Next Article in Journal
Hydroxyapatite Obtained via the Wet Precipitation Method and PVP/PVA Matrix as Components of Polymer-Ceramic Composites for Biomedical Applications
Next Article in Special Issue
Epithelial Sodium Channel Inhibition by Amiloride Addressed with THz Spectroscopy and Molecular Modeling
Previous Article in Journal
Extraction of Phenolic Compounds from Fresh Apple Pomace by Different Non-Conventional Techniques
Previous Article in Special Issue
Intramolecular H-Bond Dynamics of Catechol Investigated by THz High-Resolution Spectroscopy of Its Low-Frequency Modes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Large Amplitude Motions of Pyruvic Acid (CH3-CO-COOH)

by
María Luisa Senent
1,*,† and
Samira Dalbouha
2,3
1
Theoretical Chemistry and Physics Department Institute for the Structure of Matter IEM-CSIC, Serrano 121, 28006 Madrid, Spain
2
Laboratory of Applied Chemistry and Environment, Department of Chemistry, Faculty of Science, Ibn Zohr University, Agadir B.P. 8106, Morocco
3
Laboratory of Spectroscopy, Molecular Modeling, Materials, Nanomaterials, Water and Environment, Department of Chemistry, Faculty of Sciences, University Mohammed V, Avenue Ibn Battouta, Rabat B.P. 1014, Morocco
*
Author to whom correspondence should be addressed.
Associated Unit GIFMAN, CSIC-UHU.
Molecules 2021, 26(14), 4269; https://doi.org/10.3390/molecules26144269
Submission received: 30 June 2021 / Revised: 8 July 2021 / Accepted: 10 July 2021 / Published: 14 July 2021

Abstract

:
Torsional and rotational spectroscopic properties of pyruvic acid are determined using highly correlated ab initio methods and combining two different theoretical approaches: Second order perturbation theory and a variational procedure in three-dimensions. Four equilibrium geometries of pyruvic acid, Tc, Tt, Ct, and CC, outcome from a search with CCSD(T)-F12. All of them can be classified in the Cs point group. The variational calculations are performed considering the three internal rotation modes responsible for the non-rigidity as independent coordinates. More than 50 torsional energy levels (including torsional subcomponents) are localized in the 406–986 cm−1 region and represent excitations of the ν24 (skeletal torsion) and the ν23 (methyl torsion) modes. The third independent variable, the OH torsion, interacts strongly with ν23. The A1/E splitting of the ground vibrational state has been evaluated to be 0.024 cm−1 as it was expected given the high of the methyl torsional barrier (338 cm−1). A very good agreement with respect to previous experimental data concerning fundamental frequencies (νCAL − νEXP ~ 1 cm−1), and rotational parameters (B0CAL − B0EXP < 5 MHz), is obtained.

Graphical Abstract

1. Introduction

Pyruvic acid (CH3CO-COOH, PA) is a prevalent species in the Earth’s atmosphere [1,2] where it is a key intermediate of keto-acid reactions. Its presence has been reported in gas phase, aerosols, fogs, clouds, polar ice, and rainwater. Primary sources of pyruvic acid in the atmosphere are the reaction between the hydroxyl radical and hydrated methylgloxal, the oxidation of isoprene, and the ozonolysis of methyl vinyl ketone [3]. In regions with abundant vegetation, it can be emitted directly by plants. The presence of PA can alter the chemical and optical properties of atmospheric particles. It acts as a key intermediate in the oxidative channels of isoprene and secondary organic aerosols (SOA) formation.
Photo-oxidation of volatile organic compounds (VOC) yields products that are precursors of SOAs. The photodissociation of pyruvic acid displays a complex and significant chemistry in the atmosphere [1,2]. It absorbs light in the near-UV and by photodissociation, it decomposes forming CO2, additional products, and a reactive intermediate, methylhydroxycarbene, which produces acetaldehyde by isomerization. The photolysis represents the main destructor of the acid since it oxidized relatively slowly by peroxides and OxH radicals [2]. Vaida et al., who have published a long list of recent papers focused on PA [4,5,6] and references therein], have explored the multiphase photochemistry to elucidate its likely primary atmospheric removal pathways and the implications for SOA. Some of these works attend to the role of the conformers and the spectroscopic properties of the acid on its reactivity. Since the high overtone excitation of the OH-stretching mode in the gas phase leads to a unimolecular decarboxylation reaction, the dynamics of overtone-excited pyruvic acid (PA) was the object of a meticulous study using a combination of theoretical methods and measurements of the OH-stretch fundamental and overtone transitions with Fourier transform infrared spectrometry [7,8].
In addition to the earth atmosphere relevance, carboxylic acids are present in the volatile fraction of carbonaceous meteorites [9]. Examples are compounds such as acetic acid already observed in extraterrestrial atmospheres in the gas phase [10]. Recently, Kleimeier et al. [11] have designated pyruvic acid to be a detectable species in extraterrestrial sources. They have studied its formation by barrierless recombination of hydroxycarbonyl (HOCO) and acetyl (CH3CO) radicals in ices of acetaldehyde (CH3CHO) and carbon dioxide (CO2) modeling interstellar conditions driven by cosmic rays.
Motivated by a possible search of PA in gas phase sources, Kisiel et al. [12] measured the millimeter wave rotational spectrum in the region at 160–314 GHz and in supersonic expansion at 10–17.4 GHz. They provide parameters for various torsional states initiating the analysis on the basis of previous work predictions [13,14,15,16]. Dyllic-Brenzinger et al. [15] observed an anomalous behavior in the first excited vibrational state in the rotational spectrum that they attributed to an interaction between the two low-lying torsional modes. Meyer and Bauder [16] used a two-dimensional flexible model to predict the energies of the lowest vibrational states and the A/E splittings in the second exited ν24 torsional state. The barrier to internal rotation was established to be V3 = 965 ± 40 cal/mol. The stark effect measurements of Marstokk and Möllendal [14] yielded to μa = 2.27 ± 0.02 D, μb = 0.35 ± 0.02 D, and μ = 2.3 ± 0.03 D.
The vibrational spectrum was measured in the gas phase and in Ar matrix by Holleisntein et al. in 1978 [17]. More recently, the vibrational spectra of the two most stable conformers were measured in N2 and Ar matrices and analyzed by Reva et al. [18]. Ab initio calculations have been performed to search for stable geometries [19]. PA was studied as a benchmark to test theoretical studies of conformers of flexible molecules performed combining couple cluster (CC) and DFT methodologies [20].
In this paper, we attend to the pyruvic acid from the point of view of ab initio calculations with special attention to the far infrared region (FIR). The work describes structural and spectroscopic properties following the procedures used for molecules such as acetic acid [21], acetone [22] or biacetyl [23], which are properties or functional groups with PA. Ab initio calculations are achieved using explicitly correlated coupled cluster theory [24,25]. The far infrared region, relevant for rotational studies, is explored using a three-dimensional variational procedure [26]. Recently, this procedure has been improved and implemented for a proper classification of the energy levels in molecules with regions showing a high density of states [27]. In the procedure, the two low-lying torsional modes are considered to be independent coordinates. These two vibrational modes are strongly coupled as was early observed [15]. A third coordinate, the OH internal rotation, is defined as an independent coordinate since it interacts strongly with the ν24 mode.

2. Results

2.1. Electronic Structure Calculations

The geometrical parameters and energies of the pyruvic acid conformers were computed using the explicitly correlated coupled cluster theory with singles and doubles substitutions augmented by a perturbative treatment of triple excitations (CCSD(T)-F12 [24,25] as it is implemented in MOLPRO [28], using the default options and the AVTZ-F12 basis set. This basis set contains the aug-cc-pVTZ (AVTZ) [29] atomic orbitals and the default functions for the density fitting and the resolutions of the identity. To obtain accurate rotational constants, the core-valence electron correlation effects on the structures were introduced using CCSD(T) [30] and the cc-pCVTZ basis set (CVTZ) [31].
Two different theoretical models were applied for determining the spectroscopic parameters: The vibrational second order perturbation theory (VPT2) [32] implemented in Gaussian [33], and a variational procedure of reduced dimensionality implemented in our code ENEDIM [34,35,36]. If VPT2 is applied, the four conformers are treated as four different semi-rigid species, unable to interconvert, and the vibrations are described as small displacements around the equilibrium geometries. However, if the variational procedure is employed, the non-rigidity is taken into account. The four minima are treated together and their interconversion is considered implicitly.
For VPT2, the quadratic, cubic, and quartic terms of a force field were computed using the second order Möller-Plesset theory (MP2) and the AVTZ basis set [29]. For the variational calculations a three-dimensional potential energy surface was obtained at the CCSD(T)-F12/AVTZ-F12 level of theory applied on a grid of structures partially optimized with MP2/AVTZ. The surface was vibrationally corrected using MP2/AVTZ harmonic calculations.

2.2. The Four Conformers of Pyruvic Acid

Four equilibrium geometries of pyruvic acid, Tc, Tt, Ct, and CC have been identified based on a search with CCSD(T)-F12. All of them present a symmetry plane and can be classified in the Cs point group. The most stable geometry is the Tc (TRANS-cis) conformer represented in Figure 1. “TRANS and CIS” designate the relative orientation of the two C=O groups, whereas “trans and cis” denote the relative position of the C1 and H10 atoms.
Table 1 summarizes the properties characterizing the four conformers: The CCSD(T)-F12/AVTZ-F12 relative energies E, and the corresponding vibrationally corrected relative energies, EZPVE; the CCSD(T)-F12/AVTZ-F12 equilibrium rotational constants, the MP2/AVTZ dipole moment components along the principal axis, the high of the barriers restricting the minimum interconversion, and the three coordinates, θ1, θ2, and α, that identify the structures. They correspond to the methyl group torsion, the torsion of the C3-C1 bond, and the hydroxyl group torsion, respectively. In the variational procedure described in the next sections, θ1, θ2, and α represent the independent coordinates. They are defined as the following linear combination of internal coordinates:
θ1 = (H5C2C1C3 + H6C2C1C3 + H7C2C1C3)/3
θ2 = O8C2C1C3 + O9C2C1C3 − 180·
α = H10O9C3O8
In Table 1, the computed dipole moment components are compared with previous experimental data derived from stark effect measurements [14].
Figure 2 represents the relative stabilities of the four conformers. The CCSD(T)-F12/AVTZ-F12 structural parameters and MP2/AVTZ anharmonic vibrational frequencies are collected in Table 2 and Table 3, respectively.
The Tc form stabilizes by the formation of an intramolecular hydrogen bond (H10….O4 = 2.0030 Å). The next two conformers Tt and Ct lie in the 950–2000 cm−1 region. In previous papers [18], the higher energy conformer Cc is ignored since it is very unlikely to be observed. Although, it is described as a transition state by several levels of theory, all the MP2/AVTZ harmonic wavenumbers are real. The low stability of Cc is due to the non-bonding repulsions between H10 and its neighboring methyl group hydrogens.
The methyl torsional barrier of Cc computed using CCSD(T)-F12, is strongly affected by the repulsive interactions. We calculated it to be 852.8 cm−1, whereas the corresponding values for Tc, Tt, and Ct are 338.0, 397.8, and 540.2 cm−1, respectively (see Figure 3). The computed barrier corresponding to Tc is a very good agreement with the experimental value of 336.358(50) cm−1, obtained by Kisiel et al. [12] in the analysis of the A and E lines of the millimeter spectrum.
The torsional barriers VOH and VCC, shown in Table 1 and Figure 4 and Figure 5, restrict the t → c and T → C interconversions. They were computed from the vibrationally corrected three-dimensional potential energy surface described below. The one-dimensional cuts were obtained by fixing two coordinates at their values in the low-energy conformer involved in the pathways. With the exception of VCC located between Tt and Ct, all the barriers are higher than 3000 cm−1. The pathways emphasize the instability of the Cc that easily transforms in Ct or Tc.
Whereas the Tt → Ct process can occur at very low temperatures, the Tc → Tt pathway is impeded by a relatively high barrier. It may be inferred that the very low vibrational energy levels of the most stable conformer, Tc, can be studied considering it as a molecule with a unique minimum. However, since the conformers Tt and Ct are very close in energy (ΔE = 538 cm−1) and the barriers restricting the processes Tt → Ct and Ct → Tt are relatively low (812 and 274 cm−1), an accurate calculation of the low vibrational energy levels requires taking into consideration the tunneling effects in those barriers.
In Table 3, the Tc anharmonic wavenumbers are compared with the observed bands assigned by Hollenstein et al. [17], who measured the IR spectrum in the 4000–200 cm−1 region, and with the mid-IR vapor-phase spectrum of Plath et al. [8]. Emphasized in bold are the transitions in which important Fermi displacements are predicted by the VPT2 theory. The OH stretching fundamental denotes the presence of the intramolecular hydrogen bond in the Tc conformer. The frequency, 3426 cm−1, observed at 3463 cm−1 [8], is lower than the one of Tt (calculated to be 3561 cm−1 and observed at 3579 cm−1 [8]) and the one of Ct (3554 cm−1). In Cc, the frequency is higher than in the remaining structures (3636 cm−1) due to the proximity of the methyl group hydrogens. Two internal coordinates, the H10O9C3O8 dihedral angle and the out-of-plane C=O bending contribute to the ν21 mode, computed to be 606 cm−1 and observed by Hollenstein et al. [17] at 668 cm−1. In this paper, the H10O9C3O8 dihedral angle is identified with the OH torsional coordinate α. Direct measurements of the CH3 torsional fundamental are not available.

2.3. Ground Vibrational State: Rotational and Centrifugal Distortion Constants

The ground vibrational state rotational constants and the centrifugal distortion constants of the four conformers are shown in Table 4. The collected distortion constants are parameters of the asymmetrically reduced Hamiltonian in the I r representation [37]. They were determined using two different basis sets, AVTZ and the corresponding triple zeta basis set ignoring the diffuse functions, VTZ. For the low-lying conformer, Tc, the MP2/VTZ parameters were obtained in a very good agreement with those fitted by Kisiel et al. [12] during the assignment of the millimeter-wave spectrum using the ERHAM program [38]. Their analysis included more than 1500 lines comprising A and E internal sublevels. Given the good agreement between ab initio and experimental results for Tc, it can be expected that the MP2/VTZ predicted sets will be useful for future assignments of the remaining conformers.
The ground vibrational state rotational constants were computed using the CCSD(T)-F12 equilibrium parameters of Table 2 and the following equation, proposed and verified in previous studies [22,39,40,41].
Bi0 = Bie (CCSD(T)-F12/AVTZ-F12) + ∆Bicore (CCSD(T)/CVTZ)
+ ∆Bivib (MP2/AVTZ); i = a;.b;.c
Here, ∆Biecore takes into account the core-valence-electron correlation effect on the equilibrium parameters. It can be evaluated as the difference between Bie(CV) (calculated correlating both core and valence electrons) and Bie(V) (calculated correlating just the valence electrons). For Tc, the correction increases 15.8, 11.8, and 6.9 MHz the equilibrium values of A, B, and C, respectively. ∆Bivib represents the vibrational contribution to the rotational constants derived from the VPT2 αir vibration-rotation interaction parameters.
The rotational constants of the preferred conformer Tc are in a very good agreement with the parameters fitted by Kisiel et al. [12] (A0CAL − A0EXP = −4.96 MHz, B0CAL − B0EXP = −2.23 MHz, and C0CAL − C0EXP = −1.78 MHz). Generally, for many molecules, if Equation (2) is applied, the computed B0 and C0 are more accurate than A0 [41]. Based on the very good agreement between our calculated rotational constants for Tc conformer and their experimentally determined counterparts, such parameters for the other isomers are also plausibly very accurate.

2.4. The Far Infrared Region

The ground electronic state potential energy surface of pyruvic acid presents a total of 12 minima corresponding to four conformers since each methyl group inter-transforms into three equivalent minima. Then, our exploration of the far infrared region uses a three-dimensional variational procedure that takes into account the interconversion of the minima. The three torsional coordinates of Equation (1) that are responsible for the non-rigidity are defined to be the independent variables. The validity of the separability of the three internal rotations θ1, θ2 and α from the remaining vibrations entails a previous discussion. Although VPT2 does not represent the right theory for internal rotation studies, it provides a valuable preliminary description since the three internal rotation coordinates represent the most relevant contributions to three normal modes. The anharmonic analysis of Table 3 can supply arguments based on energies and the cubic force field. In the most stable conformer, pyruvic acid displays five out-of-plane fundamental transitions lying below 1000 cm−1. The central bond torsion and the methyl group torsion lie at very low frequencies and interact. Both motions contribute to two normal modes, ν23 and ν2424 (VPT2) = 93 cm−1; ν23 (VPT2) = 124 cm−1). In principle, it can be assumed that at least, three excited CH3 and four excited C-C torsional states could be computed using a two-dimensional procedure since the VPT2 resulting energies are lower than the next lowest frequency mode. However, above 350 cm−1, three out-of-plane fundamentals, ν22 (VPT2) = 392 cm−1, ν21 (VPT2) = 606 cm−1, and ν20 (VPT2) = 792 cm−1, have been found. Although these modes cannot really be interpreted in terms of local modes, they can be interpreted as two skeletal bending modes and to the OH torsional fundamental. It cannot be strictly stated that ν21 represents the OH torsional fundamental, since the out-of-plane bending of the neighboring C=O group has an important contribution to ν21 [17]. In addition, the OH torsional mode is strongly coupled to the C-C torsional mode, ν24. In the most stable Tc conformer, the OH hydrogen forms part of an intramolecular hydrogen bond that vanishes with the C-C torsional mode excitations. To obtain accurate low energy levels for the two low-lying modes, ν24 and ν23, the internal coordinate describing the OH torsion must be considered explicitly as an independent variable.
If the 3D model is applied, very accurate CH3 and C-C torsional levels and very unrealistic OH torsional levels will be expected. In addition, the VPT2 theory predicts a displacement of the 2ν23 overtone (∆(2ν23) ~ −5 cm−1) due to Fermi resonances with the ν16 fundamental which is not defined as an independent variable. The remaining low-lying levels seem to be free of relevant resonances.
The 3D Hamiltonian for J = 0 must be defined as [34,35,36]:
H ( θ 1 , θ 2 , α ) = i = 1 3 j = 1 3 ( q i ) B q i q j ( θ 1 , θ 2 , α ) ( q j )   + V e f f ( θ 1 , θ 2 , α ) q i = θ 1 ,   θ 2 , α   ;   q j = θ 1 ,   θ 2 , α
where Bqiqj1, θ2, α) are the kinetic energy parameters [35,36]; the effective potential is the sum of three terms:
Veff1, θ2, α) = V(θ1, θ2, α) + V’(θ1, θ2, α) + VZPVE1, θ2, α)
V(θ1, θ2, α), V’(θ1, θ2, α), and VZPVE1, θ2, α) represent the ab initio potential energy surface, the Podolsky pseudopotential, and the zero point vibrational energy correction, respectively.A grid of 332 geometries was selected to determine the Hamiltonian parameters. In all these geometries, the remaining 3Na-9 internal coordinates (Na = number of atoms) were optimized at the MP2/AVTZ level of theory, whereas three internal coordinates were fixed at the following values:
H5C2C1C3 = 0, 90, 180, and −90°
O8C2C1C3 = 0, 30, 60, …, 150, 180°
H10O9C3O8 = −180, −150,…, 0…, 150, 180°
Single point CCSD(T)-F12/AVTZ-F12 calculations were performed on the 332 optimized geometries. The total electronic energies were fitted to a symmetry adapted triple Fourier series to obtain the V(θ1, θ2, α) ab initio potential energy surface (R2 = 0.99999; σ = 0.013 cm−1). The following analytical expression was employed:
V   ( θ 1 , θ 2 ,   α ) = m = 0 n = 0 l = 0       ( A n m l c c c cos 3 m cos n θ 2 cos l α   + A n m l s s c sin 3 m θ 1 sin n θ 2 cos l α + A n m l c s s cos 3 m θ 1 sin n θ 2 sin l α + A n m l s c s sin 3 m θ 1 cos n θ 2 sin l α )
Formally identical expressions were employed for V’(θ1, θ2, α) and VZPVE1, θ2, α). The pseudopotential represents a negligible correction. However, the zero point vibrational correction VZPVE1, θ2, α) is mandatory. It was computed from the MP2/AVTZ harmonic fundamentals using all the geometries. The expansion coefficients of the final effective potential are supplied as supplementary material in Table S1. Two dimensional cuts of the potential energy surface are represented in Figure 6.
A formally identical formula was also used for the kinetic energy parameters which were computed for the 332 partially optimized geometries. Details on the procedure for obtaining the Hamiltonian parameters using ENEDIM [34,35,36]. The A 000 c c c coefficients of the kinetic parameters are given in Table 5.
The symmetry adapted triple Fourier series were employed as trial functions. To reduce dimensionality, a contracted basis set was employed for the C-C torsional coordinate performing a pre-diagonalization of the Hamiltonian matrix [27,42]. This route helps the classification of the resulting energies.
The minimum number of functions required for convergence produces a matrix of 169,455 × 169,455 elements. The symmetry factorizes it into four blocks with the dimensions as 28,243 (A1), 28,242 (A2), and 56,485 × 2 (E). Using the contracted basis set, the dimensions are reduced to 4648 (A1), 4647 (A2), and 9295 × 2 (E). This represents the 16% of the original matrix.
Each energy level splits into three subcomponents corresponding to a non-degenerate representation (A1 and A2) and to the double-degenerate representation, E. Table 6 collects torsional energies up to ~425 cm−1 over the ground vibrational state (ZPVE = 406.072 cm−1). They are compared with previous data from Ref. [16] and with computations performed in this work using VPT2. All the collected levels were assigned to the Tc conformers since the first secondary minimum lies at 877.5 cm−1. The energy levels of Ref. [16] were obtained from experimental splittings using a flexible two-dimensional model.
More than 50 torsional energy levels (including subcomponents) were localized in the 406–986 cm−1 region below the first excited state (0 0 1) of the OH torsion ((0 0 1) = 581.306 cm−1 (A2) and 581.327 cm−1 (E)). All of the first 50 energies represent excitations of the ν24 and ν23 modes. Since the classification of the resulting levels is really difficult given the large density of states in the same region, procedures developed for previous studies were employed for a right assignment. The properties of the 3D-wavefunctions and the contribution of the contracted basis functions [27,42] allow assigning the main part of the energies. The assignments of few E components (in blue in Table 6) are not conclusive.
The A1/E splitting of the ground vibrational state has been evaluated to be 0.024 cm−1 as it was expected given the high of the methyl torsional barrier (338.0 cm−1). This value is in a very good agreement with the results of Refs. [12,16] (0.7203 GHz [16], 0.72355(43) GHz) [12]. It can be emphasized that a surprising very good agreement between our calculations and the results of these previous references, was found for many vibrational and rotational properties [12,16].
The ν24 fundamental (010 000) computed to be 93 cm−1 with VPT2, presents two components at 89.754 cm−1 (A2 A1) and 89.709 cm−1 (E E) in a very good agreement with Refs. [16,17]. The methyl torsional fundamental ν23 (100 000) computed to be 123 cm−1 with VPT2, presents two components at 120.539 cm−1 (A2 A1) and 119.843 cm−1 (E E). Previous estimations placed it in the 118.2–117.7 cm−1 zone.
With respect to the results of Ref. [16], there is a disagreement for the relative order of the levels (1 1 0) and (2 0 0). The VPT2 theory predicts Fermi displacements of the 2ν23 overtone.

3. Discussion

In general, the computed spectroscopic properties corresponding to the most stable conformer of pyruvic acid (rotational and torsional parameters) are in a surprising good agreement with previous experimental data. This behavior was not expected since molecules with low methyl torsional barriers can present sudden problems causing divergences between “ab initio” results and observations. On the basis of this agreement, we predict parameters for all the conformers to be employed in other spectrum assignments.
Four equilibrium geometries of pyruvic acid outcome from the search with CCSD(T)-F12. The preferred geometry Tc stabilizes by the formation of an intramolecular hydrogen bond (H10….O4 = 2.0030 Å) that a play a role in the internal dynamics of the molecule. Two next conformers, Tt and Ct, lie between 950 and 2000 cm−1 region over Tc. The higher energy conformer Cc is very unlikely to be observed. With few exceptions, all the VOH and VCC barriers restricting the conformers interconversions are higher than 3000 cm−1. The interconversion pathways emphasize the instability of the Cc that easily transforms to Ct or Tc.
More than 50 torsional energy levels (including subcomponents) were localized in the 406–986 cm−1 region below the first excited state (0 0 1) of the OH torsional mode. They represent excitations of the ν24 and ν23 modes and all of them are assigned to the most stable conformer. The energies can be calculated with accuracy using a three-dimensional model although the VPT2 theory predicts Fermi displacements of the 2ν23 overtone caused by resonances that are not considered.
In general, there is an excellent agreement for the fundamental transitions and for the splittings between the new calculations and the previous experimental or semiempirical data. However, there is not an agreement for the relative order of the levels (1 1 0) and (2 0 0). We propose a new assignment for these two energies to be taken into consideration in future spectrum assignments.

4. Materials and Methods

The Materials and Methods are detailed in the section dedicated to the “Electronic structure calculations”.

Supplementary Materials

The following are available online at Table S1: Expansion coefficients of the potential energy surface (in cm−1).

Author Contributions

Conceptualization, M.L.S. and S.D.; methodology, M.L.S.; software, M.L.S.; validation, M.L.S. and S.D.; formal analysis, M.L.S.; investigation, M.L.S. and S.D.; resources, M.L.S.; data curation, S.D.; writing—original draft preparation, M.L.S.; writing—review and editing, M.L.S.; visualization, S.D.; supervision, M.L.S.; project administration, M.L.S.; funding acquisition, M.L.S. Both authors have read and agreed to the published version of the manuscript.

Funding

This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No. 872081. This research was supported by the Ministerio de Ciencia, Inovación y Universidades of Spain through the grants EIN2019-103072 and FIS2016-76418-P. This work has received funding from the CSIC i-coop+2018 program under the reference number COOPB20364. The author acknowledges the CTI (CSIC) and CESGA and to the “Red Española de Computación” for the grants AECT-2020-2-0008 and RES-AECT-2020-3-0011 for computing facilities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

The parameters of the torsional Hamiltonian are available from the authors.

References

  1. Yu, S. Role of organic acids (formic, acetic, pyruvic and oxalic) in the formation of cloud condensation nuclei (CCN): A review. Atmos. Res. 2000, 53, 185–217. [Google Scholar] [CrossRef]
  2. Mellouki, A.; Wallington, T.J.; Chen, J. Atmospheric chemistry of oxygenated volatile organic compounds: Impacts on air quality and climate. Chem. Rev. 2015, 115, 3984–4014. [Google Scholar] [CrossRef]
  3. Nozière, B.; Kalberer, M.; Claeys, M. The molecular identification of organic compounds in the atmosphere: State of the art and challenges. Chem. Rev. 2015, 115, 3919–3983. [Google Scholar] [CrossRef]
  4. Reed Harris, A.E.; Pajunoja, A.; Cazaunau, M.; Gratien, A.; Pangui, E.; AMonod, A.; Griffith, E.C.; Virtanen, A.; Doussin, J.-F.; Vaida, V. Multiphase photochemistry of pyruvic acid under atmospheric conditions. J. Phys. Chem. A 2017, 121, 3327–3339. [Google Scholar] [CrossRef] [PubMed]
  5. Church, J.R.; Vaida, V.; Skodje, R.T. Gas-phase reaction kinetics of pyruvic acid with OH radicals: The role of tunneling, complex formation, and conformational structure. J. Phys. Chem. A 2020, 124, 790–800. [Google Scholar] [CrossRef] [PubMed]
  6. Church, J.R.; Vaida, V.; Skodje, R.T. Kinetic study of gas-phase reactions of pyruvic acid with HO2. J. Phys. Chem. A 2021, 125, 2232–2242. [Google Scholar] [CrossRef] [PubMed]
  7. Takahashi, K.; Plath, K.L.; Skodje, R.T.; Vaida, V. Dynamics of vibrational overtone excited pyruvic acid in the gas phase: Line broadening through hydrogen-atom chattering. J. Phys. Chem. A 2008, 112, 7321–7331. [Google Scholar] [CrossRef]
  8. Plath, K.L.; Takahashi, K.; Skodje, R.T.; Vaida, V. Fundamental and overtone vibrational spectra of gas-phase pyruvic acid. J. Phys. Chem. A 2009, 113, 7294–7303. [Google Scholar] [CrossRef]
  9. Sephton, M.A. Organic compounds in carboniceous meteorites. Nat. Prod. Rep. 2002, 19, 292–311. [Google Scholar] [CrossRef]
  10. Mehringer, D.M.; Snyder, L.E.; Miao, Y.; Lovas, F. Detection and confirmation of interstellar acetic acid. Astrophys. J. 1997, 480, L71–L74. [Google Scholar] [CrossRef] [Green Version]
  11. Kleimeier, N.F.; Eckhardr, A.K.; Schreiner, P.R.; Kaiser, R.I. Interstellar formation of biorelevant pyruvic acid (CH3COCOOH). Chem 2020, 6, 3385–3395. [Google Scholar] [CrossRef]
  12. Kisiel, Z.; Pszczołkowski, L.; Białkowska-Jaworska, E.; Charnley, S.B. The millimeter wave rotational spectrum of pyruvic acid. J. Mol. Spectrosc. 2007, 241, 220–229. [Google Scholar] [CrossRef]
  13. Kaluza, C.E.; Bauder, A.; Günthard, H.H. The microwave spectrum of pyruvic acid. Chem. Phys. Lett. 1973, 22, 454–457. [Google Scholar] [CrossRef]
  14. Marstokk, K.-M.; Möllendal, H. Microwave spectrum, conformation, barrier to internal rotation and dipole moment of pyruvic acid. J. Mol. Struct. 1974, 20, 257–267. [Google Scholar] [CrossRef]
  15. Dyllick-Brenzinger, C.E.; Bauder, A.; Günthard, H.s.H. The substitution structure, barrier to internal rotation, and low frequency vibrations of pyruvic acid. Chem. Phys. 1977, 23, 195–206. [Google Scholar] [CrossRef]
  16. Meyer, R.; Bauder, A. Torsional coupling in pyruvic acid. J. Mol. Spectrosc. 1982, 94, 136–149. [Google Scholar] [CrossRef]
  17. Hollenstein, H.; Akermann, F.; Gunthard, H.H. Vibrational analysis of pyruvic acid and D-, 13C- and 18O-labelled species: Matrix spectra, assignments, valence force field and normal coordinate analysis. Spectrochim. Acta A 1978, 34, 1041–1063. [Google Scholar] [CrossRef]
  18. Reva, I.; Nunes, C.M.; Biczysko, M.; Fausto, R. Conformational switching in pyruvic acid isolated in Ar and N2 matrixes: Spectroscopic analysis, anharmonic simulation, and tunneling. J. Phys. Chem. A 2015, 119, 2614–2627. [Google Scholar] [CrossRef]
  19. Tarakeshwar, P.; Manogaran, S. An ab initio study of pyruvic acid. J. Mol. Struct. 1998, 430, 51–56. [Google Scholar] [CrossRef]
  20. Barone, V.; Biczysko, M.; Bloino, J. CC/DFT Route toward accurate structures and spectroscopic features for observed and elusive conformers of flexible molecules: Pyruvic acid as a case study. J. Chem. Theory Comput. 2015, 11, 4342–4363. [Google Scholar] [CrossRef] [Green Version]
  21. Senent, M.L. Ab initio determination of the torsional spectra of acetic acid. Mol. Phys. 2001, 99, 1311–1321. [Google Scholar] [CrossRef]
  22. Dalbouha, S.; Mogren Al-Mogren, M.; MLSenent, M.L. Rotational and torsional properties of various monosubstituted isotopologues of acetone (CH3-CO-CH3) from explicitly correlated ab initio methods. ACS Earth Space Chem. 2021, 5, 890–899. [Google Scholar] [CrossRef]
  23. Senent, M.L.; Moule, D.C.; Smeyers, Y.G.; Toro-Labbé, T.; Peñalver, F.J. A theoretical spectroscopic study of the Ã1Au(S1) ← X1Ag(S0), n → π* Transition in Biacetyl, (CH3CO)2. J. Mol. Spectrosc. 1994, 164, 66–78. [Google Scholar] [CrossRef] [Green Version]
  24. Knizia, G.; Adler, T.B.; Werner, H.-J. Simplified CCSD(T)-F12 methods: Theory and benchmarks. J. Chem. Phys. 2009, 130, 054104. [Google Scholar] [CrossRef] [PubMed]
  25. Werner, H.-J.; Adler, T.B.; Manby, F.R. General orbital invariant MP2-F12 theory. J. Chem. Phys. 2007, 126, 164102. [Google Scholar] [CrossRef] [PubMed]
  26. Senent, M.L.; Ruiz, R.; Dominguez-Gómez, R.; Villa, M. CCSD(T) study of the far-infrared spectrum of ethyl methyl ether. J. Chem. Phys. 2009, 130, 064101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Boussesi, R.; Senent, M.L. Computational analysis of the far Infrared spectral region of various deuterated varieties of Ethylene Glycol. Phys. Chem. Chem. Phys. 2020, 22, 23785–23794. [Google Scholar] [CrossRef]
  28. Werner, H.-J.; Knowles, P.J.; Manby, F.R.; Schütz, M.; Celani, P.; Knizia, G.; Korona, T.; Lindh, R.; Mitrushenkov, A.; Rauhut, G.; et al. MOLPRO, version 2012.1; a Package of ab Initio Programs; John Wiley & Sons, Ltd: Hoboken, NJ, USA, 2012; Available online: http://www.molpro.net (accessed on 4 June 2021).
  29. Kendall, R.A.; Dunning, T.H., Jr.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef] [Green Version]
  30. Knowles, P.J.; Hampel, C.; Werner, H.-J. Coupled cluster theory for high spin, open shell reference wave functions. J. Chem. Phys. 1993, 99, 5219–5227. [Google Scholar] [CrossRef]
  31. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. V. Core-valence basis sets for boron through neon. J. Chem. Phys. 1995, 103, 4572–4585. [Google Scholar] [CrossRef] [Green Version]
  32. Barone, V. Anharmonic vibrational properties by a fully automated second-order perturbative approach. J. Chem. Phys. 2005, 122, 014108. [Google Scholar] [CrossRef]
  33. Frisch, M.J.; Trucks, G.W.; Schlegel; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision, C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  34. Senent, M.L. ENEDIM, “A Variational Code for Non-Rigid Molecules”. 2001. Available online: http://tct1.iem.csic.es/PROGRAMAS.htm (accessed on 7 June 2021).
  35. Senent, M.L. Determination of the kinetic energy parameters of non-rigid molecules. Chem. Phys. Lett. 1998, 296, 299–306. [Google Scholar] [CrossRef]
  36. Senent, M.L. Ab initio determination of the roto-torsional energy levels of trans-1,3-butadiene. J. Mol. Spectrosc. 1998, 191, 265–275. [Google Scholar] [CrossRef]
  37. Watson, J.K.G. Determination of Centrifugal Distortion Coefficients of Asymmetric-Top Molecules. III. Sextic Coefficients. J. Chem. Phys. 1968, 48, 4517–4524. [Google Scholar] [CrossRef]
  38. Groner, P. Effective rotational Hamiltonian for molecules with two periodic large-amplitude motions. J. Chem. Phys. 1997, 107, 4483–4498. [Google Scholar] [CrossRef]
  39. Boussesi, R.; Senent, M.L.; Jaïdane, N. Weak intramolecular interaction effects on the low temperature spectra of ethylene glycol, an astrophysical species. J. Chem. Phys. 2016, 144, 164110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Motiyenko, R.A.; Margulès, L.; Senent, M.L.; Guillemin, J.C. Internal rotation of OH group in 4-hydroxy-2-butynenitrile studied by millimeter-wave spectroscopy. J. Phys. Chem. A 2018, 122, 3163–3169. [Google Scholar] [CrossRef]
  41. Dalbouha, S.; Senent, M.L.; Komiha, N.; Domínguez-Gómez, R. Structural and spectroscopic characterization of methyl isocyanate methyl cyanate, methyl fulminate, and acetonitrile N-oxide using highly correlated ab initio methods. R. J. Chem. Phys. 2016, 145, 124309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Gámez, V.; Senent, M.L. The formation of C3O3H6 structural isomers in the gas phase through barrierless pathways. Formation and spectroscopic characterization of methoxy acetic acid. Astrophys. J. 2021, 913, 21–37. [Google Scholar] [CrossRef]
Figure 1. The preferred geometry of pyruvic acid.
Figure 1. The preferred geometry of pyruvic acid.
Molecules 26 04269 g001
Figure 2. The four pyruvic acid conformers. The relative energies are given in Table 1.
Figure 2. The four pyruvic acid conformers. The relative energies are given in Table 1.
Molecules 26 04269 g002
Figure 3. CH3 torsional barrier.
Figure 3. CH3 torsional barrier.
Molecules 26 04269 g003
Figure 4. The CCSD(T)-F12 VOH torsional barriers. The Tc Tt and Ct Cc pathways were computed using the θ1 and θ2 values of the Tc and Ct conformers, respectively.
Figure 4. The CCSD(T)-F12 VOH torsional barriers. The Tc Tt and Ct Cc pathways were computed using the θ1 and θ2 values of the Tc and Ct conformers, respectively.
Molecules 26 04269 g004
Figure 5. The CCSD(T)-F12 VCC torsional barriers. The Tc Cc and Tt → Ct pathways were computed using the θ1 and α values of the Tc and Tt conformers, respectively.
Figure 5. The CCSD(T)-F12 VCC torsional barriers. The Tc Cc and Tt → Ct pathways were computed using the θ1 and α values of the Tc and Tt conformers, respectively.
Molecules 26 04269 g005
Figure 6. Two-dimensional cuts of the 3D potential energy surface of pyruvic acid.
Figure 6. Two-dimensional cuts of the 3D potential energy surface of pyruvic acid.
Molecules 26 04269 g006
Table 1. CCSD(T)-F12/AVTZ-F12 relative energies (E, EZPVE, in cm−1), internal rotation barriers (V3, VOH, VCC in cm−1), and equilibrium rotational constants (in MHz); MP2/AVTZ dipole moment (in D) of pyruvic acid.
Table 1. CCSD(T)-F12/AVTZ-F12 relative energies (E, EZPVE, in cm−1), internal rotation barriers (V3, VOH, VCC in cm−1), and equilibrium rotational constants (in MHz); MP2/AVTZ dipole moment (in D) of pyruvic acid.
TcTtCtCc
Calc.Exp. [Ref]Calc.Calc.Calc.
E0.0 a 960.31526.03742.6
EZPVE0.0 b 877.51416.33563.5
θ1180° 180°180°180°
θ2 180°180°
α180° 180°
Ae5537.641 5640.9465568.8435513.683
Be3606.390 3497.7673503.1283483.147
Ce2213.503 2187.8752179.0972163.620
μa2.55752.27 ± 0.02 [14]0.38090.20212.9383
μb0.1310.35 ± 0.02 [14]1.3534.43135.2582
μc0.0 0.00.00.0
μ2.56092.3 ± 0.03 [14]1.40564.43596.0235
V3338.0336.358(50) [12]397.8540.2852.8
VOH (Tc → Tt) 4852
VOH (Ct → Cc) 4383
VCC (Tc → Cc) 3871
VCC (Tt → Ct) 812
E = −342.03025 a.u.; ZPVE = 15601.9 cm−1.
Table 2. CCSD(T)-F12/AVTZ-F12 equilibrium structural parameters (distances, in Å, angles, in degrees) of pyruvic acid conformers.
Table 2. CCSD(T)-F12/AVTZ-F12 equilibrium structural parameters (distances, in Å, angles, in degrees) of pyruvic acid conformers.
TcTtCtCc
H10….O42.0030---
C1C21.49711.50081.50311.5149
C1C31.54331.54081.54761.5561
O4C11.21681.20631.20481.2031
H5C21.08701.08631.08661.0865
H6C21.09141.09081.09091.0935
H7C21.09141.09081.09091.0935
O8C31.20691.20541.19721.1922
O9C31.33631.33531.35021.3536
H10O90.97340.96700.96710.9616
C2C1C3116.8114.7117.8118.8
O4C1C3117.8120.3117.8118.2
H5C2C1109.9109.2109.0108.9
H6C2C1109.3109.7109.9110.6
H7C2C1109.3109.7109.9110.6
O8C3C1122.9122.6124.2122.3
O9C3O8124.6124.8124.4121.6
H10O9C3105.3106.5106.6110.8
H6C2C1H5−121.9−121.5−121.3−119.9
H7C2C1H5121.9121,5121.3119.9
Table 3. MP2/AVTZ anharmonic fundamentals a (in cm−1).
Table 3. MP2/AVTZ anharmonic fundamentals a (in cm−1).
TcTtCtCc
Calc.Exp. refCalc.Exp. refCalc.Calc.
ν(a’)
ν1OH st34263463 b35613579 b35543636
ν2CH3 st30723025 b3072 30673059
ν3CH3 st29712941 b2973 29732958
ν4C3=0 st17821804 b1747 17731788
ν5C1=0 st17051737 b1723 17211721
ν6CH3 b14351424 b1436 14361447
ν7C-C- st13801391 b1382 13631355
ν8CH3 b13321360 b1356 13401273
ν9COH b12141211 b1203 11621177
ν10C-OH st11351133 b1120 11021098
ν11CH3 b971~970 c962 971965
ν12C-C st756761 c731 724738
ν13skeletal b602604 c587 605606
ν14skeletal b529 517 484478
ν15OCCb387 386 401405
ν16CCCb258 245 253263
ν(a”)
ν17CH3 st3019 3024 30263005
ν18CH3 b1432 1437 14411458
ν19CH3 b10171030 b1019 10201011
ν20skeletal b792 723 713699
ν21OH tor (+C=O)606668 c601 601461
ν22skeletal b391392 c373 373357
ν23CH3 tor123 130 151187
ν24C-C tor9390 c39 191
a Emphasized in bold transitions in which important Fermi displacements are predicted. b Ref. [8]: Infrared absorption spectroscopy. c Ref. [17]: Infrared spectroscopy in an argon matrix.
Table 4. Ground vibrational state rotational constants (in MHz) and centrifugal distortion constants corresponding to the asymmetrically reduced Hamiltonian parameters (Ir representation) a.
Table 4. Ground vibrational state rotational constants (in MHz) and centrifugal distortion constants corresponding to the asymmetrically reduced Hamiltonian parameters (Ir representation) a.
TcTtCtCc
in MHz
Calc.Exp. [12]Calc.Calc.Calc.
A05530.5065535.46113 (18)5629.8415553.7525499.149
B03581.1773583.408634 (78)3473.3633474.9823449.328
C02203.0772204.858443 (63)2180.4232173.7172158.751
in kHz
ΔJ0.6454650.675114 (35)0.5779810.5837711.831225
ΔK1.2595011.49373 (49)1.2756841.3406953.880986
ΔJK−0.550295−0.77911 (12)−0.392555−0.469353−4.231851
δJ0.2522450.264241 (16)0.2211770.2252230.480955
δK0.5826140.552491 (97)0.6011050.5761545.552698
in Hz
HJ0.0001110.0001239 (86)0.0001160.000118−0.320165
HK0.0034860.00561 (48)0.0044400.010283−6.821416
HJK0.000466-0.0005230.001860−2.315587
HKJ−0.003375−0.00171 (13)−0.004398−0.0116239.438257
φJ0.0000910.0000718 (46)0.0000940.000098−0.252688
φJK0.002459 0.0039830.011783−7.667262
φK0.0007550.002441 (76)0.0007720.001065−0.425569
a The rotational constants were computed using Equation (2) and the centrifugal distortion constants using a MP2/VTZ force field.
Table 5. A 000 c c c coefficients of the kinetic energy parameters (in cm−1).
Table 5. A 000 c c c coefficients of the kinetic energy parameters (in cm−1).
A 000 c c c A 000 c c c
Bθ1θ15.4904Bθ1θ2−0.1751
Bθ2θ20.7411Bθ1α0.0376
Bαα18.5985Bθ2α0.0061
Table 6. Low-lying torsional energy levels (in cm−1) of pyruvic acid computed variationally and/or with VPT2 a. The m, n, and l quanta corresponds to the θ1, θ2, and α coordinates.
Table 6. Low-lying torsional energy levels (in cm−1) of pyruvic acid computed variationally and/or with VPT2 a. The m, n, and l quanta corresponds to the θ1, θ2, and α coordinates.
m n lVariationalVPT2Ref. [16]m n lVariationalVPT2
0 0 0
A1
E
0.000 b
0.024
-0.000
0.024
3 0 0
(3ν23)
A2
E
327.476
278.686
340
0 1 0
(ν24)
A2
E
89.754
89.709
9390.1
90.0
4 0 0
(4ν23)
A1
E
334.735
394.098
432
1 0 0
23)
A2
E
120.539
119.833
123118.3
117.7
24ν16) 348
0 2 0
(2ν24)
A1
E
178.744
178.766
186179.9
180.0
23ν16) 380
1 1 0
24ν23)
A1
E
202.976
206.570
214222.5
225.0
0 4 0
(4ν24)
A1
E
355.311
355.197
368
2 0 0
(2ν23)
A1
E
223.301
226.756
232200.5
204.5
15) 387
16) 258 22) 391
0 3 0
(3ν24)
A2
E
267.193
267.161
277 1 3 0
(3ν24ν23)
A1
E
380.268
390.184
393
1 2 0
(2ν24ν23)
A2
E
291.127
298.695
304 2 2 0
(2ν2324)
A1
E
408.508
421.158
412
2 1 0
(2ν23ν24)
A2
E
315.337
325.067
325 3 1 0
(3ν23ν24)
A1
E
425.414
368.775
425
0 0 1
21)
A2
E
581.306
581.327
680
a Emphasized in bold transitions in which important Fermi displacements are predicted (2ν23 ↔ ν16, ν24ν22 ↔ ν14). In italic and blue are energies in which the assignments are not conclusive (ZPVE = 406.072 cm−1).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Senent, M.L.; Dalbouha, S. Large Amplitude Motions of Pyruvic Acid (CH3-CO-COOH). Molecules 2021, 26, 4269. https://doi.org/10.3390/molecules26144269

AMA Style

Senent ML, Dalbouha S. Large Amplitude Motions of Pyruvic Acid (CH3-CO-COOH). Molecules. 2021; 26(14):4269. https://doi.org/10.3390/molecules26144269

Chicago/Turabian Style

Senent, María Luisa, and Samira Dalbouha. 2021. "Large Amplitude Motions of Pyruvic Acid (CH3-CO-COOH)" Molecules 26, no. 14: 4269. https://doi.org/10.3390/molecules26144269

Article Metrics

Back to TopTop