David M. G.
Williams
*a,
Nicole
Weike
b,
Manuel
Lara
c,
Kevin M.
Dunseath
a and
Alexandra
Viel
a
aUniv Rennes, CNRS, IPR (Institut de Physique de Rennes) – UMR 6251, F-35000 Rennes, France. E-mail: david.williams@univ-rennes.fr
bTheoretische Chemie, Universität Bielefeld, Postfach 100131, D-33501 Bielefeld, Germany
cDepartamento de Química Física Aplicada, Facultad de Ciencias, Universidad Autónoma de Madrid, 28049 Madrid, Spain
First published on 24th January 2025
An accurate potential energy model, explicitly designed for studying scattering and treating the spin–orbit and nonadiabatic couplings on an equal footing, is proposed for the S + Ar system. The model is based on the Effective Relativistic Coupling by Asymptotic Representation (ERCAR) approach, building the geometry dependence of the spin–orbit interaction via a diabatisation scheme. The resulting full diabatic model is used in close-coupling calculations to compute inelastic scattering cross sections for de-excitation from the S(1D2) fine structure level into the 3P multiplet. The energy grid is tuned to resolve the many resonances present and to guarantee converged thermal rates from 1 to 300 K. At temperatures above 100 K, the computed thermal rate coefficients for quenching of S(1D2) are in good agreement with results from an earlier experimental and theoretical study. The branching ratio at 296 K for de-excitation into the S(3P0) level agrees well with the value obtained by a different experiment. A discrepancy however remains between theory and experiment at lower temperatures. This is discussed in light of the interference mechanisms at play during this quenching process.
The experimental work on quenching of S(1D) by argon reported by Lara et al.5 was accompanied by a theoretical study based on close-coupling calculations with the aim of understanding further the dynamics involved. The theoretical rate coefficients obtained are in good agreement with the experimental results at higher temperatures, but overestimate them at lower temperatures. At 5.8 K for example, the theoretical rate is a factor two larger than the experimental one. It is interesting to note that similar discrepancies between theory and experiment are seen in recent studies on the electronic quenching of O(1D) in collisions with noble gases.7–9 In particular, for the case of O + Ar, the differences are also of the order of a factor of two, not just at low temperatures but throughout the range from 50 to 350 K. Implementation details of the scattering simulations notwithstanding, we infer that the differences between theory and experiment for these chemically very similar systems are likely caused by results being extremely sensitive to the underlying potential energy curves. Indeed, for S + Ar, Lara et al.5 attributed this discrepancy to the interference mechanism underlying the process. In simple terms, the quenching process can be understood as probability transfers (jumps) localized around crossing points of the various potential curves and in particular those correlating with S(1D) and S(3P). The system traverses the crossings twice, once when the atoms are approaching and once when they are receding (having been reflected by the repulsive barrier). This creates two portions of flux which have travelled along different paths and hence have different phases. The resulting interference produces oscillations in the quenching probability as a function of the energy, known as Stückelberg oscillations.10 In the S + Ar system, there are crucial avoided crossings whose intrinsic interference mechanism makes the theoretical results extremely sensitive to any small change in the position of these crossings. This poses a stringent test for any theoretical simulation of this kind of process: any small inaccuracy in the potential energy curves will have a radical effect in the interference pattern and hence in the scattering simulation. Determining the curve crossings with an accuracy of a few cm−1, particularly challenging at crossings in the repulsive walls, was therefore deemed necessary in order to better reproduce this quantum interference process. This stringent dependence on the quality of the ab initio computation motivates us to use the Effective Relativistic Coupling by Asymptotic Representation (ERCAR) approach to reinvestigate the electronic quenching of S by Ar. Indeed the ERCAR scheme is one of the rare approaches for building accurate potential energy surface (PES) models that treat vibronic and spin–orbit coupling on an equal footing. Furthermore, the PES is furnished in a form suitable for scattering studies, as demonstrated in a recent application to the collisions of H with I.11,12
This paper is organized as follows. In Section II, the theoretical background is given while Section III provides the computational details. The coupled potential curves thus obtained are presented in Section IV A, and the results of the scattering computations are discussed in Section IV B. Finally, this work is summarized in Section V.
The molecular electronic Hamiltonian Ĥe can be separated into the Coulomb Hamiltonian Ĥc, which also contains all scalar relativistic effects, and the SO Hamiltonian ĤSO as
Ĥe = Ĥc + ĤSO. | (1) |
![]() | (2) |
The molecular electronic Hamiltonian Ĥe is represented in the basis (2) and the matrix elements Wdjk(Q, R) of the resulting diabatic model read
![]() | (3) |
In the determination of the SO model, the following effective n-electron SO operator is applied to the diabatic basis states,
![]() | (4) |
More details on the ERCAR approach can be found for example in ref. 16. The particular expressions for the geometry dependent Coulomb and the (geometry independent) SO model are system dependent. The details for the S + Ar system, where R is the sole internal coordinate, are given in Section III.
![]() | (5) |
![]() | (6) |
![]() | (7) |
The partial integral cross section for a transition from level γ to level γ′ for a given J, averaged over initial states and summed over final states, is given by
![]() | (8) |
TJ = −2iKJ(1 − iKJ)−1. |
The total integral cross section is then obtained by summing the partial cross sections over all values of the total angular momentum J. Energy-dependent rate coefficients may also be defined as
![]() | (9) |
To compare with the experimental results, it is necessary to introduce thermal rates. Collisional rate coefficients kγγ′(T) at temperature T for a transition from level γ to level γ′ are given by averaging the integral cross section over a Maxwellian distribution:
![]() | (10) |
Molecular orbitals and reference wave functions were first computed by state-averaged CASSCF calculations for each data point. The active space includes the 3p orbitals of sulfur. The orbitals were diabatized for each data point with respect to a reference calculation performed at a large S–Ar separation (1890 bohr). These diabatic orbitals then form the one-electron basis in the MRCI calculations in which the 3s and 3p orbitals of both sulfur and argon were used in the active space. Correlation and Davidson correction energies were rescaled by a factor of 1.123 for the lowest-lying triplet and singlet curves to maximally resemble a CCSD(T) reference coupled cluster calculation.
Wjk(R) = WRjk(R) + WIjk(R) + WDjk(R) | (11) |
![]() | (12) |
S | Symmetry | j | R R 0j | R R 1j |
---|---|---|---|---|
3P | 3Σ− | 1 | 9.26 | 10.77 |
3Π | 2–3 | 8.88 | 10.02 | |
1D | 11Σ+ | 4 | 7.37 | 8.13 |
1Π | 5–6 | 9.35 | 10.02 | |
1Δ | 7–8 | 9.35 | 10.02 | |
1S | 21Σ+ | 9 | 7.37 | 8.13 |
To ensure eqn (12) holds, we employ a switching function defined by
![]() | (13) |
In the present case, the matrix W(R) has only one symmetry-allowed diabatic coupling element between 11Σ+ and 21Σ+ (states 4 and 9). This coupling term is of course vanishing at infinite separation of the two atoms, resulting in a diagonal W(R) matrix for large enough values of R. The diagonal matrix elements are modelled as
![]() | (14) |
![]() | (15) |
This ansatz is exact in the asymptotic limit of R → ∞ because both Ar and S are neutral atoms. Since the adiabatic and diabatic representations coincide for R > 19 bohr, cD6j was obtained via a linear fit against ab initio data on a double logarithmic scale, confirming the asymptotic behavior. We verified that no additional orders are needed in this dispersion domain to properly reproduce the ab initio data. χD(R) serves to shift and scale the distance coordinate R such that the contribution of the dispersion term vanishes for R ≤ RD0 = 15 bohr whereas WRjj(R) and WIjj(R) vanish for R ≥ RD1 = 9 bohr. The functions used for the diagonal matrix elements in the repulsion domain are similar but less constrained in their functional form. They are given by
![]() | (16) |
![]() | (17) |
By construction WRjj(R) vanishes for all R ≥ RR1j, where RR1j depends on the particular state j, as given in Table 1. The diagonal functions in the interaction domain are defined such that they vanish at each domain boundary and are given by
![]() | (18) |
Finally, the lone diabatic coupling element (j ≠ k) is modelled similarly, but adding an exponential term for both interaction and repulsion domains:
![]() | (19) |
![]() | (20) |
As stated above, WDjk(R) = 0.
All parameters for the repulsion and interaction domains have been obtained using a Levenberg–Marquardt fitting algorithm. Since only two states are coupled, most could be directly fitted against ab initio data. The coupled 2 × 2 sub-block describing the 1Σ+ states had to be diagonalized before fitting.
![]() | (21) |
Basis number | L S | S S | L Ar | S Ar | j S | j Ar | j SAr | M j SAr |
---|---|---|---|---|---|---|---|---|
1 | 1 | 1 | 0 | 0 | 2 | 0 | 2 | −2 |
2 | 1 | 1 | 0 | 0 | 2 | 0 | 2 | −1 |
3 | 1 | 1 | 0 | 0 | 2 | 0 | 2 | 0 |
4 | 1 | 1 | 0 | 0 | 2 | 0 | 2 | 1 |
5 | 1 | 1 | 0 | 0 | 2 | 0 | 2 | 2 |
6 | 1 | 1 | 0 | 0 | 1 | 0 | 1 | −1 |
7 | 1 | 1 | 0 | 0 | 1 | 0 | 1 | 0 |
8 | 1 | 1 | 0 | 0 | 1 | 0 | 1 | 1 |
9 | 1 | 1 | 0 | 0 | 0 | 0 | 0 | 0 |
10 | 2 | 0 | 0 | 0 | 2 | 0 | 2 | −2 |
11 | 2 | 0 | 0 | 0 | 2 | 0 | 2 | −1 |
12 | 2 | 0 | 0 | 0 | 2 | 0 | 2 | 0 |
13 | 2 | 0 | 0 | 0 | 2 | 0 | 2 | 1 |
14 | 2 | 0 | 0 | 0 | 2 | 0 | 2 | 2 |
15 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 |
The strategy followed to fit the three λi parameters aims at reproducing the experimental levels of the sulfur atom.21 In our model, the 5 asymptotic energies are obtained by diagonalization of the sum of the SO matrix and the three asymptotic energies of the Coulomb part. These three SO-free ab initio energies differ from the weighted average of the observed values21 by at most 70 cm−1 due to inaccuracies in the ab initio computations. As a consequence, the optimization of the λi parameters was combined with the modification of the asymptotic energies of the Coulomb model in order to match with the experimental values. Experimental transition energies are 396.055 cm−1, 573.64 cm−1, 9238.609 cm−1 and 22179.954 cm−1, respectively.21 These numbers are (by construction) perfectly reproduced by our model.
It should be noted that the inter-state SO couplings expressed in the basis given in Table 2 do not vanish at large distances. These couplings mix different terms arising from the same electron configuration and having the same total angular momentum jSAr. As a result, the physical states, characterized by the quantum number jSAr, are described by linear combinations of the basis states of Table 2. To facilitate the matching process and to compute collision cross sections, a basis transformation must be performed to diagonalize the potential matrix at large distances.
Table 3 compares the long-range interaction coefficients thus obtained with those of ref. 5. In this earlier work, ab initio electronic energies were obtained using MOLPRO from MRCI+Q calculations in C2v symmetry, with the aug-cc-pV5Z orbital basis. The energies were corrected for the basis set superposition error (BSSE) using the counterpoise procedure of Boys and Bernardi.22 Once the MRCI wavefunctions for the nine electronic states had been determined, the matrix elements of the SO part of the Breit-Pauli Hamiltonian were calculated at each internuclear distance. The long-range interactions corresponding to the 11Σ+ state were independently calculated using a multipolar expansion of the electrostatic interaction operator, treated at the second order of perturbation theory. The resulting energies were matched at large distances with their MRCI counterpart.
The values obtained in the current calculation are about 10% larger than those reported in ref. 5. We infer that this comes from the different electronic structure, as well as from the different methodology employed to obtain these coefficients. Our values for the well depths and equilibrium distances also subtly deviate from those reported in ref. 5. The largest difference in equilibrium distance is found to be 0.12 bohr for the two states of Σ+ symmetry. The well depths obtained in the current work agree with those given in ref. 5 to within 30 cm−1. The largest differences are obtained for the ground 3Π (27 cm−1 shallower) and for the 11Σ+ (25 cm−1 deeper) curves. The total root mean square error (RMSE) achieved by the fit across all ab initio data points is 28 cm−1 including all adiabatic energies up to 1 eV above their respective dissociation energy. The RMSE is dominated by the error in the repulsive walls, with differences between model and ab initio data consistently ≤1 cm−1 for large ranges of R, for all considered states. Fig. 1 shows the energies of the Coulomb model corresponding to the 3P and 1D channels. The 1S is about 22000 cm−1 above the 3P state (and thus not shown). On the scale of the figure, the ab initio energies are indistinguishable from the presented curves. The 11Σ+ state exhibits a characteristically “dented” shape between the minimum and the repulsive wall with a moderate slope. Due to this particular shape, the intersections of this curve with the two triplet states are extremely sensitive to minor changes in the electronic structure. These singlet–triplet intersections become avoided crossings when the SO coupling is taken into account. The dented shape of the 11Σ+ state is caused by subtle nonadiabatic coupling effects with the 21Σ+ state. Indeed our model shows a strong coupling between these two electronic states especially for distances less than 6.5 bohr, where the coupling strength approaches the size of the energy gap between the adiabatic energies (on the eV scale). We thus infer that the exact shape of the 11Σ+ potential curve is extremely sensitive to the electronic structure calculation and especially to the convergence of the nonadiabatic coupling. This is a key limitation, as it is far more difficult to converge electronic wave function properties rather than the adiabatic energies.
Fig. 2 shows the adiabatic energies obtained from the full ERCAR potential model, including SO coupling. The asymptotic energies are split into 5 levels corresponding to Ar(1S0) and S(3P2,1,0, 1D2, 1S0). As outlined in Section III C, experimental transition energies for atomic sulfur are reproduced by construction. Note that the S–Ar adiabatic curves are labelled by their Ω = |MjSAr| quantum number as is common in spectroscopy. For Ω = 0±, the upper index denotes whether the state is odd or even with respect to reflections in a plane which contains the atoms.
![]() | ||
Fig. 2 Adiabatic potential curves over the full energy range. The S–Ar distance is given in bohr and the energies in cm−1. Zero is defined as the asymptotic limit of the lowest curve. |
At shorter inter-atomic distances, the 3P2 level splits into Ω = 2, 1 and 0+ components with Ω = 2 being the lowest curve (see Fig. 3). The 3P1 level is composed of Ω = 1 and Ω = 0− curves while 3P0 corresponds to the second Ω = 0+. Finally, the 1D2 channel splits into another set of Ω = 2, 1, 0+ curves as shown in Fig. 4, while 1S0 forms the fourth Ω = 0+. The curves of identical Ω form avoided crossings where the former 11Σ+ intersects with the 3Π and 3Σ−. In particular the Ω = 0+ curves present two avoided crossings at distances between 4 and 5 bohr. Fig. 4 shows the aforementioned avoided crossings in greater detail. It is of particular note that the lower-lying crossing nearly coincides with the dissociation energy of the 1D2 channel, making it a dynamically accessible and extremely relevant feature of the surface. The horizontal line in Fig. 4 corresponding to the energy of the 1D2 channel is drawn to emphasize this particularity. The second crossing, at a distance about half a bohr shorter and only 400 cm−1 higher in energy, is also expected to be dynamically reachable.
![]() | ||
Fig. 3 Zoom on the low energy part of the adiabatic potential curves. The S–Ar distance is given in bohr and the energies in cm−1. |
In Fig. 5 we present the total energy-dependent rate coefficient, eqn (9), for the quenching process from the S(1D2) level down to the 3P multiplet. We also show the individual rates for quenching into the 3P0,1,2 fine structure levels. Transitions into the 3P0 level provide the largest contribution to the total quenching rate, although its relative importance compared to the contribution from the 3P2 level diminishes as the collision energy (relative to the initial S(1D2) level) increases. Quenching into the 3P1 level is very small throughout the energy range considered. This can be attributed to the absence of any coupling between the Ω = 0+ component of the asymptotic S(1D2) potential curve and the Ω = 0− component of the S(3P1) curve in the neighborhood of the relevant crossings, see Fig. 2–4. Populating the 3P1 level hence requires changing Ω through Coriolis couplings (i.e. off-diagonal elements of the centrifugal potential in the BF basis), whose action has been shown to be very small for S + Ar.5
![]() | ||
Fig. 5 Total and partial energy-dependent rate coefficients as a function of the collision energy relative to the initial S(1D2) level. |
Resonances are discernable at collision energies below 60 cm−1, as can be seen in Fig. 6 which provides a zoom on this energy region. Most are associated with the 1D2 → 3P0 transition simply because this is the dominant contribution at these energies. Occurring in different partial waves (J values), resonances are signatures of quasi-bound states supported by the effective potential in each channel resulting from the combination of the attractive well and the repulsive centrifugal barrier. The resulting potential landscape can then be quite tortuous due to the various avoided crossings in the adiabatic potential curves shown in Fig. 4. While the individual resonance structures themselves are quite narrow, they nevertheless have an important effect on the thermally averaged rates at low temperatures. It is therefore important to adequately resolve these resonances.
![]() | ||
Fig. 6 Total and partial energy-dependent rate coefficients as a function of the collision energy in the resonance region close to threshold. |
Comparing with the corresponding energy-dependent rates given in Fig. 6 of ref. 5 reveals a number of important differences. For collision energies above 500 cm−1, the current total energy-dependent rate is consistently larger than that of ref. 5, the percentage differences reaching as much as 30%. In the resonance region shown in Fig. 6, however, the results of ref. 5 tend to be larger, for example by approximately 30% between 20 and 30 cm−1 where our total energy-dependent rate has a minimum. While it may appear that there are more resonances in the current results than in those of ref. 5, this is almost certainly due to the much finer energy grid used here to adequately resolve the resonance structure. The most striking differences occur in the partial energy-dependent rates into the 3P0 and 3P2 levels. In the earlier work, these partial rates as a function of collision energy cross near 110 cm−1, so that at higher energies the main contribution comes from transitions into the 3P2 level. The authors attributed this to a crossing of potential energy curves at 6.86 bohr (labelled as C3 in their work). In the current work, however, no such crossing occurs, so that the major contribution to the quenching process is always from transitions into the 3P0 level. Indeed, at collision energies corresponding to temperatures in the vicinity of 296 K (∼206 cm−1), transitions into the 3P0 and 3P2 levels contribute roughly 82% and 17% respectively to the total energy-dependent rate.
In Fig. 7 we present the total and partial thermal rate coefficients, eqn (10), for temperatures from 5 to 300 K. The experimental and theoretical results from ref. 5 are also shown. The two experimental data points above 100 K are well reproduced by both calculations. Below 50 K, neither of the two theoretical thermal rates agree with the experimental data. While the earlier results overestimate the experimental values, with a maximum near 10 K, the current results underestimate them, displaying a minimum between 20 and 30 K. The sensitivity of the theoretical thermal rates is directly related to the fine details of the coupled potential curves used, as shown in Fig. 11 of ref. 5. We further infer that the position of the avoided crossing at the energy of the entrance channel is responsible for this. We also remark that the partial thermal rate at 296 K for transitions into the 3P0 level accounts for approximately 77% of the total rate, close to the experimental estimation6 of around 80%, while transitions into the 3P2 level contribute roughly 23%. Comparing with the branching ratios for the energy-dependent rates, we see that the thermal averaging has slightly reduced the relative contributions from transitions into S(3P0), enhancing those into S(3P2).
![]() | ||
Fig. 7 Total and partial thermal rate coefficients for quenching. The experimental data and the green curve labelled Lara et al. are from ref. 5. |
Our branching ratios thus indicate a strong propensity for quenching into the 3P0 fine structure level throughout the temperature range considered. A similar propensity was seen for quenching of O(1D) by Ar,9 which should not be surprising since oxygen lies just above sulfur in the periodic table. It is interesting to note that while the calculations in ref. 5 give similar values to ours for the total rate coefficient, their branching ratios are very different with a propensity for populating the 3P2 level at 296 K. Despite the continuing discrepancy in the rate coefficients at low temperature, our results would thus appear to be a step in the right direction.
The coupled surfaces thus obtained have been used to compute the energy-dependent and thermal rates for the quenching of S(1D2) by Ar. Since the underlying basis is constructed using eigenstates of the electronic angular momentum, the close coupling equations follow directly. Partial waves up to total angular momentum J = 120 have been used to compute quenching cross sections at collision energies up to 3000 cm−1 and hence thermal rates from 1 to 300 K. The results above 100 K are in good agreement with those of a previous theoretical and experimental study.5 At low temperatures however the discrepancy between the theoretical and the experimental data remains. In contrast to the earlier work, the computed branching ratio at 296 K for transitions into the 3P0 fine structure level is in good agreement with the experiment of Stout et al.6
The S + Ar system has a delicate electronic structure with multiple avoided crossings. We infer that at least the position of one of these is responsible for the high sensitivity of the scattering results to fine details of the potential model. Since our diabatic PES reproduces ab initio data well, we conclude that inaccuracies in the underlying data are a dominant source of error, rather than the surface itself. Further work is clearly necessary in order to reconcile theory and experiment, particularly in view of the different trends shown by the two calculations and the experiment as the temperature approaches 0 K. On the theory side, ab initio calculations going beyond those used here are clearly necessary, while additional experimental work would help confirm and extend the existing measurements.
This journal is © the Owner Societies 2025 |