Open Access Article
Robert
Skog
a and
Benjamin N.
Frandsen
*ab
aDepartment of Chemistry, Faculty of Science, University of Helsinki, Finland. E-mail: benjamin.frandsen@helsinki.fi
bAerosol Physics Laboratory, Faculty of Engineering and Natural Sciences, Tampere University, Finland
First published on 8th October 2025
Polysulfurs (SX) are thought to be an important part of the Venusian atmospheric sulfur cycle. Among these, S3, S4, and S8 have been proposed to be contributors to the enigmatic near-UV absorption feature observed between 320 and 400 nm in the atmosphere of Venus. Furthermore, the individual UV-Vis spectral signatures of polysulfur isomers remain poorly characterized, hindering the accurate photochemical modeling of these species. In this work, we use the nuclear ensemble approach to simulate the UV-Vis spectra for a comprehensive set of S2–S8 structural isomers, generating wavelength-dependent absolute absorption cross sections suitable for use in photochemical models. Benchmark calculations using a set of molecules and radicals were used to validate the chosen spectral simulation approach. The results show that the smaller S2–S4 systems have open-chain geometric isomers as their global minima, while cyclic structures are the most stable for the larger S5–S8 species. We also identify two low-energy triplet S4 isomers (cis-3S4 and trans-3S4), which may serve as important intermediates. The simulated spectra indicate that the global minimum S3 structural isomer, linear-S3, and a high-energy structural isomer of S4, trigonal-S4, absorb in the 320–400 nm window, while S8 shows negligible absorption in the 320–400 nm range. This work advances our understanding of atmospheric sulfur species beyond Earth and supports the interpretation of observations of the atmosphere of Venus and other exoplanets.
Atmospheric sulfur species are not limited to Venus. Using the James Webb Space Telescope, spectral signals matching SO2 have been recorded from the atmospheres of the hot Jupiter WASP-37b,6 the super-Neptune WASP-107b,7 the sub-Neptune GJ 3470 b,8 and the Super-Earth L 98-59 d.9 Recent infrared measurements from the atmosphere of the sub-Neptune K2-18b revealed features suggested to match dimethyl sulfide (DMS) and dimethyl disulfide (DMDS), both potential biosignatures.10 In the solar system, SO2 was confirmed on Jupiter's moon Io using the infrared spectroscopy instrumentation onboard the Voyager 1 space probe.11 Disulfur (S2) was detected by the Hubble Space Telescope in the plume of Pele, a volcano on Io.12
In the atmosphere on Earth, sulfur species play a major role in several important atmospheric processes. The most abundant species are SO2 and OCS, but others, such as DMS, hydrogen sulfide (H2S), and methanesulfonic acid (MSA) are also important.13,14 In the Venusian atmosphere SO2 ultimately forms H2SO4 through oxidation and hydrolysis, and H2SO4 is a key contributor to acid rain15 and atmospheric nucleation.16–18
Due to the relatively high mixing ratio of SO2 in the Venusian atmosphere, it is the main absorber between 200–320 nm.19 In addition, another weaker absorption feature can be seen in the 320–400 nm range.20 Unfortunately, even after half a century since its discovery, there is still no scientific consensus on the identity of this unknown absorber (UA). Numerous different species have been proposed, including Cl2,19 SCl2,21 FeCl3 in H2SO4 aerosols,21–25 the SSCO isomer of the [C, O, S, S] system,26 the cis and trans isomers of OSSO,27,28 and even complex organic species formed in H2SO4 aerosols.29 The latest UA candidate suggestion are different structural isomers of S3O.30
One proposed set of candidate compounds as the UA are polysulfurs (SX), specifically S3, S4, and S8.31,32 Using the data from Venera 14, Krasnopolsky33 estimated the mixing ratio of S3 to be (80 ± 30) pptV in the lower atmosphere of Venus between 5 and 25 km. By reanalyzing the spectra from Venera 11, Maiorov et al.34 estimated the mixing ratio of S3 to be 30–100 pptV between 3–19 km, and 0 pptV between 19–36 nm. Krasnopolsky35 also reanalyzed the spectra from Venera 11, but included S4 in the analysis. The mixing ratios of (11 ± 3) pptV at 3–10 km and (18 ± 3) pptV at 10–19 km were obtained for S3, and (4 ± 4) pptV at 3–10 km and (6 ± 2) pptV at 10–19 km for S4.35 Polysulfurs are also hypothesized to be a part of the Venusian atmospheric sulfur cycle36,37 and therefore included in several Venusian atmospheric models.36–41 A possible polysulfur production pathway was reported by Francés-Monerris et al.42 using high-level photodynamics simulations.
For S2, quite a few studies on UV-Vis absorption have been published.43–47 To the best of our knowledge, the most recent experimental UV-Vis spectrum of S2 was published by Stark et al.48 For the larger SX species, Meyer et al.49 identified the absorption features of S3 and S4 using matrix isolation spectroscopy. Later, Billmers and Smith50 published the experimental spectra of S3 and S4 from sulfur vapor at 673–903 K. Hassanzadeh and Andrews51 identified two S4 isomers using matrix isolation spectroscopy. The absorption peak at 518 nm was assigned to the cis form of the linear open-chain S4 isomer (see Fig. 1), and the peak at 560–660 nm to the cyclic-S3 (S) isomer. Based on experimental Raman spectra of sulfur vapor, Boumedien and Picquenard52 also assigned the peak at 560–660 nm to the cyclic-S3 (S) isomer, but the 518 nm peak was instead attributed to the trans form of the open-chain isomers. Finally, Steudel et al.53 measured the spectra of homocycles S6, S7, S8, S9, S10, S12, S15, and S20 in solutions of methanol or methylcyclohexane.
![]() | ||
| Fig. 1 All S2–S8 structures considered in this work, ordered by increasing zero-point vibrational corrected electronic energy. | ||
The tendency of small polysulfur species to react and form larger ones makes it difficult to experimentally study individual structures of, for example, the S4 system. Therefore, for distinguishing absorption features for individual isomers, a computational approach is appropriate. Several studies have investigated the excited states of S2 using ab initio approaches.54–59 Recently, Sarka et al.60 performed high-level calculations and spectral simulations using the R-matrix approach for S2, with isotopic effects included. The approach used was then further refined in a subsequent publication by Sarka and Nanbu.61 Peterson et al.62 explored the low-energy excited states of S3 using multireference configuration interaction methods. The S4 system was explored by Wong and Steudel,63 who calculated vertical excitations for the different S4 isomers. Based on these calculations, the authors concluded that the linear open-chain cis form of S4 was the species absorbing at 518 nm, while the trans form was the absorber at 560–660 nm. Furthermore, several studies have also calculated the potential energy surfaces (PES) for different polysulfur species.64–73
An attractive approach for computationally simulating UV-Vis spectra for different species is the nuclear ensemble approximation (NEA). When using the NEA approach more natural absorption band widths are obtained with wavelength-dependent absolute absorption cross sections, as opposed to the stick spectra from similar vertical excitation calculations. Although the approach does not recover vibrational progression, it has been used to qualitatively reproduce the absorption spectra for a variety of species, such as sulfur-containing molecules relevant to the atmospheres of both Earth and Venus,28,74 nucleobases,75 isomers of Cl2O2,76 the ClCO radical,77 gaseous mercury compounds,78 pyrrole,79 and different volatile organic compounds.80,81 The NEA approach was used to simulate the spectra for different UA candidates, such as the two OSSO isomers,28 the SSCO molecule,26 and the isomers of S3O.30 The approach is also used in AtmoSpec,82 a tool for calculating photo absorption cross sections for volatile organic compounds in the troposphere of Earth.
In this work, we use the NEA approach to simulate the UV-Vis spectra for the polysulfur species S2–S8 to produce wavelength-dependent absolute absorption cross sections for the different species to be included in photochemical models. To confirm the validity of the chosen spectral simulation approach we performed benchmark calculations for several molecules and radicals (CS2, H2S, HO2, O3, OCS, PH3, SO2, and SO3), comparing the obtained results to experimental spectra.
We performed initial test using 11 different density functionals, combined with 13 basis sets [see the SI for details]. Based on these test, we selected the overall best performing levels of theory for the benchmark calculations. The wavefunction-based method chosen was the equation-of-motion coupled cluster singles and doubles (EOM-CCSD). Farahani et al.74 also found that EOM-CCSD was a good approach for accurate spectral simulation. The density functionals chosen were the hybrid GGA functional BHandHLYP,86 the range-separated hybrid GGA functionals CAM-B3LYP87 and ωB97X-D3,88 and the range-separated double hybrid functional ωB2PLYP89 without the Tamm–Dancoff approximation (TDA). In all calculations the aug-cc-pV(T+d)Z basis set90–93 was used. EOM-CCSD calculations were performed with the Gaussian 16 Rev C.02 program.94 All other calculations were performed with the Orca 5.0.4 quantum chemistry code.95 The basis set exchange96 was used to import basis sets not implemented in the quantum chemistry packages.
The theory of the nuclear ensemble approximation for spectral simulation is described in detail in the literature.97,98 Briefly, the absorption cross section σ is calculated using eqn (1):97
![]() | (1) |
In this work, the Newton-X 2.6 (build 01) program99,100 was used for all spectral simulations. A nuclear ensemble consisting of 2000 geometries was created based on a Wigner distribution using the calculated normal modes. The ensemble size was chosen as the best compromise between minimizing the error due to statistical sampling and the computational resources needed, as shown by Farahani et al.74 The ensemble size chosen in this work is also on the larger side when compared to other works using the NEA for spectral simulation.26,30,80,101 Calculated vertical excitations were convoluted with a 0.1 eV full width at half-maximum (FWHM) normalized Gaussian function.74 The ten lowest-energy excited states were calculated for all species. All spectral simulations were done at a temperature of 0 K. The simulated spectra were visualized using Matlab.102
Geometry optimizations and frequency calculations for all S2–S8 minima and transition states were performed at the ωB97X-D/aug-cc-pV(T+d)Z107,108 level of theory, with the Int = UltraFine and Opt = tight keywords added. Calculated normal nodes were used to verify that the optimized geometries were true minima or transition states. Intrinsic reaction coordinate (IRC) calculations were used to verify that the transition states were connected to the correct minima. In addition, the S2–S4 isomers were also optimized at the CCSD/aug-cc-pV(T+d)Z level of theory. Single-point energy corrections were performed at the CCSD(T)/aug-cc-pV(T+d)Z level of theory for all optimized geometries. All density functional geometry optimizations were performed with the Gaussian 16 Rev C.02 program. Coupled cluster geometry optimizations and single point energy corrections were performed with the Cfour 2.1 program suite, with the exception of the cyclic-S4 and trigonal-3S4 structures. For these two, the Gaussian 16 Rev C.02 program was used for the CCSD/aug-cc-pV(T+d)Z geometry optimization.
Miliordos and Xantheas109 showed that S3 has partial biradical character. In addition, Frandsen et al.27 noted that some structures of the S2O2 system, which is isovalent with S4, also showed multiconfigurational character. Motivated by this, we performed additional calculations on the CCSD/aug-cc-pV(T+d)Z optimized S2–S4 geometries. Using the MRCC program,110 single-point energies were calculated at the CCSDT/aug-cc-pV(T+d)Z level of theory to gauge the effect of including canonical coupled cluster triples. Complete active space self-consistent field (CASSCF) calculations, with dynamical correlation corrections accounted for with n-electron valence state perturbation theory (NEVPT2), were performed with the aug-cc-pV(T+d)Z basis set using the Orca 6.0.1 program.111 The active spaces included all valence orbitals for the sulfur atoms, resulting in 12 electrons in 8 orbitals [CASSCF(12,8)] for S2, CASSCF(18,12) for S3, and CASSCF(24,16) for S4.
For S2–S4, the CCSD/aug-cc-pV(T+d)Z optimized geometries and normal modes were used to create a Wigner ensemble, while ωB97X-D/aug-cc-pV(T+d)Z geometries and normal modes were used for S5–S8. We used the same levels of theory for the spectral simulations of the polysulfur species as in the benchmark calculations, these being EOM-CCSD, BHandHLYP, CAM-B3LYP, ωB97X-D3, and ωB2PLYP, combined with the aug-cc-pV(T+d)Z basis set. All EOM-CCSD calculations were again performed with the Gaussian 16 Rev C.02 program, and TD-DFT calculations with the Orca 5.0.4 quantum chemistry code. Spectral simulations for the S5–S8 species were only performed using TD-DFT, as calculations using EOM-CCSD were too computationally demanding.
To concisely communicate the levels of theory calculations were performed at we will use Method2/Basis2//Method1/Basis1 notation, where Method1/Basis1 indicates the level of theory for the geometry optimization and frequency calculation, and Method2/Basis2 refers to a single-point or vertical excitation calculation performed on the optimized geometry. If the same basis set was used in both calculations (meaning Basis1 is the same as Basis2), this is shortened to Method2//Method1/Basis1.
![]() | ||
| Fig. 2 The simulated spectra for the benchmark set compared to experimental spectra. Note the different scales for the axes. The references for the experimental spectra are: CS2 – Grosch et al.112 (206–350 nm), Sunanda et al.113 (185–206 nm), and Rabalais et al.114 (180–185 nm); H2S – Wu and Chen115 (160–260 nm); HO2 – JPL 2010 recommendation116 (160–260 nm); O3 – JPL 2010 recommendation116 (120–350 nm); OCS – JPL 2010 recommendation116 (258–280 nm), Limão-Vieira et al.117 (140–258 nm); PH3 – Chen et al.118 (130–230 nm); SO2 – compilation by Manatt and Lane119 (130–340 nm); SO3 – Burkholder and McKeen120 (196–300 nm), Hintze et al.121 (140–194 nm). | ||
Based on the benchmark calculations and the numerous successful applications of the NEA,26,28,30,74–76,78,80,82 we conclude that using the NEA approach to simulate the UV-Vis spectra for the different SX species is appropriate.
| Structure | ωB97X-Da | CCSD(T)//ωB97X-Da | CCSDa | CCSD(T)//CCSDa | CCSDT//CCSDa | NEVPT2//CCSDa | |
|---|---|---|---|---|---|---|---|
| a See the end of Section 2.2 for an explanation of the convention used to indicate the level of theory used. b Calculated using NEVPT2(12,8) on top of a CASSCF(12,8) reference. c Calculated using NEVPT2(18,12) on top of a CASSCF(18,12) reference. d Calculated using NEVPT2(24,16) on top of a CASSCF(24,16) reference. e Calculation did not converge. | |||||||
| S2 | 3S2 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0b |
| 1S2 | 22.3 | 17.7 | 19.2 | 17.4 | 16.5 | 13.3b | |
| S3 | Linear-S3 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0c |
| Cyclic-S3 | 2.6 | 5.0 | 1.3 | 5.4 | 5.5 | 4.3c | |
| 3S3 | 17.2 | 20.6 | 18.6 | 21.0 | 20.4 | 21.5c | |
| S4 | cis-S4 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0d |
| trans-S4 | 8.3 | 8.5 | 8.3 | 8.4 | 7.8 | 8.2d | |
| cis-3S4 | 1.5 | 9.7 | 0.9 | 10.6 | 9.1 | 13.2d | |
| trans-3S4 | 1.8 | 10.9 | 0.9 | 11.3 | 9.8 | 13.7d | |
| Cyclic-S3 (S) | 7.6 | 12.7 | 7.2 | 13.1 | —e | 14.3d | |
| Cyclic-S4 | 8.2 | 14.5 | 5.1 | 11.3 | —e | 17.2d | |
| Trigonal-S4 | 18.5 | 23.5 | 18.9 | 23.7 | 23.2 | 25.2d | |
| Trigonal-3S4 | 38.7 | 46.5 | 39.9 | 47.7 | 45.2 | 49.2d | |
| S5 | Cyclic-S5 (A) | 0.0 | 0.0 | ||||
| 3S5 (B) | 18.2 | 23.0 | |||||
| 3S5 (C) | 18.2 | 23.1 | |||||
| S5 (D) | 34.3 | 26.9 | |||||
| S5 (E) | 30.9 | 28.8 | |||||
| S5 (F) | 31.3 | 30.8 | |||||
| S5 (G) | 31.5 | 31.3 | |||||
| S5 (H) | 47.4 | 43.7 | |||||
| S5 (I) | 57.5 | 56.9 | |||||
| 3S5 (J) | 78.0 | 97.9 | |||||
| S6 | Chair-S6 (A) | 0.0 | 0.0 | ||||
| Boat-S6 (B) | 14.3 | 13.2 | |||||
| S6 (C) | 29.3 | 23.5 | |||||
| S6 (D) | 27.0 | 24.2 | |||||
| 3S6 (E) | 27.8 | 29.1 | |||||
| S6 (F) | 32.6 | 29.3 | |||||
| 3S6 (G) | 28.7 | 32.2 | |||||
| S6 (H) | 48.2 | 43.8 | |||||
| S6 (I) | 54.2 | 51.3 | |||||
| S6 (J) | 56.5 | 53.5 | |||||
| S7 | Chair-S7 (A) | 0.0 | 0.0 | ||||
| Boat-S7 (B) | 5.8 | 4.2 | |||||
| S7 (C) | 21.2 | 19.1 | |||||
| 3S7 (D) | 29.0 | 32.5 | |||||
| 3S7 (E) | 28.9 | 33.1 | |||||
| S7 (F) | 45.0 | 35.5 | |||||
| S7 (G) | 42.2 | 36.0 | |||||
| S8 | Crown-S8 (A) | 0.0 | 0.0 | ||||
| Twisted-S8 (B) | 10.4 | 8.8 | |||||
| exo–endo-S8 (C) | 11.4 | 9.7 | |||||
| Boat-S8 (D) | 21.2 | 17.0 | |||||
| S8 (E) | 25.7 | 18.3 | |||||
| S8 (F) | 30.8 | 26.6 | |||||
| 3S8 (G) | 36.4 | 37.4 | |||||
| S8 (H) | 68.1 | 49.2 | |||||
Below we discuss the results for the different SX systems in more detail.
A high-resolution UV-Vis spectrum of S2 was recently published by Stark et al.48 Therefore, we present our simulated spectra in Fig. A9 in the SI. The simulated spectra do not reproduce the vibrational fine structure seen in the experimental spectrum due to neglecting Franck–Condon factors.
The simulated spectra for the S3 isomers are presented in Fig. 3. For the cyclic-S3 isomer, all simulated spectra are in good agreement with each other. However, with the linear-S3 isomer, a blue shift in the EOM-CCSD/aug-cc-pV(T+d)Z simulated spectrum is observed when compared to the TD-DFT spectra. To investigate this discrepancy, we performed further single-point vertical excitation calculations on the CCSD/aug-cc-pV(T+d)Z optimized geometry. The vertical excitation energies and oscillator strengths for the three lowest energy excited states were calculated using multiconfigurational random phase approximation, MC-RPA(18,12)/aug-cc-pV(T+d)Z, on top of a CASSCF(18,12)/aug-cc-pV(T+d)Z reference using the Orca 6.0.1 program.111 The MC-RPA approach was chosen based on its reasonable computational cost and relatively good performance compared to other multireference methods.129,130 Furthermore, EOM-CC3/aug-cc-pV(T+d)Z vertical excitation calculations were performed using the eT 1.9.13 program.131
![]() | ||
| Fig. 3 Simulated spectra for the two singlet S3 isomers, compared with the experimental spectrum of S3 by Billmers and Smith,50 measured at temperatures in the range of 673–903 K. The experimental spectrum was digitized using the PlotDigitizer134 tool. Note the different scales on the axes. The experimental spectrum has been scaled down by 80% for clarity in the lower part of the figure. | ||
The results from these additional calculations are presented in Table A4 in the SI. When specifically looking at the third transition (S3 ← S0), the result from the EOM-CCSD/aug-cc-pV(T+d)Z calculation is slightly better in line with the MC-RPA(18,12)/aug-cc-pV(T+d)Z and EOM-CC3/aug-cc-pV(T+d)Z results. Using the MC-RPA(18,12)/aug-cc-pV(T+d)Z results, we scale the EOM-CCSD/aug-cc-pV(T+d)Z simulated spectrum and visualize this in Fig. 3, which we recommend as our highest quality linear-S3 UV-Vis spectrum (the scaling procedure is outlined in Section A2.3 in the SI). Scaling based on the MR-RPA result was motivated by the non-negligible amount of multiconfigurational character present in linear-S3. For cyclic-S3 we recommend the EOM-CCSD/aug-cc-pV(T+d)Z spectrum.
All simulated spectra in Fig. 3 predict a lower absorption cross section than the experimental spectrum. This is especially relevant for linear-S3, as this is the minimum energy isomer. However, it should be noted that the experimental spectrum was measured at very high temperatures, 673–903 K.50 Therefore, we postulate that the primary difference between the experimental spectrum and the simulated spectra originates from the different temperatures used, such as by the inclusion of hot bands in the experimental spectrum. Because of the high temperatures used in the experimental work, we consider the simulated spectra to be representative of a lower temperature (0–300 K) absorption spectra for use in photochemical models.
S3 was one of the polysulfur species suggested as the unknown absorber (UA).32 Indeed, based on our simulated spectra, linear-S3 absorbs in the same wavelength region as the UA (320–400 nm) with a relatively large absorption cross section. But the spectral match is only one part, as there must also be an appreciable abundance of linear-S3 in the Venusian atmosphere to have an impact. The model by Zhang et al.37 predicts a S3 mixing ratio of approximately 0.01 pptV at an altitude of 58 km. However, while accurate reaction rates have been calculated for the S + S2 reaction,132 other reaction rates for polysyulfurs are still based on estimates.133 Updated reaction rates are needed for a more accurate determination.
The singlet PES for the S4 system is shown in Fig. 4. We find, like Wong and Steudel,63 that the rectangular structure (labeled TS1 in Fig. 4) is a transition state. At the ωB97X-D/aug-cc-pV(T+d)Z level of theory this transition state corresponds to an interconversion of cis-S4 to itself, which might be relevant for isotope scrambling. We also note that, despite our best efforts, we were unable to locate a transition state connecting cis-S4 to trans-S4 at the ωB97X-D/aug-cc-pV(T+d)Z level of theory. As can be seen in Table A3 in the SI, the CSF weights of cis-S4 and trans-S4 indicate that multiconfigurational approaches should be used. Therefore, this transition state might exist only on a multireference PES, or proceed via the spin triplet PES. However, based on the relatively high isomerization energies shown in Fig. 4, we conclude, as did Wong and Steudel,63 that the different singlet structural isomers are well separated and should be experimentally accessible, especially in low-temperature conditions.
The simulated spectra for the S4 singlet species are presented in Fig. 5–7. The simulated spectra for cyclic-S3 (S), cyclic-S4, and trigonal-S4 are similar. But for cis-S4 and trans-S4, the first peak in the EOM-CCSD/aug-cc-pV(T+d)Z spectra is blue-shifted compared to the spectra simulated using TD-DFT. Therefore, as with linear-S3, we performed additional vertical excitation calculations on the CCSD/aug-cc-pV(TT+d)Z optimized geometries using EOM-CC3/aug-cc-pV(T+d)Z and MC-RPA(24,16)/aug-cc-pV(T+d)Z.
![]() | ||
| Fig. 5 Simulated spectra for cis-S4 and trans-S4, compared with the scaled experimental spectrum of S4 by Billmers and Smith,50 measured at temperatures in the range of 673–903 K. The experimental spectrum was digitized using the PlotDigitizer134 tool. The experimental spectrum has been scaled up by 400% for clarity. Note the different scales on the axes. | ||
![]() | ||
| Fig. 6 Simulated spectra for cyclic-S3 (S) and cyclic-S4, compared with the experimental spectrum of S4 by Billmers and Smith,50 measured at temperatures in the range of 673–903 K. The experimental spectrum was digitized using the PlotDigitizer134 tool. Note the different scales on the axes. The experimental spectrum has been scaled down by 80% for clarity in the lower part of the figure. | ||
![]() | ||
| Fig. 7 Simulated spectra for trigonal-S4, compared with the experimental spectrum of S4 by Billmers and Smith,50 measured at temperatures in the range of 673–903 K. The experimental spectrum was digitized using the PlotDigitizer134 tool. | ||
The vertical excitation energies and oscillator strengths for the three lowest excited states of cis-S4 and trans-S4 are presented in the SI in Tables A5 and A6. It is clear that the vertical excitation energies calculated at the EOM-CCSD/aug-cc-pV(T+d)Z level of theory are in good agreement with those calculated using EOM-CC3/aug-cc-pV(T+d)Z and MC-RPA(24,16)/aug-cc-pV(T+d)Z. In contrast, TD-DFT energies differ by approximately 100 nm. We use the results from the MC-RPA(24,16)/aug-cc-pV(T+d)Z calculations and scale the EOM-CCSD/aug-cc-pV(T+d)Z spectra for both cis-S4 and trans-S4. These scaled spectra are also visualized in Fig. 5, and we recommend these as our highest quality UV-Vis spectra for cis-S4 and trans-S4 to be used in photochemical models. The EOM-CCSD/aug-cc-pV(T+d)Z spectra are recommended for the other singlet S4 isomers.
In Fig. 8 we show the simulated spectra for the spin triplet S4 species cis-3S4 and trans-3S4. In the case of cis-3S4, simulated spectra are similar, except for the ωB2PLYP/aug-cc-pV(T+d)Z spectrum. The ωB2PLYP/aug-cc-pV(T+d)Z spectrum is more uniform, while the other simulated spectra show two different peaks. The ωB2PLYP/aug-cc-pV(T+d)Z spectrum also predicts a larger absorption cross section. For trans-3S4, all simulated spectra are similar. Furthermore, as can be seen and Fig. A10 and A11 in the SI, the simulated spectra predict that the triplet S4 species weakly absorb in the near-infrared region.
Of the S4 isomers we have studied in this work, based on the simulated spectra, we find that trigonal-S4 absorbs in the same wavelength window as the UA. Furthermore, the triplet structures cis-3S4 and trans-3S4 also show some absorption between 320 and 400 nm. However, cis-3S4 and trans-3S4 are very unlikely to contribute to the enigmatic absorption, but they might be a useful addition as intermediates in photochemical models. Zhang et al.37 estimated a mixing ratio of approximately 0.1 pptV using their Venusian atmospheric model. However, the reaction rates are again based on estimates.133 In addition, atmospheric models do not differentiate between the S4 species.36–41 Based on the PES in Fig. 4, the S4 isomers are separated by quite high barriers, meaning that several isomers could form and persist. Therefore, individual reaction rates for the formation and consumption of the different isomers should be determined.
![]() | ||
| Fig. 10 The simulated spectra for chair-S6, compared to the experimental spectrum by Steudel et al.53 measured in a methylcyclohexane (C7H14) solution. The experimental spectrum was digitized using the PlotDigitizer134 tool. | ||
![]() | ||
| Fig. 11 The simulated spectra for the two lowest energy S7 conformers, chair-S7 and boat-S7, compared to the experimental spectrum by Steudel et al.53 measured in a methylcyclohexane (C7H14) solution. The experimental spectrum was digitized using the PlotDigitizer134 tool. Note the different scales on the axes. | ||
![]() | ||
| Fig. 12 The simulated spectra for crown-S8, compared to the experimental spectrum by Steudel et al.53 measured in a methylcyclohexane (C7H14) solution. The experimental spectrum was digitized using the PlotDigitizer134 tool. | ||
The cyclic-S5 structure is clearly the most stable S5 structural isomer, being 20 kcal mol−1 lower in energy than the next lowest energy S5, which is 3S5 (B). To the best of knowledge, no experimental UV-Vis spectrum for S5 has been published. As seen in Fig. 9, three of the TD-DFT methods, BHandLHYP/aug-cc-pV(T+d)Z, CAM-B3LYP/aug-cc-pV(T+d)Z, and ωB97X-D3/aug-cc-pV(T+d)Z, predict similar absorption profiles for cyclic-S5, with a maximum around 290 nm. The ωB2PLYP/aug-cc-pV(T+d)Z spectrum differs slightly from the others. Although there is a clear shoulder around 290 nm, the maximum is predicted to be at 265 nm.
Like cyclohexane, there are two different conformers of cyclic S6. Of these two, the chair conformer is the minimum, as the boat conformer is calculated to be 13.2 kcal mol−1 higher in energy at the CCSD(T)//ωB97X-D/aug-cc-pV(T+d)Z level of theory. This is in excellent agreement with the energy difference reported by Cioslowski et al.68 Other studies also report chair-S6 to be the minimum energy structure.66,71 The experimental spectrum of S6 by Steudel et al.53 seen in Fig. 10 shows a steady increase in the absorption cross section when going below 330 nm. All simulated spectra predict a lower absorption cross section than the experiment, but the steady increase is reproduced.
For S7, the general features of the experimental spectrum are similar to that of S6; at wavelengths shorter than 340 nm, the absorption cross section steadily increases. Spectral simulations were performed for two structures, these being chair-S7 and boat-S7, as the energy difference is only 4.2 kcal mol−1 at the CCSD(T)//ωB97X-D/aug-cc-pV(T+d)Z level of theory – an energy gap consistent with previous results.66,68,71 The simulated spectra in Fig. 11 show a lower absorption cross section compared to the experimental counterpart, but once again the general shape is reproduced. The overall absorption cross section of the boat-S7 conformer is predicted to be slightly higher than that of the chair-S7 conformer, even surpassing the experimental value below 270 nm.
There are several unique cyclic S8 structures. Our calculations show that the lowest energy structure is crown-S8, which is consistent with previously reported results.66–68,70,71 Contrary to the findings by Wong et al.,70 we find that the twisted-S8 is slightly lower in energy compared to the exo–endo-S8 conformer. Using simulated annealing, Jones and Hohl135 reported a S8 cluster (see Fig. 1b in ref. 135) that was only 9.9 kcal mol−1 higher in energy than crown-S8. This cluster was also reported by Wong et al.70 and Jones and Ballone.71 However, using ωB97X-D/aug-cc-pV(T+d)Z, this cluster optimizes to the S8 (E) structure, which is 18.3 kcal mol−1 higher in energy than crown-S8 at the CCSD(T)//ωB97X-D/aug-cc-pV(T+d)Z level of theory. The experimental spectrum for S8 is presented in Fig. 12, together with the simulated spectra for crown-S8. The simulated spectra predict a lower absorption cross section compared to the experimental spectrum between 320–270 nm, but reach agreement at 260 nm.
An important detail with respect to the experimental UV-Vis spectra published by Steudel et al. used for S6–S8 is that these are liquid phase spectra, measured in methylcyclohexane (C7H14). Different effects, such as clustering or solvent interactions, might affect the final spectrum. In contrast, our simulated spectra are in the gas phase. Therefore, the experimental spectra are not directly comparable to our spectra. We consider our simulated spectra to be a more appropriate representation of gas phase spectra. For the S5–S8 species, we recommend using the spectra simulated at the ωB97X-D3/aug-cc-pV(T+d)Z level of theory.
The experimental spectra S6–S8 reach down to approximately 220 nm. To construct the simulated spectra, we used the ten lowest excited states. This number of excited states limits us to about 250–260 nm. Increasing the number of excited states would extend the simulated spectra, but this was prohibitively expensive and thus not done in this work.
We estimated if spin–orbit coupling (SOC) would influence the simulated spectra in any significant way. We performed single-point vertical excitation calculations using the ωB97X-D3 functional with the aug-cc-pV(T+d)Z basis set on the ωB97X-D/aug-cc-pV(T+d)Z optimized geometries [specifically, linear-S3, cyclic-S3, cis-S4, trans-S4, cyclic-S5 (A), chair-S6 (A), chair-S7 (A), boat-S7 (B), and crown-S8 (A)]. The calculations were performed with the Orca 6.0.1 program, where SOC is implemented for TD-DFT.136 We saw shifts of less than 0.02 eV when comparing the calculated singlet transitions without SOC with those calculated including SOC. Furthermore, while including SOC allowed some triplet transitions to have nonzero oscillator strengths, the oscillator strengths were at least two orders of magnitude lower when compared to the singlet transitions. We therefore conclude that the effect of SOC is negligible.
Although S8 was suggested as a possible UA candidate,31 based on our simulated gas phase spectra crown-S8 has no appreciable absorption cross section in the UA wavelength range of 320–400 nm. We therefore refute gas phase S8 as an UA candidate.
Our results show that polysulfurs display great structural diversity. The lowest energy S2–S4 structures tend to be linear, while cyclic structures are the lowest energy structures for the larger S5–S8 species. Furthermore, we identified two low-energy triplet S4 species, cis-3S4 and trans-3S4, that could be important intermediates in polysulfur reactions.
We recommend using the MC-RPA scaled EOM-CCSD/aug-cc-pV(T+d)Z spectra for linear-S3, cis-S4, and trans-S4 when adding these species to photochemical models. For the other S3 and S4 species, the EOM-CCSD/aug-cc-pV(T+d)Z spectra are recommended. The ωB97X-D3/aug-cc-pV(T+d)Z spectra are recommended for the S5–S8 species. Finally, for S2 we recommend using previously published experimental gas phase spectra or high-level simulated spectra.48,60,61
Based on the simulated spectra, we find two structures that absorb in the same wavelength region (320–400 nm) as the unknown absorber in the Venusian atmosphere, these being linear-S3 and trigonal-S4. However, the low concentrations predicted by Venusian atmospheric models makes it unlikely that polysulfurs are a major contributor to the enigmatic absorption. A conclusive assessment could be possible by investigating the accuracy of the polysulfur reaction rates used in Venusian atmospheric models, either using experimental methods or through high-level ab initio calculations.
For future work, the simulated spectra for the S5–S8 species should be extended down to 200 nm, as the high spectral resolution channel of the VenSpec-U UV-Vis spectrometer onboard the EnVision spacecraft reaches down to 205 nm.137 Finally, with the upcoming mission to Venus by NASA (DAVINCI138 and VERITAS139) and ESA (EnVision140), investigating sulfur chemistry in the context of Venus is vital for correctly interpreting the data these missions will produce in the coming decade.
Data supporting the findings in this article can be found in the SI. The simulated spectra for all species are included in the Zenodo archive DOI: https://doi.org/10.5281/zenodo.16361268.
| This journal is © the Owner Societies 2025 |