Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 03 September 2024
Sec. Chemical Physics and Physical Chemistry
This article is part of the Research Topic Frontiers in Chemistry: 10 Years Anniversary View all 11 articles

Toward the identification of cyano-astroCOMs via vibrational features: benzonitrile as a test case

  • International Centre for Quantum and Molecular Structures, Department of Physics, College of Sciences, Shanghai University, Shanghai, China

The James Webb Space Telescope (JWST) opened a new era for the identification of molecular systems in the interstellar medium (ISM) by vibrational features. One group of molecules of increasing interest is cyano-derivatives of aromatic organic molecules, which have already been identified in the ISM on the basis of the analysis of rotational signatures, and so, are plausible candidates for the detection by the JWST. Benzonitrile considered in this work represents a suitable example for the validation of a computational strategy, which can be further applied for different, larger, and not-yet observed molecules. For this purpose, anharmonic simulations of infrared (IR) spectra have been compared with recent FTIR experimental studies. The anharmonic computations using the generalized second-order vibrational perturbation theory (GVPT2) in conjunction with a hybrid force field combining the harmonic part of revDSD-PBEP86-D3/jun-cc-pVTZ with anharmonic corrections from B3LYP-D3/SNSD show very good agreement with those in the experiment, with a mean error of 11cm1 for all fundamental transitions overall and only 2cm1 for the CN stretching fundamental at 4.49 μm. The inclusion of overtones up to three-quanta transitions also allowed the prediction of spectra in the near-infrared region, which shows distinct features due to CN overtones at the 2.26 μm and 1.52 μm. The remarkable accuracy of the GVPT2 results opens a pathway for the reliable prediction of spectra for a broader range of cyano-astroCOMs.

1 Introduction

Despite being initially considered too harsh and diluted for molecule creation and survival, it is now recognized that several chemical processes can happen in astronomic environments leading to more than 310 molecules already detected in the interstellar medium (ISM) or circumstellar shells (Endres et al., 2016). Several of these molecules have been detected in the last decade, in particular, a number of complex organic species (COMs), defined as astrophysically relevant organic molecules consisting of six or more atoms. Some COMs that contain C, H, N, O, and possibly also S atoms can be claimed as prebiotic (Fulvio et al., 2021). To date, more than 30 prebiotic molecules have been detected in Taurus molecular cloud 1 (TMC-1), a dust-enshrouded gaseous cloud located 400 light-years from the Sun in the Taurus constellation. Such progress and fast increase in new detections has become possible due to advances in instrumentation, in particular the sub-millimeter and radio domains, allowing the analysis of the lowest rotational lines where radiation can pass through dust-enshrouded clouds (Guélin and Cernicharo, 2022).

However, rotational transitions are not always suitable and accessible for a study, for instance, for exoplanet atmospheres or dense dark regions, for which vibrational spectroscopies are often the methods of choice. The increasing importance of vibrational transitions in astrochemical context is clearly represented by the James Webb Space Telescope (JWST), which, with the infrared (IR) (Frost et al., 2022) observations obtained by the mid-infrared (MIR) and near-IR (NIR) instruments integrated within the Integrated Science Instrument Module (ISIM, Near-Infrared Camera [NIRCam], Near-Infrared Spectrograph [NIRSpec], Mid-Infrared Instrument [MIRI], and Jet Propulsion Laboratory [JPL]), already provides superlative sensitivity, spectral resolution, and wavelength coverage compared to previous space telescopes, such as Herschel, that observe in the visible and ultraviolet spectra (see Figure 1). These emerging experimental data already provide new information about the major ices in molecular cloud cores just prior to their collapse to form protostars (McClure et al., 2023). Currently, the Mars 2020 Perseverance rover also searches for signs of organic matter, in the contexts of the emergence of life, as well as its consideration as a habitable planet, performing, among others, NIR studies using SuperCam (Perez et al., 2017; Eigenbrode et al., 2018; Lasne, 2021; Sharma et al., 2023; McIntosh et al., 2024). These missions yield a huge amount of new significant results, which need to be analyzed, also highlighting the urgent need for accurate reference data in the MIR-to-NIR (0.6–28 μm) range (Öberg, 2016; Nazari et al., 2021; Zapata Trujillo et al., 2023; Fortenberry, 2024b).

Figure 1
www.frontiersin.org

Figure 1. Cyano-derivatives in the ISM and the year of detection by rotational (left panel) or vibrational (right panel) features. Left panel: astro- COMs and Herschel’s view of TMC-1 (Herschel Space Observatory, 2017) (credit: ESA/Herschel/NASA/JPL-Caltech TMC-1). Right panel: OCN recently detected by the JWST and NGC 604 NIRCam Image (Webb Space Telescope, 2024) (credit: NASA/ESA/CSA/STScI).

Spectroscopic techniques are the key for the analysis of astronomical observations and the detection of molecules in the interstellar medium and other astrochemical environments, such as atmospheres or soils of exoplanets or planetary moons (Loru et al., 2022; Lemmens et al., 2022; Mackie et al., 2022; Peeters et al., 2021; K. Lemmens et al., 2023; Petrignani and Candian, 2022; Puzzarini et al., 2023). However, the investigation of the chemical composition of “astrochemical samples” is complicated due to the concomitant presence of many, possibly unknown species. The detection of molecules is based on the comparison of spectra from spectroscopic observations in space with reference experiments from the laboratory (Puzzarini et al., 2023). Unfortunately, the latter might be limited, incomplete, or difficult to obtain under appropriate conditions (Barone et al., 2015c; Fortenberry, 2024b). The optimal strategy is represented by the combination of experiments with the theoretical approaches (Barone et al., 2021; Puzzarini, 2022), supporting and/or complementing laboratory studies (Biczysko et al., 2018a; Barone and Puzzarini, 2023; Fortenberry, 2024a). An increasing role of computational spectroscopy is related to its increasing accuracy (Yang et al., 2021; Barone and Puzzarini, 2023), as well as the possibility to match different, often extreme environmental conditions possible in space (Barone et al., 2015c; Biczysko et al., 2018b; Zapata Trujillo et al., 2023; Fortenberry, 2024b).

Among all the molecules detected to date in the ISM, over 30% bear a nitrogen atom, which usually bonds to carbon, in a large fraction (over 80% of species) by the triple CN bond (Endres et al., 2016). This is in line with the increased dipole moment that facilitates detection by rotational features, as well as the abundance of the CN radical in the ISM (McGuire, 2022) and its extreme reactivity (McGuire et al., 2018; 2021) leading to the abundance of cyano-substituted derivatives (cyano-astroCOMs and CN-astroCOMs) in the ISM. Indeed, benzonitrile (BCN, cyano-benzene, C6H5CN), was the first six-membered aromatic compound detected in 2018 in the ISM (McGuire et al., 2018), toward the dark molecular cloud TMC-1, a well-studied region where most molecules were first observed. This detection was possible as the CN group attachment creates a permanent dipole moment, which is null in otherwise “silent” (poly-)aromatic hydrocarbons (PAHs). This breakthrough intensified the search for aromatic CN-astroCOMs, leading to the detection in the TMC-1 of 1- and 2-cyanonaphthalene (C10H7CN) isomers (McGuire et al., 2021) in 2021 and 2-cyanoindene (C9H7CN) (Sita et al., 2022) in 2022. Following the observation of the six-membered rings, 1- and 2-cyanocyclopentadiene (C5H5CN) (McCarthy et al., 2021; Lee et al., 2021), which contains a five-membered cycle, were also detected in the same source (see Figure 1). Further searches extending to other hetero-aromatic and bridged bicyclic compounds (Martin-Drumel et al., 2023) also aim at establishing the abundance ratio between CN-PAHs and their parent PAHs (Barone and Lazzari, 2023). To date, none of these molecules have been identified by their vibrational features, but the first cyano-compound, i.e., the OCN- anion, which is one of the most well-detected species in astrophysical ices, has been detected in the low-mass star-forming region ChameleonI by the JWST based on the peak centered at 2165.9cm1 (4.62 μm) (McClure et al., 2023) (see Figure 1).

In this work, we focus on the first observed cyano-astroCOM, i.e., benzonitrile (McGuire et al., 2018). The experimental microwave spectrum of benzonitrile was first studied simultaneously by Erlandsson (1954) and Lide (1954). Bak et al. conducted comprehensive analyses of benzonitrile utilizing infrared and centimeter-wave techniques, also including feasible mono-substituted isotopologs (Bak and Nielsen, 1960; Bak et al., 1962). This led to the first derivation of a substitution structure, rs, which was further refined by Casado et al. by incorporating multiple Q-branch transitions for the majority of mono-substituted isotopologs (Casado et al., 1971). Green and Harrison subsequently enhanced the analysis of the experimental IR spectrum, taking into account the lowest wavenumber modes (Green and Harrison, 1976). Recently, these initial spectroscopic investigations have been improved, also in view of the increasing accuracy and spectrum coverage requirements due to the astrochemical importance of BCN (Yamamoto et al., 2000; Kwon et al., 2003; Burova and Anashkin, 2007; Rajasekhar et al., 2022). The accurate equilibrium structure of benzonitrile has been determined by Rudolph et al. (2013) using two different, complementary techniques, namely, the theoretical estimate rBO and semi-experimental reSE. The most extended study is the 2022 synchrotron investigation of BCN in a very broad spectrum range, up to 90000cm1 (11.1 eV), which also included the new detection of the gas-phase IR spectra at a resolution of 0.5cm1 (Rajasekhar et al., 2022). This work, along with the high-resolution far-infrared spectra collected in the 65695cm1 range using synchrotron radiation at the SOLEIL facility (Zdanovskaia et al., 2022), represents the reference for the anharmonic computations of IR spectra.

This work focuses on the IR spectra in the MIR-NIR region, first comparing the simulated infrared spectra with the available experimental counterparts and then providing the prediction of not yet available spectral data, which can support either laboratory or astrochemical studies.

2 Computational details

In order to simulate the spectroscopic parameters of benzonitrile in its electronic ground state X̃1A1, the geometry optimization and harmonic and anharmonic vibrational computations are performed using GAUSSIAN 16 (Frisch et al., 2016). In geometry optimization, the tight convergence criteria (maximum forces and displacements smaller than 1.5×105 Hartree/Bohr and 6.0×105Å, respectively), as required for the anharmonic computations, are used. The equilibrium structure, harmonic force constants, and first-order electric dipole moment derivatives have been computed using the double-hybrid density functional revDSD-PBEP86 (Santra et al., 2019), which has been recommended for spectroscopic studies of medium-sized biomolecules (Barone et al., 2020; Yang et al., 2021; Mehta et al., 2023). These computations have been performed in conjunction with the jun-cc-pVTZ (denoted hereafter as junTZ) basis set (Papajak et al., 2011), which provides the optimal accuracy/cost ratio, as recently discussed by Xu et al. (2024). Moreover, the B3LYP (Becke, 1993)/SNSD (Barone et al., 2014) level has been used in the anharmonic computations. For both density functional theory (DFT) functionals, the dispersion correction proposed by Grimme (2011) has been added using the D3 (Grimme et al., 2010) version with Becke–Johnson (BJ) damping (Grimme et al., 2011; Najibi and Goerigk, 2018). For brevity, hereafter, the revDSD-PBE86-D3/jun-cc-pVTZ and B3LYP-D3/SNSD levels will be denoted as revDSD and B3LYP.

Computations of the third- and fourth-order derivatives of the potential energy surface have been performed at the B3LYP level by the numerical differentiation (Barone, 2005; Bloino, 2015) of analytic second-order derivatives, while the cubic electric dipole moment surfaces have been obtained through numerical differentiations of the dipole moment derivatives. The revDSD equilibrium and harmonic computations have been combined with B3LYP anharmonic computations to create a hybrid model used in spectroscopic simulations. The consistency of these two sets of data has been checked automatically, as implemented in GAUSSIAN 16. In other words, the overlap between two sets of normal modes (two different levels of theory) is defined using the linear transformation, as proposed by Duschinsky (1937):

Q=JQ+K,

where Q and Q represent the two sets of mass-weighted normal coordinates. The Duschinsky matrix J describes the projection of normal coordinate basis vectors on those of the other, allowing the automatic check of the normal mode consistency between the two levels of theory used to define the hybrid method. To ensure that the two sets of normal modes computed at different levels of theory are equivalent, a 90% cut-off for each coordinate was required.

This cost-effective (Xu et al., 2024) hybrid revDSD/B3LYP scheme has been further used to compute spectroscopic parameters using the second-order vibrational perturbation theory (VPT2) (Nielsen, 1951; Mills, 1972). The ground vibrational state rotational constants have been obtained from the revDSD equilibrium structure by adding vibrational corrections computed at the revDSD/B3LYP level, which also provided data allowing the determination of the quartic and sextic centrifugal-distortion constants (Puzzarini et al., 2010; Puzzarini, 2013). For the vibrational spectra, it is also necessary to account for the possible presence of anharmonic resonances (Amos et al., 1991; Martin et al., 1995; Barone, 2005; Vázquez and Stanton, 2007; Rosnik and Polik, 2014; Barone et al., 2014; Bloino et al., 2015; Bloino, 2015; Krasnoshchekov et al., 2015; Franke et al., 2021; Mendolicchio et al., 2021; Franke et al., 2021) by the generalized VPT2 (GVPT2) model (Bloino and Barone, 2012; Bloino et al., 2015), where nearly resonant contributions are removed from the perturbative treatment (leading to the deperturbed model, DVPT2) and treated in a second step variationally. Resonance definition and general recommendations on the applied computational procedures are described in detail in the tutorial review by Bloino et al. (2016). It should be noted that, although improved criteria to define automatic resonances have been proposed recently, they would have negligible impacts on the energies (Yang and Bloino, 2022). Overall, the GVPT2 scheme employed in this work has been successfully applied to medium-sized or larger biomolecules with up to 100 atoms (Fornaro et al., 2015; Fusè et al., 2019; Yatsyna et al., 2019; Green and Improta, 2020; Yang et al., 2021; Sheng et al., 2021), also in the astrochemical context (Zhao et al., 2021; McIntosh et al., 2024; Alberini et al., 2024), so it is a valuable tool to be employed also for significantly larger cyano-astroCOMs.

3 Results and discussion

3.1 Equilibrium structure and rotational parameters

Selected equilibrium structural parameters of benzonitrile calculated at the revDSD/junTZ level are shown in Figure 2, while Table 1 compares all equilibrium parameters with the semi-experimental equilibrium structure reSE (Pulay et al., 1978) derived in the reference (Rudolph et al., 2013) by combining the experimental ground-state rotational constants for a set of isotopologs with rovibrational corrections derived from cubic force fields determined at the B3LYP level (Piccardo et al., 2015). Moreover, Table 1 reports the best estimated theoretical structure rBO obtained using the composite scheme employing all-electron CCSD(T) and MP2 geometry optimizations, with basis sets up to the quintuple-zeta, reaching this way the complete basis set (CBS) limit, as well as computations at the CCSD(T)/ANO1 level (Rajasekhar et al., 2022). Structural parameters computed by all combinations of revDSD and B3LYP functionals with the junTZ and SNSD basis sets are provided in Supplementary Material, along with the Cartesian coordinates by revDSD/junTZ. All DFT structures agree very well with the reSE reference, among which revDSD/junTZ shows the smallest mean absolute errors (MAEs) of approximately 0.0030 for bond lengths and 0.08° for the angles, which, in terms of the largest discrepancies, correspond to approximately 0.005 and ± 0.15°, respectively. The good quality structure with MAEs of 0.0045 and 0.13°, respectively, is also observed for B3LYP/SNSD, justifying its application in the hybrid scheme. Interestingly, the revDSD/junTZ structure is closer to the reSE reference than the CCSD(T)/ANO1 structure, further proving the reliability of revDSD as a cost-effective computational model, allowing to derive accurate geometrical parameters (Ceselin et al., 2021; Barone and Lazzari, 2023).

Figure 2
www.frontiersin.org

Figure 2. Structural parameter of benzonitrile as computed at the rDSD/junTZ level, with bond lengths (black) in Å and angles (blue) in degrees.

Table 1
www.frontiersin.org

Table 1. Equilibrium parameters of benzonitrile (bond lengths in Å; angles in degrees).

The final validation is provided by the direct comparison with the experiment, i.e., spectroscopic constants from Watson’s asymmetric rotor Hamiltonian (A-reduction, Ir representation), which are given in Table 2. Interestingly, the vibrational ground-state rotational constants obtained based on the revDSD structures with anharmonic corrections computed at the revDSD/B3LYP level agree with experiment within 0.4%, which is even better than the CCSD(T)/ANO1 obtained by Zdanovskaia et al. (2022). Moreover, good agreement, again similar to the CCSD(T)/ANO1 results, is also obtained for the quartic and sextic centrifugal-distortion constants. Notably, in some cases, such as the ΦJ sextic constant, very good agreement with the most recent global fit including previous and expanded mm-wave measurements (Zdanovskaia et al., 2022) is obtained, while older experimental data reported values smaller by 50%.

Table 2
www.frontiersin.org

Table 2. Rotational spectroscopic constants for the ground vibrational state of benzonitrile.

3.2 Vibrational properties and IR spectra

Table 3 compares harmonic vibrational wavenumbers and IR intensities with those computed at the CCSD(T)/ANO1 level, showing very good agreement with the average error of approximately 6 cm1 and largest differences of approximately 18cm1, as well as qualitative agreement for IR intensities, with a MAE below 1 km/mol, and largest discrepancies of approximately ± 7 km/mol observed for the most intense bands ν18 and ν19. Thus, the good accuracy of revDSD harmonic wavenumbers is also demonstrated for benzonitrile, in accordance with what has been observed based on the comparison with CCSD(T) results with the CBS extrapolation from MP2 computations (Pietropolli Charmet et al., 2022; Tasinato et al., 2022; Xu et al., 2024). In Table 3 and following, we have adapted mode numbering, mode description, and Wilson notation, as done by Rajasekhar et al. (2022).

Table 3
www.frontiersin.org

Table 3. Harmonic wavenumbers (ω,cm1) and IR intensities (km/mol) compared with reference computed data.

Table 4 lists all fundamental anharmonic wavenumbers and IR intensities of benzonitrile computed at the rDSD/junTZ//B3PLYP/SNSD GVPT2 level, while selected overtones and combination bands are given in Table 5. The accuracy of the simulated IR spectra of benzonitrile in the 500 cm1–4,000 cm1 range can be assessed by comparing with experimental results recorded in the gas phase (Kwon et al., 2003; Rajasekhar et al., 2022). Table 4 provides both spectra, with the latter showing a higher resolution of 0.5cm1. This increased resolution allows us to identify and assign non-fundamental transitions, as shown in Figure 3. The GVPT2 computations show overall good agreement with the experiment with a MAE of approximately 11cm1 and the largest positive and negative errors of approximately ± 44cm1. Moreover, the largest errors are all related to the C-H stretching vibrations, which contribute to the broad band with some additional side peaks, which has not been assigned. Considering that our simulation agrees within 12cm1 with the most intense peak assigned as ν2, it could be expected that further analysis of experimental data, including non-fundamental transitions, would lead to some re-assignments. Excluding all C-H stretching vibrations from the statistics leads to the average errors of 8cm1 and maximum discrepancies within 31cm1. The most important result is extremely good agreement, within 2cm1, for the νCN, the fingerprint vibration of benzonitrile, which is predicted at 2227cm1 (4.49 μm). This result can be compared with the very recent study where the B3LYP/N07D quadratic force field was combined with VPT2 computations with resonances included (see Esposito et al. (2024) for details) but performed with the SPECTRO code, yielding νCN of 2298cm1. Indeed, GVPT2 B3LYP-D3/N07D computations in GAUSSIAN 16 lead to a similar result, with νCN of 2305cm1. Such a huge discrepancy of approximately 70cm1 was not expected based on previous benchmark tests, highlighting the need for a dedicated validation. We hope that the proposed GVPT2 revDSD/B3LYP methodology will allow us to distinguish between the different cyano-astroCOMs observed in the ISM using the νCN vibrations normally occurring in the broader region 2,200–2,400 cm1 (Császár and Fogarasi, 1989).

Table 4
www.frontiersin.org

Table 4. Fundamental wavenumbers (υ,cm1) and IR intensities (km/mol) computed at the GVPT2//revDSD/junTZ//B3LYP/SNSD level for benzonitrile compared with reference experimental and computed data.

Table 5
www.frontiersin.org

Table 5. Non-fundamental band wavenumbers (υ,cm1) and IR intensities (km/mol) computed at the GVPT2//revDSD/junTZ//B3LYP/SNSD level for benzonitrile compared with reference experimental data.

Figure 3
www.frontiersin.org

Figure 3. IR spectrum of benzonitrile in the range 500–4,000 cm1. Computed stick spectra were broadened by Lorentzian functions with half-width at half-maximum (HWHM) of 2 cm1. The experimental IR gas-phase spectrum of benzonitrile (Rajasekhar et al., 2022) is shown for comparison.

A direct comparison between the spectra given in Figure 3 highlights that the GVPT2 computation not only correctly predicts fundamental bands but also a pattern of five distinct non-fundamental bands in the 16501980cm1 range, allowing to correct their assignment with respect to the tentative one reported by Rajasekhar et al. (2022). These non-fundamental bands are reported in Table 5, together with those observed in the far-infrared spectra in a high-resolution FTIR experiment (Zdanovskaia et al., 2022). The overestimated intensity of the ν18 band at the 758cm1 should be noted (computed as 746cm1), while the experimental spectra show two similar intensity peaks in this range, the second one being ν19 at 686cm1. However, this discrepancy needs to be linked to the harmonic values, which already predict the intensity of ν18 as twice that of ν19. Notably, CCSD(T)/ANO-1 harmonic IR intensities yield the same pattern of these two bands as revDSD/junT. In order to provide more information about this discrepancy, a dedicated benchmark analysis, which would require appropriate numerical data on integrated intensities Charmet et al. (2013), not available at present, would be required.

Overall, the good accuracy of our simulations, for both fundamental and non-fundamental transitions, allows us to predict the spectra in the NIR region, which is shown in Figures 4A and B for 40006500cm1 and 650010000cm1, respectively (the whole 1009000cm1 spectrum is also reported in Supplementary Material). The most pronounced bands in 40404300cm1 and 60006180cm1 are related to the combinations of νCH with in-plane ring deformations and 2νCH overtones, respectively. Similarly, at the higher energies, there are combinations of νCH two quanta transitions with in-plane ring deformations (70707350cm1) and 3νCH second overtones (90009200cm1). Although it is expected that the accuracy of GVPT2 results decreases for the higher-quanta transitions, the error bars for the first and second overtones can be estimated based on the fundamental bands (Barone et al., 2015a). For instance, in the case of formaldehyde (Biczysko et al., 2018a), good accuracy within 1cm1 has been obtained for fundamentals and first overtones of ν6 and ν2, while a lower accuracy of ν4 of approximately 10cm1 transfers to errors of 20cm1 and 26cm1 for 2ν4 and 3ν4, respectively. This allows us to provide a reliable prediction of νCN overtones, with distinct 2νCN transition at 4,426±2cm1 (2.26 μm) and a significantly weaker second overtone 3νCN at 6,598±5cm1 (1.52 μm).

Figure 4
www.frontiersin.org

Figure 4. NIR spectra simulated at the GVPT2//rDSD/junTZ//B3LYP/SNSD level. The spectra were broadened by Lorentzian functions with HWHM of 2 cm1.

4 Conclusion and astrochemical implications

Good accuracy of simulated MIR spectra, confirmed by a comparison with available experimental results, allowed us to provide predictions regarding the “missing” data on relevance for the astrochemical observations, in particular, concerning the NIR region.

The availability of NIR references is important for the interpretation of data collected during the Mars 2020 (Williford et al., 2018) space mission by instruments such as SuperCam, for incoming ExoMars 2022 (ESZ-Roscosmos) (Vago et al., 2017), as well as for the JWST observations by the NIRSpec (McClure et al., 2023). The advantage of NIR is its lower spectrum congestion than MIR, which is also clearly visible in Figures 3, 4. However, these reference NIR data are often very scarce and were not available even for benzonitrile prior to this work. It has been already highlighted that anharmonic computations provide significant support in the analysis of experimental results and identification of plausible molecules by NIR features (Fornaro et al., 2020; Alberini et al., 2024).

Computational spectroscopy can also support the identification of BCN in other spectral regions, also considering effects due to the interactions with cosmic rays (Öberg, 2016; Arumainayagam et al., 2021), relevant for different astrochemical environments, from the ISM to planetary atmospheres or soil. These interactions can initiate different processes within molecules, depending on the photon energies, and can lead to electronic excitation within neutral molecules or create ions by ejecting off the valence or even inner-layer electrons. Interactions with photons can lead to the creation of new molecules, or their damage, but are also relevant for extending observable spectral ranges (Öberg, 2016).

Extensive laboratory experimental investigation of the photoabsorption spectra of benzonitrile recorded using synchrotron radiation in 3500090000cm1 (4.3–11.1 eV, 0.111–0.286 μm), which encompasses several neutral and ionic excited states, as recently reported by Rajasekhar et al. (2022). From a computational perspective, these processes can be simulated by means of vibronic computations (Bloino et al., 2016; Barone et al., 2021), which have been demonstrated to allow us to decipher a broad range of energies by the combination of two-state electronic transitions for a series of halogenated benzene (Palmer et al., 2015a; b). Moreover, first-principle spectral simulations also allow us to obtain reference data for unstable species difficult to study in the laboratories and to improve the resolution and predict spectra at a broad range of temperatures (Zhao et al., 2021). Computational spectroscopy studies combining anharmonic vibrational and vibronic simulations covering the broad range from MIR at approximately 20 μm up to even 20 nm in a high-energy photoelectron range can be extended toward other cyano-astroCOMs, supporting their possible detection.

The most relevant are those based on aromatic systems, such as 1- and 2-cyanocyclopentadiene (McCarthy et al., 2021; Lee et al., 2021), and 1- and 2-cyanonaphthalene (McGuire et al., 2021), which have been already discovered in TMC-1. However, it can be expected that similar accuracy can also be obtained for aliphatic systems, for which several conformers can be present (Barone et al., 2013; 2015b). This situation was highlighted by the recent discovery of five cyano-derivatives of propene (CH2CHCH3), based on the QUIJOTE line survey of TMC-1 (Cernicharo et al., 2022). Such computations, including electronic spectra for benzonitrile and MIR to PES spectra for other CN-astroCOM species, are deferred to subsequent works, within the framework of development of the COSY-ASTRO dataset (COSY COST Action CA21101, 2024).

Data availability statement

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

Author contributions

YX: formal analysis, investigation, validation, writing–original draft, writing–review and editing, data curation, and visualization. MB: formal analysis, investigation, validation, writing–original draft, writing–review and editing, conceptualization, funding acquisition, methodology, project administration, resources, and supervision.

Funding

The author(s) declare that financial support was received for the research, authorship, and/or publication of this article. This work was performed within COST Action CA21101 “Confined molecular systems: from a new generation of materials to the stars” (COSY) supported by COST (European Cooperation in Science and Technology). This work was supported by the National Natural Science Foundation of China (Grant No. 31870738).

Acknowledgments

MB acknowledges support from the COST Action “COSY,” in particular, discussions with Action Chair Prof. María Pilar de Lara-Castells, Action Vice-Chair Prof. Cristina Puzzarini, and WG5 “Confined Systems in Astrochemistry” Leader Prof. Lauri Halonen. MB and XY thank Prof. Julien Bloino for fruitful discussions on anharmonic computations.

Conflict of interest

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

The author(s) declared that they were an editorial board member of Frontiers, at the time of submission. This had no impact on the peer review process and the final decision.

Publisher’s note

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

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fchem.2024.1439194/full#supplementary-material

References

Alberini, A., Fornaro, T., García-Florentino, C., Biczysko, M., Poblacion, I., Aramendia, J., et al. (2024). Investigating the stability of aromatic carboxylic acids in hydrated magnesium sulfate under uv irradiation to assist detection of organics on mars. Sci. Rep. 14, 15945. doi:10.1038/s41598-024-66669-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Amos, R. D., Handy, N. C., Green, W. H., Jayatilaka, D., Willetts, A., and Palmieri, P. (1991). Anharmonic vibrational properties of ch2f2: a comparison of theory and experiment. J. Chem. Phys. 95, 8323–8336. doi:10.1063/1.461259

CrossRef Full Text | Google Scholar

Arumainayagam, C. R., Herbst, E., Heays, A. N., Mullikin, E., Farrah, M., and Mavros, M. G. (2021). “Extraterrestrial photochemistry: principles and applications,” in Prebiotic photochemistry: from urey–miller-like experiments to recent findings. Editors F. Saija, and G. Cassone (Cambridge, UK: The Royal Society of Chemistry), 9–36. doi:10.1039/9781839164354-00009

CrossRef Full Text | Google Scholar

Bak, B., Christensen, D., Dixon, W. B., Hansen-Nygaard, L., and Rastrup-Andersen, J. (1962). Benzene ring distortion by one substituent. Microwave determination of the complete structure of benzonitrile. J. Chem. Phys. 37, 2027–2031. doi:10.1063/1.1733422

CrossRef Full Text | Google Scholar

Bak, Bø., and Nielsen, J. T. (1960). Infrared absorption spectra of benzonitrile and its monodeuterated derivatives: tentative assignment of fundamentals. Z. Elektrochem. Ber. Bunsenges. Phys. Chem. 64, 560–562. doi:10.1002/bbpc.196000003

CrossRef Full Text | Google Scholar

Barone, V. (2005). Anharmonic vibrational properties by a fully automated second-order perturbative approach. J. Chem. Phys. 122, 014108. doi:10.1063/1.1824881

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., Alessandrini, S., Biczysko, M., Cheeseman, J. R., Clary, D. C., McCoy, A. B., et al. (2021). Computational molecular spectroscopy. Nat. Rev. Methods Prim. 1, 38–27. doi:10.1038/s43586-021-00034-1

CrossRef Full Text | Google Scholar

Barone, V., Biczysko, M., and Bloino, J. (2014). Fully anharmonic ir and Raman spectra of medium-size molecular systems: accuracy and interpretation. Phys. Chem. Chem. Phys. 16, 1759–1787. doi:10.1039/C3CP53413H

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., Biczysko, M., Bloino, J., Cimino, P., Penocchio, E., and Puzzarini, C. (2015a). 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. 11, 4342–4363. doi:10.1021/acs.jctc.5b00580

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., Biczysko, M., Bloino, J., Cimino, P., Penocchio, E., and Puzzarini, C. (2015b). 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. 11, 4342–4363. doi:10.1021/acs.jctc.5b00580

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., Biczysko, M., Bloino, J., and Puzzarini, C. (2013). Glycine conformers: a never-ending story? Phys. Chem. Chem. Phys. 15, 1358–1363. doi:10.1039/C2CP43884D

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., Biczysko, M., and Puzzarini, C. (2015c). Quantum chemistry meets spectroscopy for astrochemistry: increasing complexity toward prebiotic molecules. Acc. Chem. Res. 48, 1413–1422. doi:10.1021/ar5003285

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., Ceselin, G., Fusè, M., and Tasinato, N. (2020). Accuracy meets interpretability for computational spectroscopy by means of hybrid and double-hybrid functionals. Front. Chem. 8, 584203. doi:10.3389/fchem.2020.584203

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., and Lazzari, F. (2023). Hunting for complex organic molecules in the interstellar medium: the role of accurate low-cost theoretical geometries and rotational constants. J. Phys. Chem. A 127, 10517–10527. doi:10.1021/acs.jpca.3c06649

PubMed Abstract | CrossRef Full Text | Google Scholar

Barone, V., and Puzzarini, C. (2023). Gas-phase computational spectroscopy: the challenge of the molecular bricks of life. Annu. Rev. Phys. Chem. 74, 29–52. doi:10.1146/annurev-physchem-082720-103845

PubMed Abstract | CrossRef Full Text | Google Scholar

Becke, A. D. (1993). Density-functional thermochemistry. iii. the role of exact exchange. J. Chem. Phys. 98, 5648–5652. doi:10.1063/1.464913

CrossRef Full Text | Google Scholar

Biczysko, M., Bloino, J., and Puzzarini, C. (2018a). Computational challenges in astrochemistry. WIREs Comput. Mol. Sci. 8, e1349. doi:10.1002/wcms.1349

CrossRef Full Text | Google Scholar

Biczysko, M., Krupa, J., and Wierzejewska, M. (2018b). Theoretical studies of atmospheric molecular complexes interacting with NIR to UV light. Faraday Discuss. 212, 421–441. doi:10.1039/C8FD00094H

PubMed Abstract | CrossRef Full Text | Google Scholar

Bloino, J. (2015). A vpt2 route to near-infrared spectroscopy: the role of mechanical and electrical anharmonicity. J. Phys. Chem. A 119, 5269–5287. doi:10.1021/jp509985u

PubMed Abstract | CrossRef Full Text | Google Scholar

Bloino, J., Baiardi, A., and Biczysko, M. (2016). Aiming at an accurate prediction of vibrational and electronic spectra for medium-to-large molecules: an overview. Int. J. Quantum Chem. 116, 1543–1574. doi:10.1002/qua.25188

CrossRef Full Text | Google Scholar

Bloino, J., and Barone, V. (2012). A second-order perturbation theory route to vibrational averages and transition properties of molecules: general formulation and application to infrared and vibrational circular dichroism spectroscopies. J. Chem. Phys. 136, 124108. doi:10.1063/1.3695210

PubMed Abstract | CrossRef Full Text | Google Scholar

Bloino, J., Biczysko, M., and Barone, V. (2015). Anharmonic effects on vibrational spectra intensities: infrared, Raman, vibrational circular dichroism and Raman optical activity. J. Phys. Chem. A 119, 11862–11874. doi:10.1021/acs.jpca.5b10067

PubMed Abstract | CrossRef Full Text | Google Scholar

Burova, T. G., and Anashkin, A. A. (2007). Quantum-mechanical calculation of the intensity distribution in the resonance Raman spectrum of benzonitrile. Opt. Spectrosc. 102, 825–828. doi:10.1134/S0030400X07060045

CrossRef Full Text | Google Scholar

Casado, J., Nygaard, L., and Sørensen, G. O. (1971). Microwave spectra of isotopic benzonitriles. Refined molecular structure of benzonitrile. J. Molec. Struct. 8, 211–224. doi:10.1016/0022-2860(71)80056-X

CrossRef Full Text | Google Scholar

Cernicharo, J., Fuentetaja, R., Cabezas, C., Agúndez, M., Marcelino, N., Tercero, B., et al. (2022). Discovery of five cyano derivatives of propene with the QUIJOTE line survey. Astron. Astrophys. 663, L5. doi:10.1051/0004-6361/202244255

CrossRef Full Text | Google Scholar

Ceselin, G., Barone, V., and Tasinato, N. (2021). Accurate biomolecular structures by the nano-LEGO approach: pick the bricks and build your geometry. J. Chem. Theory Comput. 17, 7290–7311. doi:10.1021/acs.jctc.1c00788

PubMed Abstract | CrossRef Full Text | Google Scholar

Charmet, A. P., Stoppa, P., Tasinato, N., Giorgianni, S., Barone, V., Biczysko, M., et al. (2013). An integrated experimental and quantum-chemical investigation on the vibrational spectra of chlorofluoromethane. J. Chem. Phys. 139, 164302. doi:10.1063/1.4825380

PubMed Abstract | CrossRef Full Text | Google Scholar

COSY COST Action CA21101 (2024). COSY Confined molecular systems: from a new generation of materials to the stars CA21101 COST Action.

Google Scholar

Császár, A. G., and Fogarasi, G. (1989). Scaled quantum mechanical (SQM) force field and theoretical vibrational spectrum for benzonitrile. Spectrochim. Acta A 45, 845–854. doi:10.1016/0584-8539(89)80222-3

CrossRef Full Text | Google Scholar

[Dataset] Herschel Space Observatory (2017). Herschel’s view of the Taurus molecular cloud annotated. Available at: https://www.herschel.caltech.edu/image/nhsc2017-003a.

Google Scholar

Duschinsky, F. (1937). The importance of the electron spectrum in multi atomic molecules. concerning the franck-condon principle. Acta Physicochim. URSS 7, 551.

Google Scholar

Eigenbrode, J. L., Summons, R. E., Steele, A., Freissinet, C., Millan, M., Navarro-González, R., et al. (2018). Organic matter preserved in 3-billion-year-old mudstones at Gale crater, Mars. Science 360, 1096–1101. doi:10.1126/science.aas9185

PubMed Abstract | CrossRef Full Text | Google Scholar

Endres, C. P., Schlemmer, S., Schilke, P., Stutzki, J., and Müller, H. S. P. (2016). The cologne database for molecular spectroscopy, CDMS, in the virtual atomic and molecular data centre, VAMDC. J. Mol. Spectrosc. 327, 95–104. doi:10.1016/j.jms.2016.03.005

CrossRef Full Text | Google Scholar

Erlandsson, G. (1954). Microwave spectrum of benzonitrile. J. Chem. Phys. 22, 1152. doi:10.1063/1.1740313

CrossRef Full Text | Google Scholar

Esposito, V. J., Fortenberry, R. C., Boersma, C., Maragkoudakis, A., and Allamandola, L. J. (2024). CN stretches around 4.4 microns dominate the IR absorption spectra of cyano-polycyclic aromatic hydrocarbons. Mon. Not. R. Astron. Soc. Lett. 531, L87–L93. doi:10.1093/mnrasl/slae037

CrossRef Full Text | Google Scholar

Fornaro, T., Brucato, J. R., Poggiali, G., Corazzi, M. A., Biczysko, M., Jaber, M., et al. (2020). UV irradiation and near infrared characterization of laboratory mars soil analog samples. Front. Astron. Space Sci. 7. doi:10.3389/fspas.2020.539289

CrossRef Full Text | Google Scholar

Fornaro, T., Carnimeo, I., and Biczysko, M. (2015). Toward feasible and comprehensive computational protocol for simulation of the spectroscopic properties of large molecular systems: the anharmonic infrared spectrum of uracil in the solid state by the reduced dimensionality/hybrid vpt2 approach. J. Phys. Chem. A 119, 5313–5326. doi:10.1021/jp510101y

PubMed Abstract | CrossRef Full Text | Google Scholar

Fortenberry, R. C. (2024a). Quantum chemistry and astrochemistry: a match made in the heavens. J. Phys. Chem. A 128, 1555–1565. doi:10.1021/acs.jpca.3c07601

PubMed Abstract | CrossRef Full Text | Google Scholar

Fortenberry, R. C. (2024b). A vision for the future of astrochemistry in the interstellar medium by 2050. ACS Phys. Chem. Au 4, 31–39. doi:10.1021/acsphyschemau.3c00043

PubMed Abstract | CrossRef Full Text | Google Scholar

Franke, P. R., Stanton, J. F., and Douberly, G. E. (2021). 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 125, 1301–1324. doi:10.1021/acs.jpca.0c09526

PubMed Abstract | CrossRef Full Text | Google Scholar

Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., et al. (2016). Gaussian 16 rev. C.01. Wallingford CT: Gaussian, Inc.

Google Scholar

Frost, A. J., Lau, R. M., Burtscher, L., Packham, C., Tasker, E. J., Rees, G. A., et al. (2022). IR 2022: an infrared-bright future for ground-based IR observatories in the era of JWST. Nat. Astron. 6, 772–773. doi:10.1038/s41550-022-01733-9

CrossRef Full Text | Google Scholar

Fulvio, D., Potapov, A., He, J., and Henning, T. (2021). Astrochemical pathways to complex organic and prebiotic molecules: experimental perspectives for in situ solid-state studies. Life 11, 568. doi:10.3390/life11060568

PubMed Abstract | CrossRef Full Text | Google Scholar

Fusè, M., Mazzeo, G., Longhi, G., Abbate, S., Masi, M., Evidente, A., et al. (2019). Unbiased determination of absolute configurations by vis-à-vis comparison of experimental and simulated spectra: the challenging case of diplopyrone. J. Phys. Chem. B 123, 9230–9237. PMID: 31580674. doi:10.1021/acs.jpcb.9b08375

PubMed Abstract | CrossRef Full Text | Google Scholar

Green, J. A., and Improta, R. (2020). Vibrations of the guanine–cytosine pair in chloroform: an anharmonic computational study. Phys. Chem. Chem. Phys. 22, 5509–5522. doi:10.1039/C9CP06373K

PubMed Abstract | CrossRef Full Text | Google Scholar

Green, J. H. S., and Harrison, D. J. (1976). Vibrational spectra of benzene derivatives—XVII. Benzonitrile and substituted benzonitriles. Spectrochim. Acta A 32, 1279–1286. doi:10.1016/0584-8539(76)80166-3

CrossRef Full Text | Google Scholar

Grimme, S. (2011). Density functional theory with london dispersion corrections. WIREs Comput. Mol. Sci. 1, 211–228. doi:10.1002/wcms.30

CrossRef Full Text | Google Scholar

Grimme, S., Antony, J., Ehrlich, S., and Krieg, H. (2010). A consistent and accurate ab initio parametrization of density functional dispersion correction (dft-d) for the 94 elements h-pu. J. Chem. Phys. 132, 154104. doi:10.1063/1.3382344

PubMed Abstract | CrossRef Full Text | Google Scholar

Grimme, S., Ehrlich, S., and Goerigk, L. (2011). Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 32, 1456–1465. doi:10.1002/jcc.21759

PubMed Abstract | CrossRef Full Text | Google Scholar

Guélin, M., and Cernicharo, J. (2022). Organic molecules in interstellar space: latest advances. Front. Astron. Space Sci. 9. doi:10.3389/fspas.2022.787567

CrossRef Full Text | Google Scholar

Krasnoshchekov, S. V., Vogt, N., and Stepanov, N. F. (2015). Ab initio anharmonic analysis of vibrational spectra of uracil using the numerical-analytic implementation of operator Van Vleck perturbation theory. J. Phys. Chem. A 119, 6723–6737. doi:10.1021/acs.jpca.5b03241

PubMed Abstract | CrossRef Full Text | Google Scholar

Kwon, C. H., Kim, H. L., and Kim, M. S. (2003). Vibrational analysis of vacuum ultraviolet mass-analyzed threshold ionization spectra of phenylacetylene and benzonitrile. J. Phys. Chem. A 107, 10969–10975. doi:10.1021/jp036391w

CrossRef Full Text | Google Scholar

Lasne, J. (2021). Heterogeneous physical chemistry in the atmospheres of earth, mars, and venus: perspectives for rocky exoplanets. ACS Earth Space Chem. 5, 149–162. doi:10.1021/acsearthspacechem.0c00126

CrossRef Full Text | Google Scholar

Lee, K. L. K., Changala, P. B., Loomis, R. A., Burkhardt, A. M., Xue, C., Cordiner, M. A., et al. (2021). Interstellar detection of 2-cyanocyclopentadiene, C5H5CN, a second five-membered ring toward TMC-1. Astrophys. J. Lett. 910, L2. doi:10.3847/2041-8213/abe764

CrossRef Full Text | Google Scholar

Lemmens, A. K., Rap, D. B., Brünken, S., Buma, W. J., and Rijs, A. M. (2022). Polycyclic aromatic hydrocarbon growth in a benzene discharge explored by IR-UV action spectroscopy. Phys. Chem. Chem. Phys. 24, 14816–14824. doi:10.1039/D2CP01631A

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemmens, K., Mackie, C., Candian, A., Lee, J., Tielens, M., Anouk, M. R., et al. (2023). Size distribution of polycyclic aromatic hydrocarbons in space: an old new light on the 11.2/3.3 m intensity ratio. Faraday Discuss. 245, 380–390. doi:10.1039/D2FD00180B

PubMed Abstract | CrossRef Full Text | Google Scholar

Lide, D. R. (1954). Microwave spectrum and structure of benzonitrile. J. Chem. Phys. 22, 1577–1578. doi:10.1063/1.1740461

CrossRef Full Text | Google Scholar

Loru, D., Steber, A. L., Thunnissen, J. M. M., Rap, D. B., Lemmens, A. K., Rijs, A. M., et al. (2022). New potential candidates for astronomical searches discovered in the electrical discharge of the PAH naphthalene and acetonitrile. J. Mol. Spectrosc. 386, 111629. doi:10.1016/j.jms.2022.111629

CrossRef Full Text | Google Scholar

Mackie, C. J., Candian, A., Lee, T. J., and Tielens, A. G. G. M. (2022). Anharmonicity and the IR emission spectrum of neutral interstellar PAH molecules. J. Phys. Chem. A 126, 3198–3209. doi:10.1021/acs.jpca.2c01849

PubMed Abstract | CrossRef Full Text | Google Scholar

Martin, J. M. L., Lee, T. J., Taylor, P. M., and François, J.-P. (1995). The anharmonic force field of ethylene, c2h4, by means of accurate ab initio calculations. J. Chem. Phys. 103, 2589–2602. doi:10.1063/1.469681

CrossRef Full Text | Google Scholar

Martin-Drumel, M.-A., Spaniol, J.-T., Hölzel, H., Agúndez, M., Cernicharo, J., Moth-Poulsen, K., et al. (2023). Searches for bridged bicyclic molecules in space—norbornadiene and its cyano derivatives. Faraday Discuss. 245, 284–297. doi:10.1039/D3FD00016H

PubMed Abstract | CrossRef Full Text | Google Scholar

McCarthy, M. C., Lee, K. L. K., Loomis, R. A., Burkhardt, A. M., Shingledecker, C. N., Charnley, S. B., et al. (2021). Interstellar detection of the highly polar five-membered ring cyanocyclopentadiene. Nat. Astron. 5, 176–180. doi:10.1038/s41550-020-01213-y

CrossRef Full Text | Google Scholar

McClure, M. K., Rocha, W. R. M., Pontoppidan, K. M., Crouzet, N., Chu, L. E. U., Dartois, E., et al. (2023). An Ice Age JWST inventory of dense molecular cloud ices. Nat. Astron. 7, 431–443. doi:10.1038/s41550-022-01875-w

CrossRef Full Text | Google Scholar

McGuire, B. A. (2022). 2021 census of interstellar, circumstellar, extragalactic, protoplanetary disk, and exoplanetary molecules. Astrophys. J. Suppl. Ser. 259, 30. doi:10.3847/1538-4365/ac2a48

CrossRef Full Text | Google Scholar

McGuire, B. A., Burkhardt, A. M., Kalenskii, S., Shingledecker, C. N., Remijan, A. J., Herbst, E., et al. (2018). Detection of the aromatic molecule benzonitrile (c-C6H5CN) in the interstellar medium. Science 359, 202–205. doi:10.1126/science.aao4890

PubMed Abstract | CrossRef Full Text | Google Scholar

McGuire, B. A., Loomis, R. A., Burkhardt, A. M., Lee, K. L. K., Shingledecker, C. N., Charnley, S. B., et al. (2021). Detection of two interstellar polycyclic aromatic hydrocarbons via spectral matched filtering. Science 371, 1265–1269. doi:10.1126/science.abb7535

PubMed Abstract | CrossRef Full Text | Google Scholar

McIntosh, O., García-Florentino, C., Fornaro, T., Marabello, D., Alberini, A., Siljeström, S., et al. (2024). Undecanoic acid and L-phenylalanine in vermiculite: detection, characterization, and UV degradation studies for biosignature identification on mars. Astrobiology 24, 518–537. doi:10.1089/ast.2023.0088

PubMed Abstract | CrossRef Full Text | Google Scholar

Mehta, N., Santra, G., and Martin, J. M. (2023). Is explicitly correlated double-hybrid density functional theory advantageous for vibrational frequencies? Can. J. Chem. 101, 656–663. doi:10.1139/cjc-2022-0277

CrossRef Full Text | Google Scholar

Mendolicchio, M., Bloino, J., and Barone, V. (2021). General perturb-then-diagonalize model for the vibrational frequencies and intensities of molecules belonging to abelian and non-abelian symmetry groups. J. Chem. Theory Comput. 17, 4332–4358. doi:10.1021/acs.jctc.1c00240

PubMed Abstract | CrossRef Full Text | Google Scholar

Mills, I. M. (1972). Molecular spectroscopy: modern research (academic press, New York), chap. Vibration-rotation structure in asymmetric- and symmetric-top molecules, 115–140.

CrossRef Full Text | Google Scholar

Najibi, A., and Goerigk, L. (2018). The nonlocal kernel in van der waals density functionals as an additive correction: An extensive analysis with special emphasis on the B97M-v and B97M-V approaches. J. Chem. Theory Comput. 14, 5725–5738. doi:10.1021/acs.jctc.8b00842

PubMed Abstract | CrossRef Full Text | Google Scholar

Nazari, P., van Gelder, M. L., van Dishoeck, E. F., Tabone, B., van ’t Hoff, M. L. R., Ligterink, N. F. W., et al. (2021). Complex organic molecules in low-mass protostars on Solar System scales - II. Nitrogen-bearing species. Astron. Astrophys. 650, A150. doi:10.1051/0004-6361/202039996

CrossRef Full Text | Google Scholar

Nielsen, H. H. (1951). The vibration-rotation energies of molecules. Rev. Mod. Phys. 23, 90–136. doi:10.1103/RevModPhys.23.90

CrossRef Full Text | Google Scholar

Öberg, K. I. (2016). Photochemistry and astrochemistry: photochemical pathways to interstellar complex organic molecules. Chem. Rev. 116, 9631–9663. doi:10.1021/acs.chemrev.5b00694

PubMed Abstract | CrossRef Full Text | Google Scholar

Palmer, M. H., Ridley, T., Hoffmann, S. V., Jones, N. C., Coreno, M., de Simone, M., et al. (2015a). Interpretation of the vacuum ultraviolet photoabsorption spectrum of iodobenzene by ab initio computations. J. Chem. Phys. 142, 134302. doi:10.1063/1.4916121

PubMed Abstract | CrossRef Full Text | Google Scholar

Palmer, M. H., Ridley, T., Hoffmann, S. V., Jones, N. C., Coreno, M., de Simone, M., et al. (2015b). The ionic states of iodobenzene studied by photoionization and ab initio configuration interaction and DFT computations. J. Chem. Phys. 142, 134301. doi:10.1063/1.4916120

PubMed Abstract | CrossRef Full Text | Google Scholar

Papajak, E., Zheng, J., Xu, X., Leverentz, H. R., and Truhlar, D. G. (2011). Perspectives on basis sets beautiful: seasonal plantings of diffuse basis functions. J. Chem. Theory Comput. 7, 3027–3034. doi:10.1021/ct200106a

PubMed Abstract | CrossRef Full Text | Google Scholar

Peeters, E., Mackie, C., Candian, A., and Tielens, A. G. G. M. (2021). A spectroscopic view on cosmic PAH emission. Acc. Chem. Res. 54, 1921–1933. doi:10.1021/acs.accounts.0c00747

PubMed Abstract | CrossRef Full Text | Google Scholar

Perez, R., Parès, L. P., Newell, R., Robinson, S., Bernardi, P., Réess, J.-M., et al. (2017). The supercam instrument on the NASA Mars 2020 mission: optical design and performance. nternational Conf. Space Opt. — ICSO 2016 (SPIE) 10562, 744–752. doi:10.1117/12.2296230

CrossRef Full Text | Google Scholar

Petrignani, A., and Candian, A. (2022). “Chapter 3 - astrochemistry: ingredients of life in space,” in New Frontiers in astrobiology. Editors R. Thombre, and P. Vaishampayan (Elsevier), 49–66. doi:10.1016/B978-0-12-824162-2.00007-5

CrossRef Full Text | Google Scholar

Piccardo, M., Penocchio, E., Puzzarini, C., Biczysko, M., and Barone, V. (2015). Semi-experimental equilibrium structure determinations by employing b3lyp/snsd anharmonic force fields: validation and application to semirigid organic molecules. J. Phys. Chem. A 119, 2058–2082. doi:10.1021/jp511432m

PubMed Abstract | CrossRef Full Text | Google Scholar

Pietropolli Charmet, A., Ceselin, G., Stoppa, P., and Tasinato, N. (2022). The spectroscopic characterization of halogenated pollutants through the interplay between theory and experiment: application to R1122. Molecules 27, 748. doi:10.3390/molecules27030748

PubMed Abstract | CrossRef Full Text | Google Scholar

Pulay, P., Meyer, W., and Boggs, J. E. (1978). Cubic force constants and equilibrium geometry of methane from Hartree–Fock and correlated wavefunctions. J. Chem. Phys. 68, 5077–5085. doi:10.1063/1.435626

CrossRef Full Text | Google Scholar

Puzzarini, C. (2013). Rotational spectroscopy meets theory. Phys. Chem. Chem. Phys. 15, 6595–6607. doi:10.1039/C3CP44301A

PubMed Abstract | CrossRef Full Text | Google Scholar

Puzzarini, C. (2022). Gas-phase chemistry in the interstellar medium: the role of laboratory astrochemistry. Front. Astron. Space Sci. 8. doi:10.3389/fspas.2021.811342

CrossRef Full Text | Google Scholar

Puzzarini, C., Alessandrini, S., Bizzocchi, L., Melosso, M., and Rivilla, V. M. (2023). From the laboratory to the interstellar medium: a strategy to search for exotic molecules in space. Front. Astron. Space Sci. 10. doi:10.3389/fspas.2023.1211784

CrossRef Full Text | Google Scholar

Puzzarini, C., Stanton, J. F., and Gauss, J. (2010). Quantum-chemical calculation of spectroscopic parameters for rotational spectroscopy. Int. Rev. Phys. Chem. 29, 273–367. doi:10.1080/01442351003643401

CrossRef Full Text | Google Scholar

Rajasekhar, B. N., Dharmarpu, V., Das, A. K., Shastri, A., Veeraiah, A., and Krishnakumar, S. (2022). A spectroscopic study of benzonitrile. J. Quant. Spectrosc. Radiat. Transf. 283, 108159. doi:10.1016/j.jqsrt.2022.108159

CrossRef Full Text | Google Scholar

Rosnik, A. M., and Polik, W. F. (2014). Vpt2+k spectroscopic constants and matrix elements of the transformed vibrational Hamiltonian of a polyatomic molecule with resonances using Van Vleck perturbation theory. Mol. Phys. 112, 261–300. doi:10.1080/00268976.2013.808386

CrossRef Full Text | Google Scholar

Rudolph, H. D., Demaison, J., and Császár, A. G. (2013). Accurate determination of the deformation of the benzene ring upon substitution: equilibrium structures of benzonitrile and phenylacetylene. J. Phys. Chem. A 117, 12969–12982. doi:10.1021/jp408208s

PubMed Abstract | CrossRef Full Text | Google Scholar

Santra, G., Sylvetsky, N., and Martin, J. M. L. (2019). Minimally empirical double-hybrid functionals trained against the GMTKN55 database: revDSD-PBEP86-D4, revDOD-PBE-D4, and DOD-SCAN-D4. J. Phys. Chem. A 123, 5129–5143. doi:10.1021/acs.jpca.9b03157

PubMed Abstract | CrossRef Full Text | Google Scholar

Sharma, S., Roppel, R. D., Murphy, A. E., Beegle, L. W., Bhartia, R., Steele, A., et al. (2023). Diverse organic-mineral associations in Jezero crater, Mars. Nature 619, 724–732. doi:10.1038/s41586-023-06143-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Sheng, M., Silvestrini, F., Biczysko, M., and Puzzarini, C. (2021). Structural and vibrational properties of amino acids from composite schemes and double-hybrid dft: hydrogen bonding in serine as a test case. J. Phys. Chem. A 125, 9099–9114. doi:10.1021/acs.jpca.1c06993

PubMed Abstract | CrossRef Full Text | Google Scholar

Sita, M. L., Changala, P. B., Xue, C., Burkhardt, A. M., Shingledecker, C. N., Lee, K. L. K., et al. (2022). Discovery of interstellar 2-cyanoindene (2-C9H7CN) in GOTHAM observations of TMC-1. Astrophys. J. Lett. 938, L12. doi:10.3847/2041-8213/ac92f4

CrossRef Full Text | Google Scholar

Tasinato, N., Pietropolli Charmet, A., Ceselin, G., Salta, Z., and Stoppa, P. (2022). In vitro and in silico vibrational–rotational spectroscopic characterization of the next-generation refrigerant HFO-1123. J. Phys. Chem. A 126, 5328–5342. doi:10.1021/acs.jpca.2c04680

PubMed Abstract | CrossRef Full Text | Google Scholar

Vago, J. L., Westall, F., Coates, A. J., Jaumann, R., Korablev, O., Ciarletti, V., et al. (2017). Habitability on early mars and the search for biosignatures with the ExoMars rover. Astrobiology 17, 471–510. doi:10.1089/ast.2016.1533

PubMed Abstract | CrossRef Full Text | Google Scholar

Vázquez, J., and Stanton, J. F. (2007). Treatment of fermi resonance effects on transition moments in vibrational perturbation theory. Mol. Phys. 105, 101–109. doi:10.1080/00268970601135784

CrossRef Full Text | Google Scholar

Webb Space Telescope (2024). NGC 604 (NIRCam image). Available at: https://webbtelescope.org/contents/media/images/2024/110/01HQNV4GP6PR6E7ZSJXRRBQQDS.

Google Scholar

Williford, K. H., Farley, K. A., Stack, K. M., Allwood, A. C., Beaty, D., Beegle, L. W., et al. (2018). “Chapter 11 - the NASA mars 2020 rover mission and the search for extraterrestrial life,” in From habitability to life on Mars. Editors N. A. Cabrol, and E. A. Grin (Elsevier), 275–308. doi:10.1016/B978-0-12-809935-3.00010-4

CrossRef Full Text | Google Scholar

Xu, R., Jiang, Z., Yang, Q., Bloino, J., and Biczysko, M. (2024). Harmonic and anharmonic vibrational computations for biomolecular building blocks: benchmarking DFT and basis sets by theoretical and experimental IR spectrum of glycine conformers. J. Comput. Chem. 45, 1846–1869. doi:10.1002/jcc.27377

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamamoto, R., Ishikawa, S., Ebata, T., and Mikami, N. (2000). Vibrational spectra and relaxation of benzonitrile and its clusters using time-resolved stimulated Raman–UV double resonance spectroscopy. J. Raman Spectrosc. 31, 295–304. doi:10.1002/(sici)1097-4555(200004)31:4<295:aid-jrs542>3.0.co;2-y

CrossRef Full Text | Google Scholar

Yang, Q., and Bloino, J. (2022). An effective and automated processing of resonances in vibrational perturbation theory applied to spectroscopy. J. Phys. Chem. A 126, 9276–9302. doi:10.1021/acs.jpca.2c06460

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, Q., Mendolicchio, M., Barone, V., and Bloino, J. (2021). Accuracy and reliability in the simulation of vibrational spectra: a comprehensive benchmark of energies and intensities issuing from generalized vibrational perturbation theory to second order (GVPT2). Front. Astron. Space Sci. 8. doi:10.3389/fspas.2021.665232

CrossRef Full Text | Google Scholar

Yatsyna, V., Mallat, R., Gorn, T., Schmitt, M., Feifel, R., Rijs, A. M., et al. (2019). Competition between folded and extended structures of alanylalanine (ala-ala) in a molecular beam. Phys. Chem. Chem. Phys. 21, 14126–14132. doi:10.1039/C9CP00140A

PubMed Abstract | CrossRef Full Text | Google Scholar

Zapata Trujillo, J. C., Pettyjohn, M. M., and McKemmish, L. K. (2023). High-throughput quantum chemistry: empowering the search for molecular candidates behind unknown spectral signatures in exoplanetary atmospheres. Mon. Not. R. Astron. Soc. 524, 361–376. doi:10.1093/mnras/stad1717

CrossRef Full Text | Google Scholar

Zdanovskaia, M. A., Martin-Drumel, M.-A., Kisiel, Z., Pirali, O., Esselman, B. J., Woods, R. C., et al. (2022). The eight lowest-energy vibrational states of benzonitrile: analysis of Coriolis and Darling-Dennison couplings by millimeter-wave and far-infrared spectroscopy. J. Mol. Spectrosc. 383, 111568. doi:10.1016/j.jms.2021.111568

CrossRef Full Text | Google Scholar

Zhao, Y., Hochlaf, M., and Biczysko, M. (2021). Identification of DNA bases and their cations in astrochemical environments: computational spectroscopy of thymine as a test case. Front. Astron. Space Sci. 8. doi:10.3389/fspas.2021.757007

CrossRef Full Text | Google Scholar

Keywords: vibrational spectra, density functional theory, second-order vibrational perturbation theory, near-infrared, James Webb Space Telescope

Citation: Xu Y and Biczysko M (2024) Toward the identification of cyano-astroCOMs via vibrational features: benzonitrile as a test case. Front. Chem. 12:1439194. doi: 10.3389/fchem.2024.1439194

Received: 27 May 2024; Accepted: 12 August 2024;
Published: 03 September 2024.

Edited by:

Jean-Claude Guillemin, UMR6226 Institut des Sciences Chimiques de Rennes (ISCR), France

Reviewed by:

Barbara Michela Giuliano, Max Planck Institute for Extraterrestrial Physics, Germany
Nicola Tasinato, Normal School of Pisa, Italy

Copyright © 2024 Xu and Biczysko. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Malgorzata Biczysko, malgo.biczysko@gmail.com

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