Skip to main content

ORIGINAL RESEARCH article

Front. Phys., 04 October 2022
Sec. Chemical Physics and Physical Chemistry
This article is part of the Research Topic Structure and Dynamics of Atmospheric, Plasma and Astrochemical Molecular Processes View all 5 articles

Assigning quantum labels and improving accuracy for the ro-vibrational eigenstates of H3+ calculated using ScalIT

  • Department of Chemistry and Biochemistry, Texas Tech University, Lubbock, TX, United States

In a recent article [AIP Adv. 11, 045033 (2021)], we carried out exact quantum dynamical calculations and computed ro-vibrational energy levels and wave functions for the H3+ molecular ion up to the dissociation threshold (at J = 46) using a recently developed potential energy surface (PES) [Mol. Phys. 117, 1663 (2019)]—arguably, the most accurate to date —together with the ScalIT suite of parallel codes. In this work, we further improved the convergence accuracy and range of our ScalIT calculations for all J values up to J = 20 to a few 10–5 cm−1 (or better). In addition, we performed an ab initio assignment of the ro-vibrational energy levels, providing vibrational ‘v1, v2, |l|’ and rotational ‘J, G, U, K’ quantum labels for more than 2,200 ro-vibrational states, including every single 0 ≤ J ≤ 20 state up to and above the barrier to linearity at 10,000 cm−1. The main underlying motivation of our work is to provide a list of reliably labeled, spectroscopically accurate energy levels in a format that can be used in spectroscopic line lists, which are based on both experimental and theoretical levels. Such line lists are of huge importance in various astrochemical and astrophysical contexts.

1 Introduction

The H3+ molecular ion [1]—the smallest tri-atomic molecular system, with just three protons and two electrons—is a central molecule in molecular astrophysics and astrochemistry. It is the most common molecular ion in the Universe, serving as the main conduit of chemical reactions in outer space. H3+ can be found in the interstellar medium [2], supernova remnants [3], the atmospheres of gas giants, and exoplanets [4, 5], and also plays an important role in star formation [1]. Partially due to its simplicity, H3+ serves as a benchmark system for several different areas of science, in particular, high-resolution ro-vibrational spectroscopy experiments, accurate ab initio electronic structure calculations and potential energy surface (PES) development, high performance quantum dynamics calculations, and reaction dynamics. Despite its simplicity, the near-dissociation spectrum of H3+ [6, 7]—recorded 40 years ago!—still remains unassigned. H3+ has been studied very extensively both experimentally and computationally in the last four decades, as was recently summarized in a very nice review [1].

On the experimental side, numerous spectroscopic studies have been conducted [614]. Of course, the primary challenge with respect to labeling is that experiments provide only spectroscopic transitions, not the ro-vibrational energy levels themselves. Although symmetry and selection rules help, extracting the latter from the former remains a challenge, and has traditionally been something of a “black art.” Recently, more systematic approaches have been developed, based on graph theory and “spectroscopic networks” (SNs) [15, 16], in which the vertices represent rovibrational energy levels, and the lines represent experimentally observed spectroscopic transitions, to extract empirical energy levels directly from experimental data, with well-defined and realistic uncertainties. In particular, the MARVEL code (Measured Active Rotational–Vibrational Energy Levels) [17, 18], has been applied to ro-vibrational spectroscopic data of H3+ that were collected from 26 separate experimental sources [13]. The resultant energy levels and assignments replaced the earlier work of [8]. A MARVEL analysis was also carried out for two isotopologues, H2D+ and D2H+ [14], with the database last updated in 2019 [19]. Thus far, the number of validated, and therefore recommended, experimental quality ro-vibrational energy levels of H3+ are 652, [19] of which 259 belong to ortho-H3+ (I = 3/2) and 393 to para-H3+ (I = 1/2), with I being the quantum number of the total nuclear spin of the system.

On the theoretical side, due to its spectroscopic importance, a variety of H3+ PESs have been developed over the years [11, 1226]—with the latest two [25, 26] published only very recently. A number of ro-vibrational state calculations have also been performed in the past for this system [12, 2243], many employing empirical corrections of various kinds (e.g., empirically modified vibrational masses [44, 45]) in order to better match the available experimental data, and also to capture non-adiabatic effects [24, 25, 34, 3739, and 41]. The empirical approach might be less effective at higher energies, [35]—e.g., in the context of reactive collisions, which have also been extensively studied for H3+ [4648]. Additionally, empirical “corrections” can become a bit tricky, if there is any question as to how to match experimental and theoretical state labels.

For these reasons, we prefer a fully ab initio computational approach [43], both with regard to the ro-vibrational state calculation itself, as well as the determination of state labels. In particular, to the best of our knowledge, we are the first group to attempt a fully ab initio assignment of ro-vibrational state labels for H3+. Our first push in this direction was published in an article last year [43]; however, the set of J values considered in that work was restricted, and in addition, we did not use wave functions to help determine ro-vibrational state labels, but only D3h symmetry labels. In addition, although the calculations were very well converged (10–4 cm−1), better convergence would have allowed for a better determination of symmetry-induced vs. “accidental” degeneracies, which in turn leads to a less ambiguous state labeling, especially at higher vibrational and rotational excitation energies. All of these small deficiencies of the previous work have been rectified here, as discussed below.

As further motivation for adopting a purely ab initio approach, we point out that H3+ has always been targeted as an important benchmark system for achieving a direct spectroscopic agreement between theory and experiment. This has been a long-standing goal, which it can be argued, has only begun to be obtained fairly recently [12, 24]. Additionally, the highly accurate determination of the ro-vibrational spectrum of astrophysically relevant molecules such as H3+ is motivated by the “weed problem” [49, 50]. In the interstellar medium and planetary atmospheres, many different molecules or ions are present at the same time. As the spectra overlap, it is crucial to obtain highly accurate spectra in order to unambiguously differentiate the contributions from different species.

Therefore, creating highly accurate line lists can serve as an important tool, from both the experimental and computational points of view. The first such line list was created by [51] with 669 astronomically important lines. This was supplemented by [52], with about three million lines. The newest line list, MiZaTeP [53], contains more than 120 million lines by bringing together the experimental spectroscopic data using MARVEL [13], and theoretical levels computed with the DVR3D code [5456]. This line list also contains 17 meta-stable states, which are quantum states with very long lifetimes.

In this work, in order to facilitate the expansion of already existing H3+ line lists, we focus our efforts on further improving the convergence accuracy of our ab initio ro-vibrational energy level calculations down to a few 10–5 cm−1. This is far beyond the accuracy of the PES, and certainly much smaller than the discrepancies within the experiment. Nevertheless, such an extraordinarily high convergence accuracy is essential with respect to unambiguous state labeling, as discussed, especially at higher vibrational and rotational excitation energies. Additionally, we compute and analyze ro-vibrational wave functions to determine their “vibrational parent” states [57] as a further means of providing unambiguous labels. In the present work, we also consider all values of the rotational quantum number J, not just selected values—but only up to a maximum of J = 20. In the previous work [43], we considered higher J values, all the way up to rotational dissociation (J = 46). Here, we apply a restriction to comparatively low values, simply because the quantum label assignment (which is a primary focus of this work) becomes essentially impossible much beyond this point.

In all, we provide vibrational ‘v1, v2, |l|’ and rotational ‘J, G, K, U’ quantum labels for more than 2,200 ro-vibrational states, around 1,600 of which are new assignments complementing, and in certain cases arguably correcting, the ∼650 assignments in the MARVEL database [19]. To the best of our knowledge, no previous work has attempted to provide purely ab initio quantum label assignments for computed ro-vibrational states—certainly not to the extent that we have done here, in any event.

2 Materials and methods

In a recent article [43], we carried out the exact ro-vibrational energy level and wave function calculations for the H3+ molecular ion for selected J values up to J = 46. As most of the computational details remain unchanged, here, we only provide a brief summary of the overall computational methodology, and focus primarily on the differences from the previous work.

2.1 ScalIT

The quantum dynamical calculations presented in this article were performed using the ScalIT [5862] suite of parallel codes. ScalIT is a black-box molecular ro-vibrational spectroscopy code, which for tri- and tetratomic molecules employs an analytical kinetic energy operator expressed in (orthogonal) Jacobi coordinates. The use of direct product basis sets (DPBs) including discrete variable representations (DVRs) results in a Hamiltonian matrix with a sparse structure. For the radial coordinates, phase–space-optimized DVRs (PSO-DVRs) are used [6368] while for the bend and rotation angles, standard associated Legendre polynomial or Wigner rotation function basis sets are utilized. The Hamiltonian is diagonalized iteratively using sparse Krylov subspace methods together with several different effective numerical optimization strategies, such as the preconditioned inexact spectral transform (PIST) method [6971], optimal separable basis (OSB) preconditioning [7275], and the standard iterative quasi-minimal residual (QMR) algorithm [59, 76]. All of these methods working together ensure the effective scaling across massively parallel supercomputing clusters (up to a few thousand cores) and the ability of ScalIT to accurately compute even extremely energetically high-lying quantum states. So far, ScalIT has been used for around a dozen challenging systems, such as Ne4 and HCCH [43, 50, 68, and 7786], and via extending the capabilities of the ScalIT code through the SwitchIT [87] algorithm to accommodate more complicated Hamiltonians, even CH3CN [87].

The massively parallel capability of ScalIT is based on MPI parallelization, which makes the code uniquely qualified to compute many quantum states with high accuracy for small molecular systems. Other available ro-vibrational spectroscopy codes, e.g., DVR3D [5456], TROVE [8890], DOPI [33, 91, 92], DEWE [57, 9395], GENIUSH [96, 97], and ElVibRot [98100], in general traditionally only offer single node OPENMP parallelization (although ElVibRot has been made MPI parallel recently [101]). Also, it is worth mentioning that the use of GPUs is spreading slowly to the field of ro-vibrational molecular spectroscopy with a focus on computing ro-vibrational intensities [102].

2.2 Potential energy surface

In this work, we utilized the recently computed H3+ PES referred to as “PES75K+” [25] which is based on the Born–Oppenheimer ab initio points of the earlier “GLH3P″ PES [12, 24], the first “calibration quality” PES developed for H3+. In our previous work, we compared these two PESs and discussed the spurious asymptotic wells that appear in the GLH3P PES [43], which were causing significant numerical convergence problems for our ScalIT calculations. Although GLH3P has been used more frequently in previous computational ro-vibrational spectroscopy studies, PES75K+ has now been shown to provide more accurate energy levels higher up in the spectrum.

In any event, the H3+ PES shows this molecular ion to be quite stable. The first dissociation threshold (to H2 + H+) occurs at D0 = 35, 076 ± 2 cm−1 [25]. However, there is a much lower-lying (linear) isomerization barrier at around 10,000 cm−1.

2.3 Previous computational works

For the GLH3P PES [12, 24]—and other earlier PESs, including the PES developed by Cencek and colleagues [20, 21]—a wide range of ro-vibrational calculations have been performed [12, 2224, 2729, 31, 32, 3440, 42, and 43]. Most of these are summarized in a fairly recent review article [40]. In the last couple of years, newer PESs have been also developed, among them, the PES75K+ PES [25] is used here, and also a multi-sheet fit PES including more than one electronic state [26]. By and large, the ro-vibrational studies have focused on increasing the accuracy of numerical convergence, as well as pushing the limits of vibrational/rotational excitation. Indeed, computing ro-vibrational energy levels of H3+ near dissociation has a long history [3, 27, 29]. To date, all energy levels were computed up to dissociation using atomic [30, 33], nuclear [35], and modified hydrogen masses [35] to investigate the properties of high-lying vibrational states. More restricted calculations—both in terms of energy and rotational excitation, but also accuracy—were also carried out, up to J = 2 and 15,300 cm−1 [32], and up to J = 3 and the then-experimental limit of ∼17,000 cm−1 [24]. The dependence of Coriolis coupling on choice of “embedding” or body-fixed frame was also investigated [42].

2.4 Symmetry

Jacobi coordinates (denoted as r, R, and θ here) are usually the best choice for describing AB2 triatomic molecules. In such cases, the full G4 permutation–inversion (PI) symmetry of the molecules is fully described. Of course, H3+, with its three identical atomic nuclei, is an A3 system, whose energy levels are labeled by the G12 PI group irreducible representations (irreps) [103]. Note that G12 is isomorphic with the D3h point group—which, in any event, describes the global minimum equilibrium structure of H3+, which is an equilateral triangle with a bond length of 1.65034a0 = 0.873 322Å.

Nevertheless, since Jacobi coordinates, in contrast to hyperspherical coordinates [104], do not respect the cyclic permutation operations of the G12 PI group, this poses certain challenges for the ScalIT calculation performed here, which essentially presumes an AB2 structure. More specifically, it becomes necessary to correlate the symmetry labels from the G4 symmetry-adapted ScalIT calculations to the G12/D3h labels, using the Γ(D3h) ↓ G4 correlation table [43]. The “challenge” here actually only concerns the doubly-degenerate D3h irrep pairs, which are computed in different ScalIT calculations corresponding to different G4 irreps. In practice, one looks for identical eigenvalues across two G4 irreps, and identifies those as comprising, in reality, a single doubly-degenerate G12 irrep pair. Better convergence accuracy thus greatly improves the determination of numerically “identical” eigenvalues. Conversely, whatever pair splitting is observed numerically may be taken as an additional, independent measure of the overall numerical convergence accuracy.

The solutions of the AB2 Jacobi Hamiltonian are computed in four separate “symmetry blocks”, corresponding to the four (singly-degenerate) irreps of G4. These irreps can be labeled by two good quantum numbers, p = ±1 (associated with the exchange of any two identical nuclei) and ϵ = ±1 (the total parity). In addition, there are the two good rotational quantum numbers that can be used as completely reliable labels—i.e., the total angular momentum, J, and its projection along the space-fixed Z axis, M. The third rotational quantum number, K, associated with the projection of Ĵ along the body-fixed z axis, is technically not a good quantum number—though for H3+, it may still often be used as a state label in practice, together with other approximate labels described in Section 3.2.

Lastly, given the fermionic nature of the H atom nuclei (i.e., protons), it is worth mentioning that the Pauli principle requires the total spin-plus-spatial nuclear wave function to have a totally anti-symmetric or A2 character (in the S3 permutation subgroup of the G12 PI group). For three such particles, the eight-dimensional combined nuclear spin space representation reduces to an irrep direct sum as 4A1⊕2E. The corresponding spatial wave functions (i.e., the ro-vibrational states actually computed) are thus restricted to belonging to either the A2 or E irreps. Therefore, all A1 ro-vibrational states (including what would otherwise be the ground vibrational state), are unphysical, and must be ignored.

3 Results

3.1 Computational details

ScalIT computations were carried out using nuclear masses, just as in our previous article [43]. The computational parameters of this work are summarized in Table 1. In DVR calculations such as those performed here, very often as one increases the basis size in order to improve numerical convergence, one crosses over from a regime where the basis set truncation error dominates, to a regime where the numerical quadrature error dominates. This is indicated by a fast, variational convergence (from above) being replaced by a slow, oscillatory convergence behavior. Usually, when this crossover has occurred, it becomes computationally unfeasible to push the calculation much further through a “brute force” increase in the basis size.

TABLE 1
www.frontiersin.org

TABLE 1. The total bend-rotation angular basis sizes of each G4 symmetry block, NjKA+, NjKB+, NjKA, and NjKB, for all ScalIT calculations of H3+ performed here, from total angular momentum J = 0 to J = 20. The number of bend-angle basis functions in θ, i.e., jmax, is always equal to 36. The radial basis sizes are Nr = NR = 300, and the radial ranges (in atomic units) are rmin = 0.5, Rmin = 0.0, and rmax = Rmax = 5.0.

For extremely accurately converged ro-vibrational calculations, it is therefore necessary to ensure that the quadrature error is minimized. This requires two conditions. First, the PES must be very smooth and well-behaved—which, in the case of PES75K+ (but unlike GLH3P), has already been established. Second, the “primitive basis” calculations used to compute the PSO DVR basis representations must be performed as accurately as possible. To this end, a very large number of 1,001 primitive sinc-DVR grid points were used in the PSO DVR calculations for both of the Jacobi radial coordinates, r and R. The radial ranges used here were also wider than before [43]; here, we used rmin = 0.5 bohr, and rmax = 5.0 bohr for the r coordinate, and Rmin = 0.0 bohr, and Rmax = 5.0 bohr for the R coordinate.

Having put these measures into effect, our next task was to increase the basis sizes for the final calculation as far as possible, in hopes that an extremely high numerical convergence could be achieved prior to crossing over into the quadrature-error-dominated regime. We therefore used significantly larger radial basis sizes than in the previous calculation; i.e., Nr = 300 and NR = 300. The angular dimensions were also increased compared to our previous work [43]; specifically, the number of bend-angle basis functions in the Jacobi coordinate θ was set to jmax = 36 for every J value considered. The resultant total bend-rotation angular basis sizes for each G4 symmetry block calculation, i.e., NjKA+, NjKB+, NjKA, and NjKB, are listed in Table 1. Using these parameters, we were able to achieve better than 10–5 cm−1 numerical convergence for all ro-vibrational states with J ≤ 10.

3.2 State Labeling

The ro-vibrational calculations of H3+ for each J > 0 were carried out in four blocks corresponding to the four G4 irreps. Note that the inversion parity is linked to the value of K, with even ϵ = +1 parity corresponding to the even K values, and odd ϵ = −1 parity to the odd K values. Thus, for J = 0, we only have two even parity blocks. States with |K| mod 3 = 0 are ortho-states with a spin weighting gs = 4, while those with K not exactly divisible by 3 are para-states with gs = 2 [1]. As the convergence accuracy of our calculations is very high, the degenerate energy levels can be unambiguously identified, and therefore, it is easy to assign the D3h (i.e., G12) irrep labels even for highly vibrationally and rotationally excited states.

Next, we address the vibrational state labels. The H3+ molecular ion has two normal modes, the totally symmetric stretch mode v1 (belonging to the singly degenerate A1 irrep), and the asymmetric stretch-bend mode, v2 (belonging to the doubly degenerate E irrep). Displacements of the latter distort the A1 symmetry of the global minimum geometry, thereby producing a transition dipole moment. Also, being doubly degenerate, excitations of the v2 mode give rise to a new quantum number, the vibrational angular momentum l, adopting the values, l = {v2, v2 − 2, … − v2 + 2, − v2}. Therefore, the vibrational part of the ro-vibrational states can be described by the labels, ‘v1, v2, |l|‘—although it must be borne in mind that these quantum numbers are not perfectly “good”. Also note that the quantum number |l| is linked to the D3h irrep labels. In particular, the l = 0 vibrational states are always singly degenerate A1 states. For |l| > 0, the degenerate pair can be labeled as E′, unless |l| mod 3 = 0, for which the ± l pair splits into an A1 and an A2 state.

After first determining the D3h irrep labels, we assigned vibrational state labels to the J = 0 pure-vibrational states, which were found to be in complete agreement with earlier studies [9, 10, 12, 13, 24, 32, and 43]. For J > 0, it is advantageous to first determine the vibrational labels, indicating which “vibrational parent” state the ro-vibrational state “belongs to”. This is straightforward to do for low-vibrational and/or rotational excitations, where the energy level spacing is so high that the rotational progressions do not overlap. Higher up in energy, determining the “vibrational parents” becomes much more challenging. For J = 1, the different ro-vibrational progressions start to overlap at the 26th vibration at 10,000 cm−1. Increasing J, this threshold energy value shifts down drastically. For J = 11, even the ro-vibrational states belonging to the zero-point vibration start to overlap with the ro-vibrational progression of the first vibration. Beyond a certain point in both (v1, v2, |l|) and J, it becomes impossible to assign vibrational parents based solely on energy values and D3h symmetry labels.

In order to overcome this difficulty, the GENIUSH code [96, 97] was invoked, which is capable of semi-automatically assigning vibrational parent labels using the rigid rotor decomposition scheme (RRD) [57], based on computing wave function overlaps. To do this, the ScalIT calculations were repeated using GENIUSH, but with greatly reduced accuracy (10−2–10−3 cm−1)—which was nevertheless sufficient to match the energy levels with the ScalIT ones. The RRD overlap matrices were then computed using GENIUSH. In this manner, we were able to assign vibrational parent labels to much more highly excited ro-vibrational states—and for many more such states—than was previously possible.

We next move on to rotational state labels. The H3+ molecular ion can be characterized as an oblate symmetric top, for which rotational states can be described using the J and K rotational quantum labels as:

EJKBJ(J+1)+(CB)K2.(1)

where EJK is the relative energy of the ro-vibrational state corresponding to its vibrational parent. Although for H3+J is always a good quantum number, due to the coupling of the rotational and the vibrational (l) angular momenta in this case, it has been argued [8, 105107] that instead of using K, it is better to use G = |Kl|, which becomes a much better quantum number at low energies [8]. However, we will assign values to both.

For l = 0, G = |K|, and so the usual (2J + 1)-fold rotational progression arises. For |l| > 0, however, the±l values double the number of the rotational excited states to 2 (2J + 1). For |l| > J, there is only one rotational progression, where Gmin = (|l| − J) ≤ GGmax = (|l| + J). For |l| ≤ J, the rotational excited states can be separated into two distinct rotational progressions [8, 107], with 0 ≤ GGmax = (J + |l|), and 0GGmax=(J|l|), and they have (2Gmax + 1) and (2Gmax+1) states, respectively. These two rotational progressions often have different rotational constants and a trend similar to Eq. 1

EJGBJ(J+1)+(CB)G2(2)
EJGBJ(J+1)+(CB)G2.(3)

As it usually holds that Gmax > J, EJG can become negative—as happens, e.g., for the A2 state, (v1, v2, |l|, J, G)=(0 2 2 1 3). This behavior is similar to the negative rotational energies observed for “quasi-structural” molecules [108] such as H5+ [109], CH5+ [110112], CH4⋅H2O [113], CH4⋅CH4, and H2O⋅H2O [114].

The presence of the two progressions requires that in addition to G, a new quantum number, U [8, 107], has to be introduced. U can take the values “u”, “l”, or “m” to distinguish between upper and lower energy levels with the same (v1, v2, |l|, J, G) assignment (note, that U = “l” always refers to levels within the G′ progression). Therefore, the rotational part of the wave function can be unambiguously described by the (J, G, U) quantum labels. This, however, does not mean that we cannot assign K values to each ro-vibrational state, as from the definition of G, we can assign |K| = |G − |l‖ for the unprimed progressions, and |K| = G′ + |l| for the primed progressions. Therefore, in the end, we characterize the rotational part using the (J, G, U, K) quantum label quartet.

For J > 0, D3h irrep labels are also linked to the quantum numbers, but it is actually G which is most directly impacted (for J = 0, |l| = G). The D3h irrep is A1 or A2 if G = 0 (if l = 0, we have A1 for even and A2 for odd J values), and for G > 0, similarly to J = 0, the degenerate pair can be labeled as E unless |l| mod 3 = 0, where the ± l pair splits into A1⊕A2. For large G values, especially for G ≥ 12, the two singly degenerate levels can get closer than 10–5 cm−1; therefore, in certain cases, G values have to be taken into account when assigning D3h labels.

Finally, note that to further aid in the assignment of the rotational labels, the RRD overlaps of GENIUSH also provide insights into the value of K. This occurs through the vibrational parent being assigned (in particular, l) into the G quantum number, as the symmetric top rigid rotor functions are labeled by K. This feature of the approach helped us tremendously in carrying out the task of assigning labels—so long as the RRD overlap values were significantly large.

3.3 Ro-vibrational energy levels

The ro-vibrational energy levels reported in this article are presented in cm−1, relative to the zero-point vibrational energy, 4362.1726 cm−1. The levels were obtained for each J, ϵ, and p set of values in a separate ScalIT calculation with a single PIST spectral window, usually including 70 to 120 ro-vibrational states. The number of computed levels and the highest energy level computed is summarized in Table 3 for each J value. Note that these numbers include the unphysical states, and the doubly degenerate states are counted twice. In total, 105 vibrational energy levels were computed, up to 16,500 cm−1, significantly over the isomerization barrier. For J = 1, the computed states (around 350 in all) covered the range up to 17,300 cm−1. For 2 J 7, around 420 to 480 levels were computed for each J up to the decreasing energy limit of 15,600 to 12,600 cm−1 with the increase of J. For higher J values, around 300 levels were computed for each J. For 8 J 20, the energy range increased from 10,800 cm−1 for J = 8 to up to 16,700 cm−1 for J = 20. All of the computed levels are included in the Supplementary Material.

In Table 2, the lowest ro-vibrational energy levels of H3+ for each J up to J = 20 are compared with past [35, 43] calculations. The PES75K+ has been only used in our previous study [43] so far, using nuclear masses. The levels computed here are only slightly lower than those, by 3–5 × 10–4 cm−1 for all levels. In our previous study, we also carried out computations [43] using the GLH3P PES, just as Ref [35], as well. Those two sets of numbers are identical and slightly higher than the numbers of this work, by up to 0.07 cm−1. Ref [35] also repeated their calculations by unequal vibrational and rotational masses, which yielded eigenvalues lower than ours, by 0.8 cm−1 up to J = 13. We did not modify our masses because we prefer keeping our calculations “ab initio”. In the future, we plan to do further investigations where we include non-adiabatic effects explicitly through non-adiabatic calculations with multiple PESs. [26].

TABLE 2
www.frontiersin.org

TABLE 2. The lowest ro-vibrational energy levels of H3+, in cm−1, up to J = 20 total angular momentum and their comparison with the literature. The past calculations employed different PESs and masses (VRM stands for Unequal Vibrational and Rotational Masses) [35].

The main focus of this work is, however, to provide vibrational and rotational quantum labels for as many states as possible. In our previous study [43], although D3h irrep labels were provided for selected J values (J = 10, 20, 30, 40, 46) up to J = 46, we only provided a limited number of vibrational and rotational quantum label assignments. Only low J values 0 ≤ J ≤ 5 and J = 10 were considered, and even those were mostly restricted to up to 8,000–9,000 cm−1, so overall, below the isomerization limit. Only for J = 0 and 1 did we go above 10,000 cm−1. Here, we pushed our efforts further with the help of wave function overlaps and provided assignments for almost all states below 10,000 cm−1. In Table 3, we summarize the number of states labeled in this work for each J value separately as well as include the labeling threshold, Elab. As for most J values, not all of the levels were assigned up to the given threshold energy, we separately include the energy value up to which all states are labeled, Elab.

TABLE 3
www.frontiersin.org

TABLE 3. For each J value, the total number of states computed (comp.) and labeled (lab.). Note that the former number includes the unphysical states as well, and there the doubly degenerate states are counted twice. Ecomp and Elab are the threshold energies for computation and labeling, respectively. As not all the states were labeled up to Elab, a separate threshold was also included, up to which all the states are labeled (Elab<Elab).

During this work, 2,210 ro-vibrational levels of H3+ have been assigned (v1, v2, |l|) vibrational and (J, G, U, K) rotational quantum labels. All of the labeled energy levels and their assignments are included in the Supplementary Material. For 1,571 of these levels, quantum labels have not been assigned before, while the remaining 639 levels are part of the 652 experimental ro-vibrational levels currently included in MARVEL. For 33 of these MARVEL levels, new vibrational (v1, v2, |l|) and/or rotational (J, G, U, K) assignments have been proposed (see Table 4).

TABLE 4
www.frontiersin.org

TABLE 4. Suggested vibrational (v1, v2, |l|) and rotational (J, G, U, K) reassignments of H3+ MARVEL energy levels.

In order to obtain all the quantum labels presented here, our approach has been adapted keeping in mind the difficulties we faced in Ref. [43]. Based on the energy formula of Eqs 2, 3, one would expect a somewhat regular behavior in the shifting of the rotationally excited energy levels belonging to the same vibrational parent. However, this seems to hold only for the l = 0 cases. In Figure 1, the change of B rotational constant is illustrated as J is increased. Each rotational progression belonging to a vibrational state can be characterized by a slightly different B rotational constant, which also shifts slightly by the increase of J. As the vibrational excitation increases, this shift becomes more significant (e.g. in case of 0, 2, 0 and 0, 4, 0). However, for the l > 0 vibrational parents, this shift can be more chaotic (see Figure 2) and it is different for the two distinct rotational progressions, assigned to G and G′ (see Figures 2, 3). For the (1, 3, 1) vibrational parent, e.g., the shift seems to first be positive and then it turns negative as it is in all other cases for both progressions. The G = 0 energy levels of the first 13 vibrational states up to J = 20 are listed in Table 5, while G′ = 0 energy levels of the first 7 l > 0 vibrational states up to J = 16 are included in Table 6.

FIGURE 1
www.frontiersin.org

FIGURE 1. The shift of the B rotational constants with the increase of the J value for the first 6 l = 0 vibrational parents.

FIGURE 2
www.frontiersin.org

FIGURE 2. The shift of the B rotational constants with the increase of the J value for the first 7 l > 0 vibrational parents. The numbers included here correspond to the G progression.

FIGURE 3
www.frontiersin.org

FIGURE 3. The shift of the B rotational constants with the increase of the J value for the first 7 l > 0 vibrational parents. The numbers included here correspond to the G′ progression.

TABLE 5
www.frontiersin.org

TABLE 5. The G = 0 energy levels of the first 13 vibrational states (v1, v2, |l|) of H3+ up to J = 20.

TABLE 6
www.frontiersin.org

TABLE 6. The G′ = 0 energy levels of the first 7 l > 0 vibrational states (v1, v2, |l|) of H3+ up to J = 16.

Using the RRD method [57] of GENIUSH to compute wave function overlaps and assign vibrational parents we can push the labeling a lot further than simply relying on the energy progressions. However, after a certain point in the vibrational and/or the rotational energy excitation, the mixing of the wave functions becomes simply too much for the vibrational parents to be unambiguously identified. For low J values, up to J = 6, we were able to get past the isomerization barrier, and in most cases, we could continue on even further. As the rotational excitation increases, the highest vibrational parent we can possibly assign is also decreasing. From J = 7 to 10, we are barely reaching the barrier to linearity, while as J increases further, the rotational energy contribution is also getting bigger, therefore we are again getting past the isomerization barrier (see Table 3). Different vibrational parents also behave differently, e.g. the overlaps of the (1,2,0) state breaks down a lot sooner than those of the next few higher-lying vibrational states. The more spread out progressions are also more difficult to assign fully including all the states within the progression. Although the progression belonging to (1,0,0) can be fully identified up to J = 18, in the progression of (0,1,1) we already start missing levels at J = 16. The ground vibrational state is the only vibrational parent for which all the states were found within the progressions for each J up to J = 20. This, however, might not be possible for J > 20 values.

In certain cases, it can be observed that the G = 0 (or G′ = 0) level is not the lowest level of the progression, which seemingly results in negative rotational excitations [108]. This happens for both higher vibrational excitations [e.g., for (0,5,1) at J = 1], higher rotational excitation (e.g., for (0,1,1) at J = 11), or for both [e.g., for (0,3,3) at J = 3 and for (0,3,1) at J = 6]. In certain cases, this reversing of the energy levels occurs sooner for the G′ progression [e.g., for (0,3,1) at J = 5].

4 Discussion

In this article, we computed ro-vibrational energy levels and wave functions for the H3+ molecular ion using a recently developed PES [25] and the ScalIT suite of parallel codes. The calculations included every single J value up to J = 20, and for each J, all of the levels were computed up to the barrier of linearity or higher. The convergence accuracy of our calculations is further improved, now reaching up to a few 10–5 cm−1 (or better). Our work has also been compared to previous works using different potential energy surfaces and different masses. Among the nuclear mass computations, the numbers of the present work were the lowest, signaling that we are still operating in the “basis set truncation error-dominated regime” where the numerical convergence is variational from above.

In addition, we carried out vibrational (v1, v2, |l|) and rotational (J, G, U, K) quantum label assignments of ro-vibrational energy levels for more than 2,200 states. To enable this, GENIUSH calculations were also carried out with lower accuracy to obtain ro-vibrational wave functions, which were then used to compute wave function overlaps within the framework of the rigid rotor decomposition method. These RRD overlaps helped greatly to identify the vibrational parents. As part of our efforts, we suggested new vibrational (v1, v2, |l|) and rotational (J, G, U, K) reassignments for certain energy levels within the MARVEL database. We are hoping that the results of this work can be used to further improve previous efforts toward creating spectroscopic line lists (based on both theoretical and experimental data), through the list of all labeled energy levels as provided in the Supplementary Material of this article.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary Material; further inquiries can be directed to the corresponding author.

Author contributions

JS and BP designed the research, JS carried out the calculations and analyzed the data, JS and BP discussed the results, JS and BP wrote the manuscript, and BP provided funding.

Funding

This work was supported by the National Science Foundation (CHE-1665370), and the Robert A. Welch Foundation (D-1523).

Acknowledgments

The authors also gratefully acknowledge the Texas Tech University High Performance Computing Center for the use of the Quanah cluster and the Texas Advanced Computing Center for the use of the Lonestar5 and Frontera clusters.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fphy.2022.996001/full#supplementary-material

References

1. Miller S, Tennyson J, Geballe TR, Stallard T. Thirty years of H3+ astronomy. Rev Mod Phys (2020) 92:035003. doi:10.1103/revmodphys.92.035003

CrossRef Full Text | Google Scholar

2. Snow TP, McCall BJ. Diffuse atomic and molecular clouds. Annu Rev Astron Astrophys (2006) 44:367–414. doi:10.1146/annurev.astro.43.072103.150624

CrossRef Full Text | Google Scholar

3. Miller S, Tennyson J, Lepp S, Dalgarno A. Identification of features due to H3+ in the infrared spectrum of supernova 1987A. Nature (1992) 355:420–2. doi:10.1038/355420a0

CrossRef Full Text | Google Scholar

4. Miller S, Achilleos N, Ballester GE, Geballe TR, Joseph RD, Prange R, et al. The role of H3+ in planetary atmospheres. Philosophical Trans R Soc Lond Ser A: Math Phys Eng Sci (2000) 358:2485–502. doi:10.1098/rsta.2000.0662

CrossRef Full Text | Google Scholar

5. Koskinen TT, Aylward AD, Miller S. A stability limit for the atmospheres of giant extrasolar planets. Nature (2007) 450:845–8. doi:10.1038/nature06378

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Carrington A, Buttenshaw J, Kennedy R. Observation of the infrared spectrum of H3+ at its dissociation limit. Mol Phys (1982) 45:753–8. doi:10.1080/00268978200100591

CrossRef Full Text | Google Scholar

7. Carrington A, McNab IR, West YD. Infrared predissociation spectrum of the H3+ ion. II. J Chem Phys (1993) 98:1073–92. doi:10.1063/1.464331

CrossRef Full Text | Google Scholar

8. Lindsay C, McCall BJ. Comprehensive evaluation and compilation of H3+ spectroscopy. J Mol Spectrosc (2001) 210:60–83. doi:10.1006/jmsp.2001.8444

CrossRef Full Text | Google Scholar

9. Schiffels P, Alijah A, Hinze J. Rovibrational states of H3+. Part 1: The energy region below 9,000 cm−1 and modelling of the non-adiabatic effects. Mol Phys (2001) 101:175. doi:10.1080/00268970210158687

CrossRef Full Text | Google Scholar

10. Schiffels P, Alijah A, Hinze J. Rovibrational states of H3+. Part 2: The energy region between 9,000 cm-1 and 13,000 cm-1 including empirical corrections for the non-adiabatic effects. Mol Phys (2001) 101:189. doi:10.1080/00268970210158713

CrossRef Full Text | Google Scholar

11. Asvany O, Hugo E, Schlemmer S, Muller F, Kuhnemann F, Schiller S, et al. Overtone spectroscopy of H2D+ and D2H+ using laser induced reactions. J Chem Phys (2007) 127:154317. doi:10.1063/1.2794331

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Pavanello M, Adamowicz L, Alijah A, Zobov NF, Mizus II, Polyansky OL, et al. Precision measurements and computations of transition energies in rotationally cold triatomic hydrogen ions up to the midvisible spectral range. Phys Rev Lett (2012) 108:023002. doi:10.1103/physrevlett.108.023002

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Furtenbacher T, Szidarovszky T, Mátyus E, Fábri C, Csaszár AG. Analysis of the rotational–vibrational states of the molecular ion H3+. J Chem Theor Comput (2013) 9:5471–8. doi:10.1021/ct4004355

CrossRef Full Text | Google Scholar

14. Furtenbacher T, Szidarovszky T, Fábri C, Császár AG. MARVEL analysis of the rotational–vibrational states of the molecular ions H2D+ and D2H+. Phys Chem Chem Phys (2013) 15:10181. doi:10.1039/c3cp44610g

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Császár AG, Furtenbacher T. Spectroscopic networks. J Mol Spectrosc (2011) 266:99–103. doi:10.1016/j.jms.2011.03.031

CrossRef Full Text | Google Scholar

16. Császár AG, Furtenbacher T, Árendás P. Small molecules—big Data. J Phys Chem A (2016) 120:8949–69. doi:10.1021/acs.jpca.6b02293

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Furtenbacher T, Császár AG, Tennyson J. Marvel: Measured active rotational–vibrational energy levels. J Mol Spectrosc (2007) 245:115–25. doi:10.1016/j.jms.2007.07.005

CrossRef Full Text | Google Scholar

18. Furtenbacher T, Császár AG. Marvel: Measured active rotational–vibrational energy levels. II. Algorithmic improvements. J Quant Spectrosc Radiat Transf (2012) 113:929–35. doi:10.1016/j.jqsrt.2012.01.005

CrossRef Full Text | Google Scholar

19.Marvel online (2022). Available from: https://kkrk.chem.elte.hu/marvelonline/index.php (accessed 07 11, 2022).

20. Cencek W, Rychlewski J, Jaquet R, Kutzelnigg W. Sub-microhartree accuracy potential energy surface for H3+ including adiabatic and relativistic effects. I. Calculation of the potential points. J Chem Phys (1998) 108:2831–6. doi:10.1063/1.475702

CrossRef Full Text | Google Scholar

21. Jaquet R, Cencek W, Kutzelnigg W, Rychlewski J. Sub-microhartree accuracy potential energy surface for H3+ including adiabatic and relativistic effects. II. Rovibrational analysis for H3+ and D3+. J Chem Phys (1998) 108:2837–46. doi:10.1063/1.475703

CrossRef Full Text | Google Scholar

22. Polyansky O, Prosmiti R, Klopper W, Tennyson J. An accurate, global, ab initio potential energy surface for the H+3 molecule. Mol Phys (2000) 98:261–73. doi:10.1080/00268970009483290

CrossRef Full Text | Google Scholar

23. Velilla L, Lepetit B, Aguado A, Beswick J, Paniagua M. The H3+ rovibrational spectrum revisited with a global electronic potential energy surface. J Chem Phys (2008) 129:084307. doi:10.1063/1.2973629

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Pavanello M, Adamowicz L, Alijah A, Zobov NF, Mizus II, Polyansky OL, et al. Calibration-quality adiabatic potential energy surfaces for H3+ and its isotopologues. J Chem Phys (2012) 136:184303. doi:10.1063/1.4711756

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Mizus II, Polyansky OL, McKemmish LK, Tennyson J, Alijah A, Zobov NF. A global potential energy surface for H3+. Mol Phys (2019) 117:1663–72. doi:10.1080/00268976.2018.1554195

CrossRef Full Text | Google Scholar

26. Aguado A, Roncero O, Sanz-Sanz C. Three states global fittings with improved long range: Singlet and triplet states of H3+. Phys Chem Chem Phys (2021) 113. doi:10.1039/D0CP04100A

CrossRef Full Text | Google Scholar

27. Miller S, Tennyson J. Calculation of the high angular momentum dissociation limit for H3+ and H2D+. Chem Phys Lett (1988) 145:117–20. doi:10.1016/0009-2614(88)80161-1

CrossRef Full Text | Google Scholar

28. Kozin IN, Roberts RM, Tennyson J. Symmetry and structure of rotating H3+. J Chem Phys (1999) 111:140–50. doi:10.1063/1.479260

CrossRef Full Text | Google Scholar

29. Kostin MA, Polyansky OL, Tennyson J, Mussa HY. Rotation-vibration states of H3+ at dissociation. J Chem Phys (2003) 118:3538–42. doi:10.1063/1.1539034

CrossRef Full Text | Google Scholar

30. Munro JJ, Ramanlal J, Tennyson J, Mussa HY. Properties of high-lying vibrational states of the molecular ion. Mol Phys (2006) 104:115–25. doi:10.1080/00268970500399648

CrossRef Full Text | Google Scholar

31. Bachorz RA, Cencek W, Jaquet R, Komasa J. Rovibrational energy levels of H3+ with energies above the barrier to linearity. J Chem Phys (2009) 131:024105. doi:10.1063/1.3167795

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Alijah A. Accurate calculations and assignments of above-barrier states of up to. J Mol Spectrosc (2010) 264:111–9. doi:10.1016/j.jms.2010.09.009

CrossRef Full Text | Google Scholar

33. Szidarovszky T, Császár AG, Czakó G. On the efficiency of treating singularities in triatomic variational vibrational computations. The vibrational states of H3+ up to dissociation. Phys Chem Chem Phys (2010) 12:8373. doi:10.1039/c001124j

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Diniz LG, Alijah JRMA, Pavanello M, Adamowicz L, Polyansky OL, Tennyson J, et al. Vibrationally and rotationally nonadiabatic calculations on H3+ using coordinate-dependent vibrational and rotational masses. Phys Rev A (Coll Park) (2013) 88:032506. doi:10.1103/physreva.88.032506

CrossRef Full Text | Google Scholar

35. Jaquet R, Carrington T. Using a nondirect product basis to compute J > 0 rovibrational states of H3+. J Phys Chem A (2013) 117:9493–500. doi:10.1021/jp312027s

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Jaquet R. Investigation of the highest bound ro-vibrational states of H3+, DH2+, HD2+, D3+, and T3+: Use of a non-direct product basis to compute the highest allowed J> 0 states. Mol Phys (2013) 111:2606–16. doi:10.1080/00268976.2013.818727

CrossRef Full Text | Google Scholar

37. Mátyus E, Szidarovszky T, Császár AG. Modelling non-adiabatic effects in H3+: Solution of the rovibrational Schrödinger equation with motion-dependent masses and mass surfaces. J Chem Phys (2014) 141:154111. doi:10.1063/1.4897566

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Jaquet R, Khoma MV. Investigation of nonadiabatic effects for the vibrational spectrum of a triatomic molecule: The use of a single potential energy surface with distance-Dependent masses for H3+. J Phys Chem A (2017) 121:7016–30. doi:10.1021/acs.jpca.7b04703

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Jaquet R, Khoma MV. Investigation of non-adiabatic effects for the ro-vibrational spectrum of H3+: The use of a single potential energy surface with geometry-dependent nuclear masses. Mol Phys (2018) 116:3507–18. doi:10.1080/00268976.2018.1464225

CrossRef Full Text | Google Scholar

40. Tennyson J, Polyansky OL, Zobov NF, Alijah A, Császár AG. High-accuracy calculations of the rotation-vibration spectrum of H3+. J Phys B: Mol Opt Phys (2017) 50:232001. doi:10.1088/1361-6455/aa8ca6

CrossRef Full Text | Google Scholar

41. Jaquet R, Lesiuk M. Analysis of QED and non-adiabaticity effects on the rovibrational spectrum of H3+ using geometry-dependent effective nuclear masses. J Chem Phys (2020) 152:104109. doi:10.1063/1.5144293

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Sarka J, Poirier B, Szalay V, Császár AG. On neglecting Coriolis and related couplings in first-principles rovibrational spectroscopy: Considerations of symmetry, accuracy, and simplicity. II. Case studies for H2O isotopologues, H3+, O3, and NH3. Spectrochim Acta A Mol Biomol Spectrosc (2021) 250:119164. doi:10.1016/j.saa.2020.119164

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Sarka J, Das D, Poirier B. Calculation of rovibrational eigenstates of H3+ using ScalIT. AIP Adv (2021) 11:045033. doi:10.1063/5.0047823

CrossRef Full Text | Google Scholar

44. Moss RE. On the adiabatic and non-adiabatic corrections in the ground electronic state of the hydrogen molecular cation. Mol Phys (1996) 89:195–210. doi:10.1080/002689796174083

CrossRef Full Text | Google Scholar

45. Diniz LG, Alijah A, Mohallem JR. Core-mass nonadiabatic corrections to molecules: H2, H2+, and isotopologues. J Chem Phys (2012) 137:164316. doi:10.1063/1.4762442

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Huestic DL. Hydrogen collisions in planetary atmospheres, ionospheres, and magnetospheres. Planet Space Sci (2008) 56:1733. doi:10.1016/j.pss.2008.07.012

CrossRef Full Text | Google Scholar

47. McKemmish LK, Tennyson J. General mathematical formulation of scattering processes in atom–diatomic collisions in the RmatReact methodology. Phil Trans R Soc A (2019) 377:20180409. doi:10.1098/rsta.2018.0409

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Höveler K, Deiglmayr J, Agner JA, Schmutz H, Merkt F. The H2+ + HD reaction at low collision energies: H3+/H2D+ branching ratio and product-kinetic-energy distributions. Phys Chem Chem Phys (2021) 23:2676–85. doi:10.1039/d0cp06107g

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Huang X, Schwenke DW, Lee TJ. Highly accurate potential energy surface, dipole moment surface, rovibrational energy levels, and infrared line list for 32S16O2 up to 8,000 cm-1. J Chem Phys (2014) 140:114311. doi:10.1063/1.4868327

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Kumar P, Jiang B, Guo H, Klos J, Alexander MH, Poirier B. Photoabsorption assignments for the C̃1B2 ← X̃1A1 vibronic transitions of SO2, using new ab initio potential energy and transition Dipole surfaces. J Phys Chem A (2017) 121:1012–21. doi:10.1021/acs.jpca.6b12958

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Kao L, Oka T, Miller S, Tennyson J. A table of astronomically important ro-vibrational transitions for the H3+ molecular ion. Astrophys J Suppl Ser (1991) 77:317. doi:10.1086/191606

CrossRef Full Text | Google Scholar

52. Neale L, Miller S, Tennyson J. Spectroscopic properties of the H3+ molecule: A new calculated line list. Astrophys J (1996) 464:516. doi:10.1086/177341

CrossRef Full Text | Google Scholar

53. Mizus II, Alijah A, Zobov NF, Lodi L, Kyuberis AA, Yurchenko SN, et al. ExoMol molecular line lists – XX. A comprehensive line list for H3+. Mon Not R Astron Soc (2017) 468:1717–25. doi:10.1093/mnras/stx502

CrossRef Full Text | Google Scholar

54. Tennyson J, Kostin MA, Barletta P, Harris GJ, Polyansky OL, Ramanlal J, et al. DVR3D: A program suite for the calculation of rotation–vibration spectra of triatomic molecules. Computer Phys Commun (2004) 163:85–116. doi:10.1016/j.cpc.2003.10.003

CrossRef Full Text | Google Scholar

55. Tennyson J. Perspective: Accurate ro-vibrational calculations on small molecules. J Chem Phys (2016) 145:120901. doi:10.1063/1.4962907

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Tennyson J, Yurchenko SN. The ExoMol project: Software for computing large molecular line lists. Int J Quan Chem (2017) 117:92–103. doi:10.1002/qua.25190

CrossRef Full Text | Google Scholar

57. Mátyus E, Fábri C, Szidarovszky T, Czakó G, Allen WD, Császár AG. Assigning quantum labels to variationally computed rotational-vibrational eigenstates of polyatomic molecules. J Chem Phys (2010) 133:034113. doi:10.1063/1.3451075

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Chen W, Poirier B. Parallel implementation of efficient preconditioned linear solver for grid-based applications in chemical physics. I: Block Jacobi diagonalization. J Comput Phys (2006) 219:185–97. doi:10.1016/j.jcp.2006.04.012

CrossRef Full Text | Google Scholar

59. Chen W, Poirier B. Parallel implementation of efficient preconditioned linear solver for grid-based applications in chemical physics. II: QMR linear solver. J Comput Phys (2006) 219:198–209. doi:10.1016/j.jcp.2006.03.031

CrossRef Full Text | Google Scholar

60. Chen W, Poirier B. Parallel implementation of an efficient preconditioned linear solver for grid-based applications in chemical physics. III: Improved parallel scalability for sparse matrix–vector products. J Parallel Distrib Comput (2010) 70:779–82. doi:10.1016/j.jpdc.2010.03.008

CrossRef Full Text | Google Scholar

61. Chen W, Poirier B. Quantum dynamics on massively parallel computers: efficient numerical implementation for preconditioned linear solvers and eigensolvers. J Theor Comput Chem (2010) 9:825–46. doi:10.1142/s021963361000602x

CrossRef Full Text | Google Scholar

62. Petty C, Poirier B. Using ScalIT for performing accurate rovibrational spectroscopy calculations for triatomic molecules: A practical guide. Appl Math (Irvine) (2014) 5:2756–63. doi:10.4236/am.2014.517263

CrossRef Full Text | Google Scholar

63. Poirier B, Light JC. Phase space optimization of quantum representations: Direct-product basis sets. J Chem Phys (1999) 111:4869–85. doi:10.1063/1.479747

CrossRef Full Text | Google Scholar

64. Poirier B, Light JC. Phase space optimization of quantum representations: Three-body systems and the bound states of HCO. J Chem Phys (2001) 114:6562–71. doi:10.1063/1.1354181

CrossRef Full Text | Google Scholar

65. Poirier B. Research topic "proton transfer processes in biological reactions: a computational approach" frontiers in chemistry journal. Found Phys (2001) 31:1581–610. doi:10.1023/a:1012642832253

CrossRef Full Text | Google Scholar

66. Bian W, Poirier B. Accurate and highly efficient calculation of the O(1D)HCl vibrational bound states, using A combination of methods. J Theor Comput Chem (2003) 2:583–97. doi:10.1142/s0219633603000768

CrossRef Full Text | Google Scholar

67. Light J, Carrington T. Discrete-variable representations and their utilization Adv Chem Phys (2000) 114:263. doi:10.1002/9780470141731.ch4

CrossRef Full Text | Google Scholar

68. Chen W, Poirier B. Quantum Dynamical calculation of all rovibrational states of HO2 for total angular momentum J = 0–10. J Theor Comput Chem (2010) 9:435–69. doi:10.1142/s0219633610005815

CrossRef Full Text | Google Scholar

69. Huang S-W, Carrington T. A new iterative method for calculating energy levels and wave functions. J Chem Phys (2000) 112:8765–71. doi:10.1063/1.481492

CrossRef Full Text | Google Scholar

70. Poirier B, Carrington T. Accelerating the calculation of energy levels and wave functions using an efficient preconditioner with the inexact spectral transform method. J Chem Phys (2001) 114:9254–64. doi:10.1063/1.1367396

CrossRef Full Text | Google Scholar

71. Poirier B, Carrington T. A preconditioned inexact spectral transform method for calculating resonance energies and widths, as applied to HCO. J Chem Phys (2002) 116:1215–27. doi:10.1063/1.1428752

CrossRef Full Text | Google Scholar

72. Poirier B, Miller WH. Optimized preconditioners for Green function evaluation in quantum reactive scattering calculations. Chem Phys Lett (1997) 265:77–83. doi:10.1016/s0009-2614(96)01408-x

CrossRef Full Text | Google Scholar

73. Poirier B. Optimal separable bases and series expansions. Phys Rev A (Coll Park) (1997) 56:120–30. doi:10.1103/physreva.56.120

CrossRef Full Text | Google Scholar

74. Poirier B. Quantum reactive scattering for three-body systems via optimized preconditioning, as applied to the O+HCl reaction. J Chem Phys (1998) 108:5216–24. doi:10.1063/1.475958

CrossRef Full Text | Google Scholar

75. Poirier B. Efficient preconditioning scheme for block partitioned matrices with structured sparsity. Numer Linear Algebra Appl (2000) 7:715–26. doi:10.1002/1099-1506(200010/12)7:7/83.0.CO;2-R

CrossRef Full Text | Google Scholar

76. Freund RW, Nachtigal NM. QMR: A quasi-minimal residual method for non-hermitian linear systems. Numer Math (1991) 60:315–39. doi:10.1007/bf01385726

CrossRef Full Text | Google Scholar

77. Petty C, Chen W, Poirier B. Quantum Dynamical calculation of bound rovibrational states of HO2 up to largest possible total angular momentum, J ≤ 130. J Phys Chem A (2013) 117:7280–97. doi:10.1021/jp401154m

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Petty C, Poirier B. Comparison of J-shifting models for rovibrational spectra as applied to the HO2 molecule. Chem Phys Lett (2014) 605–606:16–21. doi:10.1016/j.cplett.2014.05.003

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Yang B, Chen W, Poirier B. Rovibrational bound states of neon trimer: Quantum dynamical calculation of all eigenstate energy levels and wavefunctions. J Chem Phys (2011) 135:094306. doi:10.1063/1.3630922

PubMed Abstract | CrossRef Full Text | Google Scholar

80. Brandon D, Poirier B. Accurate calculations of bound rovibrational states for argon trimer. J Chem Phys (2014) 141:034302. doi:10.1063/1.4887459

PubMed Abstract | CrossRef Full Text | Google Scholar

81. Yang B, Poirier B. Quantum dynamical calculation of rovibrational bound states of Ne2Ar. J Phys B: Mol Opt Phys (2012) 45:135102. doi:10.1088/0953-4075/45/13/135102

CrossRef Full Text | Google Scholar

82. Yang B, Poirier B. Rovibrational bound states of the Ar2Ne complex. J Theor Comput Chem (2013) 12:1250107. doi:10.1142/s0219633612501076

CrossRef Full Text | Google Scholar

83. Zhang Z, Li B, Shen Z, Ren Y, Bian W. Efficient quantum calculation of the vibrational states of acetylene. Chem Phys (2012) 400:1–7. doi:10.1016/j.chemphys.2012.01.010

CrossRef Full Text | Google Scholar

84. Petty C, Spada RF, Machado FB, Poirier B. Accurate rovibrational energies of ozone isotopologues up to J= 10 utilizing artificial neural networks. J Chem Phys (2018) 149:024307. doi:10.1063/1.5036602

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Sarka J, Poirier B. Comment on “Calculated vibrational states of ozone up to dissociation” [J. Chem. Phys. 144, 074302 (2016)]. J Chem Phys (2020) 152:177101. doi:10.1063/5.0002762

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Sarka J, Petty C, Poirier B. Exact bound rovibrational spectra of the neon tetramer. J Chem Phys (2019) 151:174304. doi:10.1063/1.5125145

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Sarka J, Poirier B. Hitting the trifecta: How to simultaneously push the limits of schrödinger solution with respect to system size, convergence accuracy, and number of computed states. J Chem Theor Comput (2021) 17:7732–44. doi:10.1021/acs.jctc.1c00824

CrossRef Full Text | Google Scholar

88. Yurchenko SN, Thiel W, Jensen P. Theoretical ROVibrational energies (TROVE): A robust numerical approach to the calculation of rovibrational energies for polyatomic molecules. J Mol Spectrosc (2007) 245:126–40. doi:10.1016/j.jms.2007.07.009

CrossRef Full Text | Google Scholar

89. Yurchenko SN, Yachmenev A, Ovsyannikov RI. Symmetry-adapted ro-vibrational basis functions for variational nuclear motion calculations: TROVE approach. J Chem Theor Comput (2017) 13:4368–81. doi:10.1021/acs.jctc.7b00506

CrossRef Full Text | Google Scholar

90. Yurchenko SN, Mellor TM. Treating linear molecules in calculations of rotation-vibration spectra. J Chem Phys (2020) 153:154106. doi:10.1063/5.0019546

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Czakó G, Furtenbacher T, Császár AG, Szalay V. Variational vibrational calculations using high-order anharmonic force fields. Mol Phys (2004) 102:2411. doi:10.1080/0026897042000274991

CrossRef Full Text | Google Scholar

92. Furtenbacher T, Czakó G, Sutcliffe BT, Császár AG, Szalay V. The methylene saga continues: Stretching fundamentals and zero-point energy of X˜3B1 CH2. J Mol Struct (2006) 780–781:283–94. doi:10.1016/j.molstruc.2005.06.052

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Mátyus E, Czakó G, Sutcliffe BT, Császár AG. Vibrational energy levels with arbitrary potentials using the Eckart-Watson Hamiltonians and the discrete variable representation. J Chem Phys (2007) 127:084102. doi:10.1063/1.2756518

PubMed Abstract | CrossRef Full Text | Google Scholar

94. Mátyus E, Šimunek J, Császár AG. On the variational computation of a large number of vibrational energy levels and wave functions for medium-sized molecules. J Chem Phys (2009) 131:074106. doi:10.1063/1.3187528

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Fábri C, Mátyus E, Furtenbacher T, Mihály B, Zoltáni T, Nemes L, et al. Variational quantum mechanical and active database approaches to the rotational-vibrational spectroscopy of ketene, H2CCO. J Chem Phys (2011) 135:094307. doi:10.1063/1.3625404

PubMed Abstract | CrossRef Full Text | Google Scholar

96. Mátyus E, Czakó G, Császár AG. Toward black-box-type full- and reduced-dimensional variational (ro)vibrational computations. J Chem Phys (2009) 130:134112. doi:10.1063/1.3076742

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Fábri C, Mátyus E, Császár AG. Rotating full- and reduced-dimensional quantum chemical models of molecules. J Chem Phys (2011) 134:074105. doi:10.1063/1.3533950

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Lauvergnat D, Nauts A. Exact numerical computation of a kinetic energy operator in curvilinear coordinates. J Chem Phys (2002) 116:8560. doi:10.1063/1.1469019

CrossRef Full Text | Google Scholar

99. Lauvergnat D, Nauts A. Quantum dynamics with sparse grids: A combination of smolyak scheme and cubature. Application to methanol in full dimensionality. Spectrochimica Acta A: Mol Biomol Spectrosc (2014) 119:18–25. doi:10.1016/j.saa.2013.05.068

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Nauts A, Lauvergnat D. Numerical on-the-fly implementation of the action of the kinetic energy operator on a vibrational wave function: Application to methanol. Mol Phys (2018) 116:3701–9. doi:10.1080/00268976.2018.1473652

CrossRef Full Text | Google Scholar

101. Chen A, Nauts A, Lauvergnat D. Elvibrot-mpi: Parallel quantum dynamics with smolyak algorithm for general molecular simulation (2021). doi:10.48550/arXiv.2111.13655

CrossRef Full Text | Google Scholar

102. Al-Refaie AF, Yurchenko SN, Tennyson J. GPU Accelerated intensities MPI (GAIN-MPI): A new method of computing Einstein-A coefficients. Comput Phys Commun (2017) 214:216. doi:10.1016/j.cpc.2017.01.013

CrossRef Full Text | Google Scholar

103. Zhang JZH. Theory and application of quantum molecular dynamics. Singapore: World Scientific (1999).

Google Scholar

104. T Pack R, Parker GA. Quantum reactive scattering in three dimensions using hyperspherical (APH) coordinates. Theory. J Chem Phys (1987) 87:3888–921. doi:10.1063/1.452944

CrossRef Full Text | Google Scholar

105. Hougen JT. Classification of rotational energy levels for symmetric‐top molecules. J Chem Phys (1962) 37:1433–41. doi:10.1063/1.1733301

CrossRef Full Text | Google Scholar

106. Watson JKG. Simplification of the molecular vibration-rotation Hamiltonian. Mol Phys (1968) 15:479–90. doi:10.1080/00268976800101381

CrossRef Full Text | Google Scholar

107. Watson JKG. Higher-order vibration-rotation energies of the X3 molecule. J Mol Spectrosc (1984) 103:350–63. doi:10.1016/0022-2852(84)90062-6

CrossRef Full Text | Google Scholar

108. Császár AG, Fábri C, Sarka J. Quasistructural molecules. Wires Comput Mol Sci (2020) 10:e1432. doi:10.1002/wcms.1432

CrossRef Full Text | Google Scholar

109. Fábri C, Sarka J, Császár AG. Communication: Rigidity of the molecular ion H5+. J Chem Phys (2014) 140:051101. doi:10.1063/1.4864360

PubMed Abstract | CrossRef Full Text | Google Scholar

110. Asvany O, Yamada KMT, Brünken S, Potapov A, Schlemmer S. Experimental ground-state combination differences of CH5+. Science (2015) 347:1346–9. doi:10.1126/science.aaa3304

PubMed Abstract | CrossRef Full Text | Google Scholar

111. Fábri C, Quack M, Császár AG. On the use of nonrigid-molecular symmetry in nuclear motion computations employing a discrete variable representation: A case study of the bending energy levels of CH5+. J Chem Phys (2017) 147:134101. doi:10.1063/1.4990297

PubMed Abstract | CrossRef Full Text | Google Scholar

112. Fábri C, Császár AG. Vibrational quantum graphs and their application to the quantum dynamics of CH5+. Phys Chem Chem Phys (2018) 20:16913–7. doi:10.1039/c8cp03019g

PubMed Abstract | CrossRef Full Text | Google Scholar

113. Sarka J, Császár AG, Mátyus E. Rovibrational quantum dynamical computations for deuterated isotopologues of the methane–water dimer. Phys Chem Chem Phys (2017) 19:15335–45. doi:10.1039/c7cp02061a

PubMed Abstract | CrossRef Full Text | Google Scholar

114. Metz MP, Szalewicz K, Sarka J, Tóbiás R, Császár AG, Mátyus E. Molecular dimers of methane clathrates: Ab initio potential energy surfaces and variational vibrational states. Phys Chem Chem Phys (2019) 21:13504–25. doi:10.1039/c9cp00993k

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: ab initio, ro-vibrational energy levels, quantum label assignment, high accuracy, ScalIT code

Citation: Sarka J and Poirier B (2022) Assigning quantum labels and improving accuracy for the ro-vibrational eigenstates of H3+ calculated using ScalIT. Front. Phys. 10:996001. doi: 10.3389/fphy.2022.996001

Received: 16 July 2022; Accepted: 12 August 2022;
Published: 04 October 2022.

Edited by:

Breno R. L. Galvão, Federal Center for Technological Education of Minas Gerais, Brazil

Reviewed by:

Oleg Polyansky, University College London, United Kingdom
Joseph Abebe Obu, University of Calabar, Nigeria

Copyright © 2022 Sarka and Poirier. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: János Sarka, Janos.Sarka@ttu.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.