Open Access Article
Valerii Andreichev
a,
Silvan Käser†
a,
Erica L. Bocanegrab,
Madeeha Salikb,
Mark A. Johnson
b and
Markus Meuwly
*a
aDepartment of Chemistry, University of Basel, Klingelbergstrasse 80, CH-4056 Basel, Switzerland. E-mail: m.meuwly@unibas.ch
bSterling Chemistry Laboratory, Yale University, New Haven, Connecticut 06520, USA
First published on 21st October 2025
The infrared spectroscopy and proton transfer dynamics together with the associated tunneling splittings for H/D-transfer in oxalate are investigated using a machine learning-based potential energy surface (PES) of CCSD(T) quality, calibrated against the results of new spectroscopic measurements. Second order vibrational perturbation calculations (VPT2) very successfully describe both the framework and H-transfer modes compared with the experiments. In particular, a newly observed low-intensity signature at 1666 cm−1 was correctly predicted from the VPT2 calculations. An unstructured band centered at 2940 cm−1 superimposed on a broad background extending from 2600 to 3200 cm−1 is assigned to the H-transfer motion. The broad background involves a multitude of combination bands but a major role is played by the COH-bend. For the deuterated species, VPT2 and molecular dynamics simulations provide equally convincing assignments, in particular for the framework modes. Finally, based on the new PES the tunneling splitting for H-transfer is predicted as ΔH = 35.0 cm−1 from ring polymer instanton calculations using higher-order corrections. This provides an experimentally accessible benchmark to validate the computations, in particular the quality of the machine-learned PES.
One of the most direct ways to arrive at assignment and identification of spectral patterns is through computation. Vibrational spectroscopy is an area where “experiment” and “theory/computation” meet in a most natural and direct fashion. However, fruitful interplay between these two approaches requires high-level computer methods which typically need to go beyond the harmonic approximation. It is generally believed that calculations at the coupled cluster with singles, doubles and perturbative triples (CCSD(T)) are needed for reliable computational vibrational spectroscopic work.5 Also, the nuclear dynamics needs to be described at a sufficiently advanced level. Standard normal mode analysis is often sufficient to obtain an overview of the spectral features. If quantitative agreement with experiment is sought, more advanced techniques such as vibrational perturbation theory (VPT2)6 or spectra from explicit molecular dynamics (MD) simulations are needed, though.2,7
A chemically particularly interesting motif involves hydrogen bonds and shared hydrogen atoms/protons between an acceptor and a donor. H-bonds and hydrogen/proton transfer (HT/PT) play important roles in governing molecular structure, stability, and dynamics and their energetics and dynamics are of fundamental importance in biology and chemistry.8–10 HT/PT is primarily influenced by the height of the barrier which is, however, difficult to determine reliably through direct experimentation. Possibilities include high resolution spectroscopy where the splitting of spectral lines can provide information about the barrier height,11 or nuclear magnetic resonance (NMR) experiments12,13 whereas kinetic isotope effects or shifts of bands in the vibrational spectrum alone can not be used directly to determine the energetics for PT. Proton transfer in systems containing X–H*⋯Y motifs – where X and Y are the donor and acceptor atoms, respectively, and H* is the transferring hydrogen – can lead to characteristically broadened features in vibrational spectra.14 This broadening reflects strong coupling between the X–H stretch and other framework modes of the environment and structural heterogeneity.15 The broadening can even persist down to low temperatures and cooling the species does not necessarily lead to sharper bands.16,17
In experiments, the broadening of spectral lines in X−H*⋯Y species has been extensively observed and reported in both the liquid and the gas phase.1 For example, in liquid water, the maximum in the OH-stretching spectrum shifts to the red as the temperature decreases from 47 °C to −6 °C, with concomitant increase of the intensity by 16% even without hydrogen or proton transfer to take place.18 In the high-density liquid, the line shapes in the OH stretching spectra of supercritical methanol was found to be sensitive to the hydrogen bonding network which depends on temperature.1
Spectral shifts and their broadening are even more pronounced in systems where a transferring proton H* is shared by a donor and acceptor moiety and potentially provides information about the proton transfer energetics. An empirical relationship between the position of the infrared (IR) absorption and the height of the proton transfer barrier has been found in combined computational/experimental investigations of acetylacetone19 and formic acid dimer.20 Earlier studies of the infrared spectroscopy of protonated ammonia dimer have also established that the IR-signatures are broad and correlated with the barrier height for proton transfer.21
Here, a high-level theoretical study is combined with improved spectroscopic measurements to elucidate the vibrational dynamics at play in the protonated oxalate (HO2CCO2−) anion, hereafter referred to as OxH and OxD for the hydrogenated and deuterated forms of the protonated dianion, see Fig. 1. OxH is a compelling system to study the dynamics and ensuing spectroscopic signatures of a shared proton coupled to framework modes of the surrounding molecular scaffold.15,22 Previous computational work concerned the kinetic isotope effects from electronic structure calculations,23 the solution dynamics of hydrated p-Oxa,24 and the IR spectroscopy in the gas phase using a reactive empirical energy function.22 The experimental IR spectrum features a rather diffuse/broad signal between 2600 and 2900 cm−1. One potential source is strong mechanical coupling between the high-frequency O–H stretch and (potentially) the low-frequency COH bends and/or skeletal deformations.
Studying the dynamics and spectroscopy of a transferring proton from MD simulations requires energy functions capable of describing bond-formation and bond-breaking. One such method is molecular mechanics with proton transfer (MMPT) which combines an accurate representation of the energetics for the transferring proton and a lower-level empirical force field for the remaining degrees of freedom which allows long simulation times.25–27 Using “PES morphing”28 provides flexibility to adapt characteristics such as the barrier height and position of the minima. More recently, alternative ways to represent high-dimensional potential energy surfaces have been explored. One of them is based on neural networks and was applied to bulk silicon or the H + HBr reaction.29,30 The evaluation of the neural network, once trained, is orders of magnitude faster than the underlying reference quantum chemical calculations the model was trained on. To reach the highest levels of quantum chemical theory, such as coupled cluster-level, which is required for spectroscopic studies, Transfer learning (TL) can be used. Such approaches have been shown to be data efficient for obtaining validated high quality PESs.31–40 TL is based on the notion that the topography of the energy landscape of a molecular system is very similar even at different levels of theory. This allows fine-tuning the learnable parameters of a neural network trained on a lower level of theory using small amounts of data calculated at a high level. Here, the MP2 level of theory provides a computationally efficient route to a broad coverage of the energy landscape, while the smaller high-accuracy dataset allows more precise modeling.
With such “vetted” PESs, it should be even possible to predict experimentally accessible observables. Similar to the shared proton situation in malonaldehyde or tropolone, OxH should exhibit tunneling splitting for the (quantum) motion between the two symmetrical wells. Quantum mechanical tunnelling is a key concept in chemistry and physics and is well established,41 and includes, inter alia, electron transfer in solution,42–45 enzymatic reactions,46–48 and the structure and dynamics of water.49–52 In contrast to the often challenging measurement of chemical rates with high precision, tunnelling splittings can be accurately measured through spectroscopic methods, e.g., using microwave53 or infrared spectroscopy.54 Tunneling splittings provide valuable benchmarks for validating state-of-the-art theoretical methodologies, which require both an accurate PES and an accurate treatment of the resulting nuclear Hamiltonian.55 The calculation of accurate tunneling splittings for multidimensional systems using fully quantum mechanical methods is a challenging task. The ring-polymer instanton (RPI) approach is a semiclassical approximation method for tunnelling calculations which is capable of scaling well with system size.50,56,57 However, instanton theory makes a local harmonic approximation around the optimal tunnelling pathway, resulting in deviations from quantum–mechanical benchmarks that can reach 20%.58 This error can be mitigated by perturbatively correcting the results using information about the third and fourth derivatives along the path.59 It has been demonstrated that correction of a low level potential energy surface to a higher level using TL, in combination with ring-polymer instanton, can be used to quantitatively characterise tunnelling splittings in such systems as malonaldehyde34 and larger systems consisting of 15-atom tropolone and 12-atom propiolic acid–formic acid dimer (PFD).40
The present work is structured as follows. First, the computational and experimental methods are described, followed by a validation of the machine learned PESs. Then, results on the IR spectroscopy of OxH and OxD, the H-transfer dynamics and HT tunneling splittings are presented. Finally, conclusions are drawn.
100 structures, for which reference energies (Eref), forces (Frefi,α), and dipole moments (prefα) (α = x, y, z) were determined at the MP2/aug-cc-pVTZ level of theory using MOLPRO64 and served as the reference data set. PhysNet was then trained by minimizing the loss function
![]() | (1) |
is a “nonhierarchical penalty” that regularizes the loss function.61 For training, a 80/10/10% split of the data as training/validation/test sets was used. Two independent PhysNet models were trained, which were subsequently employed for active learning.68
Adaptive sampling69 and diffusion Monte Carlo simulations (DMC)70 were run to ensure the robustness of the PES and to locate so-called “holes”. Adaptive sampling was done at 1000 K in the NVT ensemble and structures (to be included in training the next models) were detected if the predictions of the two independent models differed by more than 0.5 kcal mol−1. All DMC simulations employed a GPU implementation of the DMC algorithm33 for PhysNet and were run with 300 walkers, a step size of 5 a.u. and for a total of 60
000 steps. Unphysical structures were saved if the predicted energy was lower than the energy of the minimum energy structure. During the first active learning cycle, adaptive sampling provided 80 structures and 20 defective structures were located from DMC simulations. Ab initio MP2 data for the 100 structures was determined and added to the MP2 data set before retraining PhysNet. A second active learning cycle was performed, however, no additional deficient structures were found. The final and robust MP2-based PES constitutes the “base model”.
To improve the performance of the base model for molecular spectroscopy, TL was used to refine and elevate the PES from the MP2/aug-cc-pVTZ to the gold-standard CCSD(T)/aug-cc-pVTZ level of theory.71–73 The TL data set contains a total of 2688 structures selected semi-randomly from the original data set. It comprises 1067 structures that were sampled using MD simulations at 1000 K, 960 and 661 structures obtained from normal mode sampling around the minimum and transition state, respectively, for which energies, forces and dipole moments were determined at the CCSD(T)/aug-cc-pVTZ level of theory using MOLPRO.64 These structures and corresponding quantum chemical information were then used to fine-tune the learnable parameters of the MP2 PES with the same hyperparameters as for the low-level training except for the learning rate and the weighting hyperparameters in the loss function. The learning rate was reduced to 10−4 and wE = wF = wQ = wp = 1.
![]() | (2) |
![]() | (3) |
and βN = β/N. The first term in eqn (3) corresponds to the sum over all single bead potentials and the second term represents the harmonic springs with frequency 1/(βNħ) that connect adjacent beads. As the potential UN and the action S are related by division with (βNħ), the action S contains information along the instanton path (IP). A detailed description of the method is provided, e.g., in ref. 79 and 80.
Within standard RPI (sRPI) theory, fluctuations around the instanton path are included up to second order and the information is combined into a contribution Φ. More specifically, fluctuations around the path are determined from the Hessian matrix at each of the beads.34,57 With this, the leading-order tunneling splitting in a symmetric double-well system is
![]() | (4) |
| PhysNet (MP2) | PhysNet (CCSD(T)) | |
|---|---|---|
| MAE(E) | 0.009 | 0.005 |
| RMSE(E) | 0.047 | 0.012 |
| MAE(F) | 0.065 | 0.023 |
| RMSE(F) | 0.459 | 0.065 |
| 1 − R2 | 5.9 × 10−06 | 7.8 × 10−07 |
| Ea | 2.355 | 3.351 |
| ΔEa | 8 × 10−04 | 7 × 10−04 |
| MAE(ω) | 0.09 | 0.34 |
![]() | ||
| Fig. 2 Experimental predissociation vibrational spectrum of doubly-H2 tagged oxalate. For clarity, the spectrum in the range 2400–3400 cm−1 is expanded by a factor of 5 in the inset. The sharper feature at 4090 cm−1 is due to the nominally IR forbidden H2 fundamental, which appears −72 cm−1 below the vibrational quantum in isolated (bare) H2.90 | ||
One specific reason to revisit the previously published IR spectrum15 was to reassess the broad background between 2500 and 3200 cm−1 and establish the relative intensities of the bands across the entire spectrum. The complete spectrum also includes the nominally forbidden fundamental of the H2 stretch, (observed at 4090 cm−1, labeled νH2) which is red-shifted by −72 cm−1 when attached to the molecular anion. The IR2015 (grey dashed) and IR2025 (black solid) spectra are reported in Fig. 3A. The plateau-like shape between 2550 to 2820 cm−1 in the IR2015 spectrum features a more gradual increase in IR2025 but overall, the two spectra are remarkably similar, see Fig. S3. The spectrum from the present work, however, reveals a better defined shoulder near 1660 cm−1 on the red side of the strong band at 1696 cm−1 than was evident in the IR2015 spectrum (Fig. S3). All other features agree between the two spectra. One reason for the differences between IR2015 and IR2025 concerned normalization with respect to laser power which was much improved in IR2025.
![]() | ||
| Fig. 3 Infrared spectra of OxH. Panel A: Measured (H2-tagged) gas phase spectra IR201515 (dashed grey) and IR2025 from the present work (solid black). Panel B: Spectrum from second-order vibrational perturbation theory, VPT2,6 using Lorentzian Broadening with HWHM = 4 cm−1 and bands labelled from a to g for identification. Asterisks indicate fundamental vibrations, see also Table S3. Panels C to F: IR spectra at different simulation temperatures as indicated from the Fourier transform of the dipole–dipole autocorrelation function. Grey shaded areas highlight the position and width of the measured framework modes for direct comparison between computations and experiment. | ||
Second order vibrational perturbation theory6 provides a useful way to go beyond the harmonic approximation underlying normal mode analysis, also for H-bonded systems.91 The VPT2 spectrum from using the TL-PES is reported in Fig. 3B. Focusing first on the framework modes below 2000 cm−1 it is found that all features in the measured spectrum line up well with the VPT2 spectrum (bands a to g; fundamentals a, b1, c, e2, f indicated by asterisks). This includes the multiplett-structures around 1360 cm−1 (bands b1, b2, c, and d) and centered at 1750 cm−1 (bands e1, e2, f, and g). Most notably, a low-intensity signature e1 at 1666 cm−1 which was not clearly present in the IR2015 spectrum, is manifest in the VPT2 calculations and the IR2025 spectrum, see Fig. S3, and assigned to the ν5 + ν10 combination band, see Table S3. Furthermore, the low-frequency ν9 mode (band a) at 928 cm−1 in the IR2025 spectrum is at 937 cm−1 in the VPT2 calculations and a small feature at 1091 cm−1 from the measurements and at 1090 cm−1 from the VPT2 calculations (see entry with asterisk in Table S3, not labelled). The VPT2 spectrum using the MP2 ML-PES is only qualitatively correct, see Fig. S4, and is not suitable for quantitative analysis.
The high-frequency part of the measured spectrum shows a wide absorption (h1) between 2500 and 3200 cm−1 onto which a less broad signature h2 centered at ∼2900 cm−1 is superimposed. VPT2 calculations feature a considerable number of discrete peaks covering the frequency range consistent with the experiments. As the experiment does not probe a single conformation but rather a distribution of structures, it is of interest to obtain VPT2 frequencies for perturbed minimum energy structures. For this, the optimized oxalate structure was distorted by perturbing all bond lengths with a standard deviation of 0.01 Å followed by a VPT2 calculation for each structure. The Boltzmann-averaged spectrum, shown in Fig. S5, demonstrates that the computed signal is more representative of the measured frequency distribution than the VPT2 calculation for a single minimum energy structure.
Because the entire molecular structure was perturbed, some of the framework modes also shift away from the frequencies corresponding to the minimum energy structure, see Fig. S6. However, if only the transferring hydrogen atom is displaced away from its equilibrium position by 0.002 Å along the OH-bond vector, the VPT2 calculations still find the pattern in the H-transfer region to broaden and shift by several 10 cm−1 to the red or blue, respectively, whereas the framework modes respond primarily with respect to their relative intensities but the peak positions remain at the measured frequencies, see Fig. S7. It should be noted that VPT2 calculations use the harmonic approximation as its zeroth-order approximation to which anharmonic corrections are added. Hence, for strongly anharmonic modes or modes involving Fermi resonances a VPT2 treatment may be unreliable.32,92
H-transfer during the dynamics will also be reflected in the infrared spectroscopy. To include finite-temperature effects, MD simulations using the TL-PES were carried out for temperatures between 50 K and 900 K, see Fig. 3C–F. The OH-stretch is found at 3200 cm−1 and broadens considerably as temperature increases. The blue shift away from the maximum of H-stretch mode at 2940 cm−1 from experiments is well-understood and occurs because MD simulations only sample a narrow range around the bottom of the well without accessing the anharmonic parts of the PES, in particular for the OH-stretch mode.93 As temperature increases, broader signatures emerge which overlap with the peak positions from the VPT2 calculations and the measured spectrum.
For partial assignment of the measured peaks, power spectra for different internal coordinates were determined, see Fig. 4. Due to coupling between different degrees of freedom, straightforward and unambiguous assignments are not particularly meaningful. Two infrared signatures at 1700 and 1770 cm−1 are clearly assigned in the IR spectrum and involve the CA–OA, CA–OB, CB–OC, and CB–OD distances. The peak to the red, at 1666 cm−1 and part of the high-energy triplet, does not appear in the IR and power spectra, but is clearly visible in the VPT2 calculations which also report combination modes and hot bands, see Table S3. The triplet centered around 1360 cm−1 is also reproduced by the computed IR spectra and involves the CAOAH angle and all four CO-stretches.
![]() | ||
| Fig. 4 Power spectra for OxH molecule at 300 K for different internal coordinates (low-frequency range). | ||
In order to further investigate the nature of the broad peak in the 2600–3200 cm−1 region, MD trajectories run at 600 K were analyzed. For this, separate IR and power spectra were generated for trajectories that did or did not feature H-transfer, see Fig. 5 and Fig. S8, S9. The internal coordinates considered were the CA–OA–H angle and the CA–OB, CB–OC, CB–OD, CA–OA, and OA–H separations. As already mentioned, the blue shift of the OH-stretching vibration is well-understood.93,94 However, the broad [2600, 3200] cm−1 background from the measurements is rather convincingly associated with signatures in the CA–OA–H angle especially for trajectories featuring H-transfer. Interestingly, shifting the OH-power spectrum to the red by −260 cm−1 not only aligns the experimentally observed peak at 2940 cm−1 but also the faint measured feature around 2450 cm−1. For formic acid monomer,95 the shift required to align the computed with the measured infrared signature arising from the OH-stretch was −290 cm−1, which is consistent with the observation for the OH-stretch in oxalate.
![]() | ||
| Fig. 5 Power spectra for the COH-bend (green) and OH-stretch (orange) from simulations at 600 K that with (solid line, 362 trajectories) and without (dashed line, 68 trajectories) H-transfer during the dynamics. The indigo trace is the IR spectrum (300 K; no H-transfer). The blue dashed line is the shifted OH-stretch power spectrum to best overlap with the feature in the measured spectrum. For spectra at higher T that feature H-transfer, see Fig. 3. | ||
To further corroborate this assignment, simulations were carried out whereby the mass of the hydrogen atom ranged from 1mH to 2mH (i.e. mD) in increments of 0.2 mass units, see Fig. S10. As expected, the spectral response shifts to the red as the mass of the hydrogen atom increases. It is interesting to note that with increasing temperature in Fig. 3 the region above 2600 cm−1 gains intensity (note also the different y-scaling factors). Analysis of the MD trajectories shows that with increasing temperature the probability for H-transfer between the oxygen atoms OA and OD increases. Hence, the power spectra for different internal coordinates were separately analyzed for trajectories that did (dashed lines) and did not (solid) feature hydrogen transfer, see Fig. 5. In the region of interest (above 2600 cm−1) the power spectrum of the COH bending motion (green trace) features clear signals which are totally absent from simulations that do not feature H-transfer. The OH-stretch motion (orange) is also clearly visible and coupled to the COH-bend.
![]() | ||
| Fig. 6 Infrared spectra for OxD. The top trace reports the experimentally determined (H2-tagged) gas phase spectrum15 and the bottom trace (light coral) is the computed spectrum from second-order vibrational perturbation theory6 using the CCSD(T) ML-PES. Note that data within the dashed box have been scaled by a factor of 10 for the two traces. | ||
Next, the infrared spectra were obtained from MD simulations at 600 K together with power spectra for important internal degrees of freedom. Again, trajectories without and with D-transfer were analyzed separately, see Fig. 7A and B. The OD-stretch vibration likely corresponds to the peak observed at 2110 cm−1 in the experimental spectrum, whereas in the spectrum computed from MD simulations, this feature appears at 2320 cm−1. The shift of approximately −200 cm−1 for OxD is consistent with the isotope-corrected shift observed for OxH
. The maximum peak position at 2110 cm−1 in the experimental spectrum of OxD, similar to that in OxH, is associated with motions along the OA D-stretch and CAOA D-bending motions, see olive and blue traces Fig. 7B.
![]() | ||
| Fig. 7 Power Spectra and IR spectra for OxD (high-frequency region). Panel A: Without deuterium transfer, Panel B: with deuterium transfer. | ||
one obtains Ea = 3.84 kcal mol−1, and A = 3057.3 ns−1. This compares and is consistent with a barrier height of 3.35 kcal mol−1 from the CCSD(T)/aug-cc-pVTZ calculations.
For a first estimation of the tunneling splittings, sRPI theory was employed together with the single-precision PhysNet representation of the PES. Such a treatment is considered appropriate for cases in which anharmonic effects perpendicular to the instanton path are less pronounced and the barrier for H-transfer is not particularly low. All leading-order splitting calculations in the present work were carried out with up to N = 4096 beads at an effective temperature of Teff = ħ/kBτ = 25 K. Using the MP2-level PES, featuring a barrier height EB = 2.355 kcal mol−1, ΔH = 116.5 cm−1 was obtained. As for the higher-level TL-PES EB increases to 3.351 kcal mol−1, the tunneling splitting decreases to ΔH = 36.0 cm−1. The numerical convergence of the leading-order tunneling splittings as a function of N is reported in Table S5.
One limitation of sRPI for determining tunnelling splittings is that fluctuations around the instanton are treated within a harmonic approximation (i.e. the Hessians are used). As a remedy, a perturbatively corrected approach was developed which is based on third and fourth order derivatives of the PES along the instanton path.59 For numerical stability, pcRPI requires highly accurate higher-order derivatives. Within the context of machine learning-based PESs it is common to use single-precision arithmetics for training and evaluation, which was found to be insufficient for pcRPI.96 Hence, a double-precision TL-PES was obtained from a retrained MP2 PhysNet model also using double precision arithmetics throughout. The higher order instanton calculations were carried out at Teff = 25 K and with N = 1024 beads as pcRPI is considerably more memory intensive than sRPI, and the perturbative correction was found to converge significantly more rapidly than the leading-order term.40 The correction factor for H-transfer was cPC = 0.98 within pcRPI which leads to ΔpcRPIH = 35 cm−1. This is 1 cm−1 smaller than the tunneling splitting of 36.0 cm−1 determined on the single precision TL-PES. Assuming a conservative error estimate of 20% for sRPI theory applied to the computed tunneling splitting using the TL-PES brackets the prediction for H-transfer splitting to around ΔH ∈ [28, 42] cm−1. More accurate predictions will likely require further refinement of the TL-PES, with a specific focus on tunneling splittings34,40 which are very sensitive to the shape of the PES, in particular around the global minimum and the barrier to H-transfer.
Earlier, recent work using transfer-learned PESs to the CCSD(T) level of theory for malonaldehyde and tropolone showed excellent agreement with measured tunneling splittings.34,40,96 For malonaldehyde the measured97 splitting Δexpt.H = 21.6 cm−1 compares with a computed sRPI (pcRPI) ΔTLH = 25.3 (22.1) cm−1 on a PES transfer learned to the CCSD(T) level of theory and featuring a barrier height of EB = 3.89 kcal mol−1. For tropolone,98 on the other hand, Δexpt.H = 0.97 cm−1 is considerably smaller because EB = 6.62 kcal mol−1 at the CCSD(T) level of theory and pcRPI Δtheo.H = 0.94 cm−1. Hence increasing the barrier by ca. 2 kcal mol−1 reduces ΔH by approximately 1 order of magnitude. This is consistent with an increase to ΔH = 35.0 cm−1 for oxalate because the barrier to HT decreases by 0.5 kcal mol−1 (to EB = 3.35 kcal mol−1) compared with malonaldehyde.
A particular focus was on exploring the spectral signatures between 2400 and 3200 cm−1 for OxH and 1900 to 2500 cm−1 for OxD, respectively. The VPT2 calculations find a high density of states in these frequency ranges and slightly perturbing the structure away from the minimum energy configuration further increases the density of vibrational signatures. These calculations together with the finite-temperature MD simulations identify a center-frequency at ∼2940 and ∼2110 cm−1 as the main H-/D-transfer mode. However, unequivocal assignment of one feature to a single motional pattern is not meaningful because the fundamentals, combination and hot bands are heavily mixed.
Consistent with earlier work,15 it is found here that the COH angle strongly couples to the OH stretching frequencies. This is particularly apparent for trajectories that feature H-transfer. The VPT2 calculations combined with MD simulations at the highest level of theory reported in the literature provided assignments of the vibrational modes and correct previous mode descriptions. The mismatch in assignment with the one made for IR2015 is mostly related to using a comparatively low level of quantum chemical theory (B3LYP/6-311++G(d,p)). Using a reactive force field, the effective barrier height for H-transfer was also inferred from matching the measured IR2015 spectrum with that computed from the dipole-autocorrelation function from finite-temperature MD simulations. An effective barrier height of 4.2 kcal mol−1 was found to best reproduce the IR2015 in the region of the H-transfer mode. This compares with an electronic barrier height of 3.35 kcal mol−1 from the present work and 3.4 kcal mol−1 from previous CCSD(T)/aug-cc-pVTZ calculations.99 However, the electronic and effective barrier heights describe different physical quantities and thus can not be compared directly. The electronic EB represents the pure potential energy difference between reactant and the transition state whereas the effective barrier height scales the PES along a single (reaction) coordinate (here the HT-motion), depends on temperature, and accounts for zero-point effects.
Tunneling splittings for OxH have been determined with RPI theory. A rather conservative estimate for the tunneling splitting is ΔsRPIH ∈ [28, 42] cm−1 which, however, is likely to require improvements of the TL-PES for quantitative predictions of potential future experiments. Nevertheless, the predicted range offers a reference for future spectroscopic efforts and experimental confirmation would not only validate the RPI framework but also aid in the iterative improvement of the PES guided by experiments as has been recently done for small ionic complexes through morphing.28,100,101
In summary, an updated measured infrared spectrum for the oxalate anion was discussed in the context of state-of-the art VPT2 calculations and MD simulations using a transfer-learned CCSD(T)-level PES. Assignments were made based on inspecting the VPT2 vibrations and the power spectra. In the region of the framework modes considerable coupling between the different degrees of freedom was found. The agreement between theory and experiment and the fact that a specific framework mode was correctly located by the VPT2 calculations highlights the predictive power of the theoretical approaches, whereby vibrational motion can be reliably related to observed spectral features. Examples such as OxH and OxD are particularly valuable as they allow computer-based methods to approach “prediction mode”, finding the “right answer for the right reason”.
Supplementary information: Fig. S1 and S2 present the performance of PhysNet models trained at different levels of theory. Fig. S3–S11 provide supplementary analyses and spectral data that support and extend the results discussed in the main manuscript. Tables S1–S4 report a detailed comparison between the vibrational frequencies predicted by the models and those obtained from reference quantum chemical methods and experimental measurements, including peak assignments. Table S5 summarizes the calculated tunneling splittings together with their convergence. See DOI: https://doi.org/10.1039/d5cp03085d.
Footnote |
| † Present Address: Roche Pharma Research and Early Development, Pharmaceutical Sciences, Roche Innovation Center Basel, F. Hoffmann-La Roche Ltd, Basel, Switzerland. |
| This journal is © the Owner Societies 2025 |