. Almond, M. R. Tarbutt, D. T. Nguyen and T. C. Steimle, J. Mol. Spectrosc., 2017, 338, 81–90 CrossRef CAS.
  25. J. Lim, J. R. Almond, M. A. Trigatzis, J. A. Devlin, N. J. Fitch, B. E. Sauer, M. R. Tarbutt and E. A. Hinds, Phys. Rev. Lett., 2018, 120, 123201 CrossRef CAS PubMed.
  26. I. J. Smallman, F. Wang, T. C. Steimle, M. R. Tarbutt and E. A. Hinds, Spectroscopic Tests of Fundamental Physics, 2014, 300, 3–6 CAS.
  27. K. N. Uttam and M. M. Joshi, J. Mol. Spectrosc., 1995, 174, 290–296 CrossRef CAS.
  28. H. U. Lee and R. N. Zare, J. Mol. Spectrosc., 1977, 64, 233–243 CrossRef.
  29. V. A. Dzuba and A. Derevianko, J. Phys. B: At., Mol. Opt. Phys., 2010, 43, 074011 CrossRef.
  30. C. S. Dickinson, J. A. Coxon, N. R. Walker and M. C. L. Gerry, J. Chem. Phys., 2001, 115, 6979–6989 CrossRef CAS.
  31. T. C. Steimle, T. M. Ma and C. Linton, J. Chem. Phys., 2008, 128, 209903 CrossRef.
  32. T. C. Steimle, T. M. Ma and C. Linton, J. Chem. Phys., 2012, 137, 109901 CrossRef.
  33. H. Wang, A. T. Le, T. C. Steimle, E. A. C. Koskelo, G. Aufderheide, R. Mawhorter and J.-U. Grabow, Phys. Rev. A, 2019, 100, 022516 CrossRef CAS.
  34. T. Ma, C. Butler, J. M. Brown, C. Linton and T. C. Steimle, J. Phys. Chem. A, 2009, 113, 8038–8044 CrossRef CAS PubMed.
  35. S. G. Porsev, Y. G. Rakhlina and M. G. Kozlov, J. Phys. B: At., Mol. Opt. Phys., 1999, 32, 1113–1120 CrossRef CAS.
  36. B. K. Mani and D. Angom, PRA, 2011, 83, 012501 CrossRef.
  37. M. K. Nayak and R. K. Chaudhuri, Eur. Phys. J. D, 2006, 37, 171–176 CrossRef CAS.
  38. D. Naik, M. Sikarwar, M. K. Nayak and S. K. Ghosh, J. Phys. B: At., Mol. Opt. Phys., 2014, 47, 225103 CrossRef.
  39. P. A. B. Haase, E. Eliav, M. Iliaš and A. Borschevsky, J. Phys. Chem. A, 2020, 124, 3157–3169 CrossRef CAS PubMed.
  40. V. Vallet, L. Maron, C. Teichteil and J.-P. Flament, J. Chem. Phys., 2000, 113, 1391–1402 CrossRef CAS.
  41. C. M. Marian, Spin-Orbit Coupling in Molecules, John Wiley & Sons, Ltd, 2001, ch. 3, pp. 99–204 Search PubMed.
  42. C. Danilo, V. Vallet, J.-P. Flament and U. Wahlgren, Phys. Chem. Chem. Phys., 2010, 12, 1116–1130 RSC.
  43. A. Farhat and S. N. Abdul-Al, J. Comput. Chem., 2015, 36, 1252–1258 CrossRef CAS PubMed.
  44. W. Chmaisani and M. Korek, J. Quant. Spectrosc. Radiat. Transfer, 2018, 217, 63–72 CrossRef CAS.
  45. W. Chmaisani, N. El-Kork, S. Elmoussaoui and M. Korek, ACS Omega, 2019, 4, 14987–14995 CrossRef CAS PubMed.
  46. S. Kervazo, F. Réal, F. Virot, A. Severo Pereira Gomes and V. Vallet, Inorg. Chem., 2019, 58, 14507–14521 CrossRef CAS PubMed.
  47. W. Liu, J. Chem. Phys., 2020, 152, 180901 CrossRef CAS PubMed.
  48. M. K. Nayak and R. K. Chaudhuri, Pramana J. Phys., 2009, 73, 581–586 CrossRef CAS.
  49. M. K. Nayak and R. K. Chaudhuri, Phys. Rev. A: At., Mol., Opt. Phys., 2011, 83, 022504 CrossRef.
  50. H. J. A. Jensen, 2005, Douglas-Kroll the Easy Way, Talk at Conference on Relativistic Effects in Heavy Elements - REHE, Mülheim, Germany, April, 2005. Available at  DOI:10.6084/m9.figshare.12046158.
  51. W. Kutzelnigg and W. Liu, J. Chem. Phys., 2005, 123, 241102 CrossRef PubMed.
  52. W. Liu and D. Peng, J. Chem. Phys., 2006, 125, 044102 CrossRef PubMed.
  53. D. Peng, W. Liu, Y. Xiao and L. Cheng, J. Chem. Phys., 2007, 127, 104106 CrossRef PubMed.
  54. M. Ilias and T. Saue, J. Chem. Phys., 2007, 126, 064102 CrossRef PubMed.
  55. J. Sikkema, L. Visscher, T. Saue and M. Iliaš, J. Chem. Phys., 2009, 131, 124116 CrossRef PubMed.
  56. W. Liu and D. Peng, J. Chem. Phys., 2009, 131, 031104 CrossRef PubMed.
  57. L. Konecny, M. Kadek, S. Komorovsky, O. L. Malkina, K. Ruud and M. Repisky, J. Chem. Theory Comput., 2016, 12, 5823–5833 CrossRef CAS PubMed.
  58. B. A. Hess, C. M. Marian, U. Wahlgren and O. Gropen, Chem. Phys. Lett., 1996, 251, 365–371 CrossRef CAS.
  59. B. Schimmelpfennig, AMFI, an atomic mean-field spin–orbit integral program, University of Stockholm, Stockholm, Sweden, 1999 Search PubMed.
  60. A. Shee, T. Saue, L. Visscher and A. Severo Pereira Gomes, J. Chem. Phys., 2018, 149, 174113 CrossRef PubMed.
  61. L. Halbert, M. Lopez Vidal, A. Shee, S. Coriani and A. S. P. Gomes, J. Chem. Theory Comput., 2021, 17, 3583–3598 CrossRef CAS PubMed.
  62. R. J. Bartlett and M. Musiał, Rev. Mod. Phys., 2007, 79, 291–352 CrossRef CAS.
  63. T. Saue, L. Visscher, H. J. A. Jensen, R. Bast and A. S. P. Gomes, DIRAC, a relativistic ab initio electronic structure program, 2019, with new contributions from R. Bast, S. Dubillard, K. G. Dyall, U. Ekström, E. Eliav, T. Fleig, A. S. P. Gomes, T. U. Helgaker, J. Henriksson, M. Iliaš, Ch. R. Jacob, S. Knecht, P. Norman, J. Olsen, M. Pernpointner, K. Ruud, P. Sałek and J. Sikkema; see http://www.diracprogram.org.
  64. T. H. Dunning, J. Chem. Phys., 1989, 90, 1007–1023 CrossRef CAS.
  65. T. Helgaker, P. Joergensen and J. Olsen, Molecular Electronic-Structure Theory, John Wiley & Sons, 2000 Search PubMed.
  66. L. Visscher, Theor. Chem. Acc., 1997, 98, 68–70 Search PubMed.
  67. J. V. Pototschnig, K. G. Dyall, L. Visscher and A. Severo Pereira Gomes, Dataset: Electronic Spectra of Ytterbium Fluoride from Relativistic Electronic Structure Calculations, 2021 DOI:10.5281/zenodo.5121372.
  68. J. Thyssen, PhD thesis, University of Southern Denmark, 2001.
  69. T. Fleig, H. J. A. Jensen, J. Olsen and L. Visscher, J. Chem. Phys., 2006, 124, 104106 CrossRef PubMed.
  70. L. Visscher, E. Eliav and U. Kaldor, J. Chem. Phys., 2001, 115, 9720–9726 CrossRef CAS.
  71. A. Landau, E. Eliav, Y. Ishikawa and U. Kaldor, J. Chem. Phys., 2000, 113, 9905–9910 CrossRef CAS.
  72. A. Landau, E. Eliav, Y. Ishikawa and U. Kaldor, J. Chem. Phys., 2001, 115, 6862–6865 CrossRef CAS.
  73. E. Eliav, M. J. Vilkas, Y. Ishikawa and U. Kaldor, J. Chem. Phys., 2005, 122, 224113 CrossRef PubMed.
  74. A. Kramida, Yu Ralchenko, J. Reader and NIST ASD Team, NIST Atomic Spectra Database (ver. 5.8), [Online]. Available: https://physics.nist.gov/asd [2017, April 9]. National Institute of Standards and Technology, Gaithersburg, MD, 2020.
  75. E. Eliav, U. Kaldor and Y. Ishikawa, PRA, 1995, 52, 291–296 CrossRef CAS PubMed.
  76. C. Sur and R. K. Chaudhuri, PRA, 2007, 76, 012509 CrossRef.
  77. C. Blondel, C. Delsart and F. Goldfarb, J. Phys. B: At., Mol. Opt. Phys., 2001, 34, L281–L288 CrossRef CAS.
  78. L. A. Kaledin, M. C. Heaven and R. W. Field, J. Mol. Spectrosc., 1999, 193, 285–292 CrossRef CAS PubMed.
  79. A. Yokozeki and M. Menzinger, Chem. Phys., 1976, 14, 427–439 CrossRef CAS.
  80. H. Heiberg, O. Gropen, J. K. Laerdahl, O. Swang and U. Wahlgren, Theor. Chem. Acc., 2003, 110, 118–125 Search PubMed.
  81. X. Y. Cao, W. J. Liu and M. Dolg, Sci. China, Ser. B: Chem., 2002, 45, 91–96 CrossRef CAS.
  82. R. J. Le Roy, J. Quant. Spectrosc. Radiat. Transfer, 2017, 186, 167–178 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: A supplementary pdf file is available containing orbitals, additional information about dipole moments and additional Franck–Condon factors. See DOI: 10.1039/d1cp03701c

This journal is © the Owner Societies 2021