Next Article in Journal
Synthesis and Antibacterial Analysis of Analogues of the Marine Alkaloid Pseudoceratidine
Previous Article in Journal
Synthesis and Studies of the Inhibitory Effect of Hydroxylated Phenylpropanoids and Biphenols Derivatives on Tyrosinase and Laccase Enzymes
Previous Article in Special Issue
Structure of Diferrocenyl Thioketone: From Molecule to Crystal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Continuous Bioinspired Oxidation of Sulfides

Department of Pharmaceutical Sciences (Group of Catalysis, Synthesis and Organic Green Chemistry), University of Perugia via del Liceo, 1–06123 Perugia, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2020, 25(11), 2711; https://doi.org/10.3390/molecules25112711
Submission received: 25 May 2020 / Revised: 6 June 2020 / Accepted: 10 June 2020 / Published: 11 June 2020
(This article belongs to the Special Issue Chalcogenides: New Developments and Cutting-Edge Applications)

Abstract

:
A simple, efficient, and selective oxidation under flow conditions of sulfides into their corresponding sulfoxides and sulfones is reported herein, using as a catalyst perselenic acid generated in situ by the oxidation of selenium (IV) oxide in a diluted aqueous solution of hydrogen peroxide as the final oxidant. The scope of the proposed methodology was investigated using aryl alkyl sulfides, aryl vinyl sulfides, and dialkyl sulfides as substrates, evidencing, in general, a good applicability. The scaled-up synthesis of (methylsulfonyl)benzene was also demonstrated, leading to its gram-scale preparation.

Graphical Abstract

1. Introduction

Heteroatom oxidation, with particular emphasis on nitrogen and sulfur, is a process of great relevance in organic synthesis. For the oxidation of sulfides into their corresponding sulfoxides and sulfones, a number of procedures have been described, including stoichiometric reactions, chemocatalytic reactions, and bio-catalytic reactions [1]. The main goal of this kind of transformation is the selective preparation of the mono-oxygenated over the bis-oxygenated derivatives, and the result is strongly dependent on the protocol used for the oxidation, being directly correlated to the nucleophilicity of the sulfide and sulfoxide and the electrophilicity of the actual oxidizing species.
Arylsulfones and arylsulfoxides are useful synthetic intermediates and building blocks for the synthesis of a series of biologically active compounds [2,3].
A non-exhaustive panel of examples is presented in Figure 1: an arylsulfone moiety is present in drugs like Rofecoxib (I) [4], Laropiprant (II) [5], Dapson (III) [6], and Sulfamethoxazole (IV) [7], while examples of arylsulfoxides (even if less distributed due to their relative instability) can be found in Sulindac (V) a clinically used anti-inflammatory drug [8], in esomeprazole (VI), the S enantiomer of omeprazole [9], and in a recently reported compound (VII) with anti-HIV activity [10].
Over the last ten years, we have studied the applicability of selenium derivatives as catalysts in biomimetic oxidations, demonstrating that this approach presents a number of advantages in terms of eco-sustainability [11].
We demonstrated that as the selenium atom of the glutathione peroxidase enzyme catalyzes the oxidation of two molecules of glutathione using peroxides as stoichiometric oxidants, small organoselenium molecules can be biomimetically used as catalysts in the hydrogen-peroxide-mediated oxidation of thiols [12]. The same reactivity can be translated into the oxidation of carbon–carbon double bonds, affording epoxide intermediates that can be opened by a nucleophilic solvent, affording vicinal diols [13,14] or α-methoxy alcohols [14] when the reaction is carried out in water or in methanol, respectively. When the substrate is an alkenol or an alkenoic acid, the epoxydic intermediate is opened intramolecularly by the internal nucleophile, leading to oxidative cyclofunctionalization with the formation of cyclic hydroxylethers or hydroxylactones [15]. Furthermore, the same catalytic system has been used also to oxidize aldehydes into their corresponding carboxylic acids in “on water” conditions, or into their corresponding esters when the reaction medium is a primary or a secondary alcohol [16].
The possibility of performing these catalytic reactions in aqueous media or in biphasic systems and the beneficial effects on the reactivity and selectivity, as well as the reaction workup, product purification, and recovery and reuse of the catalyst, were recently reviewed [17].
Nowadays, continuous-flow transformations represent an innovative methodology which is gaining increasing interest in both academic and industrial research including in the pharmaceutical industry. Reactions performed under continuous flow are generally faster, safer, and more environmentally friendly than those realized in batches. Recently, great progress was also made in terms of flexibility and robustness for large-scale processes [18], as well as in bioorganic catalysis [19], photochemistry [20], and electrochemistry, combined with on-line analysis [21,22]. For these reasons, studies focused on the reinterpretation of old transformations with this new technology are of particular interest in order to increase the number of synthetic tools exploitable in flow chemistry.
After a seminal study in which we demonstrated that bioinspired selenium-catalyzed reactions can be translated into continuous mode [23] some of the present authors, using a biphasic liquid–liquid system, reported the continuous oxidative cyclofunctionalization of alkenoic acids. The reaction conditions were optimized for the synthesis and purification of a library of biologically relevant lactones using a fully automated flow setup [24]. We clearly observed that the flow conditions produced a sensible reduction of the reaction time (residence time) as a consequence of a slug flow, which created local vortex fields able to increase the mass transfer rates between the organic and the aqueous phases. Here, we report our recent results in the selenium-catalyzed oxidation of sulfides under flow conditions to selectively afford sulfoxides and sulfones in line with the principles of efficiency, safety, mild reaction conditions, and overall atom economy.
The first example of a continuous oxidation of sulfides was reported by Nunez in 2010, who used silica-supported peracid and supercritical carbon dioxide and obtained a moderate level of chemoselectivity depending on the pressure [25]. In the same period, a green method was reported for the continuous and chemoselective synthesis of a few arylsulfoxides using diluted hydrogen peroxide and the sulfonic resin Amberlite 120 H [26].
More recently, a selective method based on the use of a microflow electrocell for electrochemical oxidation was reported, in which the selectivity can be tuned by changing the applied potential [27], and a method in which a solution of sulfide in dichloromethane and 15 vol. % of CF3COOH were fluxed in a packed-bed reactor filled with Oxone®®. In the latter case, an example of selective mono-oxygenation of a disulfide was also described [28].
In our protocol, a bioinspired approach was implemented for the continuous and chemoselective oxidation of sulfides (1) into their corresponding sulfoxides (2) or sulfones (3) (Scheme 1). The reaction was carried out using SeO2 as the most atom-economical Se-based pre-catalyst, in a liquid–liquid biphasic system at room temperature with hydrogen peroxide as the green oxidant.

2. Results and Discussions

Some years ago, Drabowicz et al. reported that the synthesis of sulfoxides [29] and sulfones [30] can be achieved by starting from the corresponding sulfides, using hydrogen peroxide/selenoxide system to form peroxyselenic (IV) acid in situ as the actual stoichiometric oxidizing reagent. Even if the synthetic versatility of selenium (IV) oxide is widely reported in the literature, its use in organic synthesis, especially as a stoichiometric reagent, is often limited by its toxicity and tendency to generate volatile and malodorous side products [31]. With the aim of overcoming these limitations, we decided to explore the possibility of using SeO2 as the catalyst in a bioinspired oxidation process in the presence of hydrogen peroxide using liquid–liquid biphasic flow conditions. Preliminary investigations were carried out by following the conversion of phenylethylsulfide 1a into its sulfoxide (2a) and sulfone (3a). The effects of the flow rate and the amount of hydrogen peroxide were evaluated, looking at the conversion as well as the chemoselectivity of the reaction.
A 0.5 M solution of the substrate 1a in ethyl acetate (Solution A) and a 0.05 M solution of SeO2 in water in the presence of the different amounts of H2O2 reported in Table 1 (Solution B), were fluxed through a Y junction in a 2 mL tubular reactor, at four different flow rates corresponding to residence times of 6.5, 10, 20 and 100 min. The chemoselectivity seemed to be influenced predominantly by the amount of hydrogen peroxide (Entries 1–5 vs. Entries 6–8). In entries 3–5 the conversion yield for the formation of the sulfoxide 2a increased with the reduction of the flow rate from 0.3 mL/min to 0.1 mL/min and, in these conditions, using 2 equivalents of hydrogen peroxide, the target compound (2a) was obtained in an 85% yield and with 100% chemoselectivity. When the reaction was repeated in the same conditions but without SeO2, only the starting material was observed in the 1H-NMR spectrum of the crude product, confirming the catalytic role of selenium. Increasing the amount of hydrogen peroxide from 2 (entry 5) to 5 (entry 7) and 10 (entries 7,8) equivalents also increased the formation of the sulfone 3a, which was quantitatively obtained as a unique reaction product when the reaction was carried out in the presence of 10 equivalents of H2O2 at both 0.1 and 0.2 mL/min (entries 7,8). The proposed mechanism for the catalytic cycle is reported in Scheme 2. Selenium (IV) oxide in water is in equilibrium with the hydrated form which is particularly prone to oxidation by hydrogen peroxide, affording the corresponding peroxyselenic (IV) acid as the actual catalyst in the formation of both the sulfoxide 2 and sulfone 3, depending on the amount of the oxidant.
From these preliminary data, the reaction conditions reported in Entry 5 and Entry 8 were selected as the best ones for the evaluation of the scope. A set of sulfides (Compounds 1ag, Table 2) was converted into the corresponding sulfoxides 2af and sulfones 3ag.
The results obtained for the synthesis of sulfoxides 2ag are summarized in Table 2. All the tested aryl sulfides 1af afforded the corresponding mono-oxygenated derivatives 2af in very high yields and with complete chemoselectivity. Only in the case of the naphtylethylsulfide 1d was it necessary to use diluted solutions, due to a lower solubility leading to poor conversion into the target compound 2d (20%). In order to increase the yield, we decided to use 5 equivalents of hydrogen peroxide. In these new conditions, 1d was quantitatively converted in an 80:20 mixture of 2d/3d, from which 2d was isolated in a 75% yield after chromatographic purification by flash chromatography. In the case of the aliphatic tetrahydrothiofene 1g, it was impossible to optimize the conditions for the synthesis and purification of sulfoxide 2g, and a mixture of 1g, 2g, and 3g was obtained that proved to be neither separable by chromatography nor unequivocally quantifiable by 1H NMR.
Excellent results were also obtained for the synthesis of the sulfones 3ag, (Table 3). In all the cases, the desired products were obtained pure and in quantitative yield simply by the separation of the organic layers and the removal under vacuum of the solvent, without the need for any other purification. In the cases of substrate 1d, we used a 0.25 M solution of the starting material and a 0.025 M solution of the catalyst to overcome the solubility issues of the corresponding sulfone 3d, without any detrimental effect on the yield.
In these conditions, the aliphatic substrate 1g was quantitatively converted into the corresponding sulfone 3g, which was isolated and characterized using 1H and 13C-NMR spectroscopy.
It was observed that in the case of vinyl sulfide 1e, the proposed methodology afforded only the oxidation of the sulfur atom, and in both conditions for the synthesis of the sulfoxide 2d or the sulfone 3d, the C=C was not oxidized.
Finally, in order to test the scalability of the process, 10 mmol of sulfide 1b was processed using the conditions reported in Table 3 affording, in 200 min, quantitatively a 91:9 mixture of 3b/2b corresponding to a processing rate of 10.16 gd−1 for 3b. (See Supplementary Materials).

3. Materials and Methods

Solvents and reagents were purchased from Sigma-Aldrich (St Louis, MS, USA), Alfa Aesar (Kandel, Germany), and VWR (Milano, Italy), and used as received unless otherwise noted. Compound 1d was prepared according to the procedure reported in literature. [32] Analytical thin-layer chromatography (TLC) was performed on silica-gel-precoated (60F-254) aluminum foil sheets (Merck, Darmstadt, Germany), using short-wave UV light (TUV T8 HO 95W G13 UVC, Philips, Milano, Italy) as the visualizing agent. NMR spectroscopic (Bruker, Fällanden, Switzerland) experiments were conducted at 25 °C on Bruker DPX and DRX spectrometers operating at the specified frequencies. 1H and 13C chemical shifts (δ) are reported in parts per million (ppm), relative to TMS (δ = 0.0 ppm) and the residual solvent peak of CDCl3 (δ = 7.27 and 77.00 ppm in 1H and 13C NMR, respectively). Data are reported as chemical shift (multiplicity, coupling constants where applicable, number of hydrogen atoms). Abbreviations are: s (singlet), d (doublet), t (triplet), q (quartet), dd (doublet of doublet), dt (double of triplet), tt (triplet of triplet), m (multiplet), br. s (broad signal). Coupling constants (J) are quoted in Hertz (Hz) to the nearest 0.1 Hz. Melting points were measured using a Kofler hot-stage-microscope Thermovar (Reichert, Vienna, Austria) and are reported as uncorrected data.

3.1. Synthesis of Sulfoxides 2a–f

In a 2 mL volumetric flask, ethyl acetate was added to sulfides 1a–c,e–g (1.0 mmoL) or sulfide 1d (0.5 mmoL) to make up the required volume. (Solution A: 0.5 M (1a–c,e–g). In the case of 1d the concentration was optimized at 0.25 M.) Into a 2 mL volumetric flask, 0.1 mmol of SeO2 (11 mg) and H2O2 30 wt. % (2 mmol, 205 µL) were poured, and then water was added until the desired volume was reached. (Solution B: 0.05 M (SeO2) and 1 M (H2O2). In the case of 1d, a different concentration was optimized at 0.025 M (SeO2) and 1.25 M (H2O2)).
Solutions A and B were fluxed at the same flow rate through a multi-syringe pump apparatus (Chemyx Fusion 100, Stafford, TX, USA) equipped with two syringes (2 mL), and they were mixed in a Y-junction and flowed at 0.100 mL/min through a tubularPTFE reactor coil with internal diameter of 1 mm and internal volume of 2 mL (Bohlender GmbH, Grünsfeld, Germany) at room temperature. The system was washed with 4 mL of ethyl acetate, and the aqueous and organic layers were collected in a flask, quenched with an aqueous solution of Na2S2O3 10% (w/v) and extracted with ethyl acetate (5 mL × 3). The combined organic layers were dried over sodium sulfate and concentrated under reduce pressure, affording the target sulfoxides. When required, the unreacted starting material was removed under vacuum, and in the case of 2d, it was separated from the corresponding sulfone (3d) by flash chromatography on a silica gel column using EtOAc/petroleum ether (2:98) as eluent.

3.2. Synthesis of Sulfones 2a–g

In a 2 mL volumetric flask, ethyl acetate was added to sulfides 1a–c,e–g (1.0 mmoL) or sulfide 1d, (0.5 mmoL) to make up the required volume. (Solution A: 0.5 M (1a–c,e–g). In the case of 1d, the concentration was optimized at 0.25 M) In a 2 mL volumetric flask, 0.1 mmol of SeO2 (11 mg) and H2O2 30 wt. % (10 mmoL, 1.025 mL) were poured, and then water was added until the desired volume was reached. (Solution B: 0.05 M (SeO2) and 5 M (H2O2). In the case of 1d, a different concentration was optimized at 0.025 M (SeO2) and 2.5 M (H2O2)).
Solutions A and B were fluxed at the same flow rate through a multi-syringe pump apparatus equipped with two syringes (2 mL), and they were mixed in a Y-junction and flowed at 0.200 mL/min through a tubular PTFE reactor coil with internal diameter of 1 mm and internal volume of 2 mLat room temperature. The system was washed with 4mL of ethyl acetate, and the aqueous and organic layers were collected in a flask, quenched with an aqueous solution of Na2S2O3 10% (w/v), and extracted with ethyl acetate (5 mL × 3). The combined organic layers were dried over sodium sulfate and concentrated under reduce pressure affording the target sulfones without further purification.

3.3. Spectral Data

Figures of NMR spectra of all the synthesized compounds are reported in the Supplementary Materials
 
Ethyl(phenyl)sulfoxide (2a) [33]: pale yellow oil (131 mg, 0.85 mmoL, yield 85%).1H-NMR (400 MHz, CDCl3): δ 7.64–7.53 (m, 5H), 2.96–2.89 (m, 1H), 2.82–2.77 (m, 1H), 1.22 (t, J = 7.38 Hz, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 143.2, 131.0, 50.3, 124.2, 129.2, 6.0, ppm.
 
Methyl(phenyl)sulfoxide (2b) [33]: pale red oil (127 mg, 0.9 mmoL, yield 91%).1H-NMR (400 MHz, CDCl3): δ 7.58–7.56 (m, 2H), 7.46–7.44 (m, 3H), 2.65 (s, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 145.6, 131.1, 129.4, 123.5, 44.0 ppm.
 
Methyl(4-methylphenyl)-sulfoxide (2c) [33]: pale yellow oil (137.1 mg, 0.9 mmoL, yield 89%).1H-NMR (200 MHz, CDCl3): δ 7.58–7.54 (m, 2H), 7.37–7.32 (m, 2H), 2.72 (s, 3H), 2.44 (s, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 142.4, 141.5, 130.0, 123.5, 44.0, 21.4 ppm.
 
Ethyl(naphthalen-2-yl)sulfoxide (2d) [32]: pale yellow oil (82 mg, 0.375 mmoL, yield 75%).1H-NMR (200 MHz, CDCl3): δ 8.16 (s, 1H), 8.96–7.87 (m, 3H), 7.57-7.53 (m, 3H), 3.02–2.93 (m, 1H), 2.86–2.77 (m, 1H), 1.19 (t, J = 7.4 Hz, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 140.3, 134.4, 132.8, 129.4, 128.5, 128.1, 127.7, 127.3, 125.0, 120.0, 6.0, 49.9 ppm.
 
Phenyl(ethenyl)sulfoxide (2e) [34]: pale yellow oil (112 mg, 0.7 mmoL, yield 74%).1H-NMR (400 MHz, CDCl3): δ 7.56 -7.54 (m, 2H), 7.45-7.43 (m, 3H), 6.53 (dd, J = 9.5 Hz and J = 16.5 Hz, 1H), 6.14 (d, J = 16.4 Hz, 1H), 5.83 (d, J = 9.6 Hz, 1H) ppm. 13C-NMR (50 MHz, CDCl3): δ 143.3, 142.9, 131.3, 129.5, 124.7, 120.8 ppm.
 
Phenyl(benzyl)sulfoxide(2f) [35]: white crystals (m.p. 120–122 °C, lit 124–126 °C) [36]: (216 mg, 1 mmoL, quantitative yield).1H-NMR (200 MHz, CDCl3): δ 7.44–7.24 (m, 8H), 6.99–6.96 (m, 2H), 4.12–3.95 (m, 2H) ppm. 13C-NMR (50 MHz, CDCl3): δ 142.4 13.5, 130.6, 128.9, 128.5, 128.4, 128.2, 124.2, 63.2, ppm.
 
(Ethylsulfonyl)benzene (3a) [37]: white crystals (m.p. 39–42 °C, lit 39–41 °C) [37] (168 mg, 1 mmol, yield > 99%).1H-NMR (400 MHz, CDCl3): δ 7.87–7.82 (m, 2H), 7.61-7.47 (m, 3H), 3.06 (q, J = 6.9 Hz, 2H), 1.21 (t, J = 6.9 Hz, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 138.4, 133.8, 129.3, 128.2, 50.6, 7.5, ppm.
 
(Methylsulfonyl)benzene (3b) [37]: white solid (m.p. 84–86 °C, lit 85–87 °C) [37] (156 mg, 1 mmoL, yield > 99%). 1H-NMR (200 MHz, CDCl3): δ 7.99-7.96 (m, 2H), 7.73-7.56 (m, 3H), 3.08 (s, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 140.5, 133.7, 129.4, 127.4, 44.5 ppm.
 
1-Methyl-4-(methylsulfonyl)benzene (3c) [37]: white crystals (m.p. 85–87 °C, lit 82–85 °C) [37] (169 mg, 1 mmol, yield >99%).1H-NMR (200 MHz, CDCl3): δ 7.87-7.83 (m, 2H), 7.41-7.36 (m, 2H), 3.05 (s, 3H), 2.46 (s, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 144.7, 137.7, 130.0, 127.4, 44.6, 21.7 ppm.
 
2-(Ethylsulfonyl)naphthalene (3d) [32]: brown oil (109.8 mg, 0.5 mmoL, yield > 99%).1H-NMR (200 MHz, CDCl3): δ 8.36 (s, 1H), 7.89-7.72 (m, 4H), 7.54-7.47 (m, 2H), 3.07 (q, J = 7.0 Hz, 2H), 1.16 (t, J = 7.0 Hz, 3H) ppm. 13C-NMR (50 MHz, CDCl3): δ 7.3, 50.3, 122.5, 127.5, 127.8, 128.1, 128.3, 128.7, 129.1, 129.1, 129.4, 129.8, 130.6, 131.9, 135.0.
 
(Vinylsulfonyl)benzene (3e) [34]: white crystals m.p. (64–65 °C lit 64–65 °C) [34]: (167 mg, 1 mmol, yield > 99%).1H-NMR (200 MHz, CDCl3): δ 7.84–7.82 (m, 2H). 7.58–7.47 (m, 3H), 6.60 (dd, J = 9.8 Hz, 16.5 Hz, 1H), 6.40 (d, J = 16.5 Hz, 1H), 5.98 (d, J = 9.8 Hz, 1H) ppm. 13C-NMR (50 MHz, CDCl3): δ 139.5 138.4, 133.7, 129.4, 127.9, 127.8 ppm.
 
(Benzylsulfonyl)benzene (3f) [35]: white crystals (m.p. 147–149 °C, lit 147–149 °C) [38]: (231 mg, 1 mmol, yield > 99%).1H-NMR (200 MHz, CDCl3): δ 7.57–7.52 (m, 3H), 7.40–7.37 (m, 2H), 7.26–7.19 (m, 3H), 7.02-7.01 (m, 2H), 4.25 (s, 2H) ppm. 13C-NMR (50 MHz, CDCl3): δ 137.8, 133.8, 130.8, 128.9, 128.8, 128.7, 128.6, 128.1, 62.9 ppm.
 
Tetrahydrothiophene 1,1-dioxide (3g) [37]: pale yellow oil (120 mg, 1 mmoL, yield > 99%).1H-NMR (200 MHz, CDCl3):δ 3.08–3.01 (m, 2H), 2.27-2.20 (m, 2H).13C-NMR (50 MHz, CDCl3): δ 51.2, 22.8, ppm.

4. Conclusions

We demonstrated herein that, following a bioinspired approach, SeO2 can be conveniently used as pre-catalyst in the chemoselective preparation of sulfoxides and sulfones under flow conditions. The use of hydrogen peroxide as final oxidant allowed the safe in situ generation of perselenic acid as the actual catalyst, affording the target compounds in good to excellent yields. The proposed protocol is particularly attractive in terms of its simplicity, and further demonstrates that organoselenium catalysts can be conveniently employed in continuous oxidation reactions using liquid–liquid biphasic systems, improving the safety, efficiency, and the greenness of these processes.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/25/11/2711/s1. Figures of NMR spectra of all the synthesized compounds.

Author Contributions

Conceptualization, C.S. and F.M. (Francesca Marini) methodology, C.S. and L.S.; validation, C.S., F.M. (Francesca Marini) and L.S.; formal analysis, F.M. (Francesca Mangiavacchi) and L.S.; investigation, F.M. (Francesca Mangiavacchi) and L.C.; resources, C.S.; data curation, F.M. (Francesca Marini) and L.S.; writing—Original draft preparation, C.S. and F.M. (Francesca Mangiavacchi); writing—Review and editing, C.S., L.S., F.M. (Francesca Marini); supervision, C.S. and F.M. (Francesca Marini); project administration, C.S. and F.M. (Francesca Marini); funding acquisition, C.S. and F.M. (Francesca Marini) All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fondo per il sostegno della Ricerca di Base 2018, Project “Sviluppo di metodologie innovative per la sintesi efficiente di composti eterociclici, molecole drug-like e intermedi sintetici ad alto valore aggiunto”.

Acknowledgments

The authors thanks University of Perugia and the consortium C.I.N.M.P.I.S. for the support. This work has been undertaken under the umbrella of the Selenium Sulfur Redox and Catalysis Network (SeSRedCat).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bäckvall, J.-E. Modern Oxidation Methods, 2nd ed.; Completely rev. and enlarged ed.; Bäckvall, J.-E., Ed.; Wiley-VCH: Weinheim, Germany, 2010; ISBN 978-3-527-32320-3. [Google Scholar]
  2. Surur, A.S.; Schulig, L.; Link, A. Interconnection of sulfides and sulfoxides in medicinal chemistry. Arch. Pharm. 2018, 1800248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Xu, F.; Chen, Y.; Fan, E.; Sun, Z. Synthesis of 3-Substituted Aryl [4,5] isothiazoles through an All-Heteroatom Wittig-Equivalent Process. Org. Lett. 2016, 18, 2777–2779. [Google Scholar] [CrossRef] [PubMed]
  4. Pieters, L.; Van Dyck, S.; Gao, M.; Bai, R.; Hamel, E.; Vlietinck, A.; Lemière, G. Synthesis and Biological Evaluation of Dihydrobenzofuran Lignans and Related Compounds as Potential Antitumor Agents that Inhibit Tubulin Polymerization. J. Med. Chem. 1999, 42, 5475–5481. [Google Scholar] [CrossRef]
  5. Sturino, C.F.; O’Neill, G.; Lachance, N.; Boyd, M.; Berthelette, C.; Labelle, M.; Li, L.; Roy, B.; Scheigetz, J.; Tsou, N.; et al. Discovery of a Potent and Selective Prostaglandin D2 Receptor Antagonist, [(3R)-4-(4-Chloro- benzyl)-7-fluoro-5-(methylsulfonyl)-1,2,3,4-tetrahydrocyclopenta [b] indol-3-yl]-acetic Acid (MK-0524) . J. Med. Chem. 2007, 50, 794–806. [Google Scholar] [CrossRef]
  6. Yazdanyar, S.; Boer, J.; Ingvarsson, G.; Szepietowski, J.C.; Jemec, G.B.E. Dapsone Therapy for Hidradenitis Suppurativa: A Series of 24 Patients. Dermatology 2011, 222, 342–346. [Google Scholar] [CrossRef] [PubMed]
  7. Sunduru, N.; Salin, O.; Gylfe, Å.; Elofsson, M. Design, synthesis and evaluation of novel polypharmacological antichlamydial agents. Eur. J. Med. Chem. 2015, 101, 595–603. [Google Scholar] [CrossRef] [Green Version]
  8. Aono, Y.; Horinaka, M.; Iizumi, Y.; Watanabe, M.; Taniguchi, T.; Yasuda, S.; Sakai, T. Sulindac sulfone inhibits the mTORC1 pathway in colon cancer cells by directly targeting voltage-dependent anion channel 1 and 2. Biochem. Biophys. Res. Commun. 2018, 505, 1203–1210. [Google Scholar] [CrossRef]
  9. Spencer, C.M.; Faulds, D. Esomeprazole. Drugs 2000, 60, 321–329. [Google Scholar] [CrossRef]
  10. Seto, M.; Aikawa, K.; Miyamoto, N.; Aramaki, Y.; Kanzaki, N.; Takashima, K.; Kuze, Y.; Iizawa, Y.; Baba, M.; Shiraishi, M. Highly Potent and Orally Active CCR5 Antagonists as Anti-HIV-1 Agents: Synthesis and Biological Activities of 1-Benzazocine Derivatives Containing a Sulfoxide Moiety. J. Med. Chem. 2006, 49, 2037–2048. [Google Scholar] [CrossRef]
  11. Santoro, S.; Azeredo, J.B.; Nascimento, V.; Sancineto, L.; Braga, A.L.; Santi, C. The green side of the moon: Ecofriendly aspects of organoselenium chemistry. RSC Adv. 2014, 4, 31521–31535. [Google Scholar] [CrossRef]
  12. Tidei, C.; Piroddi, M.; Galli, F.; Santi, C. Oxidation of thiols promoted by PhSeZnCl. Tetrahedron Lett. 2012, 53, 232–234. [Google Scholar] [CrossRef]
  13. Santoro, S.; Santi, C.; Sabatini, M.; Testaferri, L.; Tiecco, M. Eco-Friendly Olefin Dihydroxylation Catalyzed by Diphenyl Diselenide. Adv. Synth. Catal. 2008, 350, 2881–2884. [Google Scholar] [CrossRef]
  14. Santi, C.; Di Lorenzo, R.; Tidei, C.; Bagnoli, L.; Wirth, T. Stereoselective selenium catalyzed dihydroxylation and hydroxymethoxylation of alkenes. Tetrahedron 2012, 68, 10530–10535. [Google Scholar] [CrossRef]
  15. Sancineto, L.; Mangiavacchi, F.; Tidei, C.; Bagnoli, L.; Marini, F.; Gioiello, A.; Scianowski, J.; Santi, C. Selenium-Catalyzed Oxacyclization of Alkenoic Acids and Alkenols. Asian J. Org. Chem. 2017, 6, 988–992. [Google Scholar] [CrossRef]
  16. Sancineto, L.; Tidei, C.; Bagnoli, L.; Marini, F.; Lenardão, E.; Santi, C. Selenium Catalyzed Oxidation of Aldehydes: Green Synthesis of Carboxylic Acids and Esters. Molecules 2015, 20, 10496–10510. [Google Scholar] [CrossRef]
  17. Santi, C.; Jacob, R.; Monti, B.; Bagnoli, L.; Sancineto, L.; Lenardão, E. Water and Aqueous Mixtures as Convenient Alternative Media for Organoselenium Chemistry. Molecules 2016, 21, 1482. [Google Scholar] [CrossRef] [Green Version]
  18. Baumann, M.; Moody, T.S.; Smyth, M.; Wharry, S. A Perspective on Continuous Flow Chemistry in the Pharmaceutical Industry. Org. Process Res. Dev. 2020. [Google Scholar] [CrossRef]
  19. Valikhani, D.; Srivastava, P.L.; Allemann, R.K.; Wirth, T. Immobilised Enzymes for Sesquiterpene Synthesis in Batch and Flow Systems. ChemCatChem 2020, 12, 2194–2197. [Google Scholar] [CrossRef]
  20. Maljuric, S.; Jud, W.; Kappe, C.O.; Cantillo, D. Translating batch electrochemistry to single-pass continuous flow conditions: An organic chemist’s guide. J. Flow Chem. 2020, 10, 181–190. [Google Scholar] [CrossRef] [Green Version]
  21. Cambié, D.; Noël, T. Solar Photochemistry in Flow. Top. Curr. Chem. 2018, 376. [Google Scholar] [CrossRef] [Green Version]
  22. Santi, M.; Seitz, J.; Cicala, R.; Hardwick, T.; Ahmed, N.; Wirth, T. Memory of Chirality in Flow Electrochemistry: Fast Optimisation with DoE and Online 2D-HPLC. Chem. A Eur. J. 2019, 25, 16230–16235. [Google Scholar] [CrossRef]
  23. Di Schino, L.; Incipini, L.; Dragone, V.; Tidei, C.; Scalera, C.; Santi, C. Green Chemistry for Enviromental Sustainability: An Example of “Bio-Logic” Approach. In Proceedings of the 1st World Sustainability Forum, Basel, Switzerland, 1–30 November 2011; MDPI: Basel, Switzerland, 2011; p. 675. [Google Scholar]
  24. Cerra, B.; Mangiavacchi, F.; Santi, C.; Lozza, A.M.; Gioiello, A. Selective continuous flow synthesis of hydroxy lactones from alkenoic acids. React. Chem. Eng. 2017, 2, 467–471. [Google Scholar] [CrossRef]
  25. Mello, R.; Olmos, A.; Alcalde-Aragonés, A.; Díaz-Rodríguez, A.; González Núñez, M.E.; Asensio, G. Oxidation of Sulfides with a Silica-Supported Peracid in Supercritical Carbon Dioxide under Flow Conditions: Tuning Chemoselectivity with Pressure. Eur. J. Org. Chem. 2010, 2010, 6200–6206. [Google Scholar] [CrossRef]
  26. Maggi, R.; Chitsaz, S.; Loebbecke, S.; Piscopo, C.G.; Sartori, G.; Schwarzer, M. Highly chemoselective metal-free oxidation of sulfides with diluted H2O2 in a continuous flow reactor. Green Chem. 2011, 13, 1121. [Google Scholar] [CrossRef]
  27. Laudadio, G.; Straathof, N.J.W.; Lanting, M.D.; Knoops, B.; Hessel, V.; Noël, T. An environmentally benign and selective electrochemical oxidation of sulfides and thiols in a continuous-flow microreactor. Green Chem. 2017, 19, 4061–4066. [Google Scholar] [CrossRef] [Green Version]
  28. Silva, F.; Baker, A.; Stansall, J.; Michalska, W.; Yusubov, M.S.; Graz, M.; Saunders, R.; Evans, G.J.S.; Wirth, T. Selective Oxidation of Sulfides in Flow Chemistry: Selective Oxidation of Sulfides in Flow Chemistry. Eur. J. Org. Chem. 2018, 2018, 2134–2137. [Google Scholar] [CrossRef]
  29. Drabowicz, J.; Mikołajczyk, M. A Facile and Selective Oxidation of Organic Sulphides to Sulphoxides with Hydrogen Peroxide/Selenium Dioxide System. Synthesis 1978, 1978, 758–759. [Google Scholar] [CrossRef]
  30. Drabowicz, J.; łyzwa, P.; Mikołajczyk, M. A New Procedure for Oxidation of Sulfides to Sulfones. Phosphorus Sulfur Relat. Elem. 1983, 17, 169–172. [Google Scholar] [CrossRef]
  31. Młochowski, J.; Wójtowicz-Młochowska, H. Developments in Synthetic Application of Selenium(IV) Oxide and Organoselenium Compounds as Oxygen Donors and Oxygen-Transfer Agents. Molecules 2015, 20, 10205–10243. [Google Scholar] [CrossRef] [Green Version]
  32. Chuo, T.H.; Boobalan, R.; Chen, C. Camphor-based schiff base of 3-endo-aminoborneol (SBAB): Novel ligand for vanadium-catalyzed asymmetric sulfoxidation and subsequent kinetic resolution. ChemistrySelect 2016, 1, 2174–2180. [Google Scholar] [CrossRef]
  33. Yuan, Y.; Shi, X.; Liu, W. Transition-Metal-Free, Chemoselective Aerobic Oxidations of Sulfides and Alcohols with Potassium Nitrate and Pyridinium Tribromide or Bromine. Synlett 2011, 2011, 559–564. [Google Scholar] [CrossRef]
  34. Gogoi, S.R.; Boruah, J.J.; Sengupta, G.; Saikia, G.; Ahmed, K.; Bania, K.K.; Islam, N.S. Peroxoniobium (V)-catalyzed selective oxidation of sulfides with hydrogen peroxide in water: A sustainable approach. Catal. Sci. Technol. 2015, 5, 595–610. [Google Scholar] [CrossRef]
  35. Fukuda, N.; Ikemoto, T. Imide-Catalyzed Oxidation System: Sulfides to Sulfoxides and Sulfones. J. Org. Chem. 2010, 75, 4629–4631. [Google Scholar] [CrossRef]
  36. Russell, G.A.; Pecoraro, J.M. Pummerer reaction of para-substituted benzylic sulfoxides. J. Org. Chem. 1979, 44, 3990–3991. [Google Scholar] [CrossRef]
  37. Yang, C.; Jin, Q.; Zhang, H.; Liao, J.; Zhu, J.; Yu, B.; Deng, J. Tetra-(tetraalkylammonium)octamolybdate catalysts for selective oxidation of sulfides to sulfoxides with hydrogen peroxide. Green Chem. 2009, 11, 1401. [Google Scholar] [CrossRef]
  38. Jain, S.L.; Sain, B. Perfluorinated resinsulphonic acid (Nafion-H®®) catalyzed highly efficient oxidations of organic compounds with hydrogen peroxide. Appl. Catal. A Gen. 2006, 301, 259–264. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from the authors.
Figure 1. Selected examples of aryl sulfones and aryl sulfoxides.
Figure 1. Selected examples of aryl sulfones and aryl sulfoxides.
Molecules 25 02711 g001
Scheme 1. The oxidation of sulfide 1 into the corresponding sulfoxide 2 and sulfone 3.
Scheme 1. The oxidation of sulfide 1 into the corresponding sulfoxide 2 and sulfone 3.
Molecules 25 02711 sch001
Scheme 2. Proposed catalytic mechanism.
Scheme 2. Proposed catalytic mechanism.
Molecules 25 02711 sch002
Table 1. Results of the preliminary screening of the flow-rate conditions obtained by fluxing Solution A (1a [0.5 M] in EtOAc) and Solution B (SeO2 [0.05 M] and H2O2 in H2O) at room temperature, each at the half-total flow rate.
Table 1. Results of the preliminary screening of the flow-rate conditions obtained by fluxing Solution A (1a [0.5 M] in EtOAc) and Solution B (SeO2 [0.05 M] and H2O2 in H2O) at room temperature, each at the half-total flow rate.
Molecules 25 02711 i001
EntryH2O2 eq [conc]Total Flow RatemL/minResidence Time (min)NMR Conversion %Selectivity 2/3
11 [0.5 M]0.36.559%95:5
21 [0.5 M]0.0210061%100:0
32 [1.0 M]0.36.560%100:0
42 [1.0 M]0.21078%100:0
52 [1.0 M]0.12085%100:0
65 [2.5 M]0.120>99%71:29
710 [5.0 M]0.120>99%0:100
810 [5.0 M]0.210>99%0:100
Table 2. Flow synthesis of sulfoxides 2af. (Each solution was fluxed at 0.05 mL/min.).
Table 2. Flow synthesis of sulfoxides 2af. (Each solution was fluxed at 0.05 mL/min.).
Molecules 25 02711 i002
SubstrateProductNMR Conversion
(Isolated Yield) 2 1
Selectivity 2/3 1
Molecules 25 02711 i003 Molecules 25 02711 i00485% (85%)100:0
1a2a
Molecules 25 02711 i005 Molecules 25 02711 i00691% (91%)100:0
1b2b
Molecules 25 02711 i007 Molecules 25 02711 i00893% (89%)100:0
1c2c
Molecules 25 02711 i009 Molecules 25 02711 i01020%
99% 2(75%)
100:0
80:20
1d2d
Molecules 25 02711 i011 Molecules 25 02711 i01280% (74%)100:0
1e2e
Molecules 25 02711 i013 Molecules 25 02711 i01499% (99%) 100:0
1f2f
Molecules 25 02711 i015 Molecules 25 02711 i016n.d. 3n.d. 3
1g2g
1 Conversion from the crude product, evaluated by 1H-NMR, into the isolated yields shown in brackets;; 2 Solution A (1d, [0.25 M] in EtOAc), Solution B (SeO2 [0.025 M] and H2O2 [1.25 M] in H2O). 3 Not determined (a non-resolvable mixture of 1g, 2g, and 3g was obtained, and it was not possible to calculate the corresponding ratio from the 1H-NMR spectrum of the crude product).
Table 3. Flow synthesis of sulfones 3ag. (Each solution was fluxed at 0.1 mL/min).
Table 3. Flow synthesis of sulfones 3ag. (Each solution was fluxed at 0.1 mL/min).
Molecules 25 02711 i017
SubstrateProductYield of 3
Molecules 25 02711 i018 Molecules 25 02711 i019> 99%
1a3a
Molecules 25 02711 i020 Molecules 25 02711 i021> 99%
1b3b
Molecules 25 02711 i022 Molecules 25 02711 i023> 99%
1c3c
Molecules 25 02711 i024 Molecules 25 02711 i025> 99% 1
1d3d
Molecules 25 02711 i026 Molecules 25 02711 i027> 99%
1e3e
Molecules 25 02711 i028 Molecules 25 02711 i029> 99%
1f3f
Molecules 25 02711 i030 Molecules 25 02711 i031> 99%
1g3g
1 To avoid precipitation of the reaction product, diluted solutions were used: Solution A (1d [0.25 M] in EtOAc), Solution B (SeO2 [0.025 M] and H2O2 [2.5 M] in H2O).

Share and Cite

MDPI and ACS Style

Mangiavacchi, F.; Crociani, L.; Sancineto, L.; Marini, F.; Santi, C. Continuous Bioinspired Oxidation of Sulfides. Molecules 2020, 25, 2711. https://doi.org/10.3390/molecules25112711

AMA Style

Mangiavacchi F, Crociani L, Sancineto L, Marini F, Santi C. Continuous Bioinspired Oxidation of Sulfides. Molecules. 2020; 25(11):2711. https://doi.org/10.3390/molecules25112711

Chicago/Turabian Style

Mangiavacchi, Francesca, Letizia Crociani, Luca Sancineto, Francesca Marini, and Claudio Santi. 2020. "Continuous Bioinspired Oxidation of Sulfides" Molecules 25, no. 11: 2711. https://doi.org/10.3390/molecules25112711

Article Metrics

Back to TopTop