Next Article in Journal
The Dopaminergic Cells in the Median Raphe Region Regulate Social Behavior in Male Mice
Previous Article in Journal
Novel Ultrastructural Insights into the Clear-Cell Carcinoma of the Pancreas: A Case Report
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Chiral Thianthrenes

by
M. John Plater
* and
William T. A. Harrison
Department of Chemistry, University of Aberdeen, Meston Walk, Aberdeen AB24 3UE, UK
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(8), 4311; https://doi.org/10.3390/ijms25084311
Submission received: 13 March 2024 / Revised: 8 April 2024 / Accepted: 9 April 2024 / Published: 13 April 2024

Abstract

:
The absolute configuration and stability of two thianthrene chiral sulfoxides has been determined by means of X-ray single-crystal structure determinations. The analyses and configurations allow verification that the diastereomeric sulfoxides are stable in solution and are not interconverting, which has been suggested in some studies of sulfoxides. The two thianthrene sulfoxides have slightly different Rf values, which allowed their separation using flash chromatography on silica. The spots run back-to-back, which posed a challenge for their separation. The pure, separated compounds in solution remain as separate, single spots on a Thin Layer Chromatography (TLC) plate.

Graphical Abstract

1. Introduction

Stereogenic and configurationally stable sulfur atoms, such as sulfoxides and sulfinate esters, have been known for a century [1,2,3]. The earliest examples of optically active forms of sulfoxides were reported in 1926 [4,5] and by 1962, a method for making enantio-enriched isomers was known [6,7,8]. From this time, chiral sulfoxides have been of interest in asymmetric synthesis [9,10,11,12] and drugs [13,14]. Numerous reviews are available which cover their synthesis and properties [15,16,17,18,19,20,21]. Notable is the instability of some chiral sulfoxides, which rearrange into racemic sulfoxides and make correct enantiopurity analysis difficult [22,23,24,25,26,27,28,29]. Kagan initially reported this observation [23], but it was not accepted widely until later, making many observations on enantiopurity in the literature doubtful [15]. Racemic thianthrene sulfoxides are rare but some are known such as 1-methylthianthrene-5-oxide [30] and 4-methylthianthrene-5-oxide [30], 2,7-dimethylthianthrene-5-oxide [31], 1-p-tolylsulfanylthianthrene-10-oxide [32] and chlorpromazine-S-oxide [33]. Asymmetric sulfide oxidation is useful for forming chiral sulfoxides [34,35,36,37,38]. Thianthrene racemates with adjacent chiral centres have been formed by the addition of thianthrene cation radical salts to cycloalkenes and alkenes [39]. Also, regiospecific alkene amino functionalisation was achieved via an electrogenerated dielectrophile [40]. Our previous heterocyclic synthesis studies with 1,2-dinitro-4,5-difluorobenzene led to X-ray single-crystal determinations of phenazines, ref. [41] phenoxazines and phenothiazines [42,43]. Heterocyclic syntheses with this building block have been expanded here to explore the X-Ray single-crystal structures and stabilities of chiral, enantiomerically pure, thianthrene sulfoxides.

2. Results and Discussion

In this work, enantiopure sulfoxides of the thianthrene framework 3 are prepared for the first time. This has a folded butterfly shape that is expected to invert rapidly, but we anticipated that making a sulfoxide from one of the sulfur atoms would fix the conformation. Attaching a chiral centre would give two diastereoisomers with different physical and chemical properties, allowing them to be separated and characterised, provided that the sulfoxide is stable enough to hold its configuration. The configuration of a chiral sulfur atom can be assigned using a sub-rule of the Cahn–Ingold–Prelog (CIP) sequence rules [44,45]. The CIP rules are based on quadri-covalent asymmetric atoms, and a lone pair of electrons, on nitrogen or sulfur, has negligible mass. An imaginary or phantom atom, of low priority, is added on here. A configuration of either R or S can then be assigned (Figure 1).
These are the first thianthrenes prepared using 1,2-difluoro-4,5-dinitrobenzene, 5. The reaction of benzenedithiol 4 with compound 5 and Na2CO3 gave 2,3-dinitrothianthrene 6 (Figure 2). Oxidation with meta-chloroperbenzoic (mcpba) acid gave an enantiomeric mixture of sulfoxides 7. The oxidation of the second sulfur atom was not observed, suggesting that one sulfoxide deactivates the other sulfur atom. Treatment with (S)-phenylethylamine 8 displaced the 3-nitro group, which is conjugated to the electron-withdrawing sulfoxide group. The two nitro groups activate each other, but the sulfoxide provides additional activation to the 3-nitro group. Only products 9 and 10 were formed in equal amounts, which is reasonable given their similar framework. They eluted as two back-to-back spots on a TLC plate with an eluent of dichloromethane/ether (90:10). Using chromatography on flash silica, they were resolved and separated, and then characterised using NMR and X-Ray single-crystal structure determinations. The proton and 13C NMR data are virtually identical, making it impossible to distinguish the two diastereoisomers. Figure 3 shows the molecular structures of both diastereoisomers 9 and 10 side by side. The yields for each step are shown in Figure 2 by the side of the compound number. The yields are calculated as the ratio of the moles of dry product divided by the moles of starting material x 100 to convert this into a percentage. The maximum possible yield is 100%. A yield of 70–80% is very good and over 90% is excellent and very high. Yields of 37% and 30% are much lower and are workable over a short scheme, but a great deal of mass is lost, which makes it difficult to carry on with further steps. For example, 500 mg of compound 6 gave 191 mg of sulfoxide 7, a 37% yield. This work was challenging because of two low yields, the back-to-back TLC spots and split material. The remaining material in each reaction had no products which were easily isolated. It was mainly baseline and intractable. Compounds 67 and 910 are crystalline solids, which dried easily in air, but compound 11, an orange/red oil, had DCM in the proton and 13 carbon NMR. This evaporated after a beaker of it was left exposed to the atmosphere for two months and the spectra re-run.
The products were characterised using X-Ray single-crystal structure determinations. A guide to crystallography is available [46]. Compound 9 (Figure 3) crystallises in the trigonal space group R3 (No. 146). The dihedral angle between the C1–C6 and C7–C12 rings is 43.56 (5)°, resulting in a ‘butterfly’ conformation for the fused ring system; atoms S1, O1 and S2 deviate from the mean plane of C4/C5/C7/C8 by −0.746 (2), −0.142 (4) and −0.560 (2) Å, respectively. As expected, S1 is pyramidal [deviation from C5, C8 and O1 = −0.6872 (9) Å], the S1–O1 bond length is 1.4890 (14) Å, indicating a significant degree of double-bond character, and the C5–S1–C8 bond angle is 97.10 (7)°. These data compare with the corresponding S atom displacement, S–O separation and C–S–C angle of 0.684 (2) Å, 1.489 (6) Å and 96.9 (3)°, respectively, in the centrosymmetric co-crystal of thianthrene 5-oxide with 1,4-di-iodotetrafluorobenzene [47]. The dihedral angle between the C1–C6 ring and the pendant C15–C20 ring in 9 is 85.96 (6)°, and the C2–N1–C13–C14 torsion angle is 154.86 (16)°. An intramolecular N1–H1O2 hydrogen bond [HO = 2.02 (3) Å, N–HO = 133 (2)°] generates an S(6) ring. The absolute structure (C13 S, S1 R) is well established and consistent with the starting materials. No significant directional intermolecular interactions were identified in the extended structure.
Compound 10 (Figure 3) crystallises in the orthorhombic space group P212121 (No. 19). The C1–C6 and C7–C12 rings of the thianthrene fused ring system subtend a dihedral angle of 52.55 (5)°, and the equivalent angle between the C1–C6 and C15–C20 rings is 81.47 (6)°. Atoms S1, O1 and S2 deviate from the C4/C5/C7/C8 plane by 0.831 (3), 0.356 (5) and 0.663 (3) Å, respectively. The C2–N1–C13–C14 torsion angle is 160.95 (19)° and an intramolecular N1–H1O2 hydrogen bond [HO = 1.97 (3) Å, N–HO = 134 (2)°] occurs. Unlike compound 9, the N–H group in compound 10 also forms a weak intermolecular hydrogen bond to the sulfoxide O atom [HO = 2.57 (3) Å, N–HO = 136 (2)°], which results in [001] chains in the extended structure (Figure 4) [48]. The absolute structure is well established with C13 S and S1 S.
Although the dihedral angles between the aryl rings of the thianthrene ring systems are similar in compounds 9 and 10, the overall molecular conformations are quite different, as illustrated in an overlay plot [49] (Figure 5), which shows that the C7–C12 ring is ‘flipped’ up or down in the two structures due to the rigid, chiral, sulfoxide moiety. Both structures are well ordered with no suggestion of disorder.
Thianthrenes 9 and 10 were also prepared in similar yields from the mcpba oxidation of thianthrene 11 (Figure 6). This was prepared from 2,3-dinitrothianthrene 6 through the displacement of one of the nitro groups. One of the sulfur atoms in compound 11, conjugated to the amine group, is more electron-rich, so it might have oxidised more readily, but the oxidation was still a low-yielding reaction. Sulfoxides 9 and 10 formed in equal amounts, so the energy pathways to their formation, as two different diastereoisomers, must be similar.

3. Materials and Methods

IR spectra were recorded on a diamond-attenuated total reflection (ATR) Fourier transform infrared (FTIR) spectrometer (Thermo Fisher Scientific, Stafford House, Boundary Way, Hemel Hempstead, UK). Ultraviolet (UV) spectra were recorded using an Evolution, UV-Vis spectrometer with EtOH as the solvent (Thermo Fisher Scientific, Stafford House, Boundary Way, Hemel Hempstead, UK). The term sh means shoulder. 1H and 13C nuclear magnetic resonance (NMR) spectra were recorded at 400 and 100.5 MHz, respectively, using a Bruker 400 spectrometer (Research Complex at Harwell, Rutherford Appleton Laboratory, Harwell, Didcot, Oxon, UK). Chemical shifts, δ, are given in ppm and measured by comparison with the residual solvent. Coupling constants, J, are given in Hz. A broad signal is abbreviated as br. High-resolution mass spectra were obtained at the University of Wales, Swansea, using an Atmospheric Solids Analysis Probe (ASAP) (positive mode) instrument: Xevo G2-S ASAP (Waters Corporation, 34 Maple Street, Milford, MA, USA). Melting points were determined on a Cole-Palmer MP-200D Stuart digital melting point apparatus (9 Orion Court, Ambuscade Road, Colmworth Business Park, St Neots, Cambridgeshire, UK). All chemicals were purchased from Sigma-Aldrich, Gillingham, UK.The crystal structures of 9 (yellow plate 0.10 × 0.05 × 0.03 mm) and 10 (yellow slab, 0.12 × 0.10 × 0.03 mm) were established using intensity data collected on a Rigaku CCD diffractometer (Mo Kα radiation, λ = 0.71073 Å for 9 and Cu Kα radiation, λ = 1.54178 Å for 10) at 100 K. (Malvern Panalytical Ltd., Barn B, 2 Cygnus Business Park, Middle Watch, Swavesey, Cambridge, UK).
The structures were routinely solved by dual-space methods using SHELXT [50] and the structural models were completed and optimized by refinement against |F|2 with SHELXL-2018 [51]. The N-bound H atoms were located in difference maps and their positions were freely refined. The C-bound H atoms were placed geometrically (C–H = 0.95–1.00 Å) and refined as riding atoms. The methyl groups were allowed to rotate, but not to tip, to best fit the electron density. The constraint Uiso(H) = 1.2Ueq(carrier) or 1.5Ueq(methyl carrier) was applied in all cases. Full details of the structures and refinements are available in the deposited cifs.
Crystal data for 9: C20H16N2O3S2, Mr = 396.47, trigonal, space group R3 (No. 146), a = 20.3224 (4) Å, c = 11.3272 (2) Å, V = 4051.38 (17) Å3, Z = 9, T = 100 K, Mo Kα radiation, λ = 0.71073 Å, R(F) = 0.039 [7914 reflections with I > 2σ(I)], wR(F2) = 0.083 (9333 reflections), Flack absolute structure parameter = 0.04 (2), CCDC deposition number 2337820.
Crystal data for 10: C20H16N2O3S2, Mr = 396.47, orthorhombic, space group P212121 (No. 19), a = 7.97576 (14) Å, b = 13.9104 (3) Å, c = 16.8489 (3) Å, V = 1869.32 (6) Å3, Z = 4, T = 100 K, Cu Kα radiation, λ = 1.54178 Å, R(F) = 0.023 [3506 reflections with I > 2σ(I)], wR(F2) = 0.060 (3578 reflections), Flack absolute structure parameter = –0.025 (6), CCDC deposition number 2337821.

Synthesis

2,3-Dinitrothianthrene 6 4,5-Difluoro-1,2-dinitrobenzene 5 (567 mg, 2.8 mmol) in EtOH (30 mL) was mixed with benzene-1,2-dithiol 4 (395 mg, 2.8 mmol) and Na2CO3 (4.0 g) then stirred at 75 °C for 20 h. The mixture was added to water (200 mL) and allowed to stand for 1 h. This was filtered and air dried for 2 days to give a bright yellow precipitate of the title compound (791 mg, 93%) as yellow crystals, 167–168 mp °C (from dichloromethane:light petroleum ether). λmax (EtOH)/nm 243 (log ε 3.7) and 287 (3.5); νmax (diamond) (cm−1) 3089w, 1532s, 1517s, 1442s, 1427s, 1349s, 1330s, 1238s, 899s, 851s, 748s, 661w, 444s and 423s; δH (400 MHz; D6DMSO) 7.45 (1H, d, J = 2.0), 7.46 (1H, d, J = 2.0); 7.63 (1H, d, J = 2.0), 7.65 (1H, d, J = 2.0) and 8.44 (2H, s); δC (100.1 MHz; D6DMSO) 125.3, 129.6, 129.9, 131.9, 141.4 and 142.3; m/z (Orbitrap ASAP) 306.9839 (M+ + H, 100%) C12H6N2O4S2H requires 306.9847.
2,3-Dinitrothianthrene-S-oxide 7 2,3-Dinitrothianthrene 6 (500 mg, 1.63 mmol) in DCM (30 mL) was treated with meta-chloroperbenzoic acid (mcpba) (564 mg, 3.27 mmol) for 24 h at rt. The DCM layer was diluted with more DCM (70 mL), extracted with dilute KOH) (4 pellets of KOH dissolved in 200 mL of H2O), dried over MgSO4, then concentrated and purified by chromatography on silica. DCM eluted the title compound (191 mg, 37%) as a pale yellow powder, mp > 200 °C (from dichloromethane:light petroleum ether). λmax (EtOH)/nm 226 (log ε 2.9) and 335 (2.1); νmax (diamond) (cm−1) 3110w, 1530s, 1436s, 1356s, 1338s, 1069s, 1033s, 902s, 849s, 756s, 537s and 469s; δH (400 MHz; CDCl3) 7.67 (1H, t, J = 8.0 and 8.0), 7.79 (1H, t, J = 8.0 and 8.0), 7.91–7.94 (2H, m), 8.48 (1H, s) and 8.78 (1H, s); δC (100.1 MHz; CDCl3) 122.2, 125.3, 126.4, 126.7, 130.1, 130.5, 131.7, 136.6, 138.8, 141.3, 143.2 and 146.2; m/z (Orbitrap ASAP) 322.9879 (M+ + H, 100%) C12H6N2O5S2H requires 322.9796.
Thianthrene 9 and Thianthrene 10 Enantiomeric 2,3-dinitrothianthrene-S-oxide 7 (30 mg, 0.093 mmol) in EtOH (30 mL) and Et3N (28 mg, 0.28 mmol) were treated with (S)-phenylethylamine 8 (34 mg, 0.28 mmol) and heated under reflux for 24 h. After cooling the mixture was diluted with water (200 mL), treated with aqHCl (10 mL, 5M) and filtered giving a clear filtrate. The precipitate was dissolved in dichloromethane (100 mL), extracted with dilute aq HCl (30 mL, 1M) then water (100 mL), dried over MgSO4 and filtered. The two products were purified by chromatography on flash silica. They run close to each other on a TLC plate eluting with 10:90 ether/dichloromethane. The column was eluted with dichloromethane then 1:100 ether/dichloromethane eluted the title compound 9 (11 mg, 30%) as yellow crystals, mp 148–149 °C (from dichloromethane:light petroleum ether). λmax (EtOH)/nm 264 (log ε 4.2), 334 (3.5) and 412 (3.5); νmax (diamond) (cm−1) 3350w, 1597s, 1557s, 1474s, 1443s, 1418s, 1335s, 1278s, 1224s, 1115s, 1084s, 1033s, 967s, 908s, 836s, 753s, 699s, 535s and 466s; δH (400 MHz; CDCl3) 1.69 (3H, d, J = 7.0), 4.72 (1H, q, J = 7.0), 6.91 (1H, s), 7.311–7.42 (5H, m), 7.45 (1H, d, J = 8.0), 7.53 (1H, d, J = 8.0), 7.56 (1H, d, J = 8.0), 7.92 (1H, d, J = 8.0), 8.65 (1H, d, J = 8.0) and 8.71 (1H, s); δC (100.1 MHz; CDCl3) 24.8, 53.7, 114.0, 125.0, 125.5, 125.6, 126.9, 127.8, 127.9, 128.7, 129.1, 129.3, 130.5, 132.0, 137.6, 140.5, 142.2 and 144.8; m/z (Orbitrap ASAP) 397.0680 (M+ + H, 100%) C20H16N2O3S2 requires 397.0681. 1:100 Ether/dichloromethane eluted the title compound 10 (11 mg, 30%) as yellow crystals, mp 194–195 °C (from dichloromethane:light petroleum ether). λmax (EtOH)/nm 265 (log ε 4.2), 334 (3.5) and 412 (3.5); νmax (diamond) (cm−1) 3350w,1598s, 1558s, 1476s, 1279s, 1229s, 1112s, 1083s, 1032s, 969s, 910s, 832s, 753s, 703s, 598s, 534s and 464s; δH (400 MHz; CDCl3) 1.69 (3H, d, J = 7.0), 4.72 (1H, q, J = 7.0), 6.93 (1H, s), 7.32–7.45 (5H, m), 7.48 (1H, d, J = 8.0), 7.55–7.59 (2H, m), 7.93 (1H, d, J = 8.0), 8.64 (1H, d, J = 8.0) and 8.70 (1H, s); δC (100.1 MHz; CDCl3) 24.8, 53.7, 114.2, 124.9, 125.6, 127.1, 127.8, 127.9, 128.8, 129.0, 129.2, 130.5, 131.9, 137.6, 140.8, 142.4 and 144.8 (one peak is overlapping); m/z (Orbitrap ASAP) 397.0675 (M+ + H, 100%) C20H16N2O3S2 requires 397.0681.
2-Nitro-3-amino(phenylethyl)thianthrene 11 2,3-Dinitrothianthrene 6 (100 mg, 0.33 mmol) in EtOH (30 mL) was treated with (S)-phenylethylamine 8 (80 mg, 0.65 mmol) and Et3N (66 mg, 0.65 mmol). The mixture was heated for 12 h. On cooling it was diluted with water and dilute aq HCl (20 mL, 1.0M) then extracted into DCM (100 mL). It was dried over MgSO4 and evaporated to dryness then purified by chromatography on flash silica. Elution with Et2O: DCM (2:98) eluted the title compound (33 mg, 27%) as an orange oil, λmax (EtOH)/nm 257 (log ε 3.8), 348 (2.9) and 446 (3.0); νmax (diamond) (cm−1) 3366w, 1599s, 1551s, 1506s, 1467s, 1446s, 1331s, 1273s, 1218s, 1088s, 1029s, 970s, 904s, 835s, 726s, 698s, 5644s, 470s and 444s; δH (400 MHz; CDCl3) 1.54 (3H, d, J = 7.0), 4.56 (1H, q, J = 7.0), 6.68 (1H, s), 7.07–7.14 (2H, m), 7.17 (1H, t, J = 8.0 and 8.0), 7.22–7.28 (5H, m), 7.32 (1H, d, J = 8.0), 8.14 (1H, s) and 8.32 (1H, d, J = 5.0); δC (100.1 MHz; CDCl3) 24.8, 53.2, 113.8, 120.6, 125.6, 125.7, 127.7, 127.8, 128.1, 128.7, 128.8, 129.2, 131.3, 133.7, 135.1, 142.9, 143.6 and 146.7; m/z (Orbitrap ASAP) 381.0741 (M+ + H, 100%) C20H16N2O2S2H requires 381.0732.
Thianthrene 9 and Thianthrene 10 2-Nitro-3-amino(phenylethyl)thianthrene 11 (30 mg, 0.079 mmol) in DCM (30 mL) was treated with mcpba (27 mg, 0.16 mmol) at rt for 24 h. The mixture was extracted with dilute KOH (50 mL, 0.1M), washed with water (50 mL), dried over MgSO4, then purified by chromatography on flash silica. DCM eluted remaining starting material then 1% ether: DCM eluted firstly compound 9 (10 mg, 32%) of identical spectroscopic properties to that previously reported in this paper, followed by compound 10 (10 mg, 32%) of identical spectroscopic properties to that previously reported in this paper.

4. Conclusions

2,3-Dinitrothianthrene 6 was prepared by means of a novel one-pot condensation of benzenedithiol 4 with 4,5-difluoro-1,2-dinitrobenzene 5. The activated fluorine atoms are displaced preferentially to a nitro group by the thiol groups. After oxidation of just one diarylthio ether to a sulfoxide, a nitro group conjugated to the sulfoxide and another nitro group was displaced by the chiral amine (S)-phenylethylamine 8, chosen because of its availability and spatial difference. Two diastereomeric products 9 and 10 were produced, which ran back-to-back on a TLC plate eluting with dichloromethane/ether (98:2). These were separated using chromatography on flash silica and fully characterised. The absolute configuration of each diastereoisomer was established using X-ray crystal structure determinations. The more polar compound has a configuration 9 (SR) and the less polar spot has a configuration 10 (SS). These different configurations are apparent in the crystal structures as they flip the unsubstituted aryl ring in different orientations. The separation of the two diastereoisomers proves their stability as they are not interconverting and helps overcome concerns from others that sulfur-based chiral centres are not sufficiently stable for synthetic studies [23,24,25,26,27,28,29]. An alternative but similar pathway was developed to access thianthrenes 9 and 10 through the reaction of 2,3-dinitrothianthrene 6 with (S)-phenylethylamine 8 followed by the mcpba selective oxidation of the more electron-rich sulfur atom. Compound 11 was an oil, presumably because of its asymmetry, but with the conjugation of the amine to the sulfoxide, it became crystalline. The yield of sulfoxide formation was similar to that obtained by the mcpba oxidation of 2,3-dinitrothianthrene 6. The sulfur oxidation yield is not influenced by an amine donor or a nitro group acceptor, but we did not observe the over-oxidation of the thianthrene ring to a sulfone or even a bis-sulfoxide, using two equivalents of mcpba. In summary, the crystal structures of these sulfoxides establish the absolute configuration of the chiral sulfur atom and prove the stability of the two diastereoisomers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25084311/s1.

Author Contributions

Methodology, M.J.P.; Software, W.T.A.H.; Formal analysis, W.T.A.H.; Investigation, M.J.P.; Writing—original draft, M.J.P.; Writing—review & editing, W.T.A.H.; Project administration, M.J.P. All authors have read and agreed to the published version of the manuscript.

Funding

The authors received no financial support for the research, authorship and/or publication of this article.

Data Availability Statement

CCDC-2337820 and CCDC-2337821 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge by emailing [email protected] or by contacting The Cambridge Crystallography Centre, 12 Union Road, Cambridge, CB2 1EZ, UK; https://www.ccdc.cam.ac.uk/ (accessed on 5 March 2024).

Acknowledgments

We thank the UK EPSRC National Mass Spectrometry Service Centre for mass spectrometric data and the UK National Crystallography Centre (University of Southampton) for the X-ray data collections. M. J. Plater performed all syntheses and obtained the characterisation data, and W. T. A. Harrison solved the crystallographic data sets. Data sets were obtained free of charge from the National Crystallography Centre, Southampton University.

Conflicts of Interest

The authors declared no potential conflicts of interest with respect to the research, authorship and/or publication of this article.

References

  1. Pellissier, H. Use of chiral sulfoxides in asymmetric synthesis. Tetrahedron 2006, 62, 5559–5601. [Google Scholar] [CrossRef]
  2. Fernandez, I.; Khiar, N. Recent Developments in the Synthesis and Utilization of Chiral Sulfoxides. Chem. Rev. 2003, 103, 3651–3705. [Google Scholar] [CrossRef] [PubMed]
  3. Carreno, M.C. Applications of Sulfoxides to Asymmetric Synthesis of Biologically Active Compounds. Chem. Rev. 1995, 95, 1717–1760. [Google Scholar] [CrossRef]
  4. Harrison, P.W.B.; Kenyon, J.; Phillips, H. CCLXX1X—The Dependence of Rotatory Power on Chemical Constitution. Part XXIX. The Resolution of Sulphoxides into their Optically Active Forms. J. Chem. Soc. 1926, 129, 2079–2090. [Google Scholar] [CrossRef]
  5. Gilman, H.; Robinson, J.; Beaber, N.J. The reaction between organomagnesium halides and the esters of some sulfur acids. J. Am. Chem. Soc. 1926, 48, 2715–2718. [Google Scholar] [CrossRef]
  6. Andersen, K.K.; Gaffield, W.; Papanikolaou, N.E.; Foley, J.; Perkins, R.I. Optically Active Sulfoxides. The Synthesis and Rotatory Dispersion of Some Diaryl Sulfoxides. J. Am. Chem. Soc. 1964, 86, 5637–5646. [Google Scholar] [CrossRef]
  7. Andersen, K.K. Configurational Relationships among Some Sulfoxides. J. Org. Chem. 1964, 29, 1953–1956. [Google Scholar] [CrossRef]
  8. Andersen, K.K. Synthesis of (+)ethyl-e-tolyl sulfoxide from (-)menthyl(-)-e-toluenesulfinate. Tetrahedron Lett. 1962, 3, 93–95. [Google Scholar] [CrossRef]
  9. Mellah, M.; Voituriez, A.; Schulz, E. Chiral Sulfur Ligands for Asymmetric Catalysis. Chem. Rev. 2007, 107, 5133–5209. [Google Scholar] [CrossRef]
  10. Mei, H.; Xie, C.; Han, J.; Soloshonok, V.A. N-tert-Butylsulfinyl-3,3,3-trifluoroacetaldimine: Versatile Reagent for Asymmetric Synthesis of Trifluoromethyl-Containing Amines and Amino Acids of Pharmaceutical Importance. Eur. J. Org. Chem. 2016, 2016, 5917–5932. [Google Scholar] [CrossRef]
  11. Pellissier, H. Chiral sulfur-containing ligands for asymmetric catalysis. Tetrahedron 2007, 63, 1297–1330. [Google Scholar] [CrossRef]
  12. Robak, M.T.; Herbage, M.A.; Ellman, J.A. Synthesis and Applications of tert-Butanesulfinamide. Chem. Rev. 2010, 110, 3600–3740. [Google Scholar] [CrossRef] [PubMed]
  13. Agranat, I.; Caner, H. Intellectual property and chirality of drugs. Drug Discov. Today 1999, 4, 313–321. [Google Scholar] [CrossRef] [PubMed]
  14. Bentley, R. Role of sulfur chirality in the chemical processes of biology. Chem. Soc. Rev. 2005, 34, 609–624. [Google Scholar] [CrossRef] [PubMed]
  15. Han, J.; Soloshonok, V.A.; Klika, K.D.; Drabowicz, J.; Wzorek, A. Chiral sulfoxides: Advances in asymmetric synthesis and problems with the accurate determination of the stereochemical outcome. Chem. Soc. Rev. 2018, 47, 1307–1350. [Google Scholar] [CrossRef]
  16. Kagan, H.B. Asymmetric Synthesis of Chiral Sulfoxides. In Organosulfur Chemistry in Asymmetric Synthesis; Toru, T., Bolm, C., Eds.; Wiley Online Library: New York, NY, USA, 2008; pp. 1–29. [Google Scholar]
  17. O’Mahony, G.E.; Kelly, P.; Lawrence, S.E.; Maguire, A.R. Synthesis of enantioenriched sulfoxides. ARKIVOC 2011, 1, 1–110. [Google Scholar] [CrossRef]
  18. Otocka, S.; Kwiatkowska, M.; Madalinska, L.; Kiełbasinski, P. Chiral Organosulfur Ligands/Catalysts with a Stereogenic Sulfur Atom: Applications in Asymmetric Synthesis. Chem. Rev. 2017, 117, 4147–4181. [Google Scholar] [CrossRef]
  19. O’Mahony, G.E.; Ford, A.; Maguire, A.R. Asymmetric oxidation of sulfides. J. Sulfur Chem. 2013, 34, 301–341. [Google Scholar] [CrossRef]
  20. Wojaczynska, E.; Wojaczynski, J. Enantioselective Synthesis of Sulfoxides: 2000−2009. Chem. Rev. 2010, 110, 4303–4356. [Google Scholar] [CrossRef]
  21. Sipos, G.; Drinkel, E.E.; Dorta, R. The emergence of sulfoxides as efficient ligands in transition metal catalysis. Chem. Soc. Rev. 2015, 44, 3834–3860. [Google Scholar] [CrossRef]
  22. Soloshonok, V.A.; Klika, K.D. Terminology related to the phenomenon ‘self-disproportionation of enantiomers’ (SDE). Helv. Chim. Acta. 2014, 97, 1583–1589. [Google Scholar] [CrossRef]
  23. Diter, P.; Taudien, S.; Samuel, O.; Kagan, H.B. Enantiomeric Enrichment of Sulfoxides by Preparative Flash Chromatography on an Achiral Phase. J. Org. Chem. 1994, 59, 370–373. [Google Scholar] [CrossRef]
  24. Girard, C.; Kagan, H.B. On diastereomeric perturbations. Can. J. Chem. 2000, 78, 816–828. [Google Scholar] [CrossRef]
  25. Kagan, H.B.; Riant, O. Advances in Asymmetric Synthesis; Elsevier B.V.: Amsterdam, The Netherlands, 1997; Volume 2, pp. 189–235. [Google Scholar]
  26. Brunel, J.-M.; Kagan, H.B. Catalytic enantioselective oxidation of sulfides with a chiral titanium complex. Bull. Soc. Chim. Fr. 1996, 133, 1109–1115. [Google Scholar]
  27. Girard, C.; Kagan, H.B. Nonlinear Effects in Asymmetric Synthesis and Stereoselective Reactions: Ten Years of Investigation. Angew. Chem. Int. Ed. 1998, 37, 2923–2959. [Google Scholar] [CrossRef]
  28. Brunel, J.-M.; Luukas, T.O.; Kagan, H.B. Nonlinear effects as ‘indicators’ in the tuning of asymmetric catalysts. Tetrahedron Asymmetry 1998, 9, 1941–1946. [Google Scholar] [CrossRef]
  29. Fenwick, D.R.; Kagan, H.B. Top. Stereochem; Denmark, S.E., Ed.; Wiley: Hoboken, NJ, USA, 1999; Volume 22, pp. 257–296. [Google Scholar]
  30. Fujita, T.; Kamiyama, H.; Osawa, Y.; Kawaguchi, H.; Kim, B.J.; Tatami, A.; Kawashima, W.; Maeda, T.; Nakanishi, A.; Morita, H. Photo SN-bond cleavage and related reactions of thianthrene sulfilimine derivatives. Tetrahedron 2007, 63, 7708–7716. [Google Scholar] [CrossRef]
  31. Fries, V.; Volk, W. Uber Thianthrene. Chem. Ber. 1909, 42, 1170–1176. [Google Scholar] [CrossRef]
  32. Naomichi, F.; Takeshi, K.; Yoji, H.; Satoshi, O. A convenient preparation of sterically crowded 1,9-disubstituted dibenzothiophenes and 3,3′-disubstituted diaryl sulfides. Heterocycles 1991, 32, 675–678. [Google Scholar]
  33. Bosch, E.; Kochi, J.K. Catalytic oxidation of chlorpromazine and related phenothiazines. Cation radicals as the reactive intermediates in sulfoxide formation. J. Chem. Soc. Perkin Trans. I 1995, 8, 1057–1064. [Google Scholar] [CrossRef]
  34. Mata, E.G. Recent advances in the synthesis of sulfoxides from sulfides. Phosphorus Sulfur Silicon Relat. Elem. 1996, 117, 231–286. [Google Scholar] [CrossRef]
  35. Tanaka, H.; Nishikawa, H.; Uchida, T.; Katsuki, T. Photopromoted Ru-Catalyzed Asymmetric Aerobic Sulfide Oxidation and Epoxidation Using Water as a Proton Transfer Mediator. J. Am. Chem. Soc. 2010, 132, 12034–12041. [Google Scholar] [CrossRef] [PubMed]
  36. Sprout, C.M.; Seto, C.T. Using Enzyme Inhibition as a High Throughput Method to Measure the Enantiomeric Excess of a Chiral Sulfoxide. Org. Lett. 2005, 7, 5099–5102. [Google Scholar] [CrossRef]
  37. Sun, J.; Zhu, C.; Dai, Z.; Yang, M.; Pan, Y.; Hu, H. Efficient Asymmetric Oxidation of Sulfides and Kinetic Resolution of Sulfoxides Catalyzed by a Vanadium−Salan System. J. Org. Chem. 2004, 69, 8500–8503. [Google Scholar] [CrossRef] [PubMed]
  38. Legros, J.; Bolm, C. Investigations on the Iron-Catalyzed Asymmetric Sulfide Oxidation. Chem.–Eur. J. 2005, 11, 1086–1092. [Google Scholar] [CrossRef] [PubMed]
  39. Qian, D.-Q.; Shine, H.J.; Guzman-Jimenez, I.Y.; Thurston, J.H.; Whitmire, K.H. Mono- and Bis-adducts from the Addition of Thianthrene Cation Radical Salts to Cycloalkenes and Alkenes. J. Org. Chem. 2002, 67, 4030–4039. [Google Scholar] [CrossRef]
  40. Holst, D.E.; Dorval, C.; Winter, C.K.; Guzei, I.A.; Wickens, Z.K. Regiospecific Alkene Aminofunctionalisation via an Electrogenerated Dielectrophile. J. Am. Chem. Soc. 2023, 145, 8299–8307. [Google Scholar]
  41. Plater, M.J.; Harrison, W.T.A. A potential iterative approach to 1,4-dihydro-N-heteroacene arrays. ChemistryOpen 2022, 11, e202100150. [Google Scholar] [CrossRef]
  42. Plater, M.J.; Harrison, W.T.A. Potential building blocks for 1,4-dihydro-N-heteroacenes. ChemistryOpen 2022, 11, e202200092. [Google Scholar] [CrossRef]
  43. Plater, M.J.; Harrison, W.T.A. New funtionalised phenoxazines and phenothiazines. ACS Omega 2023, 8, 44163–44171. [Google Scholar] [CrossRef]
  44. Cahn, R.S.; Ingold, C.K. Specification of Configuration about Quadricovalent Asymmetric Atoms. J. Chem. Soc. 1951, 612–622. [Google Scholar]
  45. Cahn, R.S.; Ingold, C.K.; Prelog, V. The Specification of Asymmetric Configuration in Organic Chemistry. Experientia 1956, 15, 81–94. [Google Scholar] [CrossRef]
  46. Clegg, W. X-Ray Crystallography, 2nd ed.; Oxford University Press (Oxford Chemistry Primer): Oxford, UK, 2015. [Google Scholar]
  47. Eccles, K.S.; Morrison, R.E.; Stokes, S.P.; O’Mahony, G.E.; Hayes, J.A.; Kelly, D.M.; O’Boyle, N.M.; Fabian, L.; Moynihan, H.A.; Maguire, A.R.; et al. Utilizing Sulfoxide Iodine Halogen Bonding for Cocrystallization. Cryst. Growth Des. 2012, 12, 2969–2977. [Google Scholar] [CrossRef]
  48. Macrae, C.F.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Shields, G.P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury: Visualization and analysis of crystal structures. J. Appl. Cryst. 2006, 39, 453–457. [Google Scholar] [CrossRef]
  49. Gans, J.S.; Shalloway, D. Qmol: A program for molecular visualization on Windows-based PCs. J. Mol. Graph. Model. 2001, 19, 557–559. [Google Scholar] [CrossRef] [PubMed]
  50. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef]
  51. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
Figure 1. Drawings of chiral sulfoxides and thianthrene.
Figure 1. Drawings of chiral sulfoxides and thianthrene.
Ijms 25 04311 g001
Figure 2. Synthesis of enantiomerically pure thianthrenes 9 and 10.
Figure 2. Synthesis of enantiomerically pure thianthrenes 9 and 10.
Ijms 25 04311 g002
Figure 3. The molecular structures of 9 (left) and 10 (right) showing 50% displacement ellipsoids. The hydrogen bonds are shown as double-dashed lines. The sulfoxides orientate in the direction of the thianthrene puckering. Oxygen is red, blue is nitrogen and yellow is sulfur.
Figure 3. The molecular structures of 9 (left) and 10 (right) showing 50% displacement ellipsoids. The hydrogen bonds are shown as double-dashed lines. The sulfoxides orientate in the direction of the thianthrene puckering. Oxygen is red, blue is nitrogen and yellow is sulfur.
Ijms 25 04311 g003
Figure 4. Fragment of a hydrogen-bonded [001] chain of molecules in the extended structure of compound 10. Note that the NH group participates in both intramolecular and intermolecular links. Oxygen is red, blue is nitrogen and yellow is sulfur..
Figure 4. Fragment of a hydrogen-bonded [001] chain of molecules in the extended structure of compound 10. Note that the NH group participates in both intramolecular and intermolecular links. Oxygen is red, blue is nitrogen and yellow is sulfur..
Ijms 25 04311 g004
Figure 5. An overlay view of 9 (red) and 10 (blue) showing the different orientations of the terminal aryl ring owing to the rigid sulfoxide chiral centre. Atoms C1–C6 in the two structures are superimposed (the blue ring comes forward and the red ring goes back).
Figure 5. An overlay view of 9 (red) and 10 (blue) showing the different orientations of the terminal aryl ring owing to the rigid sulfoxide chiral centre. Atoms C1–C6 in the two structures are superimposed (the blue ring comes forward and the red ring goes back).
Ijms 25 04311 g005
Figure 6. An alternative pathway to thianthrenes 9 and 10. Supplementary Materials are available from the electronic site below.
Figure 6. An alternative pathway to thianthrenes 9 and 10. Supplementary Materials are available from the electronic site below.
Ijms 25 04311 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Plater, M.J.; Harrison, W.T.A. Chiral Thianthrenes. Int. J. Mol. Sci. 2024, 25, 4311. https://doi.org/10.3390/ijms25084311

AMA Style

Plater MJ, Harrison WTA. Chiral Thianthrenes. International Journal of Molecular Sciences. 2024; 25(8):4311. https://doi.org/10.3390/ijms25084311

Chicago/Turabian Style

Plater, M. John, and William T. A. Harrison. 2024. "Chiral Thianthrenes" International Journal of Molecular Sciences 25, no. 8: 4311. https://doi.org/10.3390/ijms25084311

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop