ISSN : 0970 - 020X, ONLINE ISSN : 2231-5039
     FacebookTwitterLinkedinMendeley

Biocatalytic Oxidation of Sulfides to Sulfones

Shefali Dhiman

Department of Chemistry, Assistant Professor and Head, Mehr Chand Mahajan DAV College for Women, Chandigarh, Punjab, India.

Corresponding Author E-mail: shefalitac@gmail.com

DOI : http://dx.doi.org/10.13005/ojc/370136

Article Publishing History
Article Received on : 05-Dec-2020
Article Accepted on :
Article Published : 13 Feb 2021
Article Metrics
ABSTRACT:

This paper describes a method for the biocatalytic oxidation of sulfides. During the screening of microorganisms using pure cultures of Bacteria and Fungi for the oxidation of sulfides, it was observed that a number of strains of microorganisms, were able to oxidize various sulfides (1-4), but the desired sulfoxide was either not obtained or obtained only as a minor product. A close observation of the reaction showed complete oxidation and thus sulfone (5-8) formation had occurred in these cases.Sulfones are used to stabilize intermediates like α-radicals, α-anionsetc. and also used as cationic synthons in many known reactions.This prompted us to explore the sulfone synthesis by biocatalytic route.Approximately 20% of the strains tested (400 bacterial and 200 fungal) showed the formation of sulfone with conversion rate varying from 3 to 100% based on TLC analysis. There were two strains of fungi, Aspergillus ochraceus MTCC 5245 and Penicillium Funiculosum MTCC 5246 which showed excellent biocatalytic activity for oxidation sulfides to corresponding sulfones in high yield. In all these strains, the product was different from corresponding standard sulfoxide prepared by oxidation with m-chloroperbenzoic acid but well corresponded with the standard sample of sulfone prepared by oxidation of the corresponding sulfides with oxone®. The identity of sulfones in all cases was confirmed by 1H NMR.

KEYWORDS:

Biocatalysis; Biocatalytic Oxidation, Biocatalytic Activity; Oxidation Using Micro-Organisms; Sulfide Oxidation; Sulfone Synthesis

Download this article as: 

Copy the following to cite this article:

Dhiman S. Biocatalytic Oxidation of Sulfides to Sulfones. Orient J Chem 2021;37(1).


Copy the following to cite this URL:

Dhiman S. Biocatalytic Oxidation of Sulfides to Sulfones. Orient J Chem 2021;37(1). Available from: https://bit.ly/37bbav9


Introduction

Sulfones have been widely used in organic synthesis to synthesize many popular and complicated molecules.1They constitute key skeletal framework of many natural products as well as pharmaceuticals and have attracted considerable attention because of their important biological properties.2 A few medicinally privileged compounds containing sulfone as the structural motif are shown in Figure 1.  Many compounds containing sulfonyl group are known to exhibit antibiotic, antitumor, antidepressant, antiresorptive and anticancer properties.3-5

Figure 1: A few medicinally important Sulfones.

Click here to View figure

Synthesis of sulfoxides or sulfone has attracted a great deal of attention from the researchers. They are used to stabilize intermediates like α-radicals,6 α-anions7etc. They have been also used as cationic synthons in many known reactions.8There are many reagents available which can carry out the complete oxidation of sulfides to sulfoxides and sulfones.9,10 There are manyconventional oxidants available like NaBO3,11 NaClO4,12Ca(ClO)2,13 H5IO6/[MnIV-MnIV-(µ-O)3L2](PF6)2,14 KHSO5,15 HNO3,16(NH4)2Ce(NO3)6,17 NaIO4,18 MnO2,19 KMnO4,20 RuO4,21 CF3CO3H,22 dimethyldioxirane,23,24 t-C4H9O2H,25 4-Methyl morpholine is used with OsO4,26 3-ClC6H4CO3H27 and [(n-C4H9)4N]-HSO528 which resultin oxidation. The major drawback of most of the above mentioned reagents is that they were not effective for medium- to large-scale synthesis. The major factors which contribute to this are low content of effective oxygen available which would cause the desired oxidation, high cost of production and formation of unwanted side products which is not industry friendly. Singlet oxygen29 or molecular oxygen combined with reagents like CH3CH(CH3)CHO30 or CH3CH(CH3)CH2CHO and Co(acac)231 have also been  widely employed. It has been observed that aqueous hydrogen peroxide having a concentration of 60% is a versatile reagent as it forms negligible side products, is industry friendly and can be transported easily.32 These added advantages have led to the development of various H2O2 basedoxidation methods, which primarily uses tungsten (W) as catalyst systems such as H2WO4,33[C5H5N(n-C16H33)]3PO4-[W(O)(O2)2]4,34 Na2WO4+[(n-C4H9)4N]Cl,35 as well as CH3ReO3,36 CH3ReO3,37 2-NO2C6H4SeO2H38and hemoglobin.39

The oxidation of various sulfides was carried out using a mixture containing 30% H2O2 and a buffer solution of NaHCO3 in the presence of a catalyst MnSO4.H2O.40A number of sulfides which contained different functional groups were subjected to oxidation (Table 1). Aromatic sulfides such as diphenyl sulfide showed complete oxidation without the use of any catalyst using both solvents acetonitrile and DMF (Table 1, Entries 1-3). The use of benzyl phenyl sulfide as a substrate to carry out the oxidation showed that acetonitrile was the best solvent for the reaction (Table 1, Entries 4 and 5). The oxidation of  p-anisyl methyl sulfide showed that the rate of the reaction was much faster in DMF than CH3CN without the use of any catalyst. (Table 1, Entries 6-9).

Table 1: Chemical Oxidation of Various Sulfides.

Entry

Sulfide

H2O2

Solvent

Time (h)

Yield (%)

1

PhSPh

3

CH3CN

0.25

100(94)

2

PhSPh

3

DMF

0.25

100(94)

3

PhSPh

2b

CH3CN

0.25

100

4

PhSCH2Ph

5

DMF

24

55

5

PhSCH2Ph

5

CH3CN

0.25

100 (99)

6

4-OMeC6H4SMe

5

DMF

0.25

100 (70)

7

4-OMeC6H4SMe

5 b

DMF

0.25

100

8

4-OMeC6H4SMe

3.5

CH3CN

24

63

9

4-OMeC6H4SMe

3.5 b

CH3CN

24

98

10

PhSCH2CH2OH

5

DMF

0.25

72

11

PhSCH2CH2OH

5

CH3CN

0.25

100 (88)

12

PhSCH2CH2OTHP

5

DMF

0.25

80

13

PhSCH2CH2OTHP

5

CH3CN

0.25

100 (85)

14

PhSCH2CH2OTBDMS

5

DMF

0.25

72 c

15

PhSCH2CH2OTBDMS

5

CH3CN

0.25

100 (86)

16

3,5-(CF3) 2C6H3SCH2CH2OH

5

CH3CN

0.25

100 d

17

PhSCH2CO2CH2Ph

5

DMF

0.25

80 e

18

PhSCH2CO2CH2Ph

5

CH3CN

0.25

100 d

19

3-CF3C6H4SCO2CH2Ph

5

CH3CN

24

100 (93)

20

PhSCH=CH2

5

CH3CN

0.25

100 (80)

21

PhSCH2CH=CH2

5

CH3CN

0.25

95

22

(CH2= CH CH2)2S

5

CH3CN

0.25

100 (94)

23

PhCH2 SCH2CH=CH2

5

CH3CN

0.25

100 (90)

24

PhSCH2CH=C(CH3)2

5

DMF

0.25

100 (87)

25

PhSCH2CH=C(CH3)CH=C(CH3)2

5

CH3CN

0.25

100

The rate of conversion rate has been determined by spectroscopic analysis. In brackets are mentioned isolated yields after working up the reaction

There has been no use of the catalyst for the reaction.

It clearly shows sulfoxide yield in the reaction.

Only sulfoxide was formed as the oxidation product. No sulfone formation occurred.

The yield of the sulfoxide as by-product was only 20%.

However, many other examples showed acetonitrile as a better solvent to get complete oxidation product sulfone (Table 1, Entries 10-18). It was observed that both vinyl and allyl containing sulfides were completely oxidized to the corresponding sulfones (Table 1, Entries 20-23).

Another strange observation was that even reactive tri-substituted olefin containing moieties were quite inert under these strong oxidation conditions and produced the completely oxidized sulfone (Table 1, Entries 24 and 25).

Alonso et al. have demonstrated the best oxidation of various sulfides to their corresponding sulfoxides or sulfones. The oxidant that was used was 30% hydrogen peroxide without the use of any solvent.41

Materials and Methods

During the initial screening of microorganisms, it was noticed that a number of strains of microorganisms, were able to oxidize simple sulfides like butyl phenyl sulfide and benzyl phenyl sulfide. It was observed that the desired sulfoxide was either not obtained or obtained only as a minor product. In order to select a strain of micro-organisms which would cause complete oxidation of sulfide to sulfone required extensive screening of a library of microorganisms. To begin with, small molecules containing sulfides like butyl phenyl sulfide was chosen as a substrate to oxidize using different strains of micro-organisms. The bacterial cultures chosen after the initial screening were grown in a medium containing Peptone (0.5%), beef extract (0.2%), yeast extract (0.1%) and Sodium Chloride (0.5%) at a pH 7.0-7.5 for 18-24 h. The medium used for selected strains of Fungi was Peptone M (0.2%), beef extract (0.2%) and KH2PO4 (0.2%) at a pH of 6.0-6.5 for 48-72 h.42 The bacterial or fungal cells were isolated by centrifugation, followed by their washing with phosphate buffer having a pH 7.0 The cells were subjected to a bio-catalytic reaction using 2 mg of butyl phenyl sulfide at 200 rpm in a shaker for 12 h. The supernatant was extracted with 4 mL of organic solvent (ethyl acetate). The standard reagents butyl phenyl sulfoxide and sulfone were prepared by chemical oxidation with reagents like m-chloroperbenzoic acid and oxone® respectively. The bio-catalyzed butyl phenyl sulfoxide prepared was analyzed using Thin Layer Chromatography (TLC) by running parallel standard samples of sulfoxide and sulfone. The final confirmation of sulfone was done by 1H-NMR spectroscopy. The same reaction was repeated at a larger scale using 4 g of bacterial or fungal cells and 20 mM substrate concentration.

Approximately 20% of the strains tested (400 bacterial and 200 fungal) showed the formation of sulfone with conversion rate varying from 3-100% based on TLC analysis. A number of strains (such as BNI0011N, BNI0047N, BNI0025N, BNI0018N, BNI0029N and BNI0044N) isolated from sea water and grown in the presence of 5% NaCl produced sulfone at 10-60% conversion rate. Two strains designated as PF7 and PF9 were selected for further studies.

Results and Discussion

Two fungal Strains PF7 and PF9 have been previously described in connection with enantioselective oxidation of β-keto esters.42The identification of the strains has already been reported as Aspergillus ochraceus and Penicillium funiculosum, respectively by MTCC, IMTECH. The strains have been given accession numbers MTCC 5245 and MTCC 5246. Aspergillus ochraceus and Penicillium funiculosum were grown in fungal medium as described in the experimental section. The mycelia were isolated by filtration and the cells were suspended in phosphate buffer at pH 7.0 at a concentration of 0.5 g/mL and incubated with 2 mg of butyl phenyl sulfide (3) at 200 rpm at 30 oC for 12 h. The supernatant was separated by centrifugation and extracted with ethyl acetate. The formation of the sulfone 7 was confirmed by 1H NMR for which the biotransformation was repeated at 50 mg scale using 20 mM substrate concentration and 6 g washed cells. The product was assigned structures sulfone based on its 1H NMR spectral data which showed presence of methyl group as triplet (J = 7.27 Hz) at δ 0.89, two of the methylene protons as multiplet at δ 1.35-1.45 and δ 1.64-1.79 and the methylene attached to sulfur atom at δ 3.00 as triplet (J = 7.99 Hz). The aromatic protons appeared as multiplet at δ 7.51-7.88. The NMR spectral data was different from NMR data of sulfoxide prepared by oxidation with m-chloroperbenzoic acid. The CH2SO methylene group in sulfoxide resonated at δ 2.73 as triplet (J = 7.1Hz), approximately 0.27 ppm upfield and aromatic protons resonated at δ 7.43-7.60 approximately 0.15 ppm upfield compared to the biocatalyzed product. The 1H NMR spectral data, however, corresponded well with 1H NMR data of standard sample of sulfone prepared by oxidation of sulfides with oxone®.

Similarly, Aspergillus ochraceus and Penicillium funiculosum catalyzed oxidation of methyl phenyl sulfide (1), ethyl phenyl sulfide (2) and4-Chlorothioanisole (4) produced corresponding sulfones 5, 6 and 8(Scheme 1). In all these examples, the product was different from corresponding standard sulfoxide prepared by oxidation with m-chloroperbenzoic acid. In all cases, the proton attached to the α-carbon atom of the biocatalyzed product showed 0.25-0.31 ppm downfield shift and aromatic protons 0.10-0.16 ppm downfield shift compared to standard sulfoxide. NMR spectral data of sulfones corresponded with the standard sample of sulfone prepared by oxidation of the corresponding sulfides with oxone®.

Scheme 1: Aspergillus ochraceus and Penicillium funiculosum catalyzed oxidation of methyl phenyl sulfide yield = 90 – 95%.

Click here to View scheme

Methyl 3-(phenylthio) propionate (9), prepared by conjugate addition of thiophenol on methyl acrylate on biocatalyzed reaction with Aspergillus ochraceus and Penicillium Funiculosum failed to give the desired sulfoxide 10 or sulfone 11, but TLC showed complete consumption of starting material and formation of a high polarity compound. The NMR spectral data of the isolated product showed absence of ester methyl protons, instead a broad singlet was present at δ 11.12. The aromatic protons resonated at δ 7.13-7.33 as multiplet and methylene protons as a triplet (J = 7.47 Hz) each at δ 2.62 and δ 3.10. The product was therefore assigned the structure 12 (Scheme 2). No sulfoxide or sulfone of ester, 9 or acid, 12 could be isolated from the reaction mixture.

Scheme 2: sulfoxide or sulfone of ester, 9 or acid, 12 could be isolated from the reaction mixture.

Click here to View scheme

4-phenyl-4-(phenylsulfonyl)butan-2-one 15 also failed to produce any sulfoxide or sulfone, but instead gave a product, which was identified as trans-cinnamic acid 17 based on 1H NMR spectral data. In all probabilities the trans-cinnamic acid was produced via oxidation of sulfide to either sulfoxide or sulfone (16).But neither sulfoxide nor sulfone could be isolated from the reaction mixture. Oxidation of 15 with H2O2 or oxone® also resulted in the formation of 17, which indicated lability of the sulfoxide/sulfone under the reaction conditions (Scheme 3).

Scheme 3: which indicated lability of the sulfoxide/sulfone under the reaction conditions.

Click here to View scheme

Conclusion

Two strains of fungi, Aspergillus ochraceus MTCC 5245 and Penicillium funiculosum MTCC 5246 have been identified which oxidize alkyl aryl sulfides to corresponding sulfones in high yield. The method is environmentally benign as it requires no organic solvents and can be carried under mild reaction conditions. However, methyl 3-(phenylthio)propionate(9) failed to give desired sulfone 11, instead only the hydrolysis of the ester functionality occurred to give acid, 12. Oxidation of 4-phenyl-4-(phenylsulfonyl)butan-2-one (15) occurred, but sulfoxide/sulfone readily eliminated under the reaction conditions to give trans-cinnamic acid (17). In a similar way, Aspergillus ochraceus and Penicillium funiculosum catalyzed oxidation of substrates such as methyl phenyl sulfide, ethyl phenyl sulfide and4-Chlorothioanisole produced corresponding sulfones. In all these examples, the product was different from corresponding standard sulfoxide prepared by oxidation with m-chloroperbenzoic acid. In all cases, the proton attached to the α-carbon atom of the biocatalyzed product showed 0.25-0.31 ppm downfield shift and aromatic protons 0.10-0.16 ppm downfield shift compared to standard sulfoxide. NMR spectral data of sulfones corresponded with the standard sample of sulfone prepared by oxidation of the corresponding sulfides with oxone®.

References

  1. Patai, S.; Rappoport, Z.; Stirling, C. The Chemistry of Sulphones and Sulfoxides; John Wiley & Sons: Chichester 1988.
    CrossRef
  2. Liu, K. G.; Robichaud, A. J.; Bernotas, R. C.; Yan, Y.; Lo, J. R.; Zhang, M. Y.; Hughes, Z. A.; Huselton, C.; Zhang, G. M.; Zhang, J. Y.; Kowal, D. M., 2010,J. of Med.Chem., 2010,53(21), 7639-46.
    CrossRef
  3. Eldehna, W. M.; Al-Ansary, G. H.; Bua, S.; Nocentini, A.; Gratteri, P.; Altoukhy, A.; Supuran, C. T., Eur. Journal of Med. Chem.,2017,127, 521-30.
    CrossRef
  4. La Regina, G.; Coluccia, A.; Brancale, A.; Piscitelli, F.; Famiglini, V.; Cosconati, S.; Schols, D., J. of Med.Chem., 2012,55(14), 6634-8.
    CrossRef
  5. Suja, T. D.; Divya, K. V. L.; Naik, L. V.; Kumar, A. R.; Kamal, A. Coppercatalyzed Bioorganic Med. Chem. Lett.,2016,26(8), 2072-2076.
    CrossRef
  6. Paquette, L.A. Synlett, 2001, 1-12.
  7. Najera, C.; Sansano, J. M.Recent Res. Devel. Org.Chem., 1998, 2, 637-83.
  8. Chinchilla, R.; Najera, C.Recent Res. Devel.,1997, Org. Chem.,1, 437-67.
  9. Kirk-Othmer, C. E. Encyclopedia of Chemical Technology; 4th ed.; Kroschwitz, J. I., Howe-Grant, M., Eds.; Wiley: New York,1997,23, 134-46.
  10. Kirk-Othmer, W. R. Encyclopedia of Chemical Technology; 4th ed.; Kroschwitz, J. I., Howe-Grant, M., Eds.; Wiley: New York,1997,23, 217-32.
  11. Mckillop, A.; Sanderson, W.R. Tetrahedron Lett.,1995,51(22),6145-66.
    CrossRef
  12. Khurana, J. M.; Panda, A. K.; Ray, A.; Gogia, A. Org. Prep. Proc.,1996,28; 234-237.
    CrossRef
  13. Weber, J. V.; Schneider, M.; Salami, B.; Paquer, D. Recl. Trav. Chim. Pays-Bas, 1986,105, 99-102.
    CrossRef
  14. Barton, D. H. R.; Li, W.; Smith, J. A. Tetrahedron Lett.,1998,39(39),7055-8.
    CrossRef
  15. Kennedy, R. J.; Stock, A. M. J. Org. Chem.,1960,25, 1901-6.
    CrossRef
  16. Trost, B. M.; Curran, D. P. Tetrahedron Lett.,1981, 22(14), 1287-90.
    CrossRef
  17. Goheen, D. W.; Bennett, C. F. J. Org. Chem.,1961,26, 1331-3.
    CrossRef
  18. Ho, Tse- Lok.; Wong, C. M. Synthesis, 1972, 561-2.
    CrossRef
  19. Varma, R.S.; Dahiya, R. Tetrahedron Lett.,1998, 39(11),1307-8.
    CrossRef
  20. R. S.; Saini, R. K.; Meshram, H. M. Tetrahedron Lett.,1997,38(37), 6525-8.
    CrossRef
  21. Edwards, D.; Stenlake, J. B. J. Chem. Soc.,1954, 3272-4.
    CrossRef
  22. Gokel, G. W.; Gerdes, H. M.; Dishong, D. M. J. Org. Chem.,1980,45, 3634-9.
    CrossRef
  23. Djerassi, C.; Engle, R. R. J. Am. Chem. Soc.,1953,75, 3838-40.
    CrossRef
  24. Venier, C. G.; Squires, T. G.; Chen, Y.-Y.; Hussmann, G. P.; Shei, J. C.; Smith, B. F. J. Org. Chem.,1982,47, 3773-4.
    CrossRef
  25. Murray, R. W.; Jeyaraman, R., J. Org. Chem., 1985, 50, 2847-53.
    CrossRef
  26. Adam, W. and Hadjiarapoglou, L. Tetrahedron Lett.,1992, 33(4), 469-70.
    CrossRef
  27. Breton, G. W., Fields, J. D. and Kropp, P. J. Tetrahedron Lett.,1995,36(22), 3825-8.28.
    CrossRef
  28. Kaldor, S. W.; Hammond, M. Tetrahedron Lett.1991,32(38), 5043-6.
    CrossRef
  29. Paquette, L. A.; Carr, R. V. C. Org. Synth. Coll., 1990,7, 453-6.
  30. Trost, B. M.; Braslau, R. J. Org. Chem., 1988,53, 532-7.
    CrossRef
  31. Hauser, F.M.; Suzanne R.; Ellenberger, Clardy, J.C.; Bass, L.S. J.Am.Chem. Soc. 1984, 106 (8), 2458-59.
    CrossRef
  32. Mandal, A.K.; Khanna, V.; Iqbal, J. Tetrahedron Lett.,1996,37(19), 3769-72.
    CrossRef
  33. Dell’Anna, M. M.; Mastrorilli, P.; Nobile, C. F. J. Mol. Catal.,1996,108(2), 57-62.
    CrossRef
  34. Sato, K.; Aoki, M.; Noyori, R., 1998,281, 1646-7.
    CrossRef
  35. Schultz, H. S.; Freyermuth, H. B.; Buc, S. R. J. Org. Chem.,1963,28,1140-2.
    CrossRef
  36. Ishii, Y.;Tanaka, H.; Nisiyama, Y. Chem. Lett.,1994, 1-4.
  37. Stec, Z.; Zawadiak, J.; Skibinski, A.; Pastuch, G. Polish J. Chem.,1996,70, 1121-123.
  38. Adam, W.; Mitchell, C. M.; Saha-Moeller, C. R. Tetrahedron,1994,50(47), 13121-4.
    CrossRef
  39. Yamazaki, S. Bull. Chem. Soc. Jpn.,1996,69(10), 2955-9.
    CrossRef
  40. Reich, H. J.; Chow, F.; Peake, S. L. Synthesis,1978, 299-301.
    CrossRef
  41. Alonso, D. A; Najera, C.; Varea, M. Tetrahedron Lett., 2002,43, 3459-61.
    CrossRef
  42. Sangar, S.; Pal, M.; Moon, L.S.; Jolly, R.S. Bioresour. Technol.2012, 115, 102-10.
    CrossRef


Creative Commons License
This work is licensed under a Creative Commons Attribution 4.0 International License.

About The Author