1. Introduction
The spectroscopy of polar diatomic molecules, such as magnesium monofluoride (MgF) has a long history and led, in the late 1960s, to results with remarkable accuracy. Barrow and Beale [
1] resolved the rotational structure of gaseous MgF and carried out a detailed analysis resulting in the determination of molecular constants for the ground and first three electronic excited states with precise values based on the two lowest vibrational states (
). Higher excitations were measured and reported in 1971 by Novikov and Gurvich [
2]. The detailed structure of the
transitions including spin-orbit effects (the
doublet is split by about 35
) focused on the nature of the
-type doubling, and it was concluded that the doublet was inverted (negative
A parameter). A recent study [
3] does, however, conclude that the
doublet is normal, i.e., not inverted. Hyperfine-resolved measurements are also reported in Ref. [
4].
The interest in the laser cooling of MgF has resulted in a detailed study of potential energy curves and spectroscopic constants at the level of the multi-reference configuration interaction (MRCI) [
5]. These results can serve the present study as a theoretical benchmark, in addition to the experimental results [
6]. The ground-state potential energy curve can be determined also from infrared spectroscopy [
7] which leads to an extended Morse oscillator description of this curve. This allows us to focus directly on the calculation of electronic excitations. These studies (and the present work) ignore spin-orbit effects. Such effects can be calculated in some standard packages, e.g., in Molpro in the Breit–Pauli approximation on the basis of MRCI calculations [
8]. Hyperfine constants can be obtained using two-component density functional theory [
9] as implemented in TURBOMOLE [
10].
Our interest in studying the accuracy of excited-state calculations within coupled-cluster (CC) theory relates to experimental work on matrix-isolated barium monofluoride (BaF) [
11]. In this context earlier work on matrix-isolated atomic barium [
12,
13] was extended to the case of BaF in argon [
14,
15] and neon [
16]. The matrix isolation work is motivated by performing measurements of the electron electric dipole moment. While the BaF molecule is much more complicated than MgF, the use of effective core potentials makes the computational effort comparable for the two cases. The purpose of the present work is to establish expectations on accuracy with respect to methodology and basis sets in order to understand how reliable the CC methods may be in the matrix isolation environment.
The current status of describing the electronic excitations of gas-phase BaF (and its lighter homologs CaF and SrF) can be described as follows. Results from scalar relativistic CC and MRCI calculations are reported in Ref. [
17]. The all-electron Fock-space CC results for the excitation energies (which include spin-orbit splittings on the order of 400–600
) agree with the experiment at the level of about 200
. Better agreement (down to about 10
) was obtained in Ref. [
18] by including QED effects and basis set extrapolation of correlation effects calculated at the level of CCSDT(Q). A similar approach based on the four-component CC Dirac theory resulted in accurate ionization energies [
19]. Hyperfine constants for
ground and excited-state BaF molecules were obtained in Ref. [
20] and provide some confidence in the quality of the Fock-space CC wavefunctions.
To return to the topic of the present paper we note that in Ref. [
21] radiative decay rates and branching fractions of the first excited
state of
24Mg
19F were reported in a comparative experimental-theoretical work. While the theory is the relativistic Fock-space CC method, a comparison can be made with the present non-relativistic work due to the smallness of spin-orbit effects in MgF. The experimental data presented in Ref. [
21] deviate somewhat (at the
level) from the earlier work [
1]. Studies which investigate equation-of-motion (EOM) coupled cluster methods have been carried out and are typically benchmarked against full configuration interaction calculations, but the goal of this work is to compare only the theoretical
spectroscopic parameters to experimental determinations.
The layout of the paper is as follows: We begin in
Section 2 with a short summary of how to use CC methods for electronic excitations. We show an example of high-quality potential energy curves for the
and
states to demonstrate the accuracy obtainable for spectroscopic constants and to explain the difference between the vertical and adiabatic excitation energies denoted as
and
, respectively. In
Section 3 we present our results for vertical excitation energies
of MgF calculated at a fixed ground-state equilibrium separation
using three methods and two basis set families, and compare them with experimentally determined values, and with the relativistic Fock-space CC results of Ref. [
20]. We draw some conclusions in
Section 4.
2. Methodology
Electronic excitation energies for molecules can be calculated in many quantum chemistry packages with a selection of methods. For the purpose of the present work, the accuracy of time-dependent density functional theory methods is not sufficient, since we would like to reach a percent-level agreement with measured values or better. This requires the inclusion of electron correlation with a systematic treatment as provided by configuration interaction or coupled cluster methodology. The correlation energies are on the order of 0.3 Hartree (a.u.) as compared to a total energy of about 300 a.u. The excitation energies considered in this work are in the order of 0.2 a.u. (differences of total energies), and this makes it necessary to adopt a proper treatment of correlation effects by using a high-level method, and a high-level basis set.
A fast method in this realm is the CC2 method coupled with response theory [
22]. It is, e.g., implemented in TURBOMOLE [
10], alongside a faster version based on Møller–Plesset (MP) perturbation theory, namely the ADC(2) method [
23]. These allow for fast calculations based on the resolution-of-identity methods, and efficient basis sets of triple- and quadruple-zeta quality [
24], (e.g., def2-QZVPPD). Transition moments and other excited-state first-order properties are calculated quickly even for larger molecules [
25]. We will refer to this method as CC2-RPA. It is mostly restricted to not use the full symmetry, i.e., it is running with C
1 symmetry by default in TURBOMOLE. One can use the properties of the excited states, such as transition moments to obtain some idea of which excited states have been generated.
If one requires higher accuracy, one has available a general excited-state method at the level of CC theory including single and double excitations, named EOM-CCSD (EOM stands for equation of motion) [
26,
27]. Since the class of polar molecules to which MgF belongs is dominated by single excitations, it should provide reasonable accuracy. One can ask for a number of excitations for each irreducible representation. We used Psi4 [
28] to perform the calculations. The CCEOM package in Psi4 also allows us to perform higher-accuracy calculations at the level of EOM-CC3 theory. These are demanding calculations, but are considered the best there is for a general-purpose CC method for excited states [
29,
30]. The EOM-CC3 method (as implemented in Psi4) first calculates the CC3 ground state, then performs an EOM-CCSD for multiple roots (which in our work has the trend of being higher in energy than the usual EOM-CCSD roots), and then one has to pick one root at a time to calculate the EOM-CC3 value. Thus, the method is more cumbersome to use. Faster implementations of EOM-CC3 than generally available have been developed [
31].
In addition to the mentioned methods, one can also attempt so-called
-CC methods [
32]. One complements a standard ground-state CC calculation with a run where the SCF calculation is carried out for an excited state obtained by switching two orbitals (e.g., highest occupied and lowest unoccupied) to seed an SCF calculation for a Slater determinant which represents the excited state. This step may or may not work, as the SCF calculation may collapse back to the ground state. When it works, the resulting orbitals can be used in a correlated calculation at the CCSD or CCSD(T) level or higher. For the MgF molecule, we were not able to carry out the
-CC approach, since the
-SCF step failed, but it is interesting to note that in cases where the method works the excitation energies from
-CCSD and EOM-CCSD, in general, do not agree with each other [
33]. The agreement may become better at the level of the CC3 model for which comparison with
-CCSD(T) and EOM-CC3 can be made. The better agreement follows from a more complete inclusion of correlation at this level. Such results for the barium monofluoride (BaF) molecule using a pseudopotential for the inner electrons of barium for the
,
, and
states will be reported in a separate publication.
In terms of basis sets we used the def2-family at triple- and quadruple-zeta levels and made a guess at the complete basis set (CBS) limit by using an extrapolation of the form
where
and
are the energies obtained with def2-TZVPPD and def2-QZVPPD, respectively.
We also used the correlation-consistent augmented basis sets aug-cc-PVnZ with
[
34,
35]. From the calculated energies
the CBS limit can be obtained using
which follows from Equation (
2) in Ref. [
36].
For the MgF molecule, the known bond lengths for the ground and excited states in question are very close (
Å). For the comparison of methods and basis sets we restrict our calculations here to a single Mg-F separation, for which we obtain vertical excitation energies
, rather than the spectroscopic parameter
, i.e., the difference in the minimal energies between the potential energy curves. Detailed calculations of such curves at the level of four-component (relativistic) Fock space CC theory are given, e.g., in
Figure 1 of Ref. [
21], and can be used for the justification of this shortcut (at least for the lowest two excitations we can assume
). For other molecules, such as CaF, SrF and BaF one has to calculate the potential energy curves, i.e., energies as a function of
R, e.g., as given in Ref. [
17] in order to determine accurate adiabatic excitation energies. From these curves, one can then determine the spectroscopic parameters
and
, which follow from the nuclear vibrational excitation energies in accordance with [
37]
A final general comment for this section can be made: for each of the theory models and for each of the chosen basis sets the ground state is calculated at some level of precision. All CC calculations start from a common self-consistent field result, which for our calculations is based on spin-unrestricted Hartree–Fock theory. The changes in ground-state energy with a basis set at this level are not too significant. The differences between the values from these calculations are appreciable, and they arise mainly for two reasons: (i) the correlation contributions included in CC2, CCSD, and CC3 are different, and (ii) the truncated basis set introduces a significant error when computing correlation effects. Excitation energies are calculated relative to ground states which are numerically quite different. Nevertheless, they can be compared directly to each other when looking for excited-state values.
The theoretical methods for electronic excitations allow us to calculate the potential energy curves for a set of interatomic separations
R. As given by Equation (9.8) in Ref. [
37] (p. 348), the positions of the vibrational bands are determined on the basis of
and the differences in spectroscopic constants
and
(cf., Equation (
3)). If the differences in
values for the excited state vs the ground state are small, the correction due to the proper adiabatic treatment is small, and one may get away with a single calculation for
.
How comparable are the equilibrium distances
for the states of interest in the present work? The data in
Table 1 show that this is a reasonable assumption for the
and
states, but that the
and
states have minima in their potential energy curves which are shifted to progressively smaller values. It is also shown that the calculated values of
are smaller than the results derived from experimental spectra.
We adopt, for the present work, the strategy to use the experimental ground-state equilibrium separation of
Å. The calculated values in
Table 1 were obtained without taking basis set superposition error into account. For the
state, it would be straightforward to do so by subtracting atomic energies obtained with ghosted partner atoms, but for the excited states, it is not possible to do so. For computational economy reasons our procedure for obtaining excited-state energies in
Section 3 is to compute vertical excitation energies
, which overestimate
, since the relaxation to a lower value of
for the excited state is ignored. For the
state we estimate this difference to be small, i.e., within the precision goal of a few tens of
, but for the
state this increases to the order of
.
In order to illustrate the difference between the vertical excitation energy
and the true adiabatic excitation energy
we calculated the potential energy curves for the
and
states using the aug-cc-pVQZ basis, with the EOM-CC3 method.
Figure 1 shows both potential energy curves on the same vertical scale, i.e., the curve for the
state has been displaced by
= 42,113 cm
−1 to show both curves with minima at
. The data were fitted to a Morse potential,
The four parameters are obtained from calculations at values of R covering the region around the minimum, and D is not treated as a realistic dissociation energy in this work. An alternative to this fitting procedure is to use a cubic spline interpolation of the data.
Solutions for the Schrödinger equation describing the relative nuclear motion (
vibrational levels) were fitted to Equation (
3). The spectroscopic parameters are in good agreement with the values derived from the experiment. For the
state, we find
and
, while for the
state, the results are
and
(in
). The vertical excitation energy at
Å amounts to
, i.e., it overestimates the excitation energy by about
at this level of theory and basis set. The Franck–Condon factor (square of the overlap integral between the
states for
and
, respectively) is obtained as 0.744. For the two lower electronic excitations, the differences in
values are smaller, and the overestimation of the data on the basis of single-point calculations are smaller.
Figure 1.
Potential energy curves for the and states in blue and red, respectively, as obtained by the EOM-CC3 method and the aug-cc-pVQZ basis. The results for the state (red) in reality are offset relative to the (blue) by = 42,113 cm−1. The open circles show calculated results, the curves are Morse potential fits to these data. The dashed horizontal lines show the positions of the vibrational levels indicating the substantial difference in for the two states.
Figure 1.
Potential energy curves for the and states in blue and red, respectively, as obtained by the EOM-CC3 method and the aug-cc-pVQZ basis. The results for the state (red) in reality are offset relative to the (blue) by = 42,113 cm−1. The open circles show calculated results, the curves are Morse potential fits to these data. The dashed horizontal lines show the positions of the vibrational levels indicating the substantial difference in for the two states.
In terms of computational effort, we can make the following statement. All results were obtained on multi-core workstations (e.g., Apple MacBook Pro with 8 cores). The only lengthier calculations are for the CC3 method where multiple hours were required for the calculation of a single energy when using the aug-cc-pV5Z basis. Some cross-comparisons when using TURBOMOLE vs Psi4 showed no significant differences in results between equivalent calculations. Excited-state calculation using the EOM-CC3 method for a single root took about 6–8 h with the aug-cc-pVQZ basis.