Previous Article in Journal
In Silico Analysis of Curcumin and Its Analogs MS13 and MS17 Against HSF1 and HSP Family Proteins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fundamental Vibrational Frequencies and Spectroscopic Constants for Additional Tautomers and Conformers of NH2CHCO

by
Natalia Inostroza-Pino
1,
Megan McKissick
2,
Valerio Lattanzi
3,
Paola Caselli
3 and
Ryan C. Fortenberry
2,*
1
Facultad de Ingeniería, Núcleo de Astroquímica & Astrofísica, Universidad Autónoma de Chile, Av. Pedro de Valdivia 425, Providencia 7500912, Santiago, Chile
2
Department of Chemistry and Biochemistry, University of Mississippi, Oxford, MS 38677, USA
3
Center for Astrochemical Studies, Max-Planck-Institut für Extraterrestrische Physik, Gießenbachstr. 1, 85748 Garching, Germany
*
Author to whom correspondence should be addressed.
Chemistry 2025, 7(5), 140; https://doi.org/10.3390/chemistry7050140
Submission received: 18 July 2025 / Revised: 18 August 2025 / Accepted: 23 August 2025 / Published: 29 August 2025
(This article belongs to the Section Astrochemistry)

Abstract

The creation of larger prebiotic molecules in astronomical regions may require aminoketene (NH2CHCO) as an intermediate, and the two conformers of this molecule exhibit infrared vibrational frequencies with intensities larger even than the antisymmetric stretch in CO2. While the present quantum chemically computed frequencies of these fundamentals of ∼4.7 μm are in the same spectroscopic region as features from functionalized polycyclic aromatic hydrocarbons, they provide clear markers for what James Webb Space Telescope IR observations may be able to distinguish. Additionally, the IR and radioastronomical spectral characterization of two additional 2-iminoacetaldehyde, HN=CHC(=O)H, conformers are also computed as are the same data for a new carbene isomer (NH2CC(=O)H). All conformers of aminoketene and 2-iminoacetaldehyde exhibit dipole moments of more than 2.0 D, if not greater than 4.0 D, implying that they would be notable targets for radioastronomical searches. Additionally, the 2-iminoacetaldehyde conformers have a notable mid-IR C=O stretch around 1735 cm−1 slightly below the same fundamental in formaldehyde. This quantum chemical study is providing a more complete set of reference data for the potential observation of these tautomers and conformers of NH2CHCO in the laboratory or even in space.

1. Introduction

Chemical complexity leading to the molecular origins of life relies upon the creation of molecules containing the elements carbon, hydrogen, oxygen, and nitrogen. While many such molecules are known towards various astronomical sources [1,2], their complexity has grown over the recent few years. Notably, ethanolamine (NH2CH2CH2OH) has been observed towards the G+0.693 molecular cloud [3]. The proposed reaction pathway for the creation of this complicated, highly-saturated, pre-biotic molecule involves the formation of a NH2CHCO structure (aminoketene) from smaller, known interstellar molecules which themselves contain fewer than five atoms. The formation of NH2CHCO could utilize HNCCO + 2H, NH2CH + CO, and NH3 + C + CO [3,4] and may occur in shocked regions [5] where solid-phase formation supported by laboratory experiments [6,7] could then lead to gas-phase desorption and possible detection. Hence, if the formation of ethanolamine proceeds in this way, the aminoketene intermediate should be detectable in the same astronomical regions, as well. In order to potentially observe this intermediate, or any other molecule for that matter, the spectroscopic reference data for the desired compound must be on hand for comparison whether to radioastronomical observation through observatories like the IRAM 30 m telescope, the Yebes 40 m telescope, or even the Atacama Large Millimeter Array (ALMA) or through infrared observation via the James Webb Space Telescope (JWST) or any of its potential successors, e.g., the Habitable Worlds Observatory (HWO).
Previous work has shown that the C s aminoketene structure is not the lowest energy arrangement of the NCCO connectivity with three additional hydrogen atoms among the possible isomers [4,8,9]. This actually belongs to a formyl-imine structure, 2-iminoacetaldehyde or HN=CHC(=O)H, which itself exhibits four conformers all within 0.25 eV relative energy of one another [4,8]. The NH2CHCO isomer actually lies 0.36 eV above the formyl-imine minimum, and it has two conformers: the lower-energy with the amine hydrogens pointing equatorially away from the C=C=O portion of the molecule as well another conformer where the amine hydrogens are axial pointing more towards the portion of the molecule containing the heavy atoms. There are other cyclic and even zwitterionic isomers, but those are higher in energy and exhibit strain or other destabilizing factors. While the previous work has already provided high-level quantum chemical predictions for the fundamental vibrational frequencies and the rotational constants for the two lowest formyl-imine isomers as well as the equatorial aminoketene [4,8], the small energy gaps between isomers and the kinetic nature of gas-phase astrochemical reactions implies that the other conformers/isomers may also be present.
As mentioned, we have previously investigated 14 tautomers and conformers of NH2CHCO [4]. A full depiction of them can be found in Figure 1 of Ref. [8]. Following a re-evaluation of the energetic properties across the entire isomeric family, a corrigendum has been published (see Figure 1 of [8]) confirming that isomer 10 ( a n t i -(E)-2-iminoacetaldehyde) remains the lowest-energy structure, followed by its conformers 9, 7, and 8 ( a n t i -(Z)-2-iminoacetaldehyde, s y n -2-(Z)-iminoacetaldehyde, and s y n -(E)-2-iminoacetaldehyde, respectively). In that study, isomer 8 is conclusively shown to be a local minimum at the CCSD(T)-F12b/cc-pVTZ-F12 level of theory [8]. Additionally, isomer 1 and its conformer, isomer 2, are verified to constitute the next-lowest energy set within the series. Isomers 5, 6, 11, and 12 were initially considered local minima (Figure 1 of [4]); however, further analysis has revealed that they dissociate into van der Waals complexes of methanimine HN=CH2 and carbon monoxide. Consequently, these structures should not be classified as tautomers and conformers of NH2CHCO and should have been excluded from Table 1 of [8] and are excluded from the present study, as well. Isomer 3 remains a second-order saddle point, but the relative energy reported represents the most accurate value to date. Isomer 4 does not correspond to a local minimum and instead reverts to isomer 1 or 2 upon geometry optimization, depending on the initial structural guess. Similarly, isomer 14 relaxes to isomer 13; thus, both are excluded from further consideration.
While the previous study presents accurate spectroscopic data for the lowest-energy isomer and the NH2CHCO isomer hypothesized to be involved in ethanolamine synthesis, we now extend the spectral characterization to additional structures depicted in Figure 1 in this work. Specifically, conformers of the s y n -(Z)- and s y n -(E)-2-iminoacetaldehyde (HN=CHC(=O)H; isomers 7 and 8), which represent conformers of the global minimum and lie within 0.250 eV (2000 cm−1) of the lowest-energy tautomer or conformer of aminoketene, may contribute to the overall vibrational and/or rotational spectra of this system. Furthermore, the two lower-energy, a n t i isomers have been observed in condensed-phase experiments [10] as well as studied computationally [11], and similar diiminoethene conformers have been explored via quantum chemistry with a similar approach as that used here [12]. Consequently, the spectral properties of the higher-energy s y n -2-iminoacetaldehyde isomers (isomers 7 and 8, as defined in [8] and depicted in Figure 1) are computed herein. While theoretically computed spectroscopic data exist for isomers 7 and 8 [13], the present work aims to provide these at a higher level. Furthermore, new spectral data for isomer 2, as well as updated spectroscopic parameters for isomer 1, the lowest-energy conformation of aminoketene are also presented. Finally, the spectral characteristics of a previously uncharacterized aminoketene tautomer/conformer (7b) are reported in this work. Additionally, our previous work does not provide the intensities for the aminoketene fundamental frequencies, and that is addressed in this work. The revised relative energies presented in this study enhance our ability to model the interconversion and interactions of these molecular species in various environments, including diverse astrophysical regions. As such, several knowledge gaps are present for a more complete characterization of the tautomers and conformers of NH2CHCO, and this work will provide such.
As done in the previous work, the anharmonic spectroscopic data reported herein will rely upon quartic force fields (QFFs) [14,15,16,17,18,19] computed with explicitly correlated coupled-cluster theory within the F12 formalism [20,21,22]. The use of core electron correlation (“cC”) and scalar relativity (“R”) [23] adds to the accuracy of the explicit correlation which, itself, better approximates the complete basis set limit for a smaller basis set size [24,25,26,27,28,29]. As such, the so-called “F12-TcCR” QFF (so named for the F12-based energy with a “T” triple- ζ basis set along with the “cC” and “R” terms) has been shown to provide accuracies for B and C spectroscopic constants to better than 0.1% of experiment and fundamental vibrational frequencies within 0.7%, often within 1.0 cm−1 [30,31]. Furthermore, rotational constants for related nitrogen-containing hydrocarbons have been predicted to within 10 MHz of experiment for numerous molecules [32,33,34]. and the rotational constants and fundamental vibrational frequencies of other prebiotic species of astrochemical and astrobiological significance have also been provided [35,36,37,38,39]. Hence, this work will utilize this same quantum chemical approach to increase the spectroscopic reference data available for the aminoketene conformers and tautomers. These data will provide for a more complete laboratory experimental exploration of these isomers and will inform where and to what extent these isomers and conformers will be detectable with current, archival, and future multi-messenger observations [40,41].

2. Results & Discussion

2.1. Fundamental Vibrational Frequencies

As isomers 1 and 2 represent the conformers of the isomer believed to play a role in the formation of ethanolamine potentially in shocked regions, their comparison begins the discussion of the fundamental vibrational frequencies from Table 1 and Table 2. As expected and at first glance, they exhibit notably similar fundamental vibrational frequencies and corresponding intensities. Most significantly, however, are the exceptionally intense ν 4 a fundamental frequencies at 2143.4 cm−1 and 2128.4 cm−1 for isomer 1 and 2, respectively. Their respective intensities of 598 km mol−1 and 596 km mol−1 are far larger than most organic molecules exhibit, and, in fact, are more than 50% larger than even the antisymmetric stretch in carbon dioxide. These fundamentals both arise from the “antisymmetric” central carbon shuttling motion in the C=C=O portion of each isomer, again, similar to the antisymmetric stretch in CO2, and these strong intensities have been previously documented quantum chemically [42]. Additionally, they fall in the range of 4.6 μm to 4.7 μm, an IR observational region where the majority of the observed features have recently been attributed to CN and/or CD stretches within polycyclic aromatic hydrocarbons (PAHs) or CN-PAHs [43,44,45]. Hence, these molecules are exhibiting exceptional intensities in regions of current interest for explaining portions of JWST observations.
Another intense fundamental is also present for both aminoketene isomers at 748.2 cm−1 and 753.0 cm−1, again respective of isomers 1 and 2. This ν 10 a fundamental represents the umbrella inversion motion of the amine group akin to the same motion in ammonia and would be observed at 13.3 μm. Besides these sets, the fundamental frequencies exhibit in large part either small intensities or fundamental frequencies below what JWST could observe. Additionally, the isomer 2 fundamentals below 600 cm−1 are producing questionable anharmonicities. QFFs are known to exhibit such poor descriptions for quasi-degenerate conformers [46], but the use of normal coordinates from pbqff helps to minimize such improper descriptions [47,48] where they appear to influence only the low-frequency fundamentals for isomer 2. Additional considerations for a Hamiltonian employing hindered rotation would potentially also improve the descriptions of these likely, large-amplitude motions. However, the instruments onboard JWST do not perform well at frequencies of less than 400 cm−1 (25 μm) and do not operate below 333 cm−1 (30 μm) at all, making any such fundamental vibrational frequencies not relevant for application to this mission data. Their treatment is left for future work for other applications. Beyond this, the current methods are unable to fully quantify the contribution of the Darling-Denison resonances between modes 7 and 8 of isomer 2. An analysis similar to that of [49] would quantify this, but the contributions that these fundamentals would have on the overall spectrum is likely small due to the low intensities computed.
The s y n -2-iminoacetaldehyde conformers (isomers 7 and 8) also share similarities between themselves for the fundamental vibrational frequencies and corresponding intensities as shown in Table 3 and Table 4. The ν 3 a fundamental at 2844.4 cm−1 and 2751.1 cm−1, isomers 7 & 8 respectively, also has relatively large organic intensities above 70 km mol−1, but these are still nowhere near what isomers 1 and 2 exhibit. This ν 3 intensity here for isomers 7 and 8 originates from the CH stretch on the formyl moiety, but the NH stretch has nearly no intensity in either isomer with the other CH stretch intensity in between these two. The ν 4 C=O stretch just below 1800 cm−1 in both isomers also exhibits a notable intensity, but the ν 5 N=H stretch is not as intense. The ν 8 in-plane hydrogen bending at 1211.8 cm−1 for isomer 7 (Table 3) exhibits an intensity of nearly 100 km mol−1, but this drops off to roughly half this much in isomer 8. This drop occurs between isomers due to the intramolecular hydrogen bonding present in isomer 7 that is not present in isomer 8.
The lowest-frequency anharmonic fundamentals of both s y n -2-iminoacetaldehyde conformers (isomers 7 and 8) also struggle within the QFF/VPT2 framework, but their anharmonicities should be within the accuracy of the method where the harmonic fundamental frequencies are sufficient approximations for the fundamental frequencies. Again, these frequencies are below the JWST observeable range, but the harmonic frequencies computed here should be sufficient for other applications. The B3LYP/6-311+G(d,p) anharmonic frequencies are able to provide an additional estimate for these low-frequency modes.
One final set of fundamental vibrational frequencies is provided here for a new isomer. This isomer was discovered when an attempted geometry optimization of isomer 7 went awry leading us to call this isomer 7b. Its fundamental frequencies are shown in Table 5 and structure in Figure 1. This C 1 , carbene molecule lies 1.33 eV above isomer 7 which puts it 1.42 eV above the lowest-energy isomer (isomer 10) from [8]. While this energy and the highly-reactive nature of the carbene imply that such a molecule is likely short-lived in any environment in which it may briefly exist, the spectral features still may provide insights into the evolution of the NCCO molecular connectivity in that if this isomer is observable, chemistry different from the formation ethanolamine is likely taking place. Isomer 7b exhibits a number of large intensity fundamentals in Table 5 with ν 3 , ν 5 , and ν 12 all exhibiting intensities of ∼70 km mol−1 or greater with two above 115 km mol−1. The ν 4 and ν 5 C=O and C-N stretches at 1680.3 cm−1 and 1697 cm−1, respectively, imply that these features will likely combine together in some observations, especially at low resolution (potentially in archival data form older IR observatories like the Kuiper Airborne Observatory or the Spitzer Space Telescope), to generate a large feature in their lower-resolution IR spectra. Again, the anharmonic fundamental frequencies below 500 cm−1 are of questionable quality, but these harmonics and the rest of the anharmonic fundamentals should be good predictions of vibrational behavior for this molecule.

2.2. Spectroscopic Constants

The A, B, and C rotational constants for each of the isomers unique to this work are given in Table A1. The numbers in column 1 correspond to the fundamental vibrational state that exhibits those specific rotational constants; the “0” is for the pure rotational constants. For isomer 2, the R α zero-point corrected A0, B0, and C0 constants (given as row “0”) of 1.3629509 cm−1, 0.1564255 cm−1, and 0.1433041 cm−1 make this molecule near-prolate and translate into 40860 MHz, 4689 MHz, and 4296 MHz as given in Table 6. Additionally, the isomer 2 B and C rotational constants are 52 MHz and 10 MHz greater than their counterparts in isomer 1 [4], again confirming that they will have similar rotational spectra even though the A constant is more than 3000 MHz smaller here in isomer 2. However, isomers 1 and 2 have nearly the same dipole moment at 2.65 D (Table 7 and Table A2) which is expected for similar isomers with a simple shift in the positions of the hydrogen atoms. Such close correlation of dipole moments implies that the intensity of the rotational signal will not decrease if the higher energy conformer is populated.
Similarly, the rotational constants of isomers 7 and 8 show that these are also near-prolate molecules, but to a lesser degree than isomers 1 and 2. The rotational constants of isomers 7 and 8 mirror one another as shown in Table 6 and Table A1; the quartic and sextic distortion constants are reported in Table 7 and Table A2. However, the dipole moment of s y n -(E)-2-iminoacetaldehyde (isomer 8) at 4.22 D is nearly double that of isomer 7 at 2.22 D. Hence, in warmer regions (assuming thermal equilibrium in those regions) isomer 8 may likely be more readily observed than isomer 7 even though the former is 0.15 eV higher in relative energy [8]. Such a case has already been noted for two conformers of carbonic acid, where the higher-energy but higher-dipole c i s - t r a n s form has been observed while the c i s - c i s has not [50]. Differently here, though, both isomers are notably polar, implying both isomers, as well as isomers 9 and 10 computed previously [4], may all produce observable signal in any environment wherever the HN=CHC(=O)H molecule is synthesized.
Isomer 7b also exhibits a dipole moment of nearly 5 D and has similar spectroscopic constants as the 2-iminoacetaldehyde isomer constants. However, this molecule has its own unique spectroscopic signatures that will allow it to be distinguished in any observations of this molecule, especially in any potential high-resolution laboratory experiments. Finally, while the rotational constants for states ν 13 ν 15 of each isomer are given in Table 6 for completeness, the questionable behavior of the corresponding vibrational frequencies for these states likely implies that these estimates are not as accurate as the rotational constants for the higher frequency states and, especially, the zero-point.

3. Observational and Spectroscopic Considerations

2-iminoacetaldehyde is one of the simplest prebiotic molecules yet to be observed in the ISM or other astronomical regions. Its family of conformers exhibits strong IR absorption in the range of 1720 cm−1 to 1780 cm−1 and notable dipole moments for all four conformers (isomers 7–10). Case in point, Figure 2 displays the F12-TcCR QFF spectra for all four conformers. Clearly, the region from 1700 cm−1 to 1800 cm−1 (given as an inset) produces the most intense features as the ν 4 C=O stretch fundamental carries the largest intensity. The three lowest-energy conformers all cluster together in this region with the highest energy s y n -(E) conformer showcasing a blue-shifted fundamental in this region. The C=O stretch of formaldehyde also lies in this region at 1746.1 cm−1 [51], but F12-TcCR computations shift this frequency of H2CO up to 1752.0 cm−1 [30]. Hence, the C=O stretches of the 2-iminoacetaldehyde conformers computationally predicted to lie in the region of 1735 cm−1 should still be distinct from the C=O stretch of the more common formaldehyde molecule.
Additionally, Figure 2 highlights that population of the first three conformational levels will not shift the frequency of this molecule as the higher energy conformers become populated at higher temperatures. Purely thermodynamically, the a n t i -(E) conformer should be the dominant form observed in the cold ISM. From the relative energies of Ref. [8], temperatures would have to surpass 115 K before a meager 1% of the population would convert from a n t i -(E) to a n t i -(Z) (isomer 9 to isomer 10) and more than 170 K before s y n -(Z) would provide a notable contribution to the population of 2-iminoacetaldehyde. The highest energy s y n -(E) (isomer 8) conformer would require more than 500 K before its population is ever greater than 1%. Hence, the a n t i -(E) should be the primary target for observation in colder regions, but all conformers may contribute to the observed IR or JWST spectra in warmer regions if they are present.
Even with these considerations, the possible unique detection of 2-iminoacetaldehyde will likely arise from radioastronomy due to the nature of emission rotational spectroscopy and from the 2.0+ D dipole moments of the conformers of this molecular family. While the rotational constants and the fundamental vibrational frequencies of the two a n t i -2-iminoacetaldehyde conformers have been computationally explored previously [4,8], this present work is expanding that to the s y n -2-iminoacetaldehyde (Z) and (E) pair. The a n t i conformers are both very much of near-prolate character with κ values of −0.98. Due to their similarities, high resolution rotational spectroscopy would be required to separate the rotational spectra of the a n t i conformers. Most notably, the differences in the B 0 constants are less than 4.0 MHz while this grows to more than 15 MHz in the C 0 constants [4] However, the dipole moments and rotational constants of the s y n conformers are notably separated from each other and from the two a n t i conformers. Hence, they will produce other lines in the spectra as these higher-energy conformers are populated at higher temperatures. Regardless, the small number of heavy atoms (4) puts this molecule in the range of known prebiotic complex organic molecules [1,2,52]. In addition to ethanolamine, the conformers of 2-iminoacetaldehyde exhibit nearly the same skeletal structure as glycine in the N-C-C-O connectivity, missing only an additional oxygen atom bonded to the aldehyde carbon and a couple of ubiquitous hydrogen atoms.
This connectivity is maintained for the aminoketene conformers (isomers 1 and 2). Only the positions of the hydrogen atoms change. While the conformers of 2-iminoacetaldehyde are the lowest energy isomers, the NH2CHCO conformers could still be the first of these isomers synthesized in astronomical environments potentially in shocked regions [3,5]. As such, the spectroscopic constants provided in this work are vital for understanding the possible populations of these isomers astrophysically. Like isomers 9 and 10, the spectroscopic constants of isomers 1 and 2 differ by less than 50 MHz for the B constant and less than 10 MHz for the C constant. As such, the rotational signals of the two NH2CHCO conformers may combine with lower-resolution laboratory characterization or possible astronomical observation. Hence, shifts in population between the aminoketene isomers would not greatly change the observations.
While laboratory work would need to conclusively determine the rotational constants for any of these isomers, the present work should be providing rotational constants to within 10 MHz or less of experiment [30,34], greatly reducing the need to comb the spectra significantly in order to make attributions. Finally, the ratio of aminoketene to 2-iminoacetaldehyde would also provide insights into the nature of the formation of these molecules themselves, and it would also be able to give a clue as to the evolution of relatively simple prebiotic molecules.

4. Computational Details

The QFFs producing the anharmonic spectral data in this work begin with tightly converged F12 coupled cluster singles, doubles, and perturbative triples [CCSD(T)-F12b] [20,21,53] geometry optimizations employing the cc-pCVTZ-F12 basis set [54,55,56]. The optimizations and all electronic structure computations in this work (unless otherwise noted) make use of the Molpro2024.1 quantum chemistry program [57,58]. After the geometries are optimized, the pbqff program [47,59] constructs normal coordinates, executes displacements of 0.005 Å to construct the actual QFF, computes the F12-TcCR energies, sets up the Fermi resonances [60], and runs the second-order rotational and vibrational perturbation theory (VPT2) computations [61,62,63,64] in order to produce the rotational constants and vibrational frequencies. The dipole moments and fully anharmonic fundamental frequency intensities are computed with B3LYP/6-311+G(d,p) within Gaussian16 [65,66,67,68,69]. Such intensities are known to be sufficiently accurate as compared to higher-level methods and require orders of magnitude less complexity and computational time [70,71,72]. The spectra of the 2-iminoacetaldehyde conformers are produced with these intensities and F12-TcCR frequencies utilizing an artificial full-width at half-max of 15.0 cm−1. The full sets of geometries and resonances are included in the Appendix A, Appendix B and Appendix C.

5. Conclusions

A more complete spectral picture of the tautomers and conformers of NH2CHCO is produced in this work, notably showcasing that the two aminoketene conformers have strong IR features in the range of 4.6–4.7 μm (2125–2150 cm−1) potentially in competition with functionalized PAHs in astronomical observations where C≡N stretches have been reported recently [45]. Hence, these two molecules are also notable targets for their IR features as well as their radioastronomical (rotational) signatures. Since the slightly lower-energy equatorial conformer (isomer 1) has a nearly identical dipole moment as the axial conformer (isomer 2), either can be observed rotationally and would likely be observed in similar amounts. Detection of either would help to add evidence for this molecule’s role in the interstellar formation of the known prebiotic molecule ethanolamine.
The 2-iminoacetaldehyde conformers from this and previous work [4] all exhibit notable dipole moments of at least 2.0 D if not greater than 4.0 D. Their IR fundamental vibrational frequencies are also populated with intense ν 4 C=O stretch fundamentals enabling their observation at frequencies in the range of 1700 cm−1 to 1800 cm−1. The presently examined s y n -(Z) and s y n -(E) conformers of 2-iminoacetaldehyde or even the lower-energy a n t i -(E) and a n t i -(Z) conformers are likely not part of the synthesis of ethanolamine. They would be required to migrate a hydrogen atom to create the amine group. However, they may have other roles to play astrochemically given their notable mid-IR intensities, and these data provide the necessary references for their potential observation.
Finally, a new isomer, isomer 7b NH2CCOH, is also shown to be a local minimum in this work despite being a carbene. While much higher in energy that most of the rest of the isomers examined this and the previous work, it could still form from other chemical reactions. Hence, its spectral data are produced in this work, as well.

Author Contributions

Conceptualization, N.I.-P., V.L., P.C. and R.C.F.; methodology, N.I.-P., M.M. and R.C.F.; validation, N.I.-P., M.M., V.L. and R.C.F.; formal analysis, N.I.-P., M.M., V.L. and R.C.F.; investigation, M.M., R.C.F.; resources, N.I.-P., P.C. and R.C.F.; data curation, N.I.-P., M.M., V.L., P.C. and R.C.F.; writing—original draft preparation, R.C.F.; writing—review and editing, N.I.-P., M.M., V.L., P.C. and R.C.F.; visualization, N.I.-P., M.M., V.L. and R.C.F.; supervision, N.I.-P., V.L., P.C. and R.C.F.; project admistration, N.I.-P., P.C. and R.C.F.; funding acquisition, N.I.-P., V.L., P.C. and R.C.F. All authors have read and agreed to the published version of the manuscript.

Funding

NI acknowledges FONDECYT grant N°1241193 and VRIP. The work by RCF is supported by NSF grant AST-2407815 and NASA grant NNH22ZHA004 as well as the University of Mississippi’s College of Liberal Arts. Computational resources are provided by the Mississippi Center for Supercomputing Research supported in part by NSF grant OIA-1757220. PC and VL gratefully acknowledge the Max Planck Society for the financial support.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

RCF would also like to acknowledge the Falcioni family for providing him time and space to finalize this project. The authors are also thankful to Devin Matthews of Southern Methodist University for discussion of resonance treatment.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ALMAAtacama Large Millimeter Array
CCSD(T)Coupled Cluster Singles, Doubles, and Perturbative Triples
HWOHabitable Worlds Observatory
JWSTJames Webb Space Telescope
km mol−1kilometer·mole−1
cm−1wavenumber or centimeter−1

Appendix A. Cartesian Geometries (in Å)

Isomer 1
H −1.6405093651 0.0000000000 −0.7522179792
C −0.5670904669 0.0000000000 −0.6060145320
C −0.1331354564 0.0000000000 0.6456891668
O 0.2330901179 0.0000000000 1.7498731100
N 0.4066623583 0.0000000000 −1.6562135778
H 0.3168089807 −0.8191773821 −2.2407953672
H 0.3168089807 0.8191773821 −2.2407953672
Isomer 2
H −1.6449020710 0.0000000000 −0.7749380918
C −0.5781259157 0.0000000000 −0.6064104538
C −0.1511833766 0.0000000000 0.6466114844
O 0.2460075333 0.0000000000 1.7427181317
N 0.3308672541 0.0000000000 −1.7189059252
H 0.9164063373 −0.8223027220 −1.7401870804
H 0.9164063373 0.8223027220 −1.7401870804
Isomer 7
C 0.5380152852 0.0000 0.7419821849
O −0.5016102791 0.0000 1.3587300389
H 1.5115447166 0.0000 1.2597324153
C 0.5636836242 0.0000 −0.7677288505
N −0.4972491156 0.0000 −1.4700327546
H −1.3033281064 0.0000 −0.8347234184
H 1.5357136698 0.0000 −1.2577117368
Isomer 8
C 0.0000000000 −0.5486835237 −0.7582089483
H 0.0000000000 −1.5458972291 −1.2081628906
C 0.0000000000 −0.5154903515 0.7507475925
O 0.0000000000 0.4991238427 1.4009309562
H 0.0000000000 −1.5085475338 1.2367501753
N 0.0000000000 0.5377372864 −1.4216030164
H 0.0000000000 0.3401815679 −2.4220525948
Isomer 7b
O −0.4194034557 −0.1280735302 1.4780361354
C 0.4279934912 0.2427826190 0.6831532335
H 1.2281122746 0.9484203032 0.9790568234
C 0.5618420795 −0.3069894101 −0.6707968546
H −0.3706276578 −0.2602326313 −2.4454276128
N −0.3516964891 0.0990579545 −1.5029507819
H −1.1080989173 0.7333377722 −1.2567308777

Appendix B. Resonances

Note, for all resonances listed, add 1 to each as the harmonics begin counting with 0. For example, “14” listed below is actually “15”, and “0” is actually “1”.
Isomer 2
Coriolis Resonances:
8 7 C
11 10 A
11 10 B
12 10 A
14 13 A
Type 1 Fermi Resonances:
9 4
9 5
10 5
10 7
11 7
11 8
12 8
14 13
Type 2 Fermi Resonances:
5 4 2
7 3 2
8 7 3
11 9 6
11 10 6
12 9 6
12 11 7
12 11 8
13 10 8
14 1 0
14 6 5
14 7 6
14 8 6
14 11 9
14 11 10
Isomer 7
Coriolis Resonances:
8 7 A
8 7 B
11 10 C
14 13 B
Type 1 Fermi Resonances:
4 0
4 1
5 2
6 2
10 3
10 4
11 4
11 5
11 6
12 5
12 6
12 7
14 13
Type 2 Fermi Resonances:
5 3 1
5 4 1
6 3 1
6 3 2
6 4 1
6 4 2
6 5 2
7 3 1
7 3 2
11 10 3
11 10 4
13 7 5
13 7 6
13 10 6
13 10 7
13 11 7
14 8 7
14 9 7
14 10 8
14 10 9
14 11 8
14 11 9
14 12 10
14 12 11
14 13 12
Isomer 8
Coriolis Resonances:
8 7 A
8 7 B
11 10 C
12 11 A
Type 1 Fermi Resonances:
4 0
5 1
5 2
6 2
10 3
10 4
11 4
12 5
12 7
Type 2 Fermi Resonances:
5 4 1
6 3 1
6 4 1
6 4 2
6 5 2
7 3 1
7 3 2
7 4 1
11 10 3
11 10 4
12 8 4
13 6 4
13 7 6
13 10 7
14 8 7
14 9 7
14 10 8
14 10 9
14 12 11
Isomer 7b
Coriolis Resonances:
9 8 A
10 9 A
Type 1 Fermi Resonances:
3 0
3 1
4 1
6 2
10 4
10 5
11 3
11 4
12 8
12 9
12 10
13 9
14 14
Type 2 Fermi Resonances:
4 3 0
4 3 1
5 3 1
6 3 2
6 4 2
6 5 2
7 3 2
10 8 3
10 9 3
10 9 4
11 9 3
11 9 4
11 10 3
11 10 4
12 7 5
12 8 5
12 8 6
12 9 5
12 9 6
12 11 7
13 7 5
13 8 5
13 8 6
13 9 5
13 9 6
13 9 7
13 10 6
13 12 9
13 12 10
14 5 3
14 6 4
14 6 5
14 10 8
14 10 9
14 13 12

Appendix C. Spectroscopic Constants in cm−1

Table A1. The F12-TcCR rotational constants (in cm−1) for the tautomers and conformers of NH2CHCO.
Table A1. The F12-TcCR rotational constants (in cm−1) for the tautomers and conformers of NH2CHCO.
Vib.Isomer 2Isomer 7Isomer 8Isomer 7b
StateABCABCABCABC
01.36295090.15642550.14330410.79829170.21605200.16998160.86797010.20189050.16541651.00483150.18492220.1679052
11.35900940.15636800.14321200.79730420.21616970.17001280.86716220.20177160.16531311.00840020.18435250.1676343
21.35797470.15635550.14322530.79813370.21547470.16963030.86778280.20138750.16509460.99709940.18577030.1682240
31.37721800.15563370.14281760.79860170.21537990.16959690.86817430.20144090.16514621.02838360.18160240.1665233
41.35948900.15564000.14261050.79484800.21594680.16976040.86412570.20170460.16515581.00002900.18496770.1676992
51.35911700.15638500.14331740.79424850.21585450.16967470.86329390.20171420.16510200.99682890.18561490.1680231
61.34866440.15664850.14322600.79847350.21615840.16978190.86744690.20223960.16536540.99052040.18590570.1681927
71.48728060.15698010.14329410.79766190.21656810.16998490.86704680.20238040.16542881.00621750.18476290.1678530
81.21931760.15636440.14344980.80492320.21599640.16980070.88119270.20223050.16543861.00844800.18469340.1676515
91.37511640.15569490.14275390.79021590.21591430.16992460.85221020.20178050.16544271.01138960.18374560.1673406
101.36713960.15652620.14342220.79682370.21580270.17003870.86647680.20178430.16550691.00233470.18445070.1674099
111.37953520.15625890.14315260.79866300.21493800.16914130.86882710.20078250.16460341.00986990.18376990.1672491
121.36144590.15598660.14319550.79967140.21588900.16970640.86715910.20190310.16524371.00005250.18486090.1678934
131.35773940.15644940.14345540.79669410.21556320.16993330.86711150.20149870.16538501.02115710.18346600.1673613
141.37916630.15711860.14358590.80447800.21494570.16923480.87357550.20182800.16509111.01806460.18285240.1672123
151.38937520.15573690.14280350.80157970.21466010.16989740.89132320.19643690.16677081.00572300.18715740.1686394
Table A2. The Watson A-reduced F12-TcCR spectroscopic constants (in cm−1) and B3LYP/6-311+G(d,p) dipole moments (in D) for the aminoketene isomers.
Table A2. The Watson A-reduced F12-TcCR spectroscopic constants (in cm−1) and B3LYP/6-311+G(d,p) dipole moments (in D) for the aminoketene isomers.
ConstantIsomer 2Isomer 7Isomer 8Isomer 7b
Δ J 0.0000001027060.0000001909160.0000001699720.000000481225
Δ K 0.0000734637800.0000031873550.0000038099950.000052224135
Δ J K −0.000003598161−0.0000009641650.000000077775−0.000007822597
δ J 0.0000000220970.0000000534880.0000000456090.000000148854
δ k 0.0000000600070.0000003854530.000000421895−0.000003723975
Φ J 4.271 × 10 13 4.687 × 10 15 −2.733 × 10 11 9.064 × 10 12
Φ K 1.275 × 10 8 9.158 × 10 11 −2.672 × 10 8 8.709 × 10 9
Φ J K −7.108 × 10 12 5.529 × 10 12 −2.728 × 10 10 −4.758 × 10 11
Φ K J −3.874 × 10 10 −4.433 × 10 11 −2.201 × 10 9 −9.462 × 10 10
ϕ J 1.513 × 10 13 2.404 × 10 14 −1.367 × 10 11 4.180 × 10 12
ϕ J K 1.373 × 10 11 −9.394 × 10 14 −5.411 × 10 10 4.831 × 10 11
ϕ k 3.714 × 10 10 2.648 × 10 11 4.342 × 10 10 4.111 × 10 9
μ 2.652.224.224.83
μ x −1.412.180.00−0.64
μ y 0.000.00−3.562.12
μ z −2.250.40−2.26−4.29

References

  1. McGuire, B.A. 2018 Census of Interstellar, Circumstellar, Extragalactic, Protoplanetary Disk, and Exoplanetary Molecules. Astrophys. J. Suppl. Ser. 2018, 239, 17. [Google Scholar] [CrossRef]
  2. McGuire, B.A. 2021 Census of Interstellar, Circumstellar, Extragalactic, Protoplanetary Disk, and Exoplanetary Molecules. Astrophys. J. Suppl. Ser. 2021, 259, 30. [Google Scholar] [CrossRef]
  3. Rivilla, V.M.; Jiménez-Serra, I.; Martín-Pintado, J.; Briones, C.; Rodríguez-Almeida, L.F.; Rico-Villas, F.; Tercero, B.; Zeng, S.; Colzi, L.; de Vicente, P.; et al. Discovery in Space of Ethanolamine, the Simplest Phospholipid Head Group. Proc. Natl. Acad. Sci. USA 2021, 118, e2101314118. [Google Scholar] [CrossRef]
  4. Alberton, D.; Inostroza-Pino, N.; Fortenberry, R.C.; Lattanzi, V.; Endres, C.; Zamponi, J.F.; Caselli, P. Accurate ab Initio Spectroscopic Studies of Promising Interstellar Ethanolamine Iminic Precursors. Astron. Astrophys. 2024, 683, A198. [Google Scholar] [CrossRef]
  5. Zeng, S.; Rivilla, V.M.; Jiménez-Serra, I.; Colzi, L.; Martín-Pintado, J.; Tercero, B.; de Vicente, P.; Martín, S.; Requena-Torres, M.A. Amides Inventory towards the G+0.693-0.027 Molecular Cloud. Mon. Not. R. Astron. Soc. 2023, 523, 1448–1463. [Google Scholar] [CrossRef]
  6. Krasnokutski, S.A.; Chuang, K.J.; Jäger, C.; Ueberschaar, N.; Henning, T. A Pathway to Peptides in Space through the Condensation of Atomic Carbon. Nat. Astron. 2022, 6, 381–386. [Google Scholar] [CrossRef]
  7. Krasnokutski, S.A.; Jäger, C.; Henning, T.; Geffroy, C.; Remaury, Q.B.; Poinot, P. Formation of Extraterrestrial Peptides and their Derivatives. Sci. Adv. 2024, 10, eadj7179. [Google Scholar] [CrossRef]
  8. Alberton, D.; Inostroza-Pino, N.; Fortenberry, R.C.; Lattanzi, V.; Endres, C.; Zamponi, J.F.; Caselli, P. Accurate ab Initio Spectroscopic Studies of Promising Interstellar Ethanolamine Iminic Precursors (Corrigendum). Astron. Astrophys. 2024, 691, C1. [Google Scholar] [CrossRef]
  9. Fourrié, I.; Matz, O.; Ellinger, Y.; Guillemin, J.-C. Relative Thermodynamic Stability of the [C,N,O] Linkages as an Indication of the Most Abundant Structures in the ISM. Astron. Astrophys. 2020, 639, A16. [Google Scholar] [CrossRef]
  10. Drabkin, V.D.; Paczelt, V.; Eckhardt, A.K. Spectroscopic identification of interstellar relevant 2-iminoacetaldehyde. Chem. Commun. 2023, 59, 12715–12718. [Google Scholar] [CrossRef]
  11. Redondo, P.; Largo, A.; Barrientos, C. Structure and Spectroscopic Properties of Imine Acetaldehyde: A Possible Interstellar Molecule. Mon. Not. R. Astron. Soc. 2018, 478, 3042–3048. [Google Scholar] [CrossRef]
  12. McKissick, M.; Fortenberry, R.C. Anharmonic Vibrational Frequencies and Spectroscopic Constants for the Six Conformers of 1,2-Diiminoethane: A Promising Prebiotic Molecule for Astronomical Detection. ACS Earth Space Chem. 2025, 9, 403–410. [Google Scholar] [CrossRef]
  13. Redondo, P.; Sanz-Novo, M.; Largo, A.; Barrientos, C. Amino Acetaldehyde Conformers: Structure and Spectroscopic Properties. Mon. Not. R. Astron. Soc. 2019, 492, 1827–1833. [Google Scholar] [CrossRef]
  14. Barone, V.; Biczysko, M.; Puzzarini, C. Quantum Chemistry Meets Spectroscopy for Astrochemistry: Increasing Complexity toward Prebiotic Molecules. Acc. Chem. Res. 2015, 48, 1413–1422. [Google Scholar] [CrossRef]
  15. Puzzarini, C.; Barone, V. Diving for Accurate Structures in the Ocean of Molecular Systems with the Help of Spectroscopy and Quantum Chemistry. Acc. Chem. Res. 2018, 51, 548–556. [Google Scholar] [CrossRef]
  16. Fortenberry, R.C.; Lee, T.J. Computational Vibrational Spectroscopy for the Detection of Molecules in Space. Ann. Rep. Comput. Chem. 2019, 15, 173–202. [Google Scholar]
  17. Puzzarini, C.; Barone, V. The Challenging Playground of Astrochemistry: An Integrated Rotational Spectroscopy–Quantum Chemistry Strategy. Phys. Chem. Chem. Phys. 2020, 22, 6507–6523. [Google Scholar] [CrossRef] [PubMed]
  18. Fortenberry, R.C.; Lee, T.J. Vibrational and Rovibrational Spectroscopy Applied to Astrochemistry. In Vibrational Dynamics of Molecules; Bowman, J.M., Ed.; World Scientific: Singapore, 2022; pp. 235–295. [Google Scholar]
  19. Franke, P.R.; Stanton, J.F.; Douberly, G.E. How to VPT2: Accurate and Intuitive Simulations of CH Stretching Infrared Spectra Using VPT2+K with Large Effective Hamiltonian Resonance Treatments. J. Phys. Chem. A 2021, 125, 1301–1324. [Google Scholar] [CrossRef] [PubMed]
  20. Adler, T.B.; Knizia, G.; Werner, H.J. A Simple and Efficient CCSD(T)-F12 Approximation. J. Chem. Phys. 2007, 127, 221106. [Google Scholar] [CrossRef]
  21. 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]
  22. Huang, X.; Valeev, E.F.; Lee, T.J. Comparison of One-Particle Basis Set Extrapolation to Explicitly Correlated Methods for the Calculation of Accurate Quartic Force Fields, Vibrational Frequencies, and Spectroscopic Constants: Application to H2O, N2H+, NO2+, and C2H2. J. Chem. Phys. 2010, 133, 244108. [Google Scholar] [CrossRef] [PubMed]
  23. Douglas, M.; Kroll, N.M. Quantum Electrodynamical Corrections to the Fine Structure of Helium. Ann. Phys. 1974, 82, 89–155. [Google Scholar] [CrossRef]
  24. Huang, X.; Lee, T.J. A Procedure for Computing Accurate Ab Initio Quartic Force Fields: Application to HO2+ and H2O. J. Chem. Phys. 2008, 129, 044312. [Google Scholar] [CrossRef]
  25. Huang, X.; Lee, T.J. Accurate Ab Initio Quartic Force Fields for NH2 and CCH and Rovibrational Spectroscopic Constants for Their Isotopologs. J. Chem. Phys. 2009, 131, 104301. [Google Scholar] [CrossRef]
  26. Huang, X.; Taylor, P.R.; Lee, T.J. Highly Accurate Quartic Force Field, Vibrational Frequencies, and Spectroscopic Constants for Cyclic and Linear C3H3+. J. Phys. Chem. A 2011, 115, 5005–5016. [Google Scholar] [CrossRef] [PubMed]
  27. Fortenberry, R.C.; Huang, X.; Francisco, J.S.; Crawford, T.D.; Lee, T.J. The trans-HOCO Radical: Fundamental Vibrational Frequencies, Quartic Force Fields, and Spectroscopic constants. J. Chem. Phys. 2011, 135, 134301. [Google Scholar] [CrossRef]
  28. Györffy, W.; Werner, H.J. Analytical Energy Gradients for Explicitly Correlated Wave Functions. II. Explicitly Correlated Coupled Cluster Singles and Doubles with Perturbative Triples Corrections: CCSD(T)-F12. J. Chem. Phys. 2018, 148, 114104. [Google Scholar] [CrossRef]
  29. Westbrook, B.R.; Fortenberry, R.C. Anharmonic Frequencies of (MO)2 & Related Hydrides for M = Mg, Al, Si, P, S, Ca, & Ti and Heuristics for Predicting Anharmonic Corrections of Inorganic Oxides. J. Phys. Chem. A 2020, 124, 3191–3204. [Google Scholar]
  30. Watrous, A.G.; Westbrook, B.R.; Fortenberry, R.C. F12-TZ-cCR: A Methodology for Faster and Still Highly Accurate Quartic Force Fields. J. Phys. Chem. A 2021, 125, 10532–10540. [Google Scholar] [CrossRef]
  31. Inostroza-Pino, N.; Lattanzi, V.; Palmer, C.Z.; Fortenberry, R.C.; Lee, T.J.; Caselli, P.; Mardones, D. Rotational Spectroscopic Characterization of the [D2,C,S] System: An Update from the Laboratory and Theory. Mol. Phys. 2024, 122, e2280762. [Google Scholar] [CrossRef]
  32. Inostroza, N.; Huang, X.; Lee, T.J. Accurate ab initio quartic force fields of cyclic and bent HC2N isomers. J. Chem. Phys. 2011, 135, 244310. [Google Scholar] [CrossRef]
  33. Inostroza, N.; Fortenberry, R.C.; Huang, X.; Lee, T.J. Rovibrational spectroscopic constants and fundamental vibrational frequencies for isotopologues of cyclic and bent singlet HC2N isomers. Astrophys. J. 2013, 778, 160. [Google Scholar] [CrossRef]
  34. Fortenberry, R.C.; Esposito, V.J. On the Formation and Detectability of H2CNCN and Its Progenitors. Astrophys. J. 2024, 972, 35. [Google Scholar] [CrossRef]
  35. Davis, M.C.; Garrett, N.R.; Fortenberry, R.C. Confirmation of Gaseous Methanediol from State-of-the-Art Theoretical Rovibrational Characterization. Phys. Chem. Chem. Phys. 2022, 24, 18552–18558. [Google Scholar] [CrossRef]
  36. Flint, A.R.; Watrous, A.G.; Westbrook, B.R.; Patel, D.J.; Fortenberry, R.C. Gas-phase Formation and Spectroscopic Characterization of the Disubstituted Cyclopropenylidenes c-C3(C2H)2, c-C3(CN)2, and c-C3(C2H)(CN). Astron. Astrophys. 2023, 671, A95. [Google Scholar] [CrossRef]
  37. Watrous, A.G.; Westbrook, B.R.; Fortenberry, R.C. On the Detectability of Diaminomethane ((NH)2CH2). Mon. Not. R. Astron. Soc. 2024, 527, 11090–11094. [Google Scholar] [CrossRef]
  38. Westbrook, B.R.; Fortenberry, R.C. Anharmonic Rotational and Vibrational Spectroscopic Constants of NH2CH2OH. Vib. Spec. 2024, 132, 103690. [Google Scholar] [CrossRef]
  39. Fortenberry, R.C.; Esposito, V.J. Toward the IR Detection of Carbonic Acid: Absorption and Emission Spectra. Astrophys. J. 2024, 961, 164. [Google Scholar] [CrossRef]
  40. Fortenberry, R.C. A Vision for the Future of Astrochemistry in the Interstellar Medium by 2050. ACS Phys. Chem. Au 2023, 4, 31–39. [Google Scholar] [CrossRef]
  41. Fortenberry, R.C. Quantum Chemistry and Astrochemistry: A Match Made in the Heavens. J. Phys. Chem. A 2024, 128, 1555–1565. [Google Scholar] [CrossRef] [PubMed]
  42. Gupta, V.; Sharma, A.; Agrawal, S. Conformations, Chemical Reactivities and Spectroscopic Characteristics of Some Di-substituted Ketenes: An ab initio Study. Bull. Korean Chem. Soc. 2006, 27, 1297–1304. [Google Scholar] [CrossRef]
  43. Allamandola, L.J.; Boersma, C.; Lee, T.J.; Bregman, J.D.; Temi, P. PAH Spectroscopy from 1 to 5 μm. Astrophys. J. 2021, 917, L35. [Google Scholar] [CrossRef]
  44. Boersma, C.; Allamandola, L.J.; Maragkoudakis, A.; Bregman, J.D.; Lee, T.J.; Temi, P.; Esposito, V.J.; Fortenberry, R.C.; Peeters, E. JWST: Deuterated PAHs, PAH-nitriles, and PAH Overtone and Combination Bands I: Program Description and First Look. Astrophys. J. 2023, 959, 74. [Google Scholar] [CrossRef]
  45. Esposito, V.J.; Fortenberry, R.C.; Boersma, C.; Maragkoudakis, A.; Allamandola, L.J. CN Stretches around 4.4 Microns Dominate the IR Absorption Spectra of Cyano-Polycyclic Aromatic Hydrocarbons. Mon. Not. R. Astron. Soc. Lett. 2024, 531, slae037. [Google Scholar] [CrossRef]
  46. Fortenberry, R.C.; Huang, X.; Crawford, T.D.; Lee, T.J. Quartic Force Field Rovibrational Analysis of Protonated Acetylene, C2H3+, and Its Isotopologues. J. Phys. Chem. A 2014, 118, 7034–7043. [Google Scholar] [CrossRef]
  47. Westbrook, B.R.; Fortenberry, R.C. pbqff: Push-Button Quartic Force Fields. J. Chem. Theory Comput. 2023, 19, 2606–2615. [Google Scholar] [CrossRef] [PubMed]
  48. Fortenberry, R.C. Picking Up Good Vibrations through Quartic Force Fields & Vibrational Perturbation Theory. J. Phys. Chem. Lett. 2024, 15, 6528–6537. [Google Scholar] [CrossRef]
  49. Matthews, D.A.; Stanton, J.F. Quantitative Analysis of Fermi Resonances by Harmonic Derivatives of Perturbation Theory Corrections. Mol. Phys. 2009, 107, 213–222. [Google Scholar] [CrossRef]
  50. Sanz-Novo, M.; Rivilla, V.M.; Jiménez-Serra, I.; Martín-Pintado, J.; Colzi, L.; Zeng, S.; Megías, A.; López-Gallifa, Á.; Martínez-Henares, A.; Massalkhi, S.; et al. Discovery of the Elusive Carbonic Acid (HOCOOH) in Space. Astrophys. J. 2023, 954, 3. [Google Scholar] [CrossRef]
  51. Shimanouchi, T. Tables of Molecular Vibrational Frequencies, 39th ed.; National Standards Reference Data System: Washington, DC, USA, 1972; Volume 1. [Google Scholar]
  52. Rivilla, V.M.; Sanz-Novo, M.; Jiménez-Serra, I.; Martín-Pintado, J.; Colzi, L.; Zeng, S.; Megías, A.; López-Gallifa, A.; Martínez-Henares, A.; Massalkhi, S.; et al. First Glycine Isomer Detected in the Interstellar Medium: Glycolamide (NH2C(O)CH2OH). Astrophy. J. Lett. 2023, 953, L20. [Google Scholar] [CrossRef]
  53. Raghavachari, K.; Trucks, G.W.; Pople, J.A.; Head-Gordon, M. A Fifth-Order Perturbation Comparison of Electron Correlation Theories. Chem. Phys. Lett. 1989, 157, 479–483. [Google Scholar] [CrossRef]
  54. Dunning, T.H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  55. Peterson, K.A.; Adler, T.B.; Werner, H.J. Systematically Convergent Basis Sets for Explicitly Correlated Wavefunctions: The Atoms H, He, B-Ne, and Al-Ar. J. Chem. Phys. 2008, 128, 084102. [Google Scholar] [CrossRef] [PubMed]
  56. Hill, J.G.; Peterson, K.A. Correlation Consistent Basis Sets for Explicitly Correlated Wavefunctions: Valence and Core-Valence Basis Sets for Li, Be, Na, and Mg. Phys. Chem. Chem. Phys. 2010, 12, 10460–10468. [Google Scholar] [CrossRef]
  57. Werner, H.J.; Knowles, P.J.; Knizia, G.; Manby, F.R.; Schütz, M.; Celani, P.; Györffy, W.; Kats, D.; Korona, T.; Lindh, R.; et al. MOLPRO, Version 2019.2, a Package of Ab Initio Programs. 2019. Available online: http://www.molpro.net (accessed on 10 June 2020).
  58. Werner, H.J.; Knowles, P.J.; Knizia, G.; Manby, F.R.; Schütz, M. Molpro: A General-Purpose Quantum Chemistry Program Package. WIREs Comput. Mol. Sci. 2012, 2, 242–253. [Google Scholar] [CrossRef]
  59. Westbrook, B.R.; Fortenberry, R.C. High throughput anharmonic vibrational and rotational spectral computations. Ann. Rep. Comput. Chem. 2023, 19, 65–85. [Google Scholar] [CrossRef]
  60. Martin, J.M.L.; Taylor, P.R. Accurate ab Initio Quartic Force Field for trans-HNNH and Treatment of Resonance Polyads. Spectrochim. Acta A 1997, 53, 1039–1050. [Google Scholar] [CrossRef]
  61. Mills, I.M. Vibration-Rotation Structure in Asymmetric- and Symmetric-Top Molecules. In Molecular Spectroscopy—Modern Research; Rao, K.N., Mathews, C.W., Eds.; Academic Press: New York, NY, USA, 1972; pp. 115–140. [Google Scholar]
  62. Watson, J.K.G. Aspects of Quartic and Sextic Centrifugal Effects on Rotational Energy Levels. In Vibrational Spectra and Structure; During, J.R., Ed.; Elsevier: Amsterdam, The Netherlands, 1977; pp. 1–89. [Google Scholar]
  63. Papousek, D.; Aliev, M.R. Molecular Vibration-Rotation Spectra; Elsevier: Amsterdam, The Netherlands, 1982. [Google Scholar]
  64. Gaw, J.F.; Willets, A.; Green, W.H.; Handy, N.C. SPECTRO: A Program for the Derivation of Spectrscopic Constants From Provided Quartic Force Fields and Cubic Dipole Fields. In Advances in Molecular Vibrations and Collision Dynamics; Bowman, J.M., Ratner, M.A., Eds.; JAI Press, Inc.: Greenwich, CT, USA, 1991; pp. 170–185. [Google Scholar]
  65. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  66. Yang, W.T.; Parr, R.G.; Lee, C.T. Various Functionals for the Kinetic Energy Density of an Atom or Molecule. Phys. Rev. A 1986, 34, 4586–4590. [Google Scholar] [CrossRef]
  67. Lee, C.; Yang, W.T.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. By 1988, 37, 785–789. [Google Scholar] [CrossRef]
  68. Hehre, W.J.; Ditchfeld, R.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257. [Google Scholar] [CrossRef]
  69. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; 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]
  70. Huang, X.; Fortenberry, R.C.; Wang, Y.; Francisco, J.S.; Crawford, T.D.; Bowman, J.M.; Lee, T.J. Dipole Surface and Infrared Intensities for the cis- and trans-HOCO and DOCO Radicals. J. Phys. Chem. A 2013, 117, 6932–6939. [Google Scholar] [CrossRef] [PubMed]
  71. Finney, B.; Fortenberry, R.C.; Francisco, J.S.; Peterson, K.A. A Spectroscopic Case for SPSi Detection: The Third-Row in a Single Molecule. J. Chem. Phys. 2016, 145, 124311. [Google Scholar] [CrossRef] [PubMed]
  72. Yu, Q.; Bowman, J.M.; Fortenberry, R.C.; Mancini, J.S.; Lee, T.J.; Crawford, T.D.; Klemperer, W.; Francisco, J.S. The Structure, Anharmonic Vibrational Frequencies, and Intensities of NNHNN+. J. Phys. Chem. A 2015, 119, 11623–11631. [Google Scholar] [CrossRef]
Figure 1. Visual depiction for the five tautomers/conformers examined in this work (gray carbon atoms, indigo nitrogen atoms, red oxygen atoms, and white hydrogen atoms. All are of C s symmetry save for Isomer 7b.
Figure 1. Visual depiction for the five tautomers/conformers examined in this work (gray carbon atoms, indigo nitrogen atoms, red oxygen atoms, and white hydrogen atoms. All are of C s symmetry save for Isomer 7b.
Chemistry 07 00140 g001
Figure 2. The F12-TcCR simulated IR spectra for all four conformers of 2-iminoacetaldehyde with the mid IR in inset.
Figure 2. The F12-TcCR simulated IR spectra for all four conformers of 2-iminoacetaldehyde with the mid IR in inset.
Chemistry 07 00140 g002
Table 1. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (I; in km mol−1) of aminoketene isomer 1.
Table 1. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (I; in km mol−1) of aminoketene isomer 1.
ModeSym.Harmonic aAnharmonic aIB3LYP b
1 a 3611.43435.883600.2
2 a 3530.43374.713519.1
3 a 3171.43036.0193159.4
4 a 2192.32143.45982187.4
5 a 1671.21622.0231652.7
6 a 1437.71402.331380.7
7 a 1224.11188.4101208.0
8 a 1182.01154.4101191.9
9 a 1033.81006.731046.8
10 a 821.2748.27706.4
11 a 656.4641.3116667.7
12 a 599.0582.342600.3
13 a 531.5537.929559.4
14 a 254.9249.84207.4
15 a 218.2216.930157.2
a From [4]. b B3LYP/6-311+G(d,p) anharmonic frequencies.
Table 2. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of aminoketene isomer 2.
Table 2. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of aminoketene isomer 2.
ModeSym.HarmonicAnharmonicIB3LYP a
1 a 3625.13442.583584.4
2 a 3537.63382.513511.2
3 a 3210.83079.9203113.8
4 a 2173.32128.45962208.3
5 a 1664.61621.5221660.4
6 a 1392.81392.731420.5
7 a 1213.01160.911205.2
8 a 1212.31193.0101179.8
9 a 1057.31023.831026.8
10 a 757.0753.03770.2
11 a 679.5673.0123650.9
12 a 586.4257.643607.6
13 a 536.1446.826553.8
14 a 210.9219.65254.5
15 a 154.030212.7
a B3LYP/6-311+G(d,p) anharmonic frequencies.
Table 3. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of s y n -(Z)-2-iminoacetaldehyde isomer 7.
Table 3. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of s y n -(Z)-2-iminoacetaldehyde isomer 7.
ModeSym.HarmonicAnharmonicIB3LYP a
1 a 3396.03224.223380.2
2 a 3134.72971.5323092.0
3 a 2979.02844.4762922.8
4 a 1774.01739.7951793.7
5 a 1664.21620.7271679.5
6 a 1437.91408.051424.5
7 a 1395.31365.5131387.7
8 a 1239.51211.8701238.7
9 a 1152.71106.4651157.4
10 a 1031.41004.211041.1
11 a 899.6867.134872.8
12 a 781.3800.317777.9
13 a 678.1643.515676.6
14 a 301.4291.317295.8
15 a 181.156.012183.8
a B3LYP/6-311+G(d,p) anharmonic frequencies.
Table 4. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of s y n -(E)-2-iminoacetaldehyde (isomer 8).
Table 4. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of s y n -(E)-2-iminoacetaldehyde (isomer 8).
ModeSym.HarmonicAnharmonicIB3LYP a
1 a 3471.33315.013450.4
2 a 3061.82932.2613009.3
3 a 2943.12751.1672883.3
4 a 1788.41769.11051811.7
5 a 1677.01666.5211691.3
6 a 1436.21451.9221429.4
7 a 1395.31405.011395.7
8 a 1210.01195.3541206.8
9 a 1115.71112.761127.3
10 a 1026.01049.791039.3
11 a 903.5907.738879.2
12 a 767.8803.745765.5
13 a 636.3647.646636.1
14 a 295.1417.51297.0
15 a 27.6368.3
a B3LYP/6-311+G(d,p) anharmonic frequencies.
Table 5. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of 7b.
Table 5. The F12-TcCR fundamental vibrational frequencies (in cm−1) and B3LYP/6-311+G(d,p) intensities (in km mol−1) of 7b.
ModeSym.HarmonicAnharmonicIB3LYP a
1a3603.93422.4223543.8
2a3433.03221.6163376.9
3a2927.42694.8642843.4
4a1713.71680.3141710.6
5a1692.91697.02321690.2
6a1539.31510.291559.9
7a1393.01377.241381.6
8a1217.61191.5101212.5
9a1010.2991.424999.7
10a959.3957.245952.0
11a829.4828.031820.6
12a796.4787.1119810.3
13a437.2484.929429.0
14a381.6425.733367.6
15a115.2189.126147.7
a B3LYP/6-311+G(d,p) anharmonic frequencies.
Table 6. The F12-TcCR rotational constants (in MHz) for the aminoketene tautomers and conformers.
Table 6. The F12-TcCR rotational constants (in MHz) for the aminoketene tautomers and conformers.
Vib.Isomer 2Isomer 7Isomer 8Isomer 7b
StateABCABCABCABC
040,860.24004689.51854296.148823,932.18316477.07605095.920226,021.08906052.52494959.061930,124.09055543.82815033.6713
140,742.07684687.79474293.387723,902.57866480.60465096.855525,996.86876048.96044955.962130,231.07755526.74895025.5499
240,711.05734687.42004293.786523,927.44646459.76905085.388526,015.47396037.44544949.411629,892.28805569.25355043.2286
341,287.95694665.78094281.563923,941.47676456.92705084.387226,027.21076039.04634950.958530,830.16475444.30304992.2429
440,756.45494665.96984275.355223,828.94366473.92225089.288825,905.83686046.95184951.246329,980.11525545.19215027.4955
540,745.30264688.30444296.547623,810.97106471.15515086.719525,880.90006047.23964949.633429,884.17865564.59475037.2058
640,431.94154696.20394293.807523,937.63336480.26585089.933326,005.40386062.99074957.530029,695.05455573.31275042.2903
744,587.55074706.14504295.849023,913.30226492.54835096.019125,993.40916067.21184959.430730,165.64185539.05245032.1063
836,554.22204687.68684300.516824,130.99056475.40925090.496926,417.49266062.71794959.724530,232.51055536.96885026.0655
941,224.95264667.61574279.654323,690.07676472.94795094.211425,548.61916049.22724959.847430,320.69745508.55455016.7450
1040,985.81414692.53744299.689423,888.17366469.60225097.632025,976.32106049.34114961.772030,049.23835529.69295018.8225
1141,357.42494684.52404291.607023,943.31446443.67915070.728626,046.78126019.30794934.685830,275.13805509.28305014.0019
1240,815.12134676.36064292.893123,973.54556472.18945087.669925,996.77586052.90274953.881529,980.81975541.99045033.3175
1340,704.00324690.23504300.684723,884.28836462.42225094.472225,995.34886040.77914958.117630,613.51975500.17235017.3656
1441,346.36554710.29714304.597024,117.64376443.91005073.531726,189.13466050.65124949.306730,520.80895481.77705012.8986
1541,652.42064668.87484281.141224,030.75496435.34795093.395926,721.19735889.03014999.662830,150.81705610.83775055.6820
Table 7. The Watson A-reduced F12-TcCR quartic and sextic spectroscopic constants (in MHz and B3LYP/6-311+G(d,p) dipole moments (in D) for the aminoketene isomers.
Table 7. The Watson A-reduced F12-TcCR quartic and sextic spectroscopic constants (in MHz and B3LYP/6-311+G(d,p) dipole moments (in D) for the aminoketene isomers.
ConstantIsomer 2Isomer 7Isomer 8Isomer 7b
Δ J 0.00310.00570.00510.0144
Δ K 2.20240.09560.11421.5656
Δ J K −0.1079−0.02890.0023−0.2345
δ J 0.00070.00160.00140.0045
δ k 0.00180.01160.0126−0.1116
Φ J 1.280 × 10 8 1.405 × 10 10 −8.193 × 10 7 2.717 × 10 7
Φ K 3.822 × 10 4 2.745 × 10 6 −8.010 × 10 4 2.611 × 10 4
Φ J K −2.131 × 10 7 1.658 × 10 7 −8.178 × 10 6 −1.426× 10 6
Φ K J −1.161 × 10 5 −1.329 × 10 6 −6.598 × 10 5 −2.837 × 10 5
ϕ J 4.536 × 10 9 7.207 × 10 7 −4.098 × 10 7 1.253 × 10 7
ϕ J K 4.116 × 10 7 −2.816 × 10 9 −1.622 × 10 5 1.448 × 10 6
ϕ k 1.113 × 10 5 7.939 × 10 7 1.302 × 10 5 1.232 × 10 4
μ 2.652.224.224.83
μ x −1.412.180.00−0.64
μ y 0.000.00−3.562.12
μ z −2.250.40−2.26−4.29
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Inostroza-Pino, N.; McKissick, M.; Lattanzi, V.; Caselli, P.; Fortenberry, R.C. Fundamental Vibrational Frequencies and Spectroscopic Constants for Additional Tautomers and Conformers of NH2CHCO. Chemistry 2025, 7, 140. https://doi.org/10.3390/chemistry7050140

AMA Style

Inostroza-Pino N, McKissick M, Lattanzi V, Caselli P, Fortenberry RC. Fundamental Vibrational Frequencies and Spectroscopic Constants for Additional Tautomers and Conformers of NH2CHCO. Chemistry. 2025; 7(5):140. https://doi.org/10.3390/chemistry7050140

Chicago/Turabian Style

Inostroza-Pino, Natalia, Megan McKissick, Valerio Lattanzi, Paola Caselli, and Ryan C. Fortenberry. 2025. "Fundamental Vibrational Frequencies and Spectroscopic Constants for Additional Tautomers and Conformers of NH2CHCO" Chemistry 7, no. 5: 140. https://doi.org/10.3390/chemistry7050140

APA Style

Inostroza-Pino, N., McKissick, M., Lattanzi, V., Caselli, P., & Fortenberry, R. C. (2025). Fundamental Vibrational Frequencies and Spectroscopic Constants for Additional Tautomers and Conformers of NH2CHCO. Chemistry, 7(5), 140. https://doi.org/10.3390/chemistry7050140

Article Metrics

Back to TopTop