Abstract
Formyl radical (HCOâ¢) and hydroxycarbonyl radical (HOCOâ¢) are versatile building blocks in the formation of biorelevant complex organic molecules (COMs) in interstellar medium. Understanding the chemical pathways for the formation of HCO⢠and HOCO⢠starting with primordial substances (e.g., CO and CO2) is of vital importance in building the complex network of prebiotic chemistry. Here, we report the efficient formation of HCO⢠and HOCO⢠in the photochemistry of hydroxidooxidosulfur radical (HOSOâ¢)âa key intermediate in SO2 photochemistryâin interstellar analogous ices of CO and CO2 at 16âK through hydrogen atom transfer (HAT) reactions. Specifically, 266ânm laser photolysis of HOSO⢠embedded in solid CO ice yields the elusive hydrogenâbonded complexes HCOâ¢Â·Â·Â·SO2 and HOCOâ¢Â·Â·Â·SO, and the latter undergoes subsequent HAT to furnish CO2···HOS⢠under the irradiation conditions. Similar photo-induced HAT of HOSO⢠in solid CO2 ice leads to the formation of HOCOâ¢Â·Â·Â·SO2. The HAT reactions of HOSO⢠in astronomical CO and CO2 ices by forming reactive acyl radicals may contribute to understanding the interplay between the sulfur and carbon ice-grain chemistry in cold molecular clouds and also in the planetary atmospheric chemistry.
Similar content being viewed by others
Introduction
Sulfur is the tenth most abundant element in the universe, and sulfur chemistry plays vital importance not only in the biological systems and atmosphere on the Earth but also in interstellar medium (ISM). According to astronomical observations, the two sulfur oxides SO2 and SO have been found to be abundant in molecular clouds1,2,3,4, which are mainly formed through condensation of gas-phase molecules at the surface of dust grains (mostly amorphous silicates) with an onion-like structure at the temperatures of about 10â20âK5,6. The inner layer of the grains mainly consists of hydrogenated ice (H2O) with low concentrations of other H-containing species such as CH3OH, NH3, and CH4. The outer layer is made up of dehydrogenated ices with dominant compositions of CO, CO2, N2, O2, and SO2, and low concentrations of H2O may also be present in the outer layer of the icy mantle. The icy mantle at the surface of cosmic dust grains are the most important carriers of prebiotic molecules, and the composition of the mantles are largely affected by the exchanges between solid ice and gas-phase and also the photochemistry promoted by cosmic irradiations, including UV and X-ray photons from nearby stars. Therefore, the study about the chemical composition of the icy grains and the complex reaction networks is crucial for understanding the evolution of the molecular clouds7,8,9.
Carbon monoxide (CO) is the most abundant composition in the outer layer of icy grains in interstellar medium, and CO-abundant ices have also been found at the surface of many cold interstellar bodies, including comets, icy moons, and planets in the outer solar system. Therefore, the chemistry of CO through successive hydrogen atom addition reactions in the CO-rich outer layer of the interstellar icy grains may play a key role for the formation of complex organic molecules (COMs), which are probably building blocks for the origin of life10,11,12. Recently, it has been shown that simple radicals bearing elements C, N, O, P, or S play vital importance in prebiotic synthesis13. Reaction networks of these radicals in interstellar ice grains and the corresponding geochemical scenarios may help in unveiling the chemical evolution and origins of life. For instance, formyl radical (HCOâ¢) and hydroxycarbonyl radical (HOCOâ¢) are important intermediates in atmospheric and combustion chemistry14, and they are also versatile building blocks in the interstellar formation of biorelevant COMs such as formic acid (HC(O)OH)15, glyoxylic acid (HC(O)C(O)OH)16, and pyruvic acid (CH3C(O)C(O)OH)17 in low temperature (<30âK) interstellar CO and CO2 ices doped with H-containing species CH4 and H2O through barrierless radical-radical association reactions, in which the reactive acyl radicals can be generated through the hydrogenation of carbon oxides with the H-containing molecules embedded in the same ice layer at cosmic radiations18,19. As a simple organic species, HCO⢠has been observed in many interstellar clouds such as DR 21, Sgr B2, and NGC 202420,21, and its radical recombination reaction with â¢CH2OH in producing COMs during the phase transition of interstellar CO ices doped with CH3OH and H2O at a typical dense cloud temperature of about 10âK has been recently disclosed22,23,24. The cationic form of HOCO⢠has been also identified in star-forming regions such as SgrB2(OH) and lowâmass protostar IRAS 04368â+â2557 in L152725,26.
In sharp contrast to the extensively explored mechanisms for the formation of COMs through the photoreactions of H-containing species (e.g., CH3OH, NH3, and CH4) via the intermediacy of organic radicals such as HCOâ¢, HOCOâ¢, and CH3O⢠in interstellar icy grain mantles, the ice-grain photochemistry of the typical dehydrogenated molecules such as SO, SO2, and the derived sulfur-containing radicals HOS⢠and HOSO⢠in astronomical CO and CO2 ices remains barely investigated. On the other hand, the photochemistry of SO and SO2 also has great impact on the sulfur cycle in planetary atmospheres due to the formation and evolution of hazes and clouds in the upper atmospheres of Solar system planets such as Earth27, Venus28, Jupiter29, and the moon Io30. Among these sulfur oxides, SO2 is one of the most common pollutant in the Earthâs atmosphere. Therefore, the SO2 photochemistry has been the focus of enormous attention due to the important role in sulfur cycle by forming sulfuric acid and sulfate aerosols. According to the recent modeling studies31,32, the tropospheric photochemistry of SO2 in the presence of water proceeds mainly through the formation of HOSO⢠and hydroxyl radical (â¢OH) after absorption of light in the near UVâvis range (250â340ânm). In contrast, the UV photolysis (190â220ânm) of SO2 in the gas phase leads to fragmentation by yielding SO33, which is an interstellar species that has been detected in the atmospheres of Venus34 and Io35. Chemically, SO is more reactive and it dimerizes easily to yield elemental sulfur and SO2 via the intermediacy of OSSO, and the dimer has been recognized as a candidate species that contributes to the mysterious nearâUV absorption (320â400ânm) in the yellowish atmosphere of Venus36,37,38.
Herein, we report an experimental study on the photochemistry of the astrochemically relevant sulfur-containing species HOSOâ¢, HOSâ¢, SO2, SO, and OSSO in solid CO and CO2 ices at 16âK (Fig. 1). In addition to the molecular complexes formed between HOSO⢠and the carbon oxides, the photo-induced hydrogen atom transfer (HAT) to form new complexes consisting of acyl radicals (HCO⢠and HOCOâ¢) and sulfur oxides (SO2 and SO) has been observed. It is noteworthy that weakly bonded molecular complexes consisting of interstellar species have been considered as potent contributors to the rich chemistry in low-temperature giant molecular clouds39.
Results and Discussion
Isolation of HOSO⢠in CO and CO2 ices
Thanks to the strong âcage effectâ of the solid host matrix materials (e.g., Ne and Ar) at low temperatures (<30âK), the matrix isolation technique has been broadly applied in trapping highly unstable intermediates, including weakly bonded moleculeâradical complexes such as â¢OH···CO18, â¢C6H5···H2O40, â¢OC6H5···H2O41, â¢OH···H2O42, and â¢NH2···H2O43. Recently, it has been shown that HOSO⢠can be efficiently generated in the gas phase through highâvacuum flash pyrolysis (HVFP, ca. 700â°C) of CHF2S(O)OH (Fig. 2a), and photolysis of HOSO⢠(Fig. 2b) in solid Arâmatrix at 10âK yields isomeric HSO2â¢, fragments SO2/H⢠together with the caged radical complex â¢OH···OS44. The absence of free fragments â¢OH and OS in the matrix suggests that they can hardly escape from the rigid matrix cages. When the pyrolytic generation of HOSO⢠was performed in the presence of CO by using a 1:20:1000 mixture of CHF2S(O)OH/CO/Ar, the IR spectrum of the isolated species (Supplementary Fig. 1) shows the appearance of new IR bands in the range of 3500â3350âcmâ1 for OâH stretching vibrations (ν(OH)), implying complex formation between HOSO⢠and CO.
When using neat CO as the matrix host material, the IR spectrum of the isolated pyrolysis products at 16âK (Fig. 2c) clearly shows the absence of all the IR bands for free HOSO⢠(Fig. 2b), while the bands for the complex CO···HOSO⢠become dominant. The ν(OH) mode in HOSO⢠shifts from 3545.3âcmâ1 in Arâmatrix to 3396.8âcmâ1 in COâmatrix, corresponding to a redâshift (Îν) of â148.5âcmâ1. It is comparable with the shift of the ν(OH) mode in COâmatrix isolated HOCO⢠(Îνâ=ââ146.9âcmâ1) comparing to its IR spectrum in Arâmatrix45. The assignment of CO···HOSO⢠is supported by the good agreement with the theoretically calculated shift of â140âcmâ1 at the B3LYPâGD3(BJ)/def2âTZVP level (Table 1). In line with a stable hydrogen-bonded structure through OC···H bond in the complex, the stretching mode for the terminal Sâ=âO moiety (ν(Sâ=âO)) is less perturbed than the SâO stretching mode (ν(SâO)) as indicated by the shifts of â3.6 and +25.3âcmâ1, respectively. In contrast, the weak deformation mode δ(SOH) exhibits a large blueâshift of +44.2âcmâ1, which is in agreement with the calculated shift of +53âcmâ1 at the CCSD(T)/augâccâpV(Tâ+âD)Z level. Note there are several weak satellite bands around the main peaks for the fundamental modes of the complex, they probably arise from the less stable matrix sites or the less abundant complexes of HOSO⢠consisting two or more CO molecules.
By analogy, deposition of the pyrolysis products of a 1:50:1000 mixture of CHF2S(O)OH/CO2/Ar at 16âK leads to the formation of the complex CO2···HOSO⢠(Supplementary Fig. 2). Consistent with the CCSD(T) calculated red-shift of â73âcmâ1 (B3LYP: â84âcmâ1) for the ν(OH) mode in the complex, the band for HOSO⢠shifts from 3545.3âcmâ1 in Arâmatrix to 3478.7âcmâ1 in CO2âmatrix (Table 1), corresponding to a shift of â66.6âcmâ1. Concomitantly, the ν(SâO) mode undergoes a blue-shift by +22.4âcmâ1 (CCSD(T):â+â21âcmâ1; B3LYP:â+â24âcmâ1).
Weak hydrogenâbonding interactions of HOSO⢠with carbon oxides also affect its UV-vis absorption. Recently, a broad absorption centered at 270ânm (λmax) has been observed for HOSO⢠in Ar-matrix, corresponding to the transition from the ground state (X2A) to the C2A/D2A excited states46. As shown in Fig. 3, the major absorption of HOSO⢠in COâmatrix appears in the range of 350â240ânm as a weak band, and its assignment is ascertained with the photochemistry that HOSO⢠can be efficiently depleted by UVâlight irradiation at 266ânm47. This characteristic absorption for HOSO⢠is also observable in CO2âmatrix. In sharp contrast to the appearance of the absorption of HSO2⢠in the range of 320âââ500ânm after the irradiation (266ânm) of HOSO⢠in Arâmatrix, same photolysis in COâ and CO2âmatrixes results in the occurrence of weaker absorptions in the range of 300âââ400ânm, implying the formation of new species arising from the photoreactions of HOSO⢠with the carbon oxides.
Photochemistry of HOSO⢠in CO and CO2 ices
To unravel the photochemistry of HOSO⢠in COâ and CO2âmatrixes, the IR spectra for the 266ânm laser photolysis products were recorded and the resulting difference spectra reflecting the reactions of HOSO⢠are depicted in Fig. 4. In contrast to the photodissociation of HOSO⢠to Hâ¢/SO2 and the caged complex â¢OH···SO in Arâmatrix (Fig. 4a), its photolysis in solid CO (Fig. 4b) yields CO2, OCS, HCOâ¢, HOCOâ¢, H2CO, HOSâ¢, and SO2. It is noteworthy that the IR frequencies for all these species shift slightly in comparison to those observed for the corresponding species in COâmatrixes, indicating weak interactions between the neighboring counterpart species formed after the bimolecular reaction of HOSO⢠with CO inside the rigid COâmatrix cages. For instance, the two characteristic IR bands of tâHOCO⢠for the ν(OâH) and ν(C=O) modes at 3456 and 1833 cmâ1 in COâmatrix shift to 3311.0 and 1831.7âcmâ1 due to complexation with SO (Supplementary Table 1). Concomitantly, the IR band of the counterpart SO at 1139.5âcmâ1 (COâmatrix) undergoes blueâshift to 1140.0âcmâ1, and the deformation mode δ(COH) exhibits a larger blueâshift of +12.8âcmâ1. The presence of the less stable conformer câHOCO⢠is evidenced by the band at 1795.2âcmâ1, and it is also slightly redâshifted comparing to the band at 1797 cmâ1 for câHOCO⢠in COâmatrix18. Conformational conversion of câHOCO⢠to the lowerâenergy tâHOCO⢠happens upon subsequent irradiation at 532ânm. However, the previously reported45 spontaneous transformation of c-HOCOâ¢âââtâHOCO⢠in N2âmatrix (4.5âK) via quantum mechanical tunneling was not observed in CO-matrix (16âK), which is consistent with the frequently observed environmental effects on the tunneling processes in lowâtemperature matrixes48.
Consistent with the photodecomposition of HOSO⢠(â Hâ¢â+âSO2) in Arâmatrix, its photolysis in solid CO also causes HâO bond fragmentation followed by COâtrapping of the mobile hydrogen atoms to afford HCOâ¢, which acts as a hydrogen donor through weak interaction with the counterpart SO2 in the same CO-matrix cage by forming complex HCOâ¢Â·Â·Â·SO2. In this complex, the ν(CâH) mode shifts to 2493.8âcmâ1 in comparison to the same mode at 2488 and 2483âcmâ1 for HCO⢠in COâ and Arâmatrixes49. The ν(Câ=âO) and δ(COH) modes in HCO⢠and the two stretching modes of SO2 display small red-shifts (Supplementary Table 2) in the complex. The changes of the two stretching modes of SO2 in HCOâ¢Â·Â·Â·SO2 (Îνasymâ=ââ8.2âcmâ1; Îνsymâ=ââ0.3âcmâ1) are smaller than those observed in other SO2âcontaning complexes such as H2O2···SO2 (Îνasymâ=ââ12.9âcmâ1; Îνsymâ=ââ2.1âcmâ1)50. Further combination of HCO⢠with mobile hydrogen atoms in the matrix during the laser photolysis affords H2CO, and it is also perturbed by the neighboring molecule (SO2) as evidenced by the appearance of two IR bands at 1736.7 and 1734.9âcmâ1 for the ν(Câ=âO) mode in H2CO. Note, formation of HCO⢠and H2CO has been previously observed in a solid CO/H2 ice mixture at 8âK after irradiation with ultraâhigh vacuum UV (~160ânm) light19.
The mechanism for the formation of HOS⢠during the photolysis of HOSO⢠in CO-matrix is intriguing. A plausible pathway is the HAT in HOCOâ¢Â·Â·Â·SO (âââCO2···HOSâ¢). In the CO2···HOS⢠complex, the ν(OâH) mode shifts by â154.3âcmâ1 in comparison with the same mode observed in HOS⢠in solid paraâH251, while the ν(SâO) mode displays a blue-shift of +8.9âcmâ1 (Supplementary Table 3). Accordingly, the two IR bands for CO2 at 2346.7 and 659.3âcmâ1 exhibit shoulders that can be assigned to the complex (Fig. 5). Particularly, the bending mode δ(CO2) in the complex appears as a broad band in the range of 665â650âcmâ1 due to removal of vibrational degeneracy upon complexation with HOSâ¢. Similar splitting for the nondegenerate δ(CO2) modes has been observed in other CO2âcomplexes such as HKrCCH···CO252.
The spectroscopic identification of the photoproducts of HOSO⢠in COâmatrix is further supported by 18Oâisotope labeling experiments. Using 18Oâlabeled sulfinic acid as the precursor, a 1:1:1:1 mixture of HOSOâ¢, H18OSOâ¢, HOS18Oâ¢, and H18OS18O⢠can be generated, and their distinction can be assured with the revolved IR bands at 1049.7, 1045.2, 1043.4, and 1039.5âcmâ1 for the δ(SOH) mode in the four isotopologues, however, the remaining IR fundamental modes appear as doublets due to closeness of the isotopic shifts (Fig. 4d). The photochemistry of these isotopologues in CO-matrix (Fig. 4e) provides useful information for probing the reaction mechanism. The sole formation of HCO⢠(without HC18Oâ¢) and a 1:2:1 mixture of SO2, OS18O, and S18O2 among the photoproducts confirms the route for the straightforward HAT from HOSO⢠to CO. The absence of the IR bands for HOC18O⢠and H18OC18O⢠rules out the possibility for the formation of HOCO⢠via direct hydrogenation of CO2, since the singly 18Oâenriched CO2 is present in the same matrix. In contrast, the association of the photolytically generated â¢OH/â¢18OH with CO yields the experimentally observed HOCO⢠and H18OCO⢠in 1:1 ratio.
In the IR difference spectrum for the 18O-labeling experiments (Fig. 4e), each band for the ν(OâH) and ν(SâO) modes in HOS⢠splits into doublet due to the additional presence of H18OSâ¢, the corresponding 16/18O-isotopic shifts â10.8 and â31.5âcmâ1 show good agreement with the calculated values â8 and â31âcmâ1, respectively. Assuming a bimolecular reaction of HOCO⢠and SO for the formation of HOS⢠and CO2, the observation of more CO2 than HOS⢠(based on the experimental and calculated IR band intensities) in the photochemistry indicates that there is an alternative pathway for producing CO2. Given the generation of HO⢠and SO in the photolysis of HOSOâ¢, the photochemistry of SO in CO ice was also studied. Co-condensation of gaseous SO in the presence of CO (ca. 1:1000) at 16âK also yields OSSO, which can be completely destroyed by UVâlight irradiation at 365ânm (Supplementary Fig. 3). The formation of SO2 and OCS coincides with the photodissociation of OSSO to SO2 and sulfur atom with subsequent CO-trapping reaction (Sâ+âCOâââOCS). When changing the irradiation source to a 193ânm laser, depletion of SO occurs by forming CO2 and OCS (Supplementary Fig. 3), and the mechanism can be reasonably explained that monomeric SO dissociates to sulfur and oxygen atoms followed by association reactions with CO. Therefore, formation of CO2 and OCS in the photochemistry of HOSO⢠in solid CO is very likely caused by the photofragmentation of the initially generated SO. This mechanism is also consistent with the sole observation of none-isotopically labeled OCS and 1:1 mixture of CO2/OC18O in the 18Oâlalebing experiments (Fig. 4e). Additionally, traces of HCO⢠form during laser irradiation of the impurity H2O in solid CO.
The photolytic depletion of HOSO⢠in COâmatrix at 266ânm is extremely fast, and the depletion becomes much less efficient under subsequent UVâlight irradiation at 365ânm (Supplementary Fig. 4 and 5). Prolonged 266ânm laser irradiation (36âmin) leads to complete depletion of HOSOâ¢. Further irradiation at 365ânm promotes reverse HAT from HOCO⢠to SO2 by reforming HOSO⢠and CO2. The photosensitivity of HOCO⢠towards the UVâlight (365ânm) strongly indicates that the aforementioned weak absorptions in the UVâvis spectra in the range of 300âââ400ânm (Fig. 3) belongs to this carbonyl radical. It is also consistent with the observation in the IR spectra that HOCO⢠forms after the photolysis of HOSO⢠in COâmatrix (Fig. 4). The absorption for HCO⢠above 400ânm53 was not observed in the UVâvis spectra due to low intensity.
The photoâinduced HAT of HOSO⢠also occurs in CO2-doped Ar-matrix at 16âK (Fig. 4c), yielding a new molecular complex HOCOâ¢Â·Â·Â·SO2 (Supplementary Table 1). The strongest band for the ν(Câ=âO) mode in the complex locates at 1818.9âcmâ1, and it is lower than the same mode in HOCOâ¢Â·Â·Â·SO and HOCO⢠at 1831.7 and 1833 cmâ1, respectively. The very weak δ(COH) mode undergoes small blueâshift to 1278.7âcmâ1 comparing to the frequencies of 1261 and 1265.8âcmâ1 for HOCO⢠in solid COâmatrix and in HOCOâ¢Â·Â·Â·SO. In contrast, the two SO2 stretching modes at 1350.9 and 1146.4âcmâ1 are red-shifted in comparison to free SO2 at 1355.0 and 1152.2âcmâ1. Unlike the generation of two conformers of HOCO⢠in the photochemistry of HOSO⢠in COâmatrix (Fig. 4b), only one conformer was generated in the photolysis of HOSO⢠in CO2âdoped Arâmatrix (Supplementary Fig. 2). Additional formation of OCS among the photolysis products implies the trapping reaction of sulfur atoms (SOâââSâ+âO) by CO2, however, the elusive intermediate OSCO54 was not observed. Formation of OCS and HCO⢠was also observed during the 193ânm laser irradiation of a mixture of H2O and SO2 in COâmatrix (Supplementary Fig. 6).
Calculated structures of molecule-radical complexes
Weakly bonded complexes consisting of simple radicals (e.g., HO⢠and HOOâ¢) and small molecules (e.g., CO, CO2, and H2O) play important roles in gasâphase chemistry, as they may serve as key intermediates in clusters formation in atmospheric and astrochemical processes at the lowâtemperature surface of dust and ice grains39. Therefore, the structures and reactivity of these moleculeâradical complexes have been the focus of comprehensive experimental and computational studies. Despite the importance of HOSOâ¢, HOCOâ¢, and HCO⢠in gas-phase chemistry has been increasingly recognized, their complexes remain scarcely investigated. Using the UCCSD(T)/augâccâpV(Tâ+âD)Z method, the structures, energies, and bonding properties for the new molecule-radical complexes involved in the photochemistry of HOSO⢠in CO and CO2 ices were calculated (Fig. 6).
In sharp contrast to a favorable cisâplanar structure for free HOSOâ¢, the HOSO moiety in CO···HOSO⢠and CO2···HOSO⢠are nonplanar. The global minimum of CO···HOSO⢠prefers OH···CO linkage and the hydrogen atom is dramatically tilted out of the OSO plane by 33.9°. The corresponding hydrogen bond length is 2.168âà , which is shorter than the hydrogen bond in â¢OH···CO (2.341âà )55. The binding energy (Deâ=â3.9âkcalâmolâ1) is higher than those in â¢OH···CO (Deâ=â2.3âkcalâmolâ1)56 and â¢OH···OS (Deâ=â3.2âkcalâmolâ1)57. The isomer bearing a OH···OC linkage between HOSO⢠and CO is less stable by ca. 5âkcalâmolâ1. The hydrogen atom in CO2···HOSO⢠is also tilted from the OSO plane by a dihedral angle of 30.5°, and the CO2 molecule is slightly bent by 1.8°. The shorter hydrogen bond in CO2···HOSO⢠(1.990âà ) than that in CO···HOSO⢠(2.168âà ) coincides with the larger stabilizing interaction (Deâ=â5.3âkcalâmolâ1). The noncovalent interaction (NCI) analysis suggests strong attractive hydrogen bond interactions and simultaneous weak repulsive interactions of the terminal Sâ=âO moiety with CO and CO2 in these complexes (Fig. 6).
Unlike the distortion of the complexed cisâHOSOâ¢, the HOCO moiety in the hydrogen-bonded HOCOâ¢Â·Â·Â·SO and HOCOâ¢Â·Â·Â·SO2 complexes keeps a favorable transâplanar configuration. The shorter hydrogen bond in the HOCOâ¢Â·Â·Â·SO (1.893âà ) than that in the latter (1.909âà ) is also consistent with a higher stabilizing energy of 6.1âkcalâmolâ1 (5.4âkcalâmolâ1 in HOCOâ¢Â·Â·Â·CO2). The binding energy in HOCOâ¢Â·Â·Â·SO is larger than that in the hydrogen-bonded H2O···SO complex (Deâ=â3.1âkcalâmolâ1)58, while the latter remains yet experimentally unobserved. The weak interaction between HCO⢠and SO2 in HCOâ¢Â·Â·Â·SO2 (Deâ=â3.5âkcalâmolâ1) is facilitated by forming a fiveâmembered ring through concerted contacts of hydrogen bond CHâOS (2.590âà ) and chalcogen bond COâSO (2.978âà ). Similar fiveâmembered ring structure has also been predicted for the detectable HO2â¢Â·Â·Â·SO2 complex (Deâ=â4.6âkcalâmolâ1)59. The hydrogen bond interaction in the planar CO2···HOS⢠complex (Deâ=â4.5âkcalâmolâ1) is stronger than the intermolecular O···C contact in the Tâshaped van der Waals complex H2O···CO2 (Deâ=â2.8âkcalâmolâ1)60.
Energetically, HCOâ¢Â·Â·Â·SO2 and HOCOâ¢Â·Â·Â·SO are higher than CO···HOSO⢠by 37.5 and 27.1âkcalâmolâ1, respectively, whereas the secondary HAT in HOCOâ¢Â·Â·Â·SO to form CO2···HOS⢠is highly exothermic by releasing â51.5âkcalâmolâ1. Therefore, the overall process for the oxidation of CO to CO2 by reaction with HOSO⢠is exothermic by â14.0âkcalâmolâ1, which is comparable with the energy (â20.0âkcalâmolâ1) for the â¢OH radical promoted oxidation14. The HAT process in CO2···HOSO⢠to form HOCOâ¢Â·Â·Â·SO2 is endothermic by 38.5âkcalâmolâ1, which is lower than the HâO bond dissociation energy in HOSO⢠(44.1âkcalâmolâ1)57.
Implications in interstellar sulfur chemistry
Our experimental results demonstrate that organic radicals HCO⢠and HOCO⢠can be produced by UV-irradiation of HOSO⢠in astronomical CO and CO2 ices via the hydrogen atom transfer (HAT) reactions of the initially formed molecular complexes at 16âK, and the acyl radicals also form stable molecular clusters with sulfur oxides through strong hydrogen bonding interactions. Considering the facile formation of HOSO⢠from the photoreactions of SO2 with H2O, the generation of the two important building blocks HCO⢠and HOCO⢠from the photoreactions mimics the chemical evolution network of the dehydrogenated sulfur-containing molecules SO and SO2 in the outer layer of the CO-dominant interstellar icy grains in molecular clouds at a typical temperature of about 10âK. Alternatively, other radicals (e.g., â¢CH3, â¢CH2OH, and â¢CN) derived from the photoreactions of the interstellar carbon- or nitrogen-containing molecules (e.g., CH4, CH3OH, and HCN) in the cryogenic astronomical ices may also be present and react further with the acyl radicals through barrierless radical-radical association reactions to form more complex organic molecules (COMs). Additionally, the formation of OCS in the photochemistry of HOSOâ¢, SO and SO2 in CO and CO2 ices may also contribute to understanding the interstellar sulfur chemistry, since OCS not only serves as a prebiotic activating agent for amino acid polymerization in forming peptides under mild conditions in aqueous solution61,62,63, also it involves in the reduction of CO264 and acts as a condensing agent in phosphate chemistry65.
Clearly, the HAT processes of HOSO⢠may serve as the link connecting the chemistry of SO2 and the chemistry of carbon oxides (CO and CO2) in interstellar ices, although the ubiquity of HAT in chemistry, biology, and industry has been well recognized66. The uncovered chemical network for the formation of the astrochemically relevant organic radicals HCO⢠and HOCO⢠in the simple systems containing the primordial substrates (e.g., H2O, SO, SO2, CO, and CO2) might aid in disclosing the intriguing mechanism for the chemical evolution of biomolecules such as organic acids in dense molecular clouds, where barrierless radical-radical reactions at low temperatures (10â20âK) are assumed to happen spontaneously for the formation of COMs67,68. Furthermore, the spectroscopic identification and photochemistry of the new complexes consisting of the astrochemically relevant radicals HOSOâ¢, HCOâ¢, HOCOâ¢, and HOS⢠will help in understanding the chemical composition and abundances in the interstellar medium, since molecular complexes are known to contribute to the formation of interstellar media and nucleation of aerosols in diverse planetary atmospheres, such as the COârich interstellar comet 2I/Borisov69 and SO2ârich atmosphere of Venus70.
Aside from the role in ice-grain photochemistry, the planetary atmospheric chemistry of sulfur-containing species also attracts enormous interest due to importance in astrochemical reactions and planetary geology. Particularly, the photochemistry of SO2 plays a fundamental role in the sulfur cycle in the Venusian atmosphere. In addition to the primary contribution from the volcanic eruptions71, photolysis of H2SO4 vapor is another source of SO2 as the planet is completely enshrouded by the acid droplets clouds. As demonstrated by previous field measurements72 and modeling34, photolysis of H2SO4 vapor yields SO3 and H2O, followed by further photodecomposition of SO3 to SO2 and SO. This simple photochemistry can enhance the abundances of SO2 (66â±â5 ppb) and SO (31â±â4 ppb) in the cold (ca. â80â°C) mesosphere of Venus at altitudes of 85â100âkm73, where the CO abundances (ca. 30 ppm) vary with altitude partially due to reactions with sulfur compounds (e.g., SO2, SO, and S2)74. Hence, the disclosed photochemical reactions between sulfur oxides and carbon oxides in the presence of H2O (an average abundance of 30 ppm at 30â45âkm altitude)75 may affect the composition of Venusian atmosphere. Recently, Limaye et al.76 proposed that the lower cloud layer of Venus (50â70âkm) is an important target for study, since biorelevant organic acids might be generated through the iron-catalyzed metabolic redox reactions of the abundant CO2, CO, H2O, and SO2 under favorable chemical and physical conditions. The observed photoreactions of sulfur-containing species with CO2 and CO at low temperatures suggests that the hydrogenation of carbon oxides to the organic radicals HCO⢠and HOCO⢠for the formation of organic acids might be facilitated by HAT reactions with the derived radicals HOSO⢠and HSOâ¢.
Methods
Sample preparation
Difluoromethylsulfinic acid (CHF2S(O)OH) was synthesized by reaction of hydrogen chloride (HCl) with sodium difluromethylsulfinate (CHF2S(O)ONa)44. Specifically, freshly dried HCl (10âmmol) was condensed into a glass vessel containing solid CHF2S(O)ONa (0.14âg, 1âmmol) at â196â°C (liquid nitrogen bath), and the mixture was warmed to â110â°C (cold ethanol bath) and kept for overnight reaction. Then, the reaction mixture was slowly warmed to ca. â60â°C and the volatile part was pumped (10 pa) through the vacuum line consisting of two successive cold U-traps at â60 (cold ethanol bath) and â196â°C, and the acid CHF2S(O)OH (ca. 50âmg, 0.5âmmol) was obtained in the first cold trap. The 18O-enriched CHF2S(O)OH was synthesized through hydrolysis of CHF2S(O)Cl with water (18O, 97%, Eurisotop), from which a 1:1:1:1 mixture of CHF2S(O)OH, CHF2S(18O)OH, CHF2S(O)18OH, and CHF2S(18O)18OH (Supplementary Fig. 7) was obtained according to the IR spectrum of its decomposition product HOSO⢠and also the IR spectrum of its reaction product (SO2) with CO (Fig. 4e).
Matrixâisolation spectroscopy
Matrix infrared (IR) spectra are recorded using an FTâIR spectrometer (Bruker 70âV) in a reflectance mode with a transfer optic. A KBr beam splitter and liquid-nitrogen-cooled mercury cadmium telluride (MCT) detector are used in the midâIR region (5000â450âcmâ1). For each spectrum, 200 scans at a resolution of 0.5âcmâ1 are coâadded. Matrix ultraviolet-visible (UVâvis) spectra in the range of 190â800ânm are recorded using a Perkin Elmer Lambda 850+ spectrometer with a scanning speed of 1ânmâsâ1.
For the preparation of the matrix, gaseous sample is mixed by passing matrix gas (Ar, CO or CO2) through a cold Uâtrap (â10â°C) containing ca. 50âmg of the acid precursor (CHF2S(O)OH). Then, the mixture of acid vapor in matrix gas (a ratio of ca. 1:1000) is passed through an aluminum oxide furnace (2.0âmm, i.d.:1.0âmm), which can be heated over a length of â¼25âmm by a tantalum wire (o.d. 0.4âmm, resistance 0.4 Ω) and immediately deposited (2âmmolâhâ1) in a high vacuum (â¼10â5 Pa) onto a goldâplated copper block matrix support (10âK for Arâmatrix, 16âK for COâmatrix) for IR spectroscopy and onto a CaF2 window (16âK) for UVâvis spectroscopy using closedâcycle helium cryostats (Sumitomo Heavy Industries, SRDKâ408D2âF50H) inside the vacuum chambers. For the preparation of the CO2- or CO-doped Ar-matrix, a premix of Ar with CO2 or CO (a ratio of 1:20) was used as the matrix gas. Temperatures at the second stage of the cold heads are controlled and monitored using East Changing TC 290 (IR spectroscopy) and LakeâShore 335 (UVâvis spectroscopy) digital cryogenic temperature controller silicon diodes (DTâ670). The voltage and current use in the pyrolysis experiments are 7âV and 2.9âA, respectively. Photolysis experiments were performed using the Nd3+:YAG laser (266ânm, MPLâFâ266, 10âmW) and UV lamp (365ânm, 24âW).
Quantum chemistry calculation
Molecular geometries of stationary points for the monomers and complexes were first calculated at the B3LYPâGD3(BJ)/def2-TZVP77 level of theory with the Gaussian 16 software package78. The dispersion correction using the D3 version of Grimmeâs dispersion with Becke-Johnson damping, GD3(BJ)79, is necessary to obtain reliable structures for the hydrogen-bonded complexes. Then, the structures were further optimized at the CCSD(T)/augâccâpV(Tâ+âD)Z80 level of theory with MOLPRO (ver. 2019.1) software81. The nonâcovalent interactions (NCIs) analyses are carried out at the B3LYPâGD3(BJ)/def2âTZVP level of theory with Baderâs quantum atomsâinâmolecules (QTAIM)82 and Johnsonâs NCI analyses83 on basis of the CCSD(T)/augâccâpV(Tâ+âD)Z level of theory optimized structures.
Data availability
All data to evaluate the conclusion in the paper are available in the main text and/or the Supplementary Materials. The atomic coordinates generated in this paper have been deposited in the ZENODO database. (https://zenodo.org/record/7262494#.Y1zQn3ZBy3B).
References
Rydbeck, O. E. H. et al. Observations of SO in dark and molecular clouds. Astrophys. J. 235, L171âL175 (1980).
Irvine, W. M., Good, J. C. & Schloerb, F. P. Observations of SO2 and HCS+ in cold molecular clouds. Astron. Astrophys. 127, L10âL13 (1983).
Cernicharo, J. et al. Collisional excitation of sulfur dioxide in cold molecular clouds. Astron. Astrophys. 531, A103 (2011).
Vidal, T. H. G. et al. On the reservoir of sulphur in dark clouds: chemistry and elemental abundance reconciled. Mon. Not. R. Astron. Soc. 469, 435â447 (2017).
Boogert, A. C. A., Gerakines, P. A. & Whittet, D. C. B. Observations of the icy universe. Annu. Rev. Astron. Astrophys. 53, 541â581 (2015).
Hama, T. & Watanabe, N. Surface processes in interstellar amorphous solid water: adsorption, diffusion, tunneling reactions, and nuclear-spin conversion. Chem. Rev. 113, 8783â8839 (2017).
Peimbert, M. & Serrano, A., Torres-Peimbert. Interstellar matter and chemical evolution. Science 224, 345â350 (1984).
Shannon, R. J., Bltiz, M. A., Goddard, A. & Heard, D. E. Accelerated chemistry in the reaction between the hydroxyl radical and methanol at interstellar temperatures facilitated by tunnelling. Nat. Chem. 5, 745â749 (2013).
Wotos, A. et al. Synthetic connectivity, emergence, and self-regeneration in the network of prebiotic chemistry. Science 369, 345â350 (2020).
Ãberg, K. I. Photochemistry and astrochemistry: photochemical pathways to interstellar complex organic molecules. Chem. Rev. 116, 9631â9663 (2020).
Chuang, K.-J. et al. Production of complex organic molecules: H-atom addition versus UV irradiation. Mon. Not. R. Astron. Soc. 467, 2552â2565 (2017).
Oba, Y., Tomaru, T., Lamberts, T., Kouch, A. & Watanabe, N. An infrared measurement of chemical desorption from interstellar ice analogues. Nat. Astron. 2, 228â232 (2018).
Lim, R. W. J. & Fahrenbach, A. C. Radicals in prebiotic chemistry. Pure Appl. Chem. 92, 1971â1986 (2020).
Francisco, J. S. HOCO radical chemistry. Acc. Chem. Res. 43, 1519â1526 (2010).
Zhu, C., Turner, A. M., Abplanalp, M. J. & Kaiser, R. I. Formation and highâorder carboxylic acids (RCOOH) in interstellar analogous ices of carbon dioxide (CO2) and methane (CH4). Astrophys. J. Suppl. Ser. 234, 11â13 (2018).
Eckhardt, A., Bergantini, A., Singh, S. K., Schreiner, P. R. & Kaiser, R. I. Formation of glyoxylic acid in interstellar ices: a key entry point for prebiotic chemistry. Angew. Chem. Int. Ed. 58, 5663â5667 (2019).
Kleimeier, N. F., Eckhardt, A. N., Schreiner, P. R. & Kaiser, R. I. Interstellar formation of biorelevant pyruvic acid (CH3COCOOH). Chem 6, 3385â3395 (2020).
Ryazantsev, S. V., Duarte, L., Feldman, V. & Khriachtchev, L. VUV photochemistry of the H2O···CO complex in noble gas matrixes: Formation of OH···CO complex and the HOCO radical. Phys. Chem. Chem. Phys. 19, 356â365 (2017).
Chuang, K.âJ. et al. H2 chemistry in interstellar ices: hydrogenation in UV irradiation CO: H2 ice mixtures. Astron. Astrophys. 617, A87 (2018).
Snyder, L. E., Hollis, J. M. & Ulich, B. L. Radio detection of the interstellar formyl radical. Astrophys. J. 208, L91âL94 (1976).
Liszt, H. S., Gerin, M. & Lucas, R. HCO, câC3H and CF+: three new molecules in diffuse, translucent and âSpiralâArmâ clouds. Astron. Astrophys. 564, A64 (2014).
He, J., Toriello, F. E., Emtiaz, S. M., Henning, T. & Vidali, G. Phase transition of interstellar CO ice. Astrophys. J. Lett. 915, L23 (2021).
He, J., Góbi, S., Ragupathy, G., Tarczay, G. & Henning, T. Radical recombination during the phase transition of interstellar CO ice. Astrophys. J. Lett. 931, L1 (2022).
Chuang, K.-J., Fedoseev, G., Ioppolo, S., van Dishoeck, E. F. & Linnartz, H. Hâatom addition and abstraction reactions in mixed CO, H2CO and CH3OH ices â an extended view on complex organic molecule formation. Mon. Not. R. Astron. Soc. 455, 1702â1712 (2016).
Sakai, N., Sakai, T., Aikawa, Y. & Yamamoto, S. Detection of HCO2+ toward the lowâmass protostar IRAS 04368 + 2557 in L1527. Astrophys. J. 675, L89âL92 (2008).
Vastel, C., Ceccarelli, C. V., Lefloch, B. & Bachiller, R. Abundance of HOCO+ and CO2 in the outer layers of the L1544 prestellar core. Astron. Astrophys. 591, L2 (2016).
Charlson, R. J., Lovelock, J. E., Andreae, M. O. & Warren, S. G. Oceanic phytoplankton, atmospheric sulphur, cloud albedo and climate. Nature 326, 655â661 (1987).
Marcq, E., Bertaux, J. âL., Montmessin, F. & Belyaev, D. Variations of sulphur dioxide at the cloud top of Venusâs dynamic atmosphere. Nat. Geosci. 6, 25â28 (2013).
Porco, C. C. et al. Cassini imaging of Jupiterâs atmosphere, satellites, and rings. Science 299, 1541â1547 (2003).
Zolotov, M. Y. & Fegley, B. Jr. Eruption conditions of Pele volcano on Io inferred from chemistry of its volcanic plume. Geophys. Res. Lett. 27, 2789â2792 (2000).
MartinsâCosta, M. T. C., Anglada, J. M., Francisco, J. S. & RuizâLópez, M. F. Photochemistry of SO2 at airâwater interface: A source of OH and HOSO radicals. J. Am. Chem. Soc. 140, 12341â12344 (2018).
RuizâLópez, M. F., MatinsâCosta, M. T. C., Anglada, J. M. & Francisco, J. S. A new mechanism of acid rain generation of HOSO at the airâwater interface. J. Am. Chem. Soc. 141, 16564â16568 (2019).
Whitehill, A. R. et al. Vibronic origin of sulfur massâindependent isotope effect in photoexcitation of SO2 and the implications to the early Earthâs atmosphere. Proc. Natl Acad. Sci. USA 110, 17697â17702 (2013).
Zhang, X. et al. Photolysis of sulphuric acid as the source of sulphur oxides in the mesosphere of Venus. Nat. Geosci. 3, 834â837 (2010).
Russell, C. T. & Kivelson, M. G. Detection of SO in Ioâs exosphere. Science 287, 1998â1999 (2000).
Wu, Z. et al. The nearâUV absorber OSSO and its isomers. Chem. Commun. 54, 4517â4520 (2018).
Pinto, J. P. et al. Sulfur monoxide dimer chemistry as a possible source of polysulfur in the upper atmosphere of Venus. Nat. Commun. 12, 175 (2021).
Frandsen, B. N., Wennberg, P. O. & Kjaergaard, H. G. Identification of OSSO as a nearâUV absorber in the Venusian atmosphere. Geophys. Res. Lett. 43, 11146â11155 (2016).
Klemperer, W. & Vaida, V. Molecular complexes in close and far away. Proc. Natl Acad. Sci. USA 103, 10584â10588 (2006).
Mardyukov, A. SanchezâGarcia, E., CrespoâOtero, R. & Sander, W. Interaction and reaction of the phenyl radical with water: A source of OH radicals. Angew. Chem. Int. Ed. 48, 4804â4807 (2009).
Sander, W., Roy, S. & Polyak, I. RamireâAnguita, J. M. & SanchezâGarcia, E. The phenoxyl radicalâwater complexâa matrix isolation and computational study. J. Am. Chem. Soc. 134, 8222â8230 (2012).
Ohshima, Y., Sato, K., Sumiyoshi, Y. & Endo, Y. Rotational spectrum and hydrogen bonding of the H2OâHO radical complex. J. Am. Chem. Soc. 127, 1108â1109 (2005).
Ennis, C. P., Lane, J. R., Kjaergaard, H. G. & McKinley, A. J. Identification of the water amidogen radical complex. J. Am. Chem. Soc. 131, 1358â1359 (2009).
Chen, C. et al. Capture of the sulfur monoxideâhydroxyl radical complex. J. Am. Chem. Soc. 142, 2175â2179 (2020).
Ryazantsev, S. V., Feldman, V. I. & Khriachtchev, L. Conformational switching of HOCO radical: Selective vibrational excitation and hydrogenâatom tunneling. J. Am. Chem. Soc. 139, 9551â9557 (2017).
CarmonaâGarcÃa, J. et al. Photochemistry and nonâadiabatic photodynamics of the HOSO radical. J. Am. Chem. Soc. 143, 10836â10841 (2021).
Lu, B. et al. Spectroscopic characterization of HSO2⢠and HOSO⢠intermediates involved in SO2 geoengineering. J. Phys. Chem. A 125, 10615â10621 (2021).
Schreiner, P. R. Tunneling control of chemical reactions: The third reactivity paradigm. J. Am. Chem. Soc. 139, 15276â15283 (2017).
Ewing, G. E., Thompson, W. E. & Pimentel, G. C. Infrared detection of the formyl radical HCO. J. Chem. Phys. 32, 927â932 (1960).
Pehkonen, S., Lundell, J., Khriachtchev, L., Pettersson, M. & Räsänen, M. Matrix isolation and quantum chemical studies on the H2O2âSO2 complex. Phys. Chem. Chem. Phys. 6, 4607â4613 (2004).
Góbi, S., Csonka, I. P., Bazsó, G. & Tarczay, G. Successive hydrogenation of SO and SO2 in solid paraâH2: formation of elusive small oxoacids of sulfur. ACS Earth Space Chem. 5, 1180â1195 (2021).
Ryazantsev, S., Tyurin, D. A., Nuzhdin, K. B., Feldman, V. I. & Khriachtchev, L. The HKrCCH···CO2 complex: an ab initio and matrixâisolation study. Phys. Chem. Chem. Phys. 21, 3656â3661 (2019).
Flad, J. E., Brown, S. S., Burkholder, J. B., Stark, H. & Ravishankara, A. R. Absorption cross sections for the A2Aâ (0,90,0) â X2Aâ (0,01,0) band of the HCO radical. Phys. Chem. Chem. Phys. 8, 3636â3642 (2006).
Chiang, H.-J., Wang, N.-S., Tsuchiya, S., Lee, Y.âP. & Lin, M. C. Reaction dynamics of O(1D,3P) + OCS with timeâresolved fourier transform infrared spectroscopy and quantum chemical calculations. J. Phys. Chem. A 113, 13260â13272 (2009).
Caracciolo, A. et al. Combined experimentalâtheoretical study of the OH + CO â H + CO2 reaction dynamics. J. Phys. Chem. Lett. 9, 1229â1236 (2018).
Li, J. et al. Communication: A chemically accurate global potential energy surface for the HO + CO â H + CO2 reaction. J. Chem. Phys. 136, 041103 (2012).
Qin, J. & Li, J. An accurate full-dimensional potential energy surface for the reaction OH + SO â H + SO2. Phys. Chem. Chem. Phys. 23, 487â497 (2021).
Misiewicz, J. P., Noonan, J. A., Turney, J. M. & Schaefer, H. F. III The non-covalently bound SO···H2O system, including an interpretation of the differences between SO···H2O and O2···H2O. Phys. Chem. Chem. Phys. 20, 28840â28847 (2018).
Wang, B. & Hou, H. Theoretical investigations on the SO2 + HO2 reaction and the SO2âHO2 radical complex. Chem. Phys. Lett. 410, 235â241 (2005).
Andersen, J., Heimdal, J., Mahler, D. W., Nelander, B. & Larsen, R. W. Communication: THz absorption spectrum of the CO2âH2O complex: observation and assignment of intermolecular van der Waals vibrations. J. Chem. Phys. 140, 091103 (2014).
Leman, L., Orgel, L. & Ghadiri, M. R. Carbonyl sulfide-mediated prebiotic formation of peptides. Science 306, 283â286 (2004).
Frenkel-Pinter, M., Samanta, M., Ashkenasy, G. & Leman, L. J. Prebiotic peptides: Molecular hubs in the origin of life. Chem. Rev. 120, 4707â4765 (2020).
Nair, N. N., Schreiner, E. & Marx, D. Peptide synthesis in aqueous environments: the role of extreme conditions on amino acid activation. J. Am. Chem. Soc. 130, 14148â14160 (2008).
Heinen, W. & Lauwers, A. M. Organic sulfur compounds resulting from the interaction of iron sulfide, hydrogen sulfide and carbon dioxide in an anaerobic aqueous environment. Orig. Life Evol. Biosphere 2, 131â150 (1996).
Biron, J.-P. & Pascal, R. Amino acid N-carboxyanhydrides:â Activated peptide monomers behaving as phosphate-activating agents in aqueous solution. J. Am. Chem. Soc. 126, 9198â9199 (2004).
Kumar, M., Sinha, A. & Francisco, J. S. Role of double hydrogen atom transfer reactions in atmospheric chemistry. Acc. Chem. Res. 49, 877â883 (2016).
Krasnokutski, S. A. Did life originate from lowâtemperature areas of the Universe? Low. Temp. Phys. 47, 199 (2021).
Turner, A. M. & Kaiser, R. I. Exploiting photoionization reflectron time-of-flight mass spectrometry to explore molecular mass growth processes to complex organic molecules in interstellar and Solar system ice analogs. Acc. Chem. Res. 53, 2791â2805 (2020).
Bodewits, D. et al. The carbon monoxideârich interstellar comet 21/Borisov. Nat. Astron. 4, 867â871 (2020).
Esposito, L. W. Rising sulphur on Venus. Nat. Geosci. 6, 2â21 (2013).
Smrekar, S. E. et al. Recent hotspot volcanism on Venus from VIRTIS emissivity data. Science 328, 605â608 (2010).
Vaida, V., Kjaergaard, H. G., Hintze, P. E. & Donaldson, D. J. Photolysis of sulfuric acid vapor by visible solar radiation. Science 299, 1566â1568 (2003).
Bézard, B. & de Bergh, C. Composition of the atmosphere of Venus below the clouds. J. Geophys. Res. 112, E04S07 (2007).
Fegley, B. Jr., Zolotov, M. Y. & Lodders, K. The oxidation state of the lower atmosphere and surface of Venus. Icarus 125, 416â439 (1997).
Pérez-Hoyos, S. et al. Venus upper clouds and the UV absorber from MESSENGER/MASCS observations. J. Geophys. Res. Planets 123, 145â162 (2018).
Limaye, S. S. et al. Venusâ spectral signatures and the potential for life in the clouds. Astrobioloy 18, 1181â1198 (2018).
Becke, A. D. Densityâfunctional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 98, 5648â5652 (1993).
Frisch, M. J. et al. Gaussian 16, revision C.01; Gaussian, Inc.: Wallingford CT, 2016.
Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFTâD) for the 94 elements HâPu. J. Chem. Phys. 132, 154104 (2010).
Bartlett, R. J. Coupledâcluster approach to molecular structure and spectra: a step toward predictive quantum chemistry. J. Phys. Chem. 93, 1697â1708 (1989).
Werner, H.-J. et al. MOLPRO, a package of ab initio programs, ver. 2019.1. http://www.molpro.net.
Bader, R. F. W. A quantum theory of molecular structure and its applications. Chem. Rev. 91, 893â928 (1991).
Johnson, E. R. et al. Revealing noncovalent interactions. J. Am. Chem. Soc. 132, 6498â6506 (2010).
Acknowledgements
This work is supported by the National Natural Science Foundation of China (Grant 22025301 and Grant 22206027) and China Postdoctoral Science Foundation (Grant 2021M700802).
Author information
Authors and Affiliations
Contributions
X.Z. conceived project and designed the experiments. X.L. performed the synthesis and spectroscopic measurements. B.L., L.W., J.X. and B.Z. also contributed to the spectroscopic measurements and analysis of the spectral data. T.T. carried out the quantum chemical calculations. X.Z., X.L. and J.S.F. discussed the results and drafted the manuscript. All authors commented on the manuscript.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks György Tarczay and the other, anonymous, reviewer for their contribution to the peer review of this work. Peer reviewer reports are available.
Additional information
Publisherâs note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the articleâs Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the articleâs Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Li, X., Lu, B., Wang, L. et al. Unraveling sulfur chemistry in interstellar carbon oxide ices. Nat Commun 13, 7150 (2022). https://doi.org/10.1038/s41467-022-34949-4
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-022-34949-4