Next Article in Journal
[(2-Chlorophenyl)-(4-fluorophenyl)methylene]-(4-fluorophenyl)amine
Previous Article in Journal
Bis [4,4′-(1,3-Phenylenebis(azanylylidene))-bis(3,6-di-tert-butyl-2-oxycyclohexa-2,5-dien-1-one)-bis(dimethylsulfoxide)nickel(II)]
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

Benzo[d][1,3]oxathiole-2-thione

EaStCHEM School of Chemistry, University of St Andrews, North Haugh, St Andrews KY16 9ST, Fife, UK
*
Author to whom correspondence should be addressed.
Molbank 2024, 2024(4), M1891; https://doi.org/10.3390/M1891
Submission received: 7 September 2024 / Revised: 28 September 2024 / Accepted: 29 September 2024 / Published: 30 September 2024
(This article belongs to the Section Structure Determination)

Abstract

:
Although known for 120 years, the title compound has not been adequately characterised before. In this paper, it is fully characterised by 1H and 13C NMR and IR spectroscopy and its X-ray structure has been determined for the first time.

Graphical Abstract

1. Introduction

The fundamental heterocyclic compound benzo[d][1,3]oxathiole-2-thione 1 was first described in 1904 [1] when it was obtained as a side-product in the steam distillation of the hydroxyphenyl xanthate 2, which itself was obtained by diazotisation of o-aminophenol and a reaction with potassium ethyl xanthate (Scheme 1). A more convenient synthesis was reported in 1953 [2], involving the treatment of o-mercaptophenol 3 with thiophosgene in aqueous alkali. In these early studies, the only characterisation was by melting point and elemental analysis. More recently, a third route to 1 involving reaction of 3 with Me2N+=CCl2 Cl (“Viehe’s salt”) in pyridine followed by treatment with NaSH was described, and in this case 13C (but not 1H) NMR data were given for the product [3]. There have only been a few further mentions of compound 1 in the literature, including its reaction with benzyne to give the isomeric benzo[d][1,3]-dithiol-2-one by an interesting cycloaddition–cycloreversion mechanism [4], and its reduction with diisobutylaluminium hydride to regenerate 3 [5]. In this paper, we present the 1H NMR and IR spectra for 1, a full assignment of the NMR data, and determination of its structure by X-ray diffraction.

2. Results

Compound 1 was readily prepared from 3 by reaction with thiophosgene in aqueous sodium hydroxide, as reported [2], and was obtained as orange crystals after column chromatography.
The 1H NMR spectrum, which has not been documented before, turned out to be exceptionally complex for a simple ortho-disubstituted benzene derivative (Figure 1). This was, however, successfully simulated using the chemical shift values in Table 1 and the coupling constants in Table 2 to give agreement of all 30 signals to within 0.001 ppm (Figure 1). The most obvious discrepancy is that the signals appearing as doublets at 7.345/7.343 and 7.326/7.325 in the simulation are incompletely resolved in the experimental spectrum, but shoulders are visible on these signals. The precise match achieved for even the smallest peaks to the right of the spectrum gives confidence in the derived parameters. Once the data have been extracted, the reason for the complexity is clear: all four protons are coupling to each other and have similar chemical shifts, with two of them very close indeed, leading to significant second-order effects.
While the simulation yielded a viable set of chemical shifts and coupling constants, unambiguous assignment of these to the protons in the structure required consideration of the 13C NMR data and particularly HSQC and HMBC studies. The observed chemical shifts (see Supplementary Materials) are consistent with those reported earlier [3], although the assignment for the quaternary ring-junction carbons C-3a and C-7a is opposite to that made in the earlier publication (“153.3 C-3a, 126.8 C-7a”). We believe this may be due to erroneous numbering there since the signal at 155.1 is clearly the carbon directly joined to oxygen (C-7a). This value is also consistent with that of 153.9 for C-7a, which we recently reported for the corresponding cyclic thiosulfite benzo[d][1,2,3]oxadithiole 2-oxide [6]. The HSQC spectrum allowed unambiguous association of the CH carbon signals to the relevant protons (Table 1) but consideration of the HMBC results was key to arriving at a final assignment, as summarised in Figure 2. In particular, the carbon signals at 155.1 and 121.2 were linked on HMBC to the proton signal at 7.403, the carbon signal at 112.2 was linked to the proton signal at 7.337, the carbon signal at 128.0 was linked to the proton signal at 7.412, and the carbon signal at 126.0 was linked to the proton signal at 7.400. Taken together, this pattern is fully consistent with the assignment of Figure 2, with the HMBC spectrum consistently showing correlations for 3JC–H but not 2JC–H.
Since compound 1 formed suitable crystals after chromatography, we were able to determine its structure by X-ray diffraction (Figure 3). The molecular structure is completely planar, and the heterocyclic ring features the expected long C–S bonds to accommodate the large sulfur atom with the interior angle at sulfur just over 90° (Table 3).
Weak non-classical C-H···O hydrogen bonds between C(4) and O(1) (O···H 2.545(1) Å, C···O 3.3830(19) Å) link adjacent molecules along the (1 0 0) axis (Figure 4). These chains form sheets in the [0 1 0] plane via C-H···S contacts (H···S 2.9238(4) Å, C···S 3.7561(16) Å). This results in adjacent molecules forming head-to-tail stacks along the (1 0 0) axis (Figure 4), although at distances too long for π-π stacking to play a significant role.
As far as we are aware, this is the first 1,3-benzoxathiole-2-thione to be crystallographically characterised and a search of the Cambridge Structural Database (CSD) showed that only a few comparable structures have been determined before (Figure 5). There are only two previous 1,3-oxathiole-2-thione structures, 4 and 5 [7], while for the saturated 1,3-oxathiolane-2-thione ring system, there are only four structures: the tetramethyl compound 6 [8], the limonene-derived bicyclic structure 7 [9], the amino compound 8, which is of interest for ring-opening polymerisation [10], and the chiral phthalimido compound 9 [11]. As might be expected, the heterocyclic rings of 4 and 5 are essentially planar like 1 but removal of the double bond leads to a significant degree of twisting in structures 69.
In summary, we have been able to fully assign the 1H and 13C NMR spectra for compound 1 for the first time. The X-ray structure of 1 features a completely planar molecule with an internal angle at the ring sulfur atom of just over 90°.

3. Experimental

3.1. General Experimental Details

Melting points were recorded using a Reichert hot-stage microscope (Reichert, Vienna, Austria) and are uncorrected. NMR spectra were obtained using a Bruker AVII-400 instrument (Bruker, Billerica, MA, USA). Spectra were run with internal Me4Si as the reference and chemical shifts are reported in ppm to the high frequency of the reference. NMR spectra were processed, and simulations produced using iNMR reader, version 6.3.3 (Mestrelab Research, Santiago de Compostela, Spain). The IR spectrum was run on a Shimadzu IRAffinity instrument using the ATR technique.

3.2. Synthesis of Benzo[d][1,3]oxathiole-2-thione 1

2-Mercaptophenol (0.50 g, 4.0 mmol) was dissolved in water (6 mL) containing sodium hydroxide (0.4 g, 10 mmol) and stirred while thiophosgene (0.38 mL, 0.57 g, 5.0 mmol) was added. After stirring at RT for 18 h, the mixture was extracted with CH2Cl2 (2 × 5 mL) and the extracts were dried and evaporated. Column chromatography of the residue (SiO2, Et2O/hexane 1:1) gave the product (0.10 g, 15%) as orange-red crystals, mp 93–94 °C (lit. [2] 97–98 °C). IR (ATR): νmax /cm–1 1450, 1314, 1175 (S=O), 1153, 1084, 1022, 1009, 741, 656. 1H NMR and 13C NMR (CDCl3): see Table 1 and Table 2 and Figure 2.

3.3. X-ray Structure Determination of 1

X-ray diffraction data were collected at 100 K using a Rigaku FR-X Ultrahigh Brilliance Microfocus RA generator/confocal optics with XtaLAB P200 diffractometer [Mo Kα radiation (λ = 0.71073 Å)]. Data were collected (using a calculated strategy) and processed (including a multi-scan absorption correction) using CrysAlisPro [12]. The structure was solved by dual-space methods (SHELXT) [13] and refined by full-matrix least-squares against F2 (SHELXL-2019/3) [14]. Non-hydrogen atoms were refined anisotropically, and hydrogen atoms were refined using a riding model. All calculations were performed using the Olex2 interface [15].
Crystal data for C7H4OS2, M = 168.22 g mol–1, yellow prism, crystal dimensions 0.12 × 0.09 × 0.03 mm, triclinic, space group P–1 (No. 2), a = 6.8478(3), b = 7.4095(4), c = 7.9830(4) Å, α = 64.683(5), β = 83.960(4), γ = 70.423(4) °, V = 344.63(3) Å3, Z = 2, Dcalc = 1.621 g cm–3, T = 100 K, goodness of fit on F2 1.064, 7573 reflections measured, 1615 unique (Rint = 0.0426), which were used in all calculations. The final R1 [I > 2σ(I)] was 0.0287 and wR2 (all data) was 0.0805. Data have been deposited at the Cambridge Crystallographic Data Centre as CCDC 2382164. The data can be obtained free of charge from the Cambridge Crystallographic Data Centre via http://www.ccdc.cam.ac.uk/structures (accessed on 6 September 2024).

Supplementary Materials

The following are available online: CIF file for 1; IR, 1H and 13C NMR data for 1.

Author Contributions

L.C. prepared the compound and recorded the spectra, D.B.C. and A.P.M. collected the X-ray data and solved the structure; R.A.A. designed the study, analysed the data, and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The X-ray data are at CCDC, as stated in the paper, and the spectroscopic data are in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Friedländer, P.; Mauthner, F. Zur Kenntnis der Schwefelfarbstoffe. Z. Farben-Text.-Ind. 1904, 3, 333–337. [Google Scholar]
  2. Greenwood, D.; Stevenson, H.A. Benz-1:3-oxathioles, benz-1:4-oxathien, and αω-bisarylthioalkanes. J. Chem. Soc. 1953, 1514–1519. [Google Scholar] [CrossRef]
  3. Copeland, C.; Stick, R.V. The reaction of N-dichloromethylene-N,N-dimethylammonium chloride (Viehe’s salt) with thiophenols and phenols: A synthesis of unsymmetrical diaryl carbonates. Aust. J. Chem. 1984, 37, 1483–1487. [Google Scholar] [CrossRef]
  4. Jordis, U. Hydride reduction of 1,3-benzodithiole-2-thiones and -selones. J. Chem. Res (S) 1986, 432, J. Chem. Res. (M) 1986, 3401–3413. [Google Scholar]
  5. Nakayama, J.; Kimata, A.; Taniguchi, H.; Takahashi, F. Reactions of benzyne with 1,3-benzodithiole-2-thione and related compounds: Formation of novel tetracyclic sulfonium salts and their reactions leading to dibenzo-1,3,6,-trithiocin derivatives. Bull. Chem. Soc. Jpn. 1996, 69, 2349–2354. [Google Scholar] [CrossRef]
  6. Aitken, R.A.; Cordes, D.B.; Goyal, A.; McKay, A.P. Benzo[d][1,2,3]oxadithiole 2-Oxide. Molbank 2024, 2024, M1803. [Google Scholar] [CrossRef]
  7. Savarimuthu, S.A.; Prakash, D.G.L.; Thomas, S.A.; Gandhi, T. DBU-Mediated intermolecular 5-exo-dig cyclization of propargyl alcohols and carbon disulfide to [1,3]-oxathiole-2-thiones. ChemistrySelect 2018, 3, 13087–13090. [Google Scholar] [CrossRef]
  8. Diebler, J.; Spannenberg, A.; Werner, T. Regio- and stereoselective synthesis of dithiocarbonates under ambient and solvent-free conditions. ChemCatChem 2016, 8, 2027–2030. [Google Scholar] [CrossRef]
  9. Diebler, J.; Spannenberg, A.; Werner, T. Atom economical synthesis of di- and trithiocarbonates by the lithium tert-butoxide catalyzed addition of carbon disulfide to epoxides and thiiranes. Org. Biomol. Chem. 2016, 14, 7480–7489. [Google Scholar] [CrossRef] [PubMed]
  10. Krishnamurthy, S.; Yoshida, Y.; Endo, T. Cationic ring-opening polymerization of a five membered cyclic dithiocarbonate having a tertiary amine moiety. Polym. Chem. 2021, 13, 267–274. [Google Scholar] [CrossRef]
  11. Okada, M.; Nishiyori, R.; Kaneko, S.; Igawa, K.; Shirakawa, S. KI–tetraethylene glycol complex as an effective catalyst for the synthesis of cyclic thiocarbonates from epoxides and CS2. Eur. J. Org. Chem. 2018, 2018, 2022–2027. [Google Scholar] [CrossRef]
  12. CrysAlisPro; v1.171.43.109a; Rigaku Oxford Diffraction, Rigaku Corporation: Tokyo, Japan, 2023.
  13. Sheldrick, G.M. SHELXT—Integrated space-group and crystal structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  14. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  15. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of compound 1.
Scheme 1. Synthesis of compound 1.
Molbank 2024 m1891 sch001
Figure 1. Experimental (top) and simulated (bottom) 1H NMR spectrum of 1 using values from Table 1 and Table 2.
Figure 1. Experimental (top) and simulated (bottom) 1H NMR spectrum of 1 using values from Table 1 and Table 2.
Molbank 2024 m1891 g001
Figure 2. Final assignment of 1H (red) and 13C (blue) chemical shifts for compound 1.
Figure 2. Final assignment of 1H (red) and 13C (blue) chemical shifts for compound 1.
Molbank 2024 m1891 g002
Figure 3. The molecular structure of 1 showing the numbering system used and probability ellipsoids at 50% level.
Figure 3. The molecular structure of 1 showing the numbering system used and probability ellipsoids at 50% level.
Molbank 2024 m1891 g003
Figure 4. The crystal packing of 1 showing intermolecular interactions.
Figure 4. The crystal packing of 1 showing intermolecular interactions.
Molbank 2024 m1891 g004
Figure 5. Crystallographically characterised 1,3-oxathiole-2-thiones (4, 5) and 1,3-oxathiolane-2-thiones (6–9) with CSD RefCodes.
Figure 5. Crystallographically characterised 1,3-oxathiole-2-thiones (4, 5) and 1,3-oxathiolane-2-thiones (6–9) with CSD RefCodes.
Molbank 2024 m1891 g005
Table 1. 1H and 13C chemical shifts (ppm) for 1.
Table 1. 1H and 13C chemical shifts (ppm) for 1.
PositionδCδH
2201.7
3a126.6
4121.27.412
5126.07.337
6128.07.403
7112.27.400
7a155.1
Table 2. 1H–1H coupling constants (Hz) in 1.
Table 2. 1H–1H coupling constants (Hz) in 1.
Compound3JH4–H54JH4–H65JH4–H73JH5–H64JH5–H73JH6–H7
18.01.80.37.41.78.5
Table 3. Heterocyclic ring dimensions for 1 (Å, °).
Table 3. Heterocyclic ring dimensions for 1 (Å, °).
Bond Length Angle
S(3)–C(2)1.7399(16)C(3a)–S(3)–C(2)91.42(7)
C(2)–S(1)1.6293(16)S(3)–C(2)–O(1)111.77(11)
C(2)–O(1)1.3625(18)S(3)–C(2)–S(1)125.58(10)
O(1)–C(7a)1.3891(17)S(1)–C(2)–O(1)122.65(12)
C(7a)–C(3a)1.393(2)C(2)–O(1)–C(7a)112.89(11)
C(3a)–S(3)1.7422(15)O(1)–C(7a)–C(3a)114.29(13)
C(7a)–C(3a)–S(3)109.61(11)
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

Aitken, R.A.; Cordes, D.B.; Cottineau, L.; McKay, A.P. Benzo[d][1,3]oxathiole-2-thione. Molbank 2024, 2024, M1891. https://doi.org/10.3390/M1891

AMA Style

Aitken RA, Cordes DB, Cottineau L, McKay AP. Benzo[d][1,3]oxathiole-2-thione. Molbank. 2024; 2024(4):M1891. https://doi.org/10.3390/M1891

Chicago/Turabian Style

Aitken, R. Alan, David B. Cordes, Lauryne Cottineau, and Aidan P. McKay. 2024. "Benzo[d][1,3]oxathiole-2-thione" Molbank 2024, no. 4: M1891. https://doi.org/10.3390/M1891

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