Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-28T09:26:31.276Z Has data issue: false hasContentIssue false

Searching for biosignatures by their rotational spectrum: global fit and methyl group internal rotation features of dimethylsulphoxide up to 116 GHz

Published online by Cambridge University Press:  12 August 2022

Assimo Maris
Affiliation:
Department of Chemistry ‘Giacomo Ciamician’, Alma Mater Studiorum-University of Bologna, Via Francesco Selmi 2, I-40126 Bologna, Italy Interdepartmental Centre for Industrial Aerospace Research (CIRI Aerospace), Alma Mater Studiorum-University of Bologna, Via Baldassarre Canaccini 12, I-47121 Forlì, Italy
Laura B. Favero
Affiliation:
Istituto per lo Studio dei Materiali Nanostrutturati, Sezione di Bologna, Consiglio Nazionale delle Ricerche (CNR), Via Gobetti 101, I-40129 Bologna, Italy
Wentao Song
Affiliation:
Department of Chemistry ‘Giacomo Ciamician’, Alma Mater Studiorum-University of Bologna, Via Francesco Selmi 2, I-40126 Bologna, Italy
Dingding Lv
Affiliation:
Department of Chemistry ‘Giacomo Ciamician’, Alma Mater Studiorum-University of Bologna, Via Francesco Selmi 2, I-40126 Bologna, Italy
Luca Evangelisti
Affiliation:
Interdepartmental Centre for Industrial Aerospace Research (CIRI Aerospace), Alma Mater Studiorum-University of Bologna, Via Baldassarre Canaccini 12, I-47121 Forlì, Italy Department of Chemistry ‘Giacomo Ciamician’, Alma Mater Studiorum-University of Bologna, Via S. Alberto 163, I-48123 Ravenna, Italy
Sonia Melandri*
Affiliation:
Department of Chemistry ‘Giacomo Ciamician’, Alma Mater Studiorum-University of Bologna, Via Francesco Selmi 2, I-40126 Bologna, Italy Interdepartmental Centre for Industrial Aerospace Research (CIRI Aerospace), Alma Mater Studiorum-University of Bologna, Via Baldassarre Canaccini 12, I-47121 Forlì, Italy
*
Author for correspondence: Sonia Melandri, E-mail: sonia.melandri@unibo.it
Rights & Permissions [Opens in a new window]

Abstract

The identification and quantification of molecules in interstellar space and atmospheres of planets in the solar systems and in exoplanets rely on spectroscopic methods and laboratory work is essential to provide the community with the spectral features needed to analyse cosmological observations. Rotational spectroscopy in particular, with its intrinsic high resolution, allows the unambiguous identification of biomolecular building blocks and biosignature gases which can be correlated with the origin of life or the identification of habitable planets. We report the extension of the measured rotational transition frequencies of dimethylsulphoxide and its 34S and 13C isotopologues in the millimetre wave range (59.6–78.4 GHz) by use of an absorption spectrometer based on the supersonic expansion technique. Hyperfine patterns related to the methyl group internal rotation were analysed in the microwave range region (6–18 GHz) with a Pulsed Jet Fourier Transform spectrometer at extremely high resolution (2 kHz) and reliable predictions up to 116 GHz are provided. The focus on sulphur-bearing molecules is motivated by the fact that sulphur is largely involved in the intra- and inter-molecular hydrogen bonds in proteins and although it is the 10th most abundant element in the known Universe, understanding its chemistry is still a matter of debate. Moreover, sulphur-bearing molecules, in particular dimethylsulphoxide, have been indicated as possible biosignature gases to be monitored in the search of habitable exoplanets.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press

Introduction

The search for gaseous molecules in space is a guiding goal in the astrophysical community. The molecules searched for are of increasing complexity in the pursue of understanding the chemical steps that lead to the development of complex molecules, biomolecular building blocks and ultimately life. Another goal is the search for molecules produced by living organisms, called biosignatures, and accumulated in the atmospheres of planets and exoplanets to assess their habitability.

As regards the detection of biosignatures, the early paradigm for exoplanet life detection is the argument that only life can maintain gases severely out of thermodynamic equilibrium in the atmosphere. In particular, the detection of oxygen (O2) and methane (CH4) (Lederberg, Reference Lederberg1965; Lovelock, Reference Lovelock1965) was suggested as the most robust atmospheric evidence of life on Earth.

In general, biosignature gases can be those created directly by metabolic redox reactions or they can be ‘secondary’ metabolism byproducts: gases produced by life but not as byproducts of their central chemical functions. Such gases are usually related to stress or signalling in living beings and although expected to be produced in small quantities, their specificity makes them valuable as possible biosignatures as their detection is not likely to be a false positive (Seager et al., Reference Seager, Schrenk and Bains2012). Some are also produced in sufficient amounts to affect atmospheric chemistry (such as isoprene or dimethyl sulphide) and they could possibly be detected remotely on planets and exoplanets. Examples include nitrous oxide (N2O) (Des Marais et al., Reference Des Marais, Harwit, Jucks, Kasting, Lin, Lunine, Schneider, Seager, Traub and Woolf2002), dimethyldisulphide (CH3SSCH3) (Pilcher, Reference Pilcher2003), methyl chloride (CH3Cl) (Segura et al., Reference Segura, Kasting, Meadows, Cohen, Scalo, Crisp, Butler and Tinetti2005) and dimethyl sulphide (CH3SCH3, DMS) and other sulphur gases (Domagal-Goldman et al., Reference Domagal-Goldman, Meadows, Claire and Kasting2011).

Recently, it has been pointed out that although atmospheric composition will be the primary observable that could imply the presence of life (Seager, Reference Seager2013) the identification of a biosignature in a planet's atmosphere requires an understanding of the possible compositions of abiotic atmospheres (Shahar et al., Reference Shahar, Driscoll, Weinberger and Cody2019).

This last factor is important as the atmosphere is tied to the dynamics of the planet and some secondary metabolism gases are produced by geological processes as well as biological ones.

As an example, we can consider a sulphur-containing molecule, namely dimethylsulphoxide (CH3SOCH3, DMSO, Fig. 1). DMSO is the product of oxidation of DMS, it can be directly produced by some microorganisms from DMS or it can be produced in the atmosphere through a nonbiological chemical reaction which oxidizes DMS to DMSO, dimethyl sulphone (CH3SO2CH3) and finally, methyl sulphate (CH3SO3H) (Seager et al., Reference Seager, Schrenk and Bains2012). In any case the detection and quantification of a molecule like DMSO would be important either as a biosignature itself or in the construction of appropriate models of chemical reactivity in the atmosphere of planets and exoplanets.

Fig. 1. Sketch of the DMSO molecule and principal inertial axis.

The most informative method to investigate the chemical composition of the atmospheres of planets and exoplanets is high-resolution spectroscopy, which for exoplanets must rely on remote-sensing techniques (Burrows, Reference Burrows2014). One of the possible approaches is via rotational spectroscopy in the microwave-to submillimetre range undertaken from the ground with radiotelescopes (see e.g. Greaves et al., Reference Greaves, Richards, Bains, Rimmer, Sagawa, Clements, Seager, Petkowski, Sousa-Silva, Ranjan, Drabek-Maunder, Fraser, Cartwright, Mueller-Wodarg, Zhan, Friberg, Coulson, Lee and Hoge2021). As regards the interstellar space and star-forming regions, the unequivocal identification of the molecules in astronomical surveys is based on their rotational laboratory spectra and more than 200 molecules have been identified in space (CDMS, 2001; Müller et al., Reference Müller, Thorwirth, Roth and Winneswisser2001, Reference Müller, Schlöder, Stutzki and Winnewisser2005). In addition to the characterization of the species, the features of molecular spectra also yield detailed physical information on the gas and its surroundings (Herbst and van Dishoeck, Reference Herbst and van Dishoeck2009).

Laboratory rotational spectroscopy data in the microwave, millimetre and submillimetre wavelengths are necessary to perform those searches and rotational spectroscopy performed in supersonic expansions has been proved to be particularly suitable to investigate complex spectra generated by molecules (Calabrese et al., Reference Calabrese, Gou, Maris, Melandri and Caminati2016) and weakly bound molecular complexes (Velino et al., Reference Velino, Melandri and Caminati2004; Favero et al., Reference Favero, Giuliano, Maris, Melandri, Ottaviani, Velino and Caminati2010). Those experiments allow us to identify multiple conformers of a flexible molecule (Vigorito et al., Reference Vigorito, Calabrese, Paltanin, Melandri and Maris2017a), different tautomers (Sanchez et al., Reference Sanchez, Giuliano, Melandri, Favero and Caminati2007; Melandri et al., Reference Melandri, Evangelisti, Maris, Caminati, Giuliano, Feyer, Prince and Coreno2010) as well as and different isotopologues (Calabrese et al., Reference Calabrese, Vigorito, Feng, Favero, Maris, Melandri, Geppert and Caminati2014). The study of the different isotopologues of a molecule is of importance for different reasons. From a structural point of view, the analysis of the changes in the rotational constants upon isotopic substitution allows the determination of the position of the substituted atoms without any a priori assumptions (Evangelisti et al., Reference Evangelisti, Perez, Seifert, Pate, Dehghany, Moazzen-Ahmadi and McKellar2015). If the collection of a complete set of data including all the isotopes is not possible to achieve, a partial structure can be obtained by refining a structure obtained as a result of quantum chemical calculations to reproduce all experimental data (Uriarte et al., Reference Uriarte, Melandri, Maris, Calabrese and Cocinero2018). From an astrochemical point of view the detected abundance of different isotopologues of a molecular species can give additional information on reaction mechanisms of formation (Neill et al., Reference Neill, Steber, Muckle, Zaleski, Lattanzi, Spezzano, McCarthy, Remijan, Friedel, Widicus Weaver and Pate2011) or on the environment in which the molecule is found or formed (Imai et al., Reference Imai, Sakai, López-Sepulcre, Higuchi, Zhang, Oya, Watanabe, Sakai, Ceccarelli, Lefloch and Yamamoto2018).

We concentrate here on DMSO, which as stated above, can be considered an important secondary metabolic product or the product of an abiogenic mechanism occurring in the atmosphere or in geological processes. Also, DMSO is an important substance commonly used as a polar aprotic solvent, antifreeze fluid and cryo-protectant owing to its beneficial properties including low toxicity and environmental compatibility (Akkök et al., Reference Akkök, Liseth, Hervig, Ryningen, Bruserud and Ersvær2009). In this paper, we report our investigation on DMSO: the extension of the analysis of the rotational spectra of the parent species and that of the mono-substituted 13C and 34S isotopologues to the 59.6–78.4 GHz frequency range observed with a Free Jet Absorption Millimetre Wave (FJ-AMMW) spectrometer and the measurements of some transitions in the lower frequency region (6–18 GHz) with a high-resolution cavity-based Fourier transform microwave (FTMW) spectrometer. The extension of the measurement of the rotational spectrum to relatively high frequencies (>40 GHz) is very important, in particular for remote sensing using telescopes, since many of them, including ALMA, work in higher frequency regions. Even if the theoretical model used to analyse the spectrum in the low frequency region is correct, the uncertainties on the parameters reflect on the predictions thus precise experimental frequencies are essential for the identification of molecular transition in astronomical surveys or complex spectra containing the signals of many other molecules.

Experimental methods

Two different spectrometers working in the microwave and millimetre frequency regions have been utilized in this experiment. DMSO was acquired from Sigma-Aldrich (purity >99%) and used without further purification and argon or helium, purchased from SIAD (Società Italiana Acetilene e Derivati), were used as carrier gas.

The rotational spectrum of DMSO was recorded in the millimetre wave region (59.6–78.4 GHz) using a stark-modulated FJ-AMMW spectrometer with a resolution of about 300 kHz and estimated uncertainties for the frequency measurements of about 50 kHz. The main features of the spectrometer have been previously described (Calabrese et al., Reference Calabrese, Maris, Evangelisti, Favero, Melandri and Caminati2013, Reference Calabrese, Vigorito, Maris, Mariotti, Fathi, Geppert and Melandri2015; Vigorito et al., Reference Vigorito, Calabrese, Melandri, Caracciolo, Mariotti, Giannetti, Massardi and Maris2018). In this experiment the sample was maintained at room temperature and a stream of argon (P 0 = 20 kPa) was flowed over it and the mixture was expanded to about P b = 0.5 Pa through a 0.3 mm diameter pinhole nozzle. The rotational temperature of the molecules in the jet has been estimated to be 5–10 K from relative intensity measurements of rotational transitions at different rotational quantum numbers.

The measurement of the rotational spectrum in the 6–18 GHz region was carried out in a COBRA-type, cavity-based, pulsed supersonic-jet FTMW spectrometer (Balle and Flygare, Reference Balle and Flygare1981; Grabow et al., Reference Grabow, Stahl and Dreizler1996) previously described (Caminati et al., Reference Caminati, Evangelisti, Feng, Giuliano, Gou, Melandri and Grabow2016). The sample of DMSO was cooled to 273 K and helium was flowed over it at a stagnation pressure of about 3 bar resulting in a mixture of about 1%. The mixture was then expanded through a solenoid valve (General Valve, Series 9, nozzle diameter 0.5 mm) into a Fabry–Pérot cavity. During the expansion, the molecules can reach quite low-rotational temperatures (a few Kelvins) and the most stable forms can be trapped at their energy minimum when certain conditions are satisfied. To determine the position of the spectral lines, Fourier transformation of the time-domain signal perform 8k data points, recorded at a sampling interval of 100 ns.

Results and discussion

In terms of chemical structure, the rigid molecule of DMSO has idealized C s symmetry and a trigonal pyramidal molecular geometry (Fig. 1). DMSO is characterized by two equivalent methyl rotors and a Cs frame, therefore, its molecular symmetry group is G18WV in Altmann's notation (Smeyers and Niño, Reference Smeyers and Niño1987) and C3 × C3v+ in Dreizler's approach (Dreizler, Reference Dreizler1961). Due to the coupling of the methyl internal rotation to the overall rotation of the molecule, the rotational levels are split into a multiplet of four states labelled according to the irreducible representation in the above groups. Since, as Dreizler states, μ a and μ b transform as the AA2 representation, in the vibronic ground state only four kinds of transitions are allowed: AA1–AA2, AE–AE, (EbA1 + EaA1)–(EbA2 + EaA2) and (EaE + EbE)–(EaE + EbE), with spin statistical weights 16, 8, 8 and 32, respectively. In the following, we will use a simplified notation, referring to the AA1 and AA2 states as AA, to (EbA1 + EaA1) and (EbA2 + EaA2) as EA and to (EaE + EbE) and (EaE + EbE) as EE. A good general graphical representation of the energy levels for a two tops system has been recently published (Dindić and Nguyen, Reference Dindić and Nguyen2021). Due to this splitting of the rotational levels, DMSO exhibits a spectral splitting because of hindered internal rotation of its methyl rotors that leads to a triplets AA, AE (or EA) and EE splitting pattern, where A and E are the methyl torsional state symmetry labels (see e.g. Vigorito et al., Reference Vigorito, Paoloni, Calabrese, Evangelisti, Favero, Melandri and Maris2017b).

The first rotational spectroscopy study on DMSO dates back to 1964 (Dreizler and Dendl, Reference Dreizler and Dendl1964) and several other studies followed focusing on the determination of the structure through the study of isotopologues (Feder et al., Reference Feder, Dreizler, Rudolph and Typke1969; Typke, Reference Typke1978; Kretschmer, Reference Kretschmer1995) and the dynamic effects such as the centrifugal distortion and the methyl internal rotation (Dreizler and Dendl, Reference Dreizler and Dendl1965a, Reference Dreizler and Dendl1965b; Fliege et al., Reference Fliege, Dreizler and Typke1983). The analysis of the high resolution rotational spectrum of the parent species is reported in Fliege et al. (Reference Fliege, Dreizler and Typke1983) who explored the microwave spectrum of DMSO using a Fourier transform spectrometer from 8 to 18 GHz. They report the fit of the new measured lines together with the lines previously measured in the 6–40 GHz frequency range at a lower resolution. The observed transitions involve the dipole moment component along the b principal axis of rotation (μ b-type) and the frequency values used are the weighted means of the internal rotation fine structure components. Also, seven other transitions involving the coupling of the radiation with the c principal axis (μ c-type) were measured by radiofrequency-double resonance with pump frequency adjustment and methyl internal rotation splittings are also given. Measurement of the high frequency (154–660 GHz) transitions of the parent species are reported by Margulès et al. (Reference Margulès, Motiyenko, Alekseev and Demaison2010) for the ground state and by Cuisset et al. (Reference Cuisset, Drumel, Hindle, Mouret and Sadovskií2013) for the five lowest vibrational states.

As regards the parent species, we extended the spectral investigation to the 59.6–78.4 GHz frequency range and measured some transition lines in the 6–18 GHz frequency range using a higher resolution set-up. The AA transition lines reported by Fliege et al. (Reference Fliege, Dreizler and Typke1983) were fitted using a semi-rigid Hamiltonian in the S-reduction and the III l representations (Watson Reference Watson1977), suitable for oblate tops such as DMSO, by using the CALPGM suite of programs (Pickett, Reference Pickett1991) and the predictions were used to search for lines in the higher frequency range. Several μ b R-type lines with rotational quantum numbers J ranging from 4 to 9, and pseudo-quantum numbers K a from 0 to 5 and K c from 0 to 4 were measured. In the same way, starting from the lines measured by Feder et al. (Reference Feder, Dreizler, Rudolph and Typke1969), the transitions for the monosubstituted 34S and 13C isotopologues were observed in natural abundance and measured in the higher frequency range (accuracy 50 kHz). Some lines both for the parent species and its 34S and 13C monosubstituted isotopologues could be also detected with our cavity-based FTMW spectrometer where they could be measured with a higher accuracy (2 kHz). All newly measured AA lines and those reported in the previous works were analysed in a global semi-rigid fit where each set of lines was weighted accordingly to its measured accuracy. The measured frequencies of the parent species are listed in Table 1 and those of the isotopologues are reported in Table 2, whereas the spectroscopic parameters obtained from the fitting procedure are reported in Table 3.

Table 1. Experimental transition frequencies ν and deviations δν = ν(obs) − ν(calc) of internal rotation components of DMSO. Calculated frequencies n(calc) are obtained using the spectroscopic parameters listed in Table 4

a From E. Fliege et al. (Reference Fliege, Dreizler and Typke1983).

Table 2. Experimental transition frequencies of AA lines of DMSO (34S) and DMSO (13C)

Table 3. Spectroscopic parameters for the AA state of DMSO, DMSO (34S) and DMSO (13C) (S-reduction, III l representation)

a Standard error in parenthesis in units of the last digits.

b Constants fixed to zero as they result undetermined.

c HJ is undetermined from the fit and has been fixed to zero.

d Number of transitions.

e Standard deviation of the fit.

While the set of AA lines can be fit by using a semi-rigid Hamiltonian, the full set of AA, AE, EA and EE of transition lines needs a global fitting approach which takes into account the coupling effect of the internal rotation of the methyl group on the rotational transitions. For this purpose, the XIAM program (Hartwig and Dreizler, Reference Hartwig and Dreizler1996) was used. This program is based on the combined axis method (Woods Reference Woods1966) and directly supplies the V 3 barrier to internal rotation, the angles between the internal rotation axis and the principal axes and the moment of inertia of the internal top. A total of 195 transitions (including many of those from Fliege et al., Reference Fliege, Dreizler and Typke1983) were included in the global fit yielding the values of the rotational constants, the quartic and some sextic centrifugal distortion constants and the internal rotation parameters such as the value of the barrier hindering the motion and the orientation of the internal rotors. The standard deviation of the global fit is 5 kHz which is very close to the best accuracy obtained in the different measurements. In the XIAM program the S-reduction can be used but only the I r representation is available and not the III l one which would be more suitable for an oblate rotor such as DMSO. In the two representations the centrifugal distortion constants have different values and this is the reason why, in order to compare the AA-fit and the global-fit, we performed different fittings using the two representations for the AA state of the parent species. The comparison of the corresponding spectroscopic parameters reported in Table 4 shows that they are in very good agreement. Predictions from this fit to the 59.6–78.4 GHz region have confirmed that the internal rotation splittings are not resolvable in this region with the absorption spectrometer (resolution 300 kHz).

Table 4. Spectroscopic parameters for the global fit of DMSO compared to the fitting of AA lines both in I r representation

a Standard error in parenthesis in units of the last digits.

b HJ is undetermined from the fit and has been fixed to zero.

c Fixed to the value obtained from the rz structure (Feder et al., Reference Feder, Dreizler, Rudolph and Typke1969).

d δ  = ∠(ai), where i is the internal rotation axis of the methyl tops. ε (angle between the projection of i on the bc plane and the b axis) fixed to zero as undetermined from the fit.

e Derived parameters.

f Number of measured transitions in the fit.

g Standard deviation of the fit.

The obtained results are in good agreement with those reported by Fliege et al. (Reference Fliege, Dreizler and Typke1983) but the accuracy in the determination of the centrifugal distortion constants is better and so is the standard deviation of the fit which is five times better. The experimental results can also be compared to those obtained by in the optimization of the DMSO structure by quantum chemical methods with the Gaussian 16 program (Gaussian, Inc., Wallingford, CT, USA) with the MP2/aug-cc-pVTZ.

The calculated values of the rotational constants (A = 7000.6, B = 6989.0, C = 4243.1 all in MHz) and of the distortion constants in the S reduction and III l representation (DJ = 6.1, DJK = −9.1, DK = 4.1, d 1 = −0.25, d 2 = −0.23 all in kHz) are also in good agreement with the experimentally determined ones.

The moment of inertia and orientation of the methyl group was deduced from the structure reported in Feder et al. (Reference Feder, Dreizler, Rudolph and Typke1969). While the angle δ  = ∠(ai), between the i internal rotation axis of the methyl tops and the a principal axis of overall rotation could be fitted, the ε angle between the projection of i on the bc plane and the b axis was fixed to zero as it was undetermined in the fit. This results in experimentally derived values of ∠(ai) = 32.56(26)/147.44(26), ∠(bi) = 57.44(26) (°) for the two methyl tops and a fixed ∠(ci) angle of 90°.

As regards the methyl group internal rotation, the MP2/aug-cc-pVTZ potential energy surface of the methyl torsion around the SC bond in DMSO was calculated by changing the dihedral angle of τ = HC − SC on the regular grid with Δτ = 10° and is shown in Fig. 2. The calculated data marked in red bullets are well reproduced with the threefold function: V(τ) = ½V 3[1 + cos(3τ)], which is shown as a green line where V 3 = 11.6 kJ mol−1. This value is also in agreement with the experimentally determined value of 12.37(4) kJ mol−1.

Fig. 2. Ab initio (MP2-aug-cc-pVTZ) methyl internal rotational potential energy surface for DMSO.

The V 3 barrier hindering the methyl torsion in DMSO is thus much higher than the ones found for other molecules which contain a methyl group attached to a sulphur atom such as methylmercaptan (5.31(1) kJ mol−1) (Kojima, Reference Kojima1960), DMS (8.8030(5) kJ mol−1) (Pierce and Hayashi, Reference Pierce and Hayashi1961; Jabri et al., Reference Jabri, Van, Nguyen, Mouhib, Kwabia Tchana, Manceron, Stahl and Kleiner2016) and dimethyl disulphide (6.44 kJ mol−1) (Hartwig et al., Reference Hartwig, Kretschmer and Dreizler1995). This could be due to electronic effects related to the presence of the oxygen atom or to an attractive interaction between the oxygen atoms and the methyl hydrogens which can increase the barrier to internal rotation of this group.

Based on the obtained fittings, an overall prediction in the 0–1000 GHz range calculated at five different temperatures which could be found in different astronomical objects is reported in Fig. 3. While the measurements are performed up to 78.4 GHz, the predictions are quite reliable in the adjacent frequency range covered by band 3 of ALMA (84–116 GHz) for lines with rotational quantum number J less than 12. In the bottom panel of Fig. 3, an expanded view of this frequency range can be found where the frequencies and intensities of main lines calculated at different temperatures are reported. Here, we can see that the most intense transitions are somehow the same ones at different temperatures. Moreover, a thorough analysis of the predictions showed that in this frequency range the internal rotation splittings are units or tens of kilohertz thus, even if very small, they might be observable at the resolution of some astronomical surveys which can show linewidths as low as 10 kHz (see for example Loomis et al., Reference Loomis, Burkhardt, Shingledecker, Charnley, Cordiner, Herbst, Kalenskii, Lee, Willis, Xue, Remijan, McCarthy and McGuire2021). For a quantitative analysis, the predicted frequencies with their uncertainties and intensities at different temperature are reported in Table 5 while the internal rotation components are given as supplementary material (Table S1). The AA predicted frequencies from our fit are accurate within ten kilohertz as assessed by performing some fits which included lines measured at higher frequencies by Margulès et al. (Reference Margulès, Motiyenko, Alekseev and Demaison2010).

Fig. 3. Top panel: Spectral predictions in the 0–1000 GHz frequency range for DMSO at eight different temperatures. Bottom panel: Expanded spectral window overlapping with the frequency region covered by the ALMA telescope (band 3, 84–116 GHz).

Table 5. Selected predicted transitions in the Alma band 3 spectral region (84–116 GHz): predicted frequency (MHz), estimated uncertainty (MHz) and base 10 logarithm of the integrated intensity in units of nm2 MHz at different temperatures

a The intensities were calculated using the partition function from the semirigid approximation. The following values were used: Q(5 K) = 133.3339, Q(10 K) = 374.6724, Q(25) = 1475.3676, Q(50) = 4168.2031, Q(75) = 7655.3187, Q(100) = 11 785.2365, Q(150) = 21 651.4854, Q(300) = 61 250.6707.

Conclusions

In this work, rotational spectroscopy coupled to supersonic expansions has been exploited to characterize the rotational spectrum of DMSO and its 34S and 13C monosubstituted isotopologues. The lines were measured in the 59.6–78.4 GHz frequency range with an absorption spectrometer while a cavity-based FTMW spectrometer has allowed to re-measure DMSO transitions split by the methyl rotor internal rotation in the 6–18 GHz region with unprecedented high resolution and accuracy. A global fit of 195 lines has allowed for an accurate modelling of the internal rotation lines and the accurate determination of the spectroscopic parameters obtained by applying appropriate models. The molecular parameters related to the geometry of the molecule or its internal dynamics (methyl torsional barrier) could also be used to model the ro-vibrational spectrum of DMSO.

The measured rotational lines can be confidently used for the analysis of astronomical surveys for the unequivocal identification of DMSO and its isotopologues in different astronomical environments such as the interstellar medium, star-forming regions and planetary atmospheres. A natural place where one would try is astronomical objects where organic molecules containing sulphur have been detected. For example, in giant molecular clouds like Sagittarius B2 (Müller et al., Reference Müller, Belloche, Xu, Lees, Garrod, Walters, Van Wijngaarden, Lewen, Schlemmer and Menten2016) or protostars (Majumdar et al., Reference Majumdar, Gratier, Vidal, Wakelam, Loison, Hickson and Caux2016). Tentative detection of sulphur-containing molecules has also been attempted in prestellar cores (Maris et al., Reference Maris, Calabrese, Favero, Evangelisti, Usabiaga, Mariotti, Codella, Podio, Balucani, Ceccarelli, LeFloch and Melandri2019) and molecular clouds (Song et al., Reference Song, Maris, Rivilla, Fortuna, Evangelisti, Lv, Rodríguez-Almeida, Jiménez-Serra, Pintado and Melandri2022). Besides remote sensing, the value of rotational spectroscopy as a means of exact identification and quantification of molecules could also be exploited in the search of biosignatures using instruments for in situ analysis.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S1473550422000271.

Acknowledgements

We acknowledge financial support from the 2020AFB3FX_003 PRIN national project and the CINECA award under the ISCRA initiative, for the availability of high-performance computing resources and support. D. L. and W. S. acknowledge the China Scholarship Council (CSC) for financial support. This work was supported by the Italian MIUR (Attività Base di Ricerca) and the University of Bologna (Ricerca Fondamentale Orientata).

Author contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

References

Akkök, ÇA, Liseth, K, Hervig, T, Ryningen, A, Bruserud, ØH and Ersvær, E (2009) Use of different DMSO concentrations for cryopreservation of autologous peripheral blood stem cell grafts does not have any major impact on levels of leukocyte-and platelet-derived soluble mediators. Cytotherapy 11, 749760.CrossRefGoogle Scholar
Balle, TJ and Flygare, WH (1981) Fabry–Perot cavity pulsed Fourier transform microwave spectrometer with a pulsed nozzle particle source. Review of Scientific Instruments 52, 3345.CrossRefGoogle Scholar
Burrows, AS (2014) Spectra as windows into exoplanet atmospheres. Proceedings of the National Academy of Sciences of the United States of America 111, 1260112609.CrossRefGoogle ScholarPubMed
Calabrese, C, Maris, A, Evangelisti, L, Favero, LB, Melandri, S and Caminati, W (2013) Keto-enol tautomerism and conformational landscape of 1,3-cyclohexanedione from its free jet millimeter-wave absorption spectrum. Journal of Physical Chemistry A 117, 1371213718.CrossRefGoogle ScholarPubMed
Calabrese, C, Vigorito, A, Feng, G, Favero, LB, Maris, A, Melandri, S, Geppert, WD and Caminati, W (2014) Laboratory rotational spectrum of acrylic acid and its isotopologues in the 6–18.5 GHz and 52–74.4 GHz frequency ranges. Journal of Molecular Spectroscopy 295, 3743.CrossRefGoogle Scholar
Calabrese, C, Vigorito, A, Maris, A., Mariotti, S, Fathi, P, Geppert, WD and Melandri, S (2015) Millimeter wave spectrum of the weakly bound complex CH2=CHCN⋅H2O: structure, dynamics, and implications for astronomical search. Journal of Physical Chemistry A 119, 1167411682.CrossRefGoogle Scholar
Calabrese, C, Gou, Q, Maris, A, Melandri, S and Caminati, W (2016) Conformational equilibrium and internal dynamics of E-anethole: a rotational study. The Journal of Physical Chemistry B 120, 65876591.CrossRefGoogle ScholarPubMed
Caminati, W, Evangelisti, L, Feng, G, Giuliano, BM, Gou, Q, Melandri, S and Grabow, J-U (2016) On the Cl⋯C halogen bond: a rotational study of CF3Cl–CO. Physical Chemistry Chemical Physics 18, 1785117855.CrossRefGoogle Scholar
CDMS (2001) The Cologne Database for Molecular Spectroscopy (n.d.). Available at www.astro.uni-koeln.de/cdms/molecules.Google Scholar
Cuisset, A, Drumel, M-AM, Hindle, F, Mouret, G and Sadovskií, DA (2013) Rotational structure of the five lowest frequency fundamental vibrational states of dimethylsulfoxide. Chemical Physics Letters 586, 1015.CrossRefGoogle Scholar
Des Marais, DJ, Harwit, MO, Jucks, KW, Kasting, JF, Lin, DNC, Lunine, JI, Schneider, J, Seager, S, Traub, WA and Woolf, NJ (2002) Remote sensing of planetary properties and biosignatures on extrasolar terrestrial planets. Astrobiology 2, 153181.CrossRefGoogle ScholarPubMed
Dindić, C and Nguyen, HVL (2021) Microwave spectrum of two-top molecule: 2-acetyl-3-methylthiophene. Chemphyschem 22.23, 24202428CrossRefGoogle Scholar
Domagal-Goldman, SD, Meadows, VS, Claire, MW and Kasting, JF (2011) Using biogenic sulfur gases as remotely detectable biosignatures on anoxic planets. Astrobiology 11, 419–441.CrossRefGoogle Scholar
Dreizler, H (1961) Gruppentheoretische betrachtungen zu den mikrowellenspektren von molekülen mit zwei behindert drehbaren dreizählig-symmetrischen molekülgruppen. Zeitschrift für Naturforsch A 16, 13541367.CrossRefGoogle Scholar
Dreizler, H and Dendl, G (1964) Rotations spektrum, r 0-struktur und dipolmoment von dimethvlsulfoxyd. Zeitschrift für Naturforsch A 19, 512514.CrossRefGoogle Scholar
Dreizler, H and Dendl, G (1965a) Erfahrungen bei der analyse der zentrifugalaufweitung in rotationsspektren I. Dimethylsulfoxyd. Zeitschrift für Naturforsch A 20, 3037.CrossRefGoogle Scholar
Dreizler, H and Dendl, G (1965b) Bestimmung des Hinderungspotentials der internen Rotation aus dem Mikrowellenspektrum von Dimethylsulfoxyd. Zeitschrift für Naturforsch A 20, 14311440.CrossRefGoogle Scholar
Evangelisti, L, Perez, C, Seifert, NA, Pate, BH, Dehghany, M, Moazzen-Ahmadi, N and McKellar, ARW (2015) Theory vs. experiment for molecular clusters: spectra of OCS trimers and tetramers. The Journal of Chemical Physics 142, 104309.CrossRefGoogle ScholarPubMed
Favero, LB, Giuliano, BM, Maris, A, Melandri, S, Ottaviani, P, Velino, B and Caminati, W (2010) Features of the C-H…N weak hydrogen bond and internal dynamics in pyridine-CHF3. Chemistry (Weinheim an Der Bergstrasse, Germany) 16, 17611764.Google Scholar
Feder, W, Dreizler, H, Rudolph, HD and Typke, V (1969) rs-Struktur von dimethylsulfoxid im vergleich zur r0-struktur. Zeitschrift für Naturforsch A 24, 266278.CrossRefGoogle Scholar
Fliege, E, Dreizler, H and Typke, V (1983) Centrifugal distortion and ground state internal rotation analysis of the microwave spectrum of dimethylsulfoxide. Zeitschrift für Naturforsch A 38, 668675.CrossRefGoogle Scholar
Grabow, JU, Stahl, W and Dreizler, H (1996) A multioctave coaxially oriented beam-resonator arrangement Fourier-transform microwave spectrometer. Review of Scientific Instruments 67, 40724084.CrossRefGoogle Scholar
Greaves, JS, Richards, AMS, Bains, W, Rimmer, PB, Sagawa, H, Clements, DL, Seager, S, Petkowski, JJ, Sousa-Silva, C, Ranjan, S, Drabek-Maunder, E, Fraser, HJ, Cartwright, A, Mueller-Wodarg, I, Zhan, Z, Friberg, P, Coulson, I, Lee, E and Hoge, J (2021) Phosphine gas in the cloud decks of Venus. Nature Astronomy 5, 655664.CrossRefGoogle Scholar
Hartwig, H and Dreizler, H (1996) The microwave spectrum of trans-2,3-dimethyloxirane in torsional excited states. Zeitschrift für Naturforsch 51a, 923932.CrossRefGoogle Scholar
Hartwig, H, Kretschmer, U and Dreizler, H (1995) The 33S nuclear quadrupole hyperfine structure in the rotational Spectrum of 32S, 33S dimethyl disulfide. Zeitschrift für Naturforsch 50 a, 131136.CrossRefGoogle Scholar
Herbst, E and van Dishoeck, EF (2009) Complex organic interstellar molecules. Annual Review of Astronomy and Astrophysics 47, 427480.CrossRefGoogle Scholar
Imai, M, Sakai, N, López-Sepulcre, A, Higuchi, AE, Zhang, Y, Oya, Y, Watanabe, Y, Sakai, T, Ceccarelli, C, Lefloch, B and Yamamoto, S (2018) Deuterium fractionation survey toward protostellar sources in the perseus molecular cloud: HNC case. Astrophysical Journal 869, 51.CrossRefGoogle Scholar
Jabri, A, Van, V, Nguyen, HVL, Mouhib, H, Kwabia Tchana, F, Manceron, L, Stahl, W and Kleiner, I (2016) Laboratory microwave, millimeter wave and far-infrared spectra of dimethyl sulfide. Astronomy & Astrophysics 589, A127.CrossRefGoogle Scholar
Kojima, T (1960) Microwave spectrum of methyl mercaptan. Journal of the Physical Society of Japan 15, 12841291.CrossRefGoogle Scholar
Kretschmer, U (1995) The 33S nuclear quadrupole hyperfine coupling in the rotational spectrum of 33S dimethylsulfoxide. Zeitschrift für Naturforsch 50a, 666668.CrossRefGoogle Scholar
Lederberg, J (1965) Signs of life: criterion-system of exobiology. Nature 207, 913.CrossRefGoogle ScholarPubMed
Loomis, RA, Burkhardt, AM, Shingledecker, CN, Charnley, SB, Cordiner, MA, Herbst, E, Kalenskii, S, Lee, KLK, Willis, ER, Xue, C, Remijan, AJ, McCarthy, MC and McGuire, BA (2021) An investigation of spectral line stacking techniques and application to the detection of HC11N. Nature Astronomy 5, 88–196.CrossRefGoogle Scholar
Lovelock, JE (1965) A physical basis for life detection experiments. Nature 207, 568570.CrossRefGoogle ScholarPubMed
Majumdar, L, Gratier, P, Vidal, T, Wakelam, V, Loison, J-C, Hickson, K and Caux, E (2016) Detection of CH3SH in protostar IRAS 16293-2422. Monthly Notices of the Royal Astronomical Society 458, 185911865.CrossRefGoogle Scholar
Margulès, L, Motiyenko, RA, Alekseev, EA and Demaison, J (2010) Choice of the reduction and of the representation in centrifugal distortion analysis: A case study of dimethylsulfoxide. Journal of Molecular Spectroscopy 260, 2329.CrossRefGoogle Scholar
Maris, A, Calabrese, C, Favero, LB, Evangelisti, L, Usabiaga, I, Mariotti, S, Codella, C, Podio, L, Balucani, N, Ceccarelli, C, LeFloch, B and Melandri, S (2019) Laboratory measurements and astronomical search for thioacetamide. ACS Earth and Space Chemistry 3, 15371549.CrossRefGoogle Scholar
Melandri, S, Evangelisti, L, Maris, A, Caminati, W, Giuliano, BM, Feyer, V, Prince, KC and Coreno, M (2010) Rotational and core level spectroscopies as complementary techniques in tautomeric/conformational studies: the case of 2-mercaptopyridine. Journal of the American Chemical Society 132, 1026910271.CrossRefGoogle ScholarPubMed
Müller, HSP, Thorwirth, S, Roth, DA and Winneswisser, G (2001) The cologne database for molecular spectroscopy, CDMS. Astronomy & Astrophysics 370, L49L52.CrossRefGoogle Scholar
Müller, HSP, Schlöder, F, Stutzki, J and Winnewisser, G (2005) The cologne database for molecular spectroscopy, CDMS: a useful tool for astronomers and spectroscopists. Journal of Molecular Structure 742, 215227.CrossRefGoogle Scholar
Müller, H, Belloche, A, Xu, L-H, Lees, R, Garrod, R, Walters, A, Van Wijngaarden, J, Lewen, F, Schlemmer, S and Menten, K (2016) Exploring molecular complexity with ALMA (EMoCA): alkanethiols and alkanols in Sagittarius B2(N2). Astronomy & Astrophysics 587, A92.CrossRefGoogle Scholar
Neill, JL, Steber, AL, Muckle, MT, Zaleski, DP, Lattanzi, V, Spezzano, S, McCarthy, MC, Remijan, AJ, Friedel, DN, Widicus Weaver, SL and Pate, BH (2011) Spatial distributions and interstellar reaction processes. Journal of Physical Chemistry A 115, 64726480.CrossRefGoogle ScholarPubMed
Pickett, HM (1991) Fitting and prediction of vibration-rotation spectra with spin interactions. Journal of Molecular Spectroscopy 148, 371377.CrossRefGoogle Scholar
Pierce, L and Hayashi, M (1961) Microwave spectrum, dipole moment, structure, and internal rotation of dimethyl sulfide. The Journal of Chemical Physics 35, 479.CrossRefGoogle Scholar
Pilcher, CB (2003) Biosignatures of early earths. Astrobiology 3, 471486.CrossRefGoogle ScholarPubMed
Sanchez, R, Giuliano, BM, Melandri, S, Favero, LB and Caminati, W (2007) Gas-phase tautomeric equilibrium of 4-hydroxypyrimidine with its ketonic forms: a free jet millimeterwave spectroscopy study. Journal of the American Chemical Society 129, 62876290.CrossRefGoogle ScholarPubMed
Seager, S (2013) Exoplanet habitability. Science (New York, N.Y.) 340, 577581.CrossRefGoogle ScholarPubMed
Seager, S, Schrenk, M and Bains, W (2012) An astrophysical view of earth-based metabolic biosignature gases. Astrobiology 12, 6182.CrossRefGoogle ScholarPubMed
Segura, A, Kasting, JF, Meadows, V, Cohen, M, Scalo, J, Crisp, D, Butler, RAH and Tinetti, G (2005) Biosignatures from Earth-like planets around M dwarfs. Astrobiology 5, 706725.CrossRefGoogle ScholarPubMed
Shahar, A, Driscoll, P, Weinberger, A and Cody, G (2019) What makes a planet habitable? Science (New York, N.Y.) 364, 434435.CrossRefGoogle ScholarPubMed
Smeyers, YG and Niño, A (1987) Character tables and symmetry eigenvectors for two C3v rotor molecular systems. Journal of Computational Chemistry 8 380388.CrossRefGoogle Scholar
Song, W, Maris, A, Rivilla, VM, Fortuna, F, Evangelisti, L, Lv, D, Rodríguez-Almeida, L, Jiménez-Serra, I, Pintado, JM and Melandri, S (2022) Microwave- and millimeter-wave spectra of five conformers of cysteamine and their interstellar search. Astronomy & Astrophysics 661, A129.CrossRefGoogle Scholar
Typke, V (1978) Calculations on the r z-structure of dimethylsulfoxide. Zeitschrift für Naturforsch 33, 842847.CrossRefGoogle Scholar
Uriarte, I, Melandri, S, Maris, A, Calabrese, C and Cocinero, EJ (2018) Shapes, dynamics, and stability of β-ionone and its two mutants evidenced by high-resolution spectroscopy in the gas phase. The Journal of Physical Chemistry Letters 9, 14971502.CrossRefGoogle ScholarPubMed
Velino, B, Melandri, S and Caminati, W (2004) Internal motions of the rare gas atom in dimethyl ether–krypton. Journal of Physical Chemistry A 108, 42244227.CrossRefGoogle Scholar
Vigorito, A, Calabrese, C, Paltanin, E, Melandri, S and Maris, A (2017a) Regarding the torsional flexibility of the dihydrolipoic acid's pharmacophore: 1,3-propanedithiol. Physical Chemistry Chemical Physics 19, 496502.CrossRefGoogle Scholar
Vigorito, A, Paoloni, L, Calabrese, C, Evangelisti, E, Favero, LB, Melandri, S and Maris, A (2017b) Structure and dynamics of cyclic amides: the rotational spectrum of 1,3-dimethyl-2-imidazolidinone. Journal of Molecular Spectroscopy 342, 3844.CrossRefGoogle Scholar
Vigorito, A, Calabrese, C, Melandri, S, Caracciolo, A, Mariotti, S, Giannetti, A, Massardi, M and Maris, A (2018) Millimeter-wave spectroscopy and modeling of 1,2-butanediol: laboratory spectrum in the 59.6–103.6 GHz region and comparison with the ALMA archived observations. Astronomy & Astrophysics 619, A140.CrossRefGoogle Scholar
Watson, JKG. (1977) Vibrational Spectra and Structure, vol. VI. Oxford, New York: Elsevier Science Publishers B.V.Google Scholar
Woods, R (1966) A general program for the calculation of internal rotation splittings in microwave spectroscopy. Journal of Molecular Spectroscopy 21, 424.CrossRefGoogle Scholar
Figure 0

Fig. 1. Sketch of the DMSO molecule and principal inertial axis.

Figure 1

Table 1. Experimental transition frequencies ν and deviations δν = ν(obs) − ν(calc) of internal rotation components of DMSO. Calculated frequencies n(calc) are obtained using the spectroscopic parameters listed in Table 4

Figure 2

Table 2. Experimental transition frequencies of AA lines of DMSO (34S) and DMSO (13C)

Figure 3

Table 3. Spectroscopic parameters for the AA state of DMSO, DMSO (34S) and DMSO (13C) (S-reduction, IIIl representation)

Figure 4

Table 4. Spectroscopic parameters for the global fit of DMSO compared to the fitting of AA lines both in Ir representation

Figure 5

Fig. 2. Ab initio (MP2-aug-cc-pVTZ) methyl internal rotational potential energy surface for DMSO.

Figure 6

Fig. 3. Top panel: Spectral predictions in the 0–1000 GHz frequency range for DMSO at eight different temperatures. Bottom panel: Expanded spectral window overlapping with the frequency region covered by the ALMA telescope (band 3, 84–116 GHz).

Figure 7

Table 5. Selected predicted transitions in the Alma band 3 spectral region (84–116 GHz): predicted frequency (MHz), estimated uncertainty (MHz) and base 10 logarithm of the integrated intensity in units of nm2 MHz at different temperatures

Supplementary material: File

Maris et al. supplementary material

Table S1

Download Maris et al. supplementary material(File)
File 170.6 KB