Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

A study of the atmospherically relevant reaction between dimethyl sulphide (DMS) and Cl2 in the absence and presence of water using electronic structure methods

Lydia Rhyman *ab, Edmond P. F. Lee c, Ponnadurai Ramasami *ab and John M. Dyke *c
aComputational Chemistry Group, Department of Chemistry, Faculty of Science, University of Mauritius, Réduit 80837, Mauritius. E-mail: lyd.rhyman@gmail.com; p.ramasami@uom.ac.mu
bCentre For Natural Product Research, Department of Chemical Sciences, University of Johannesburg, Doornfontein, Johannesburg 2028, South Africa
cSchool of Chemistry and Chemical Engineering, University of Southampton, Highfield, Southampton SO17 1BJ, UK. E-mail: jmdyke@soton.ac.uk

Received 2nd June 2025 , Accepted 10th July 2025

First published on 11th July 2025


Abstract

The thermodynamics and mechanisms of the atmospherically relevant reaction between dimethyl sulphide (DMS) and molecular chlorine (Cl2) were investigated in the absence and presence of a single water molecule, using electronic structure methods. Stationary points on the reaction surfaces were located using density functional theory (DFT) with the M06-2X functional and aug-cc-pVTZ (aVTZ) basis sets. Then single point energy calculations were carried out using the UM06-2X/aVTZ optimised stationary point geometries, with aug-cc-pVnZ basis sets (n = T and Q), using the domain-based local pair natural orbitals coupled cluster [DLPNO-UCCSD(T)] method, to give DLPNO-CCSD(T)/CBS//M06-2X/aVTZ relative energies. The reaction can proceed in three ways depending on the initial van der Waals complex formed i.e. via DMS + Cl2·H2O, DMS·H2O + Cl2, or DMS·Cl2 + H2O. It was found that based on computed equilibrium constants for complex formation and estimated initial concentrations of DMS, Cl2 and H2O in the atmosphere that [DMS·H2O] and [Cl2·H2O] are likely to be much greater than [DMS·Cl2] under atmospheric conditions. It was found that both with and without water the reaction can proceed by two pathways (i) formation of the products CH3SCH2Cl + HCl + (H2O) via a covalently bound intermediate (CH3)2SCl2(H2O) and (ii) formation of the products via a cis-CH3SClCH2:HCl (H2O) intermediate, where (H2O) applies to the with-water case. Although the pathways and mechanisms are similar in the without- and with-water cases, the relative energies of the transition states are significantly lower and the potential energy diagram is much more complex in the with-water case. However, under tropospheric conditions the overall DMS + Cl2 rate coefficient is unlikely to be affected by the presence of water as the concentrations of DMS·H2O and Cl2·H2O are estimated to be much lower than the concentrations of DMS, Cl2 and H2O. This work extends our earlier study of the reaction of DMS with atomic chlorine (Cl) with and without water (L. Rhyman et al., Phys. Chem. Chem. Phys. 2023, 25, 4780–4793).


1. Introduction

In this work the atmospherically important reaction between molecular chlorine and dimethyl sulphide (DMS), reaction (1) is studied,
 
CH3SCH3 + Cl2 → CH3SCH2 + HCl(1)
in the absence and presence of a single water molecule.

The sulphur cycle in the earth's atmosphere has been the subject of intensive investigation in recent years because of the need to assess the contribution of anthropogenically produced sulphur to acid rain, visibility reduction and climate modification.1–4 Anthropogenic emissions of sulphur to the atmosphere are dominated by SO2 whereas natural (biogenic) sulphur emissions are thought to be dominated by DMS derived from oceanic phytoplankton initiated by ultraviolet radiation from the sun.5–8 At present, anthropogenic emissions of sulphur dominate; however, these emissions are predominantly in the northern hemisphere. In the southern hemisphere, and in particular from the southern oceans, natural emissions are extremely important. Typical day-time and night-time [DMS] levels in the troposphere are 120 and 50 pptv respectively.8

The main DMS oxidation reactions in the atmosphere are DMS + OH during the day and DMS + NO3 at night. Subsequent oxidation leads to formation of species such as SO2, H2SO4 and CH3SO3H (methane sulfonic acid or MSA). These species may contribute significantly to the acidity of the atmosphere and in the case of sulphuric acid to cloud formation. Molecular chlorine has been observed in coastal marine air. This is produced at night, as well as during the day, from heterogeneous reactions of ozone with wet sea-salt and is enhanced by the presence of ferric ions.9–13 Night-time Cl2 mixing ratios are higher than those during the day because of Cl2 photolysis during the day. Comparison of recent experimentally measured Cl2 concentrations in the lower atmosphere shows that typical day-time and night-time [Cl2] levels are 5 and 50 pptv, respectively.14–19 The DMS + Cl2 reaction, which occurs mainly at night, provides another route for DMS loss and hence SO2 production. This could contribute to explaining the discrepancy between known DMS decay rates and observed SO2 production rates. Water is the third most abundant species in the troposphere behind only N2 and O2 with concentrations of up to 7.4 × 1017 molecules cm−3 (100% relative humidity; 0.03 atm.).12 It has been demonstrated that water can change the rate coefficient of a reaction, by forming complexes with the reagents, products and transition states and lower their energies, and in this way, the activation barrier for a reaction may be reduced.20–27

We have previously studied the reaction of DMS with molecular chlorine (DMS + Cl2) using UV photoelectron spectroscopy (PES), infrared matrix isolation spectroscopy and electronic structure calculations.28–30 It was found that this reaction proceeds through the formation of a covalent reaction intermediate ((CH3)2SCl2), in which sulphur is four coordinate. This then decomposes into the final products, monochlorodimethylsulphide (CH3SCH2Cl) and hydrogen chloride (HCl). Also, using PES as the detection technique, the room temperature rate coefficient of DMS + Cl2 has been measured as (3.4 ± 0.7) × 10−14 cm3 molecule−1 s−1,28 four orders of magnitude lower than the DMS + Cl room temperature rate coefficient. One objective of our studies on DMS reactions of atmospheric importance is to investigate the effect of water on the reactions DMS + Cl and DMS + Cl2 using electronic structure methods. The DMS + Cl reaction has recently been investigated in this way.31 It was found that this reaction proceeds via four channels and, although water changes the mechanisms of these channels significantly, the presence of water was found not to affect the overall reaction rate coefficient under atmospheric conditions. Following on from this study, the aim of this present work is to study the reaction DMS + Cl2 with and without water using electronic structure methods to establish the reaction energetics and mechanisms, and to determine if the energies of the transition states relative to the reagents are changed significantly when a single water molecule is present. In the earlier, mainly experimental studies,28–30 a schematic potential energy diagram was constructed for the DMS + Cl2 reaction. Minimum energy structures and transition states were located at the MP2/aug-cc-pVDZ level. In the present work, electronic structure calculations are performed at a higher level on the DMS + Cl2 reaction in the absence and presence of a single water molecule to obtain improved relative energies, thermochemical values and structures of the energy minima and transition states.

2. Computational details

Electronic structure calculations were carried out to optimise the geometries of the reactants, reactant complexes, transition state structures (TSs), product complexes and products. As in our previous DMS + Cl work,31 the M06-2X functional was used with aug-cc-pVTZ (aVTZ) basis set.32 These computations were performed in the spin unrestricted formalism. The M06-2X functional was chosen because it has been shown to perform particularly well for TS structures and reaction barrier heights in benchmark studies on main group compounds.33–37 Harmonic vibrational frequencies were calculated to verify the nature of the stationary points (minima and TSs), and to provide the zero-point energy (ZPE) and the thermodynamic contributions to the enthalpy and free energy changes. Intrinsic reaction coordinate (IRC) calculations were also performed to ensure that a given TS connects with the desired minima.38,39

Fixed point calculations were carried out using the UM06-2X/aVTZ optimised stationary point geometries, with the aug-cc-pVnZ basis sets (n = T and Q), using the domain-based local pair natural orbitals coupled cluster (DLPNO-UCCSD(T)) approach.40–44 These were DLPNO-UCCSD(T)/aVnZ//UM06-2X/aVTZ (n = T and Q) calculations; the total energies obtained were extrapolated to the CBS limit employing the two parameter formula (eqn (2)):-45,46

 
E(x) = ECBS + Ax−3(2)
where x = 3 (aVTZ) and 4 (aVQZ).

This DLPNO-CCSD(T) method employs localised orbitals and obtains the correlation energy as a sum over the correlation energies of electron pairs. It recovers a large part of the CCSD(T) correlation energy at low computational cost.

All DFT computations were carried out using Gaussian 1647 running on SEAGrid.48–51 The DLPNO-CCSD(T) single point calculations were performed using the ORCA package.52,53

3. Results and discussion

When one water molecule is added to DMS + Cl2, the reaction could proceed in three ways depending on the initial (van der Waals) complex formed i.e. DMS + Cl2·H2O, DMS·H2O + Cl2, DMS·Cl2 + H2O. The computed minimum energy structures of the three van der Waals complexes Cl2·H2O, DMS·H2O, DMS·Cl2 at the M06-2X/aVTZ level are shown in Fig. 1.
image file: d5cp02065d-f1.tif
Fig. 1 Optimised minimum energy geometries of Cl2·H2O, DMS·H2O and DMS·Cl2.

It was found that a common feature of the DMS + Cl2 reaction with and without added water was that in each case the reaction can proceed along two pathways:- (i) the reactants to the covalently bound intermediate (CH3)2SCl2 (H2O) and on to the reaction products CH3SCH2Cl + HCl + (H2O) and (ii) on a separate surface, the reactants to cis-CH3SClCH2:HCl (H2O), and on to the products CH3SCH2Cl + HCl + (H2O), where (H2O) applies to the with-water case. The computed minimum energy structures of (CH3)2SCl2·H2O and cis-CH3SClCH2:HCl·H2O at the M06-2X/aVTZ level are shown in Fig. 2.


image file: d5cp02065d-f2.tif
Fig. 2 Optimised minimum energy geometries of (CH3)2SCl2·H2O and cis-CH3SClCH2:HCl·H2O.

The schematic potential energy diagram obtained at the DLPNO-CCSD(T)/CBS//M06-2X/aVTZ level for the DMS + Cl2 reaction, in the absence of water, is shown in Fig. 3. The M06-2X/aVTZ and DLPNO-CCSD(T)/CBS//M06-2X/aVTZ relative electronic energies (ΔE, including zero-point energy (ZPE) corrections), relative enthalpies (ΔHϕf,[thin space (1/6-em)]298K) and relative free energies (ΔGϕf,[thin space (1/6-em)]298K) of the energy minima and transition states located are shown in Table 1. Comparison of the results shown in Fig. 3 with the results obtained in the earlier, mainly experimental work28 at the MP2/aug-cc-pVDZ level (summarised in ref. 28) shows that the earlier results are incomplete with some stationary points not located, and the results are more approximate than the results presented here. In both cases pathways (i) and (ii) described above were identified. A summary of the relative energies of the common stationary points obtained in this work and in the lower level work of ref. 28 is given in Table 1. In this table, it can be seen that the present M06-2X/aVTZ and DLPNO-CCSD(T)/CBS results agree very well with each other, which supports their reliability, with the earlier lower-level MP2 energy values being mostly lower than the DLPNO-CCSD(T) values.


image file: d5cp02065d-f3.tif
Fig. 3 Energy profile for the reaction of DMS + Cl2 in the absence of water using the DLPNO-CCSD(T)/CBS//M06-2X/aVTZ method. DLPNO-CCSD(T)/CBS//M06-2X/aVTZ relative electronic energies (ΔE) including ZPE are shown in the figure.
Table 1 DMS + Cl2 relative electronic energies (ΔE, kcal mol−1 incl. ZPE contributions), relative enthalpies (in brackets, ΔHϕf,[thin space (1/6-em)]298K kcal mol−1)a and relative free energies (in italics, ΔGϕf,[thin space (1/6-em)]298K, kcal mol−1)
M06-2X/aVTZ DLPNO-CCSD(T)/CBS//M06-2X/aVTZ MP2/aug-cc-pVDZ (ref. 28)
a The thermal correction to the enthalpy obtained using the M06-2X/aVTZ method were added to the DLPNO-CCSD(T) single point energy (ZPE is included in the electronic energies). b Not reported in the literature, we located the TS3 using the MP2/aVDZ method.
DMS·Cl2 −5.5 −4.9 −9.47
(−5.1) (−4.6)
2.3 2.8
TS1 11.6 10.1 4.5
−10.8 −9.4
21.7 20.3
DMSCl2 −16.3 −17.7 −26.29
(−16.5) (−17.9)
−7.1 −8.5
TS2 6.5 6.1 2.75
(−5.9) (−5.5)
15.9 15.5
CH3SClCH2:HCl 1.6 −1.6 Not located
(−1.7) (−1.4)
9.7 6.6
TS3 18.9 16.0 8.35b
(−19.4) (−16.6)
25.3 22.4
CH3SCH2Cl:HCl −28.6 −29.5 −42.15
(−28.3) (−29.3)
−20.9 −21.8
CH3SCH2 Cl + HCl −27.6 −28.8 −32.06
(−27.3) (−28.4)
−27.3 −28.5
TS4 26.1 25.2 Not located
(−25.6) (−24.7)
34.8 36.0
DMSCl:Cl 9.1 7.7 2.43
(−8.8) (−7.4)
18.4 17.1
TS5 14.7 14.8 (14.19 quoted in ref. 28)
(−14.1) (−14.3) 11.38b
24.2 24.4 (TS3 of ref. 28)
cis-CH3SClCH2:HCl −3.9 −6.5 −14.69
(−3.8) (−6.5)
4.9 2.2
TS6 −1.7 −2.6 −10.83
(−2.2) (−3.0) (TS4 of ref. 28)
7.3 6.5
cis-CH3SCH2Cl:HCl −31.3 −32.2 −42.15
(−31.1) (−31.9)
−23.8 −24.7
TS7 −26.5 −27.1 Not located
(−26.9) (−27.6)
−17.8 −18.5


For formation of (CH3)2SCl2 from the reactants, quoting the higher level DLPNO ΔE values (with the lower level ΔE values from ref. 28 shown in italics in brackets), the reactants first form a van der Waals complex DMS·Cl2 (at −4.9 (−9.5) kcal mol−1) which converts via a transition state (TS1 at 10.1 (4.5) kcal mol−1) to (CH3)2SCl2 (at −17.7 (−26.3) kcal mol−1). In comparison, cis-CH3SClCH2:HCl (at −6.5 (−14.7) kcal mol−1) is formed from the van der Waals complex DMS·Cl2via passage over a TS (TS4 at 25.2 kcal mol−1 (not located in ref. 28)) to give DMSCl:Cl (at +7.7 (2.4) kcal mol−1) which converts via TS5 (at 14.8 (11.4) kcal mol−1) to give cis-CH3SClCH2:HCl (see Fig. 3). The route from the covalently bound intermediate (CH3)2SCl2 to the products is over TS2 (at 6.1 (2.7) kcal mol−1) to CH3SClCH2:HCl (at −1.6 kcal mol−1) (not located in ref. 28), and then over TS3 (at 16.0 (8.3) kcal mol−1) to the product complex (trans-CH3SCH2Cl:HCl at −29.5 (−42.1) kcal mol−1) and on to the products CH3SCH2Cl + HCl (at −28.8 (−32.0) kcal mol−1). The route from the intermediate cis-CH3SCH2Cl:HCl to the products is via TS6 (at −2.6 (−10.8) kcal mol−1) to cis-CH3SCH2Cl:HCl (at −32.2 (−42.1) kcal mol−1) then via TS7 (at −27.1 kcal mol−1 (not located in ref. 28)) to the product complex trans-CH3SCH2Cl:HCl (at −29.5 (−42.1) kcal mol−1) and on to the separate products CH3SCH2Cl + HCl (at −28.8 (−32.0) kcal mol−1). As can be seen from Fig. 3, for the pathway via (CH3)2SCl2, the highest barrier (rate determining) is TS3 (at 16.0 kcal mol−1) whereas for the pathway via cis-CH3SClCH2:HCl the highest barrier is TS4 (at 25.2 kcal mol−1).

The schematic reaction profiles of the DMS + Cl2 reaction in the presence of one water molecule are shown in Fig. 4 and 5. Fig. 4 shows profiles which involve reaction via the hydrated covalently bound intermediate (CH3)2SCl2·H2O (DMS·H2O·Cl2 in Fig. 4) and Fig. 5 shows profiles which involve reaction via the hydrated intermediate cis-CH3SClCH2:HCl (CH3SClCH2:HCl-4 in Fig. 5). Overall, most of the minima and transition states seen in Fig. 3 (in the absence of water) for pathways (i) and (ii) have counterparts in Fig. 4 and 5, respectively (with water present).


image file: d5cp02065d-f4.tif
Fig. 4 Energy profiles for the reaction of DMS + Cl2 + H2O via the hydrated covalently bound intermediate (CH3)2SCl2·H2O using the M06-2X/aVTZ and DLPNO-CCSD(T)/CBS//M06-2X/aVTZ methods. The M06-2X/aVTZ relative electronic energies (ΔE) including ZPE are reported in the figure, with the DLPNO-CCSD(T)/CBS//M06-2X/aVTZ values shown in brackets; values are in kcal mol−1. Note the DMS·Cl2 complex (at −4.9 kcal mol−1) connects with DMS·Cl2·H2O (at −9.4 kcal mol−1) but this only correlates with TS4·H2O-2 (20.1 kcal mol−1 in Fig. 5; DLPNO values quoted in brackets).

image file: d5cp02065d-f5.tif
Fig. 5 Energy profiles for the reaction of DMS + Cl2 + H2O using the M06-2X/aVTZ and DLPNO-CCSD(T)/CBS//M06-2X/aVTZ methods via the hydrated intermediate cis-CH3SClCH2:HCl. The M06-2X/aVTZ relative electronic energies (ΔE) including ZPE are reported in the figure, with the DLPNO-CCSD(T)/CBS//M06-2X/aVTZ values shown in brackets; values are in kcal mol−1.

As stated earlier, when one water molecule is added to DMS + Cl2, the reaction could proceed via DMS + Cl2·H2O, DMS·H2O + Cl2, DMS·Cl2 + H2O. As shown in Fig. 4 and 5, the relative energies of these, relative to DMS + Cl2 + H2O, are DMS + Cl2·H2O (−0.5 kcal mol−1) > DMS·H2O + Cl2 (−3.7 kcal mol−1) > DMS·Cl2 + H2O (−4.9 kcal mol−1) (all values quoted are the higher level DLPNO ΔE values). DMS + Cl2·H2O and DMS·H2O + Cl2 correlate with a DMS·H2O:Cl2 minimum at −8.5 kcal mol−1, whereas DMS·Cl2 + H2O correlates with another DMS·Cl2:H2O minimum at −9.4 kcal mol−1. As shown in Fig. 4, both minima (at −8.5 and −9.4 kcal mol−1) convert via TS1·H2O (at +6.1 kcal mol−1) to the solvated covalently bound intermediate (CH3)2SCl2·H2O (at −22.8 kcal mol−1). In contrast, on the cis-CH3SClCH2:HCl·H2O pathway (Fig. 5), only the DMS·H2O:Cl2 minimum at −8.5 kcal mol−1 converts via TS4·H2O-1 to a DMSCl·ClH2O minimum energy structure at +2.3 kcal mol−1 which converts via TS5·H2O-1 (at +4.7 kcal mol−1) to a hydrated cis-CH3SClCH2·HCl minimum at −12.2 kcal mol−1 and the DMS·Cl2:H2O minimum at −9.4 kcal mol−1 converts via TS4·H2O-2 (at +20.1 kcal mol−1) to a DMSCl·H2O:Cl structure at +2.9 kcal mol−1 which converts via TS5·H2O-2 at 11.3 kcal mol−1 to CH3SClCH2:HCl·H2O-3 at −12.0 kcal mol−1. For the pathway via the intermediate (CH3)2SCl2 in Fig. 3 each stationary point has an equivalent in Fig. 4. Significantly, taking the largest barriers in the no-water case, TS3 at +16.0 kcal mol−1 is lowered to TS3·H2O:HCl at −6.7 kcal mol−1 (although there is a higher TS3 on a separate surface; TS3:HCl·H2O at 10.7 kcal mol−1). Similarly, the cis-CH3SClCH2:HCl pathway, TS4 (Fig. 3) is lowered from +25.2 kcal mol−1 to TS4·H2O-1 at +6.3 kcal mol−1 (although there is a higher TS4 on a separate surface; TS4·H2O-2 at +20.1 kcal mol−1) and TS6 is lowered from −2.6 kcal mol−1 to TS6·H2O-1 at −11.1 kcal mol−1 (Fig. 5). Also, it can be seen that TS6·H2O-1 converts directly to a product complex (CH3SCH2Cl·HCl·H2O-3 at −35.5 kcal mol−1) (there is no TS7 to trans-CH3SCH2Cl:HCl as in the DMS + Cl2 case with no water).

The computed relative electronic energies ΔE, as well as the relative (ΔHϕf,[thin space (1/6-em)]298K) and (ΔGϕf,[thin space (1/6-em)]298K) values, for the minima and transition states located in the with-water case are listed in Table 2.

Table 2 DMS + Cl2 + H2O Relative electronic energies (ΔE, kcal mol−1, incl. ZPE contributions), relative enthalpies (in brackets, ΔHϕf,[thin space (1/6-em)]298K, kcal mol−1)a and relative free energies (in italic, ΔGϕf,[thin space (1/6-em)]298K, kcal mol−1)
M06-2X/aVTZ DLPNO-CCSD(T)/CBS//M06-2X/aVTZ
a The thermal correction to the enthalpy obtained using the M06-2X/aVTZ method were added to the DLPNO-CCSD(T) single point energy (ZPE is included in the electronic energies).
DMS·H2O + Cl2 −4.4 −3.7
(−4.5) (−3.8)
2.5 3.2
DMS + Cl2·H2O −0.8 −0.5
(−0.5) (−0.2)
3.3 3.5
DMS·Cl2 + H2O −5.5 −4.9
(−5.1) (−4.6)
2.3 2.8
DMS·H2O:Cl2 −9.6 −8.5
(−9.5) (−8.4)
6.9 8.0
DMS·Cl2:H2O −9.2 −9.4
(−8.8) (−9.1)
5.5 5.2
TS1·H2O 6.9 6.1
(−6.2) (−5.4)
24.9 24.1
DMS·H2O·Cl2 −22.5 −22.8
(−22.9) (−23.2)
−5 −5.3
TS2·H2O-1 3.3 5.7
(−1.7) (−4.1)
22.2 24.5
CH3SClCH2:HCl·H2O −4.4 −6.4
(−4.6) (−6.6)
12.2 10.1
TS3:HCl·H2O 12.8 10.7
(−13.1) (−10.9)
27.9 25.7
CH3SCH2Cl·HCl·H2O-1 −37.4 −38
(−37.6) (−38.3)
−23.2 −23.8
TS2·H2O-2 2.1 2.4
(−1.5) (−1.8)
19.6 19.9
CH3SClCH2·H2O:HCl-1 −5.7 −7.7
(−5.7) (−7.7)
10.6 8.6
CH3SClCH2·H2O:HCl-2 −11.0 −12.6
(−10.9) (−12.6)
5.6 3.9
TS3·H2O:HCl −6.0 −6.7
(−6.8) (−7.6)
11.7 10.9
CH3SCH2Cl·HCl·H2O-2 −35.3 −35.4
(−35.1) (−35.2)
−19.9 −20.0
TS4·H2O-1 6.3 6.3
(−5.0) (−5.0)
25.1 25.1
DMSCl:Cl·H2O 1.9 2.3
(−1.0) (−1.5)
19.9 20.4
TS5·H2O-1 3.6 4.7
(−2.8) (−3.8)
21.6 22.6
CH3SClCH2:HCl·H2O-3 −10.6 −12
(−10.9) (−12.3)
6.4 5
TS6·H2O-1 −10.9 −11.1
(−11.8) (−12.0)
7.0 6.8
CH3SCH2Cl·HCl·H2O-3 −35.3 −35.5
(−35.1) (−35.3)
−20.0 −20.2
TS4·H2O-2 20.3 20.1
(−19.9) (−19.7)
36.8 36.5
DMSCl·H2O:Cl 3.5 2.9
(−3.3) (−2.7)
20.5 19.9
TS5·H2O-2 10.4 11.3
(−10.0) (−10.9)
27.2 28
CH3SCH2Cl·HCl·H2O-4 −10.8 −12.2
(−10.9) (−12.3)
6.2 4.8
TS6·H2O-2 −9.5 −9.3
(−10.2) (−10.0)
8.0 8.2
CH3SCH2Cl·HCl·H2O-5 −35.8 −36
(−35.5) (−35.7)
−21.2 −21.4
CH3SCH2Cl + HCl·H2O −31.5 −32.7
(−31.7) (−32.9)
−25.6 −26.7
CH3SCH2Cl + HCl + H2O −27.6 −28.8
(−27.3) (−28.4)
−27.3 −28.5


As already described, the minimum energy geometries and formation energies (ΔE) relative to their reagents were computed for the van der Waals complexes DMS·H2O, Cl2·H2O and DMS·Cl2 at the UM06-2X/aVTZ and DLPNO-CCSD(T)/CBS//M06-2X/aVTZ levels. Also, the standard free energies (ΔGϕf,[thin space (1/6-em)]298K) and equilibrium constants (Keq, 298 K) were computed for the formation of these complexes. This was done in order to determine the approximate relative concentrations of these complexes under typical tropospheric conditions. The computed ΔE, ΔGϕf,[thin space (1/6-em)]298K and Keq values are shown in Table 3 at the two levels of theory used. The following typical day-time/night-time concentrations of DMS, Cl2 and H2O were assumed, using values mentioned earlier.

Table 3 Computed reaction energy (ΔE, kcal mol−1) without ZPE correction, reaction free energy (ΔG298Kϕ, kcal mol−1) and equilibrium constant (Keq) at 298 K for formation of (a) DMS·H2O, (b) Cl2·H2O and (c) DMS·Cl2
(a) (a) (a) (b) (b) (b) (c) (c) (c)
ΔE ΔG298Kϕ K eq ΔE ΔG298Kϕ K eq ΔE ΔG298Kϕ
UM06/2X/aVTZ −5.86 2.54 1.4 × 10−2 −1.43 3.26 4.1 × 10−3 −5.98 2.27 2.2 × 10−2
DLPNO-CCSD(T)/CBS//M06-2X/aVTZ −5.17 3.22 4.3 × 10−3 −1.16 3.52 2.6 × 10−3 −5.44 2.81 8.7 × 10−3


[DMS] 120/50 ppt (2.95 × 109/1.23 × 109 molecules cm−3), [Cl2] 5/50 ppt (1.23 × 108/1.23 × 109 molecules cm−3) and [H2O] 7.38 × 1017/7.38 × 1017 molecules cm−3 (ref. 8, 12, 14–19 and 31) With these estimated concentrations and the computed equilibrium constants for complex formation (Table 3), the estimated concentrations of the complexes at 298 K in the troposphere were in the order (Table 4) (a) [DMS·H2O] > (b)[Cl2·H2O] ≫ (c) [DMS·Cl2], with computed day-time/night-time values (in molecules cm−3, at the DLPNO-CCSD(T)/CBS//M06-2X/aVTZ level) of [DMS·H2O] =3.8 × 105/1.6 × 105, [Cl2·H2O] = 9.6 × 103/9.6 × 104, [DMS·Cl2] = 1.3 × 10−4/5.3 × 10−4 (see Table S2, ESI).

Table 4 Estimated day-time and night-time concentration ratios of [DMS·H2O], [Cl2·H2O] and [DMS·Cl] in the troposphere obtained using the appropriate Keq values in Table 3 and estimated concentrations in Table S1 (ESI)
UM06-2X/aVTZ DLPNO-UCCSD(T)/CBS//UM06-2X/aVTZ UM06-2X/aVTZ DLPNO-UCCSD(T)/CBS//UM06-2X/aVTZ
Day-time Day-time Night-time Night-time
[DMS·H2O]/[Cl2·H2O] 0.82 × 102 0.39 × 102 3.4 1.58
[DMS·H2O]/[DMS·Cl2] 3.8 × 109 2.9 × 109 3.8 × 108 2.9 × 108
[DMS·Cl2]/[Cl2·H2O] 2.1 × 10−8 1.3 × 10−8 0.86 × 10−8 0.58 × 10−8


Clearly, DMS·H2O is the dominant complex under atmospheric conditions with Cl2·H2O slightly lower and DMS·Cl2 much lower. [DMS·H2O] and [Cl2·H2O] are comparable at night but [Cl2·H2O] is an order of magnitude lower than [DMS·H2O] during the day. However, given that the estimated concentrations of DMS·H2O and Cl2·H2O are much lower than the estimated concentrations of DMS, Cl2 and H2O, water will have only a minor effect on the overall rate coefficient under typical tropospheric conditions.

To demonstrate how these numbers were derived, the day-time [DMS·H2O] is taken as an example. For the reaction

 
DMS + H2O → DMS·H2O(3)
the equilibrium constant, Keq at 298 K, has been calculated as 4.3 × 10−3 at the DLPNO-CCSD(T)/CBS//M06-2X/aVTZ level (see Table 3). Then
 
Keq = [DMS·H2O]/Pϕ/(([DMS]/Pϕ) [H2O]/Pϕ)(4)
 
= [DMS·H2O]Pϕ/([DMS]·[H2O])(5)
where Pϕ is the standard pressure of 1 atmosphere (2.46 × 1019 molecues cm−3).

Using the day-time values for [DMS] and [H2O] listed in Table S1 (ESI) (and quoted above), with Keq= 4.3 × 10−3 gives [DMS·H2O] = 3.80 × 105 molecules cm−3.

As already stated, for formation of DMS·H2O, Cl2·H2O and DMS·Cl2 the computed KeqGϕf,[thin space (1/6-em)]298K) values at the highest level (DLPNO-CCSD(T)/CBS) are 4.3 × 10−3 (3.22 kcal mol−1), 2.6 × 10−3 (3.52 kcal mol−1) and 8.7 × 10−3 (2.81 kcal mol−1) (Table 3). These Keq values need to be larger (i.e. the ΔGϕf,[thin space (1/6-em)]298K values need to be more negative) in order for water to have any significant effect on the overall observed reaction rate. In order for the concentrations of DMS·H2O and Cl2·H2O to be comparable with those of DMS and Cl2, the DLPNO-CCSD(T)/CBS Keq values for formation of these two complexes would have to increase by at least 3 orders of magnitude (ΔGϕf,[thin space (1/6-em)]298K values for these complex formation reactions would have to decrease by ∼4.2 kcal mol−1 to ∼−1.0 kcal mol−1). It is important to note that for these complexation reactions, ΔS is negative and hence –TΔS makes a positive contribution to ΔG (in ΔG = ΔHTΔS). Hence, a drop in temperature with altitude, as occurs in the troposphere up to ∼15 km at the tropopause (typically from ∼298 to ∼220 K) will give rise to more negative values of ΔG. However, it can be seen from eqn (5) that [DMS·H2O] depends on [H2O]. Similarly [Cl2·H2O] depends on [H2O]. Balloon-borne infrared emission measurements54 show that water concentrations show a significant decrease with altitude being highest in the first 2 km of the atmosphere. At 3 km (typical temp. Ttyp = 284 K) the [H2O] values are approximately half of the 1 km (Ttyp = 287 K) values and at 5 km (Ttyp = 280 K) [H2O] values are an order of magnitude lower than the 1 km value. Hence although the temperature drop with height would favour more negative ΔG values for the complexation reactions (a), (b) and (c) (and hence give larger Keq values and larger complex concentrations), the large decrease of [H2O] with height up to the tropopause, particularly in the first 5 km, is more significant and will dominate over the effect of a decrease in temperature, giving lower complex concentrations at heights greater than 2 km than those below 2 km. This is illustrated by the calculated values of [DMS·H2O] at the altitudes of 0, 1, 3 and 5 km of 3.8 × 105, 2.4 × 105, 1.4 × 105 and 0.25 × 105 molecules cm−3 obtained with [H2O] values at these altitudes from ref. 54 of 7.38 × 1017, 3.49 × 1017, 1.92 × 1017 and 0.31 × 1017 molecules cm−3 and computed DLPNO-CCSD(T)/CBS Keq values at the temperatures at these altitudes of 4.3 × 10−3, 5.9 × 10−3, 6.3 × 10−3 and 6.9 × 10−3.

As can be seen from Fig. 3–5, on including water in the DMS + Cl2 reaction, the energy profiles are changed significantly. For example, for pathway (i) via the intermediate (CH3)2SCl2, TS1 (10.1 kcal mol−1) and TS3 (16.0 kcal mol−1) are lowered significantly in the presence of water to TS1·H2O (6.1 kcal mol−1) and TS3·H2O·HCl (−6.7 kcal mol−1) respectively and there are two routes to the products from TS1·H2O (6.1 kcal mol−1) via TS3·H2O:HCl (−6.7 kcal mol−1) and via TS3:HCl·H2O (10.7 kcal mol−1). The highest barrier in pathway (i) in the lower energy route in the presence of water is TS1·H2O (6.1 kcal mol−1) compared to the highest barrier in the absence of water at TS1 (10.1 kcal mol−1). For pathway (ii) via a cis-CH3SClCH2:HCl intermediate the very high barrier at TS4 without water (25.2 kcal mol−1) is lowered to TS4·H2O-1 (6.3 kcal mol−1), which is the highest barrier for this pathway. This is comparable in energy to the highest barrier of the lower energy route for pathway (i), via TS1·H2O (6.1 kcal mol−1). Hence in the presence of water both pathways (i) and (ii) contribute to the overall reaction whereas the absence of water only pathway (i) via the intermediate (CH3)2SCl2 contributes. (Cartesian coordinates for all stationary points located are summarised in Table S3, ESI).

4. Conclusions

The present investigation of the DMS + Cl2 reaction at the DLPNO-CCSD(T)/CBS//M06-2X/aVTZ level provides complete energy level diagrams for the reaction pathways which were not present in the earlier, lower-level MP2/aug-cc-pVDZ calculations. Potential energy diagrams computed for this reaction in the absence and presence of one water molecule at this higher level indicate that two pathways are present in both cases; they are (i) formation of the products CH3SCH2Cl + HCl + (H2O) via a covalently bound intermediate (CH3)2SCl2·(H2O) and (ii) formation of the products via a cis-CH3SClCH2:HCl (H2O) intermediate where (H2O) applies to the with-water case. In general, the potential energy diagram is much more complex in the with-water case than in the no-water case and activation energy barriers are significantly reduced.

Three pathways were considered in the presence of water:-

DMS + Cl2·H2O, DMS·H2O + Cl2, DMS·Cl2 + H2O

It was found that the reaction can proceed via all three pathways. However, as it is expected that [DMS·Cl2] will be much lower than [DMS·H2O] and [Cl2·H2O] in the atmosphere, reaction via DMS·Cl2 + H2O is likely to be insignificant. Also, as the estimated concentrations of Cl2·H2O, DMS·H2O and DMS·Cl2 are lower than those of DMS, H2O and Cl2 in the lower troposphere under typical atmospheric conditions, H2O will have only a minor effect on the overall DMS + Cl2 rate coefficient in the atmosphere.

It is expected that in the absence of water pathway (i) will be dominant whereas in the presence of water both pathways (i) and (ii) will contribute to the overall reaction.

Conflicts of interest

There are no conflicts to declare.

Data availability

Data supporting the results and findings of this work are included in the manuscript and ESI.

Acknowledgements

This work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant number OCI-1053575. LR and PR would also like to acknowledge computing facilities from the Centre for High Performance Computing of South Africa. LR also acknowledges the postdoctoral fellowship from the Higher Education Commission (formerly known as Tertiary Education Commission) of Mauritius.

References

  1. R. J. Charlson, J. E. Lovelock, M. O. Andreae and S. G. Warren, Nature, 1987, 326, 655–661 CrossRef CAS.
  2. S. F. Watts, Atmos. Environ., 2000, 34, 761–779 CrossRef CAS.
  3. T. S. Bates, B. K. Lamb, A. Guenther, J. Dignon and R. E. Stoiber, J. Atmos. Chem., 1992, 14, 315–337 CrossRef CAS.
  4. H. Berresheim, P. H. Wine and D. D. Davis, in Composition, Chemistry and Climate of the Atmosphere, ed R. Nostrand, New York, 1995, p. 251 Search PubMed.
  5. S. E. Schwarz, Nature, 1988, 336, 441–445 CrossRef.
  6. C. F. Cullis and M. M. Hirschler, Atmos. Environ., 1980, 14, 1263–1278 CrossRef CAS.
  7. T. S. Bates, J. D. Cline, R. H. Gammon and S. R. Kelly-Hansen, J. Geophys. Res., 1987, 92, 2930–2938 CrossRef CAS.
  8. A. Bandy, D. C. Thornton, B. W. Blomquist, S. Chen, T. P. Wade, J. C. Ianni, G. M. Mitchell and W. Nader, Geophys. Lett., 1996, 23, 741–744 CrossRef CAS.
  9. M. O. Andrea and P. J. Crutzen, Science, 1997, 276, 1052–1058 CrossRef.
  10. R. Vogt, P. J. Crutzen and R. Sander, Nature, 1996, 383, 327–330 CrossRef CAS.
  11. B. J. Finlayson-Pitts, Anal. Chem., 2010, 82, 770–776 CrossRef CAS PubMed.
  12. B. Finlayson-Pitts and J. R. Pitts, Chemistry of the Upper and Lower Atmosphere, Academic Press, San Diego, CA, 2000 Search PubMed.
  13. J. D. Raff, B. Njegic, W. L. Chang, M. S. Gordon, D. Dabdub, B. Gerber and B. J. Finlayson-Pitts, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 13647–13654 CrossRef PubMed.
  14. C. B. Faxon, J. K. Bean and L. H. Ruiz, Atmosphere, 2015, 6, 1487–1506 CrossRef CAS.
  15. B. D. Finley and E. S. Saltzman, Geophys. Res. Lett., 2006, 33, L11809 CrossRef.
  16. M. J. Lawler, B. D. Finley, W. C. Keene, A. A. P. Pszenny, K. A. Read, R. von Glasow and E. S. Saltzman, Geophys. Res. Lett., 2009, 36, L08810 CrossRef.
  17. J. Liao, L. G. Huey, Z. Liu, D. J. Tanner, C. A. Cantrell, J. J. Orlando, F. M. Flocke, P. B. Shepson, A. J. Weinheimer, S. R. Hall, K. Ullmann, H. J. Beine, Y. Wang, E. D. Ingall, C. R. Stephens, R. S. Hornbrook, E. C. Apel, D. Riemer, A. Fried, R. L. Mauldin III, J. N. Smith, R. M. Staebler, J. A. Neuman and J. B. Nowak, Nat. Geosci., 2014, 7, 91–94 CrossRef CAS.
  18. T. P. Riedel, T. H. Bertram, T. A. Crisp, E. J. Williams, B. M. Lerner, A. Vlasenko, S. M. Li, J. Gilman, J. de Gouw, D. M. Bon, N. L. Wagner, S. S. Brown and J. A. Thornton, Environ. Sci. Technol., 2012, 46, 10463–10470 CrossRef CAS PubMed.
  19. C. W. Spicer, E. G. Chapman, B. J. Finlayson-Pitts, R. A. Plastridge, J. M. Hubbe, J. D. Fast and C. M. Berkowitz, Nature, 1998, 394, 353–356 CrossRef CAS.
  20. R. J. Busek, J. S. Francisco and J. M. Anglada, Int. Rev. Phys. Chem., 2011, 30, 335–369 Search PubMed.
  21. V. Vaida, J. Chem. Phys., 2011, 135, 020901 CrossRef PubMed.
  22. V. Vaida, H. G. Kjaergaard, P. E. Hulze and D. J. Donaldson, Science, 2003, 299, 1566–1568 CrossRef CAS PubMed.
  23. V. Vaida, H. G. Kjaergaard and K. J. Feierabend, Int. Rev. Phys. Chem., 2003, 22, 203–219 Search PubMed.
  24. K. Morokuma and C. Muguruma, J. Am. Chem. Soc., 1994, 116, 10316–10317 CrossRef CAS.
  25. R. S. Zhu and M. C. Lin, Chem. Phys. Lett., 2002, 354, 217–226 CrossRef CAS.
  26. R. R. Li, J. R. A. Gorse, M. C. Sauer and S. Gordon, J. Phys Chem, 1980, 84, 819–821 CrossRef.
  27. M. A. Allodi, M. E. Dunn, J. Livada, K. N. Kirshner and G. C. Shields, J. Phys. Chem. A, 2006, 110, 13283–13289 CrossRef CAS PubMed.
  28. J. M. Dyke, M. V. Ghosh, D. J. Kinnison, G. Levita, A. Morris and D. E. Shallcross, Phys. Chem. Chem. Phys., 2005, 7, 866–873 RSC.
  29. J. M. Dyke, M. V. Ghosh, M. Goubet, E. P. F. Lee and G. Levita, Chem. Phys., 2006, 324, 85–95 CrossRef CAS.
  30. S. Beccaceci, J. S. Ogden and J. M. Dyke, Phys. Chem. Chem. Phys., 2010, 12, 2075–2082 RSC.
  31. L. Rhyman, E. P. F. Lee, P. Ramasami and J. M. Dyke, Phys. Chem. Chem. Phys., 2023, 25, 4780–4793 RSC.
  32. R. A. Kendall, T. H. Dunning Jr. and R. J. Harrison, J. Chem. Phys., 1992, 96, 6796–6806 CrossRef CAS.
  33. Y. Zhao and D. G. Truhlar, Acc. Chem. Res., 2008, 41, 157–167 CrossRef CAS PubMed.
  34. J. J. Zheng, Y. Zhao and D. G. Truhlar, J. Chem. Theory Comput., 2009, 5, 808–821 CrossRef CAS PubMed.
  35. L. Goerigk and S. Grimme, Phys. Chem. Chem. Phys., 2011, 13, 6670–6688 RSC.
  36. X. F. Xu, I. M. Alecu and D. G. Truhlar, J. Chem. Theory Comput., 2011, 7, 1667–1676 CrossRef CAS PubMed.
  37. M. Bursch, J. M. Mewes, A. Hansen and S. E. Grimme, Angew. Chem., Int. Ed. 2022, 61, e202205735 Search PubMed.
  38. C. Gonzalez and H. B. Schlegel, J. Chem. Phys., 1989, 90, 2154–2161 CrossRef CAS.
  39. C. Gonzalez and H. B. Schlegel, J. Phys. Chem., 1990, 94, 5523–5527 CrossRef CAS.
  40. C. Riplinger and F. Neese, J. Chem. Phys., 2013, 138, 034106 CrossRef PubMed.
  41. C. Riplinger, B. Sandhoefer, A. Hansen and F. Neese, J. Chem. Phys., 2013, 139, 134101 CrossRef PubMed.
  42. C. Riplinger, P. Pinski, U. Becker, E. F. Valeev and F. Neese, J. Chem. Phys., 2016, 144, 024109 CrossRef PubMed.
  43. M. Saitow, U. Becker, C. Riplinger, E. F. Valeev and F. Neese, J. Chem. Phys., 2017, 146, 164105 CrossRef PubMed.
  44. Y. Guo, C. Riplinger, U. Becker, D. G. Liakos, Y. Minenkov, L. Cavallo and F. Neese, J. Chem. Phys., 2018, 148, 011101 CrossRef PubMed.
  45. T. Helgaker, W. Klopper, H. Koch and J. Noga, J. Chem. Phys., 1997, 106, 9639–9646 CrossRef CAS.
  46. A. Halkier, T. Helgaker, P. Jorgensen, W. Klopper, H. Koch, J. Olsen and A. K. Wilson, Chem. Phys. Lett., 1998, 286, 243–252 CrossRef CAS.
  47. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian16, Revision C.01, Gaussian, Inc., Wallingford CT, 2016 Search PubMed.
  48. S. Pamidighantam, E. Nakandala, C. Abeysinghe, S. R. Wimalasena, S. Yodage, S. Marru and M. Pierce, Int. Conf. Comput. Sci., 2016, 80, 1927–1937 Search PubMed.
  49. N. Shen, Y. Fan and S. Pamidighantam, J. Comput. Sci., 2014, 5, 576–589 CrossRef.
  50. R. Dooley, K. Milfeld, C. Guiang, S. Pamidighantam and G. Allen, J. Grid. Comput., 2006, 4, 195–208 CrossRef.
  51. K. Milfeld, C. Guiang, S. Pamidighantam and J. Giuliani, Proceedings of the 2005 Linux Clusters: The HPC Revolution, Apr. 2005 Search PubMed.
  52. F. Neese, Mol. Sci., 2012, 2, 73–78 CrossRef CAS.
  53. F. Neese, F. Wennmohs, U. Becker and C. Riplinger, J. Chem. Phys., 2020, 152, 224108 CrossRef CAS PubMed.
  54. L. Palchetti, G. Bianchini, B. Carli, U. Cortesi and S. Del Bianco, Atmos. Chem. Phys., 2008, 8, 2885–2894 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5cp02065d

This journal is © the Owner Societies 2025
Click here to see how this site uses Cookies. View our privacy policy here.