Next Article in Journal
Towards “Green” ANFO: Study of Perchlorates and Inorganic Peroxides as Potential Additives
Next Article in Special Issue
Simmons–Smith Cyclopropanation: A Multifaceted Synthetic Protocol toward the Synthesis of Natural Products and Drugs: A Review
Previous Article in Journal
Assessment of Purity, Stability, and Pharmacokinetics of NGP-1, a Novel Prodrug of GS441254 with Potential Anti-SARS-CoV-2 Activity, Using Liquid Chromatography
Previous Article in Special Issue
Diastereoselective Formal 1,3-Dipolar Cycloaddition of Trifluoroethyl Amine-Derived Ketimines Enables the Desymmetrization of Cyclopentenediones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DBDMH-Promoted Methylthiolation in DMSO: A Metal-Free Protocol to Methyl Sulfur Compounds with Multifunctional Groups

1
School of Chemistry, South China Normal University, GDMPA Key Laboratory for Process Control and Quality Evaluation of Chiral Pharmaceuticals, Guangzhou Key Laboratory of Analytical Chemistry for Biomedicine, Key Laboratory of Theoretical Chemistry of Environment, Ministry of Education, Guangzhou 510006, China
2
College of Pharmacy, Gannan Medical University, Ganzhou 341000, China
3
School of Mechanical and Automotive Engineering, South China University of Technology, Guangzhou 510640, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(15), 5635; https://doi.org/10.3390/molecules28155635
Submission received: 26 June 2023 / Revised: 14 July 2023 / Accepted: 24 July 2023 / Published: 25 July 2023
(This article belongs to the Special Issue Bioactive Heterocyclic Chemistry)

Abstract

:
Organic thioethers play an important role in the discovery of drugs and natural products. However, the green synthesis of organic sulfide compounds remains a challenging task. The convenient and efficient synthesis of 5-alkoxy-3-halo-4-methylthio-2(5H)-furanones from DMSO is performed via the mediation of 1,3-dibromo-5,5-dimethylhydantoin (DBDMH), affording a facile route for the sulfur-functionalization of 3,4-dihalo-2(5H)-furanones under transition metal-free conditions. This new approach has demonstrated the functionalization of non-aromatic Csp2-X-type halides with unique structures containing C-X, C-O, C=O and C=C bonds. Compared with traditional synthesis methods using transition metal catalysts with ligands, this reaction has many advantages, such as the lower temperature, the shorter reaction time, the wide substrate range and good functional group tolerance. Notably, DMSO plays multiple roles, and is simultaneously used as an odorless methylthiolating reagent and safe solvent.

Graphical Abstract

1. Introduction

Sulfur is one of the most fundamental elements in the life system in the form of proteins and amino acids [1], and its rich valence states [2] are the chemical basis for its extensive use in medical drugs [3], pesticides [4] and organic luminescent materials [5,6]. In particular, their universal role in biological metabolism makes them crucial for organisms from the ocean to the terrestrial system [7]. Therefore, among the 362 sulfur-containing drugs listed by the US Food and Drug Administration (FDA), sulfide and its derivatives are one of a series of leading pharmaceutical active compounds [8]. For example, several nitro- and amino-bound sulfides are biologically active ingredients and are widely used to treat inflammation, Alzheimer’s disease, HIV, breast cancer, malaria and fungal-related diseases. In a word, among numerous sulfur-containing compounds, thioethers are an important structural component existing in many biological and pharmaceutical molecules [9] (some important examples with the structure of the methyl sulfur group and nitrogen-containing heterocyclic ring are shown in Figure 1 [10,11]).
In addition, thioethers are also important intermediates in organic synthesis, which can be transformed into sulfones [12] and sulfoxides [13], or used as the substrates in the Sonogashira reaction [14] and other reactions [15]. At present, the construction methods of methyl sulfur compounds are mainly the reduction of sulfoxides [16,17] and the reaction of aryl thiols with iodides [18] or dimethyl carbonate [19]. Among the methods mentioned above, some sources of sulfur have obvious defects. In particular, for the more commonly used reaction of thiols (thiophenols) [20,21], these sulfur reagents bring many shortcomings, such as unpleasant odors, toxicity and harsh reaction conditions. Therefore, the development and utilization of more stable, environmentally friendly and economical sulfur reagents for the synthesis of sulfide compounds still have important significance.
Dimethyl sulfoxide (DMSO) is widely used as a high-quality solvent because of its high solubility for many organic and inorganic compounds [22,23]. In addition to being used as a solvent, DMSO has also been used as a multifunctional, inexpensive and safe reagent as a carbon [24,25], sulfur [26,27] and oxygen source [28,29] in many reactions. However, due to the low activity of reaction substrates (especially for chlorides) or the fact that these reactions often require a higher temperature and a longer reaction time, the synthesis research progress in using DMSO as a sulfur source with halogen-containing compounds to construct molecules with the methyl sulfur group is relatively slow. For example, in 2011, Cheng’s group [30] reported the copper-mediated methylthiolation of aryl halide and DMSO (Scheme 1a). Later, Mal’s group [31] further optimized the Cu-mediated method via the methylthiolation of halogenated aromatic hydrocarbons and DMSO by changing the reaction conditions (Scheme 1b).
In recent years, our group reported the methylthiolation reaction of the non-aromatic CSP2-X compound 3,4-dihalo-2(5H)-furanone with DMSO, further expanding the construction scope of methyl sulfur compounds when using halides and DMSO as substrates (Scheme 1c) [32]. Even so, it is a pity that, in these methods mentioned above, there are still shortcomings, such as the higher reaction temperature, the longer reaction time and the use of transition metal catalysis and ligands. Therefore, the green synthesis methods for preparing methyl sulfur compounds from readily available starting materials still need to be explored.
Sulfoxides are also a class of extremely important compounds because they are not only widely used as intermediates in synthesis but also as chiral ligands in various asymmetric catalysis [33,34]. In addition, sulfoxides can also be applied in many fields such as fine chemicals, pharmaceuticals, pesticides and functional materials [35,36]. On the basis of our interest in the synthesis of sulfur-containing compounds [37,38,39], especially the construction of a series of sulfur-containing compounds through the reaction of 3,4-dihalo-2(5H)-furanone with various sulfur-containing reagents (Scheme 2) [40,41,42,43], we aimed to introduce the sulfoxide group at the 4-position of 3,4-dihalo-2(5H)-furanone instead of the 4-halo group by controlling reaction conditions, in order to make the whole work of 2(5H)-furanone chemistry more systematical. However, as an unexpected result, a thioether structure instead of the sulfoxide group was found (Scheme 2).
It is worth noting that different 2(5H)-furanone compounds have been developed in many fields such as modern organic synthesis [44], drug molecular design [45] and natural product total synthesis [46] due to their potential high biological activities, such as antiviral [47] or anti-HIV activity [48]. In particular, the thioetherified 2(5H)-furanone can be a kind of potential drug molecule with good anticancer activity, as Li’s group reported before [49]. Therefore, it is quite important to develop a new method for the more efficient synthesis of the methylthioetherified 3,4-dihalo-2(5H)-furanone 3 (Scheme 1, this work). In addition, as a kind of compound containing multiple functional groups, the structure can be further modified in the next development and the obtained potential bioactive molecules may play an important role in drug development and design.
Thus, herein, on the basis of our group’s previous research on various functionalization reactions of 3,4-dihalo-2(5H)-furanones for bioactive compounds [50,51], we developed a simple, mild and transition metal-free method to realize the greener methylthiolation of 3,4-dihalo-2(5H)-furanones by further optimizing the conditions of the methylthiolation reaction from 3,4-dihalo-2(5H)-furanone and DMSO (Scheme 2, this work).

2. Results and Discussion

2.1. Optimization of Reaction Conditions

At the beginning of this study, 5-methoxy-3,4-dibromo-2(5H)-furanone 1a was used as a model substrate to screen the reaction conditions in DMSO (Table 1).
Firstly, we optimized the activating agent (entries 1–3). Obviously, compared with N-bromosuccinimide (NBS) and N-chlorosuccinimide (NCS), when using 1,3-dibromo-5,5-dimethylhydantoin (DBDMH) as an activator, the obtained effect is the best, and the corresponding yield is as high as 87% (entry 3).
Next, based on the above results, we chose DBDMH as the activator to further optimize the dosage of DBDMH (entries 3–6). It can be found that, when its dosage is 1.5 equivalents, the conversion rate of this reaction is up to 93% (entry 5). Therefore, we selected 1.5 equiv. DBDMH to further optimize the temperature required for the reaction (entry 5 vs. entries 7–8). As can be seen from a series of temperature exploration experiments, 80 °C is more conducive to the reaction (entry 5).
Finally, we also explored the time required for the reaction (entry 5 vs. entries 9–10). It is obvious that prolonging or reducing the reaction time is not beneficial for improving the yield, and the best time is still 5 h (entry 5).
Thus, taking the reaction of 5-methoxy-3,4-dibromo-2(5H)-furanone 1a in DMSO 2 as an example, the relatively ideal reaction conditions are as follows: using 0.4 mmol 5-methoxy-3,4-dibromo-2(5H)-furanone 1a to react with 2 mL DMSO at 80 °C for 5 h, the isolated yield of product 3a can reach 93%.

2.2. Investigation into the Range of Reaction Substrates

With the optimal conditions in hand, the substrate scope of this transformation was assessed. The results have been summarized, and are shown in Table 2.
Firstly, the tolerance of 3,4-dihalo-2(5H)-furanone substrates with different substituents R1 at the 5-position is examined when the halogen on the furanone ring is the bromine atom. Although the substituent groups at the 5-position on the furanone ring are different, the reaction can proceed smoothly (3a3j, 74–93%). In particular, the reaction is well tolerated even for the substitution of a strong electron-withdrawing group of biphenyl, giving a 46% yield of product 3k.
In addition, as expected, the yield decreased slightly with the extension of the carbon chain and the enhanced steric hindrance effect (e.g., 3a vs. 3d vs. 3h, 93% vs. 88% vs. 78%). Even so, it is satisfied that for the menthoxy group with large steric hindrance, the corresponding target product 3j can be obtained with an isolated yield of 74%.
When the halogen is chlorine on the furanone ring, due to the activity difference between the bromine and chlorine atoms, the yield of the corresponding target products 3l3t is between 49 and 72%, and the steric hindrance effect of the substitution group at the 5-position is similar.
It is noteworthy that although the yield of the electron-withdrawing group was relatively low (e.g., Ph- in 3t, 49%) compared with the yield data of the known compounds reported before [32], most of them were improved, especially for the cases where the previous yield was relatively low.

2.3. Structural Characterization Analysis

Firstly, the compounds synthesized herein were characterized using nuclear magnetic resonance (NMR) technology. From the 1H NMR spectra of the target compounds (please see the Supplementary Materials), it can be seen that the 1H NMR data of compounds 3a3t are consistent with the corresponding hydrogens in these target products. Similarly, the 13C NMR test results are also consistent.
In particular, for the synthesized new compounds, their spectra of high-resolution mass spectrometry (HR-MS) were also tested. Taking the target compound 3c as an example, it can be found that HR-MS can also correspond well with the structure of compound 3c (Figure 2). In short, from the analysis of the obtained HR-MS results in combination with other results, it is confirmed that DMSO can indeed react smoothly with 3,4-dihalo-2(5H)-furanones that are substituted with different alkoxy groups at the 5-position, giving the anticipated products.
Importantly, although nuclear NMR and mass spectrometry tests have confirmed the expected structure, in order to further determine the product structure, the single-crystal structure of 3a (CCDC 2270564) [52] is also obtained (the detailed data can be seen in Table 3, and the corresponding molecular structure of 3a can be seen in Table 2), which fully proves the structure of the anticipated product. Thus, the structures of the serial compounds are well characterized.

2.4. Mechanism Investigation and Gram-Scale Experiment

To have a deeper understanding into the reaction process, we performed two control experiments accordingly. Firstly, we added two equiv. radical scavenger 2,2,6,6-tetra- methylpiperidin-1-yloxyl (TEMPO) to the reaction system. It was found that the corresponding compound 3a can also be obtained in a 91% yield (Scheme 3a). And, compared with the reaction situation under the standard conditions, it is clear that the yield has almost no effect. This fully demonstrates that the reaction may not be involved in the pathway of radical participation.
Subsequently, the reaction under standard conditions for 5 h but without the addition of the activator DBDMH could not proceed smoothly (Scheme 3b). This indicates that DBDMH may be crucial in this transformation.
Compared with other brominating agents, such as NBS, or N-bromoacetamide, DBDMH as a special brominating agent has many advantages, e.g., high active bromine content, good storage stability and economic use [53,54,55]. In addition, acting as an oxidant in chemical synthesis, it also can be widely used in various transformations [56,57].
Thus, based on these relevant properties of DBDMH and the above-mentioned control experimental results, we proposed a possible reaction mechanism (Scheme 4), referring to the relevant literature reported before [58,59]:
Initially, with the promotion of DBDMH and heating, DMSO is successfully decomposed into dimethyl sulfide [60]. Subsequently, 5-alkoxy-3,4-dihalo-2(5H)-furanone 1 is attacked by dimethyl sulfide to form the intermediate A. Next, after a nucleophilic attack to intermediate A by halides in the reaction system, the target product 3 is obtained.
In order to demonstrate the feasibility of synthetic applications of this transformation, a gram-scale experiment was carried out, and the reaction was performed with 5 mmol dosage. As shown in Scheme 5, when using 1.35 g 1a to react with DMSO under standard conditions, the reaction can still be efficiently carried out, giving 1.06 g of the target compound 3a with an excellent yield (89%).
Therefore, the experimental results show that the novel method of this interesting halide 1 and DMSO 2 under the simple and mild reaction conditions without the participation of transition metal is indeed successful, which is very important in the actual production for the drug development from potential bioactive compounds.

3. Materials and Methods

3.1. General Information

The spectra of 1H and 13C NMR were collected using an AVANCE NEO-600 in CDCl3, using tetramethylsilane (TMS) as an internal standard. High-resolution mass spectra (HR-MS) were obtained using a MAT 95XP mass spectrometer. Single-crystal X-ray analysis was obtained using Agilent Gemini E. Reactions were monitored using thin-layer chromatography (TLC) and visualized via UV light at 254 nm.
All reagents and solvents were purchased from the commercial sources and used without further purification.

3.2. Experimental Procedure for Intermediate Compounds 1

Different intermediate 5-alkoxy(aryloxy)-3,4-dihalo-2(5H)-furanones 1 were synthesized according to the procedure in the literature [43]. As shown in Scheme 6, after the slow addition of 1–2 drops of concentrated H2SO4 into the mixture of mucobromic acid or mucochloric acid (20 mmol) and the corresponding alcohol (30 mL) in a three-neck flask, the obtained mixture was heated to continue refluxing for 36–72 h.
Once the reaction was completed, the reaction mixture was quenched via the saturated solution of sodium chloride and extracted with ethyl acetate. Then, the organic layer was dried over anhydrous sodium sulfate solid. Finally, after filtration, the evaporation of the solvents under reduced pressure gave the crude product, which was further purified via column chromatography on silica gel to obtain intermediate 1.

3.3. Experimental Procedure for Compounds 3a3t

As shown in Scheme 7, 3,4-dihalo-2(5H)-furanone compound 1 (0.40 mmol) and DBDMH (0.60 mmol) were mixed in DMSO (2 mL), and the mixture was stirred at 80 °C for 5 h.
After the completion of the reaction, the reaction mixture was quenched with the saturated solution of sodium chloride (15 mL) and extracted with ethyl acetate (3 × 15 mL). Then, the organic layer was dried over anhydrous sodium sulfate solid. Finally, after filtration, the evaporation of the solvents under reduced pressure gave the crude product, which was further purified via column chromatography on silica gel to afford the desired product 3.

3.4. Structural Characterization Data of Compounds 3a3t

The structures of the serial compounds 3a3t were systematically characterized via NMR, HR-MS, etc., and the corresponding data are summarized in the following.
(1) 3-Bromo-5-methoxy-4-methylthiofuran-2(5H)-one (3a), yellow solid, m.p.: 84.7–85.9 °C (86.7–87.8 °C [32]), 89 mg, 93%; 1H NMR (600 MHz, CDCl3), δ, ppm: 2.60 (s, 3H, SCH3), 3.54 (s, 3H, OCH3), 5.90 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 54.8, 101.9, 105.6, 160.0, 164.9; ESI-HRMS, m/z: calcd for C6H8BrO3S [M + H]+: 238.9372, found: 238.9369.
(2) 3-Bromo-5-ethoxy-4-methylthiofuran-2(5H)-one (3b), yellowish oil, 85 mg, 84% (83% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 1.31 (t, J = 7.2 Hz, 3H, CH3), 2.61 (s, 3H, SCH3), 3.73–3.92 (m, 2H, OCH2), 5.94 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 15.0, 64.6, 101.3, 105.4, 160.3, 165.0; ESI-HRMS, m/z: calcd for C7H10BrO3S [M + H]+: 252.9529, found: 252.9524.
(3) 3-Bromo-4-methylthio-5-propoxyfuran-2(5H)-one (3c), yellowish oil, 92 mg, 86%; 1H NMR (600 MHz, CDCl3), δ, ppm: 0.98 (t, J = 6.0 Hz, 3H, CH3), 1.67–1.74 (m, 2H, CH2), 2.61 (s, 3H, SCH3), 3.61–3.81 (m, 2H, OCH2), 5.95 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 10.5, 13.2, 22.7, 70.4, 101.4, 105.4, 160.3, 165.0; ESI-HRMS, m/z: calcd for C8H12BrO3S [M + H]+: 266.9685, found: 266.9683.
(4) 3-Bromo-5-isopropoxy-4-methylthiofuran-2(5H)-one (3d), yellowish oil, 94 mg, 88% (79% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 1.31 (d, 3H, J = 6.0 Hz, CH3), 1.33 (d, 3H, J = 6.0 Hz, CH3), 2.60 (s, 3H, SCH3), 4.13–4.18 (m, 1H, OCH), 5.96 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 22.0, 23.2, 73.7, 100.5, 105.6, 160.3, 165.2.
(5) 3-Bromo-5-butoxy-4-methylthiofuran-2(5H)-one (3e), yellowish oil, 92 mg, 82%; 1H NMR (600 MHz, CDCl3), δ, ppm: 0.94 (t, J = 7.2 Hz, 3H, CH3), 1.38–1.44 (m, 2H, CH2), 1.62–1.67 (m, 2H, CH2), 2.60 (s, 3H, SCH3), 3.64–3.83 (m, 2H, OCH2), 5.93 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 13.7, 19.2, 31.4, 68.5, 101.4, 105.5, 160.2, 165.0; ESI-HRMS, m/z: calcd for C9H14BrO3S [M + H]+: 280.9842, found: 280.9837.
(6) 3-Bromo-4-methylthio-5-pentyloxyfuran-2(5H)-one (3f), yellowish oil, 93 mg, 79%; 1H NMR (600 MHz, CDCl3), δ, ppm: 0.92 (t, J = 6.0 Hz, 3H, CH3), 1.32–1.39 (m, 4H, 2CH2), 1.65–1.70 (m, 2H, CH2), 2.61 (s, 3H, SCH3), 3.64–3.84 (m, 2H, OCH2), 5.94 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 14.0, 22.3, 28.1, 29.0, 68.8, 101.4, 105.4, 160.3, 165.1; ESI-HRMS, m/z: calcd for C10H16BrO3S [M + H]+: 294.9998, found: 294.9998.
(7) 3-Bromo-5-heptyloxy-4-methylthiofuran-2(5H)-one (3g), yellowish oil, 84 mg, 76%; 1H NMR (600 MHz, CDCl3), δ, ppm: 0.89 (t, J = 7.2 Hz, 3H, CH3), 1.25–1.38 (m, 8H, 4CH2), 1.59–1.68 (m, 2H, CH2), 2.60 (s, 3H, SCH3), 3.63–3.82 (m, 2H, OCH2), 5.93 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 14.2, 22.6, 25.9, 28.9, 29.3, 31.7, 68.8, 101.4, 105.5, 160.3, 165.0; ESI-HRMS, m/z: calcd for C12H20BrO3S [M + H]+: 323.0311, found: 323.0305.
(8) 3-Bromo-5-cyclohexyloxy-4-methylthiofuran-2(5H)-one (3h), yellowish oil, 95 mg, 78% (78% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 1.19–1.59 (m, 6H, 3CH2), 1.74–2.01 (m, 4H, 2CH2), 2.60 (s, 3H, SCH3), 3.81–3.86 (m, 1H, OCH), 6.00 (s, H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.3, 23.9, 24.0, 25.3, 32.1, 33.2, 79.3, 100.4, 105.5, 160.5, 165.3.
(9) 5-Benzyloxy-3-bromo-4-methylthiofuran-2(5H)-one (3i), yellowish oil, 93 mg, 74% (75% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 2.49 (s, 3H, SCH3), 4.70–4.87 (m, 2H, OCH2), 5.97 (s, 1H, CH), 7.36–7.41 (m, 5H, ArH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 70.5, 100.0, 105.5, 128.8, 128.9, 134.9, 160.5, 165.0.
(10) 3-Bromo-5-menthoxy-4-methylthiofuran-2(5H)-one (3j), yellowish oil, 107 mg, 74% (71% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 0.85–0.91 (m, 10H, CH, 3CH3), 1.24–1.34 (m, 3H, CH, CH2), 1.57–1.73 (m, 3H, CH, CH2), 1.89–2.31 (m, 2H, CH2), 2.63 (s, 3H, SCH3), 3.99–4.03 (m, 1H, OCH), 5.88 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.6, 14.1, 18.8, 19.6, 26.6, 28.0, 37.1, 44.9, 47.7, 49.5, 88.0, 102.9, 106.0, 160.5, 165.2.
(11) 5-([1,1′-Biphenyl]-4-yloxy)-3-bromo-4-methylthiofuran-2(5H)-one (3k), yellowish oil, 69 mg, 46% (39% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 2.66 (s, 3H, SCH3), 6.41 (s, 1H, CH), 7.26 (d, J = 9.0 Hz, 2H, ArH), 7.38 (t, J = 7.2 Hz, 1H, ArH), 7.42–7.52 (m, 2H, ArH), 7.58 (d, J = 8.4 Hz, 2H, ArH), 7.61 (d, J = 8.4 Hz, 2H, ArH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.6, 99.5, 106, 117.2, 127.0, 127.4, 128.7, 128.9, 137.6, 140.1, 154.9, 159.8, 164.6.
(12) 3-Chloro-5-methoxy-4-methylthiofuran-2(5H)-one (3l), yellowish oil, 56 mg, 72% (60% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 2.61 (s, 3H, SCH3), 3.54 (s, 3H, OCH3), 5.89 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.1, 55.0, 100.9, 117.1, 155.4, 164.4.
(13) 3-Chloro-5-ethoxy-4-methylthiofuran-2(5H)-one (3m), yellowish oil, 52 mg, 63% (58% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 1.31 (t, J = 7.2 Hz, 3H, CH3), 2.60 (s, 3H, SCH3), 3.73–3.92 (m, 2H, OCH2), 5.93 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 15.0, 64.6, 101.3, 105.5, 160.2, 165.0.
(14) 3-Chloro-4-methylthio-5-propoxyfuran-2(5H)-one (3n), yellowish oil, 55 mg, 62%; 1H NMR (600 MHz, CDCl3), δ, ppm: 0.97 (t, J = 6.0 Hz, 3H, CH3), 1.66–1.73 (m, 2H, CH2), 2.61 (s, 3H, SCH3), 3.60–3.81 (m, 2H, OCH2), 5.92 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 10.5, 13.1, 22.7, 70.5, 100.4, 117.0, 155.7, 164.6; ESI-HRMS, m/z: calcd for C8H12ClO3S [M + H]+: 223.0190, found: 223.0187.
(15) 3-Chloro-5-isopropoxy-4-methylthiofuran-2(5H)-one (3o), yellowish oil, 50 mg, 56% (55% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 1.31 (d, 3H, J = 6.0 Hz, CH3), 1.33 (d, 3H, J = 6.0 Hz, CH3), 2.61 (s, 3H, SCH3), 4.12–4.17 (m, 1H, OCH), 5.94 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.1, 22.0, 23.2, 73.8, 99.5, 117.1, 155.8, 164.8.
(16) 5-Butoxy-3-chloro-4-methylthiofuran-2(5H)-one (3p), yellowish oil, 67 mg, 71%; 1H NMR (600 MHz, CDCl3), δ, ppm: 0.94 (t, J = 7.2 Hz, 3H, CH3), 1.37–1.45 (m, 2H, CH2), 1.62–1.68 (m, 2H, CH2), 2.61 (s, 3H, SCH3), 3.64–3.85 (m, 2H, OCH2), 5.91 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.1, 13.7, 19.1, 31.4, 68.7, 100.4, 117.0, 155.6, 164.6; ESI-HRMS, m/z: calcd for C9H14ClO3S [M + H]+: 237.0347, found: 237.0346.
(17) 3-Chloro-4-methylthio-5-pentyloxyfuran-2(5H)-one (3q), yellowish oil, 69 mg, 69%; 1H NMR (600 MHz, CDCl3), δ, ppm: 0.91 (t, J = 6.0 Hz, 3H, CH3), 1.33–1.38 (m, 4H, 2CH2), 1.64–1.69 (m, 2H, CH2), 2.61 (s, 3H, SCH3), 3.63–3.84 (m, 2H, OCH2), 5.91 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.1, 13.9, 22.3, 28.1, 29.0, 69.0, 100.4, 117.0, 155.6, 164.6; ESI-HRMS, m/z: calcd for C10H16ClO3S [M + H]+: 251.0503, found: 251.0502.
(18) 3-Chloro-5-cyclohexyloxy-4-methylthiofuran-2(5H)-one (3r), yellowish oil, 54 mg, 52% (51% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 1.22–1.50 (m, 6H, 3CH2), 1.75–2.01 (m, 4H, 2CH2), 2.61 (s, 3H, SCH3), 3.80–3.85 (m, 1H, OCH), 5.98 (s, 1H, CH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 23.9, 24.0, 25.3, 32.1, 33.2, 79.3, 99.4, 117.1, 155.9, 164.8.
(19) 5-Benzyloxy-3-chloro-4-methylthiofuran-2(5H)-one (3s), yellowish oil, 54 mg, 50% (52% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 2.52 (s, 3H, SCH3), 4.71–4.88 (m, 2H, OCH2), 5.96 (s, 1H, CH), 7.37–7.41 (m, 5H, ArH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.2, 70.6, 98.9, 117.1, 128.8, 128.9, 135.0, 155.8, 164.5.
(20) 3-Chloro-4-methylthio-5-phenoxyfuran-2(5H)-one (3t), yellowish oil, 50 mg, 49% (46% [32]); 1H NMR (600 MHz, CDCl3), δ, ppm: 2.63 (s, 3H, SCH3), 6.35 (s, 1H, CH), 7.15–7.18 (m, 3H, ArH), 7.35–7.39 (m, 2H, ArH); 13C NMR (150 MHz, CDCl3), δ, ppm: 13.4, 98.5, 116.9, 124.4, 130.0, 155.3, 155.6, 164.1.
The detailed 1H, 13C NMR and spectra for all compounds 3a3t are provided in the Supplementary Materials.

4. Conclusions

In conclusion, we have disclosed a methylthiolating reaction of 5-alkoxy (or 5-aryoxy)-substituted 3,4-dihalo-2(5H)-furanone with DMSO. DMSO cannot only be used as a reaction raw material, but also as a solvent for the reaction. In addition, this transformation without any transition metal catalysts only requires a lower temperature and a shorter time. In particular, the simple reaction system is easy to operate, giving a better yield, even for the gram-scale reaction. This successful investigation provides a valuable reference for the introduction of the methyl sulfur group in organic synthesis.
Notably, due to the marked acetal and lactone ring structure of the reaction substrate with different functional groups such as C-X (X = Cl or Br), C-O, C=O and C=C bonds, the simple and green method will be attractive for synthesizing potentially bioactive methyl sulfur compounds with multifunctional groups.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules28155635/s1, which contains the 1H and 13C NMR spectra for all compounds 3a3t.

Author Contributions

Conceptualization, K.Y., J.-Y.L. and Z.-Y.W.; methodology, Y.-J.Z. and Y.-G.F.; formal analysis, Z.-J.C.; data curation, H.-Q.L.; writing—original draft preparation, Y.-J.Z.; writing—review and editing, K.Y., J.-Y.L. and Z.-Y.W.; project administration, Z.-Y.W.; funding acquisition, Z.-Y.W. and K.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Guangdong Basic and Applied Basic Research Foundation (no. 2021A1515012342), National Natural Science Foundation of China (20772035) and Natural Science Foundation of Jiangxi Province (no. 20224BAB203010).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data supporting the findings of this study are available within the paper and within its Supplementary Materials published online.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Dong, B.L.; Lu, Y.R.; Zhang, N.; Song, W.H.; Lin, W.Y. Ratiometric imaging of cysteine level changes in endoplasmic reticulum during H2O2-induced redox imbalance. Anal. Chem. 2019, 91, 5513–5516. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Yu, Y.; Wu, S.-F.; Zhu, X.-B.; Yuan, Y.F.; Li, Z.; Ye, K.-Y. Electrochemical sulfoxidation of thiols and alkyl halides. J. Org. Chem. 2022, 87, 6942–6950. [Google Scholar] [CrossRef]
  3. Feng, M.H.; Tang, B.Q.; Liang, S.H.; Jiang, X.F. Sulfur containing scaffolds in drugs: Synthesis and application in medicinal chemistry. Curr. Top. Med. Chem. 2016, 16, 1200–1216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bie, F.S.; Liu, X.J.; Cao, H.; Shi, Y.J.; Zhou, T.L.; Szostak, M.; Liu, C.W. Pd-catalyzed double-decarbonylative aryl sulfide synthesis through aryl exchange between amides and thioesters. Org. Lett. 2021, 23, 8098–8103. [Google Scholar] [CrossRef] [PubMed]
  5. Stepien, M.; Gonka, E.; Zyla, M.; Sprutta, N. Heterocyclic nanographenes and other polycyclic heteroaromatic compounds: Synthetic routes, properties, and applications. Chem. Rev. 2017, 117, 3479–3716. [Google Scholar] [CrossRef]
  6. Chen, S.-H.; Luo, S.-H.; Xing, L.-J.; Jiang, K.; Huo, Y.-P.; Chen, Q.; Wang, Z.-Y. Rational design and facile synthesis of dual-state emission fluorophores: Expanding functionality for the sensitive detection of nitroaromatic compounds. Chem. Eur. J. 2022, 28, e202103478. [Google Scholar] [CrossRef]
  7. Mondal, S.; Di Tommaso, E.M.; Olofsson, B. Transition-metal-free difunctionalization of sulfur nucleophiles. Angew. Chem. Int. Ed. 2023, 62, e202216296. [Google Scholar] [CrossRef] [PubMed]
  8. Denes, F.; Schiesser, C.H.; Renaud, P. Thiols, thioethers, and related compountandds as sources of C-centred radicals. Chem. Soc. Rev. 2013, 42, 7900–7942. [Google Scholar] [CrossRef] [PubMed]
  9. Merad, J.; Matyasovsky, J.; Stopka, T.; Brutiu, B.R.; Pinto, A.; Drescher, M.; Maulide, N. Stable and easily available sulfide surrogates allow a stereoselective activation of alcohols. Chem. Sci. 2021, 12, 7770–7774. [Google Scholar] [CrossRef]
  10. Ahmad, B.S.; Yaqoob, B.M.; Ahmad, R.S.; Salman, J.; Ahmad, B.K.; Naveed, A.Q. I2-DMSO promoted deaminative coupling reactions of glycine esters: Access to 5-(methylthio)pyridazin-3(2H)-ones. Org. Lett. 2023, 25, 2382–2387. [Google Scholar]
  11. Dunbar, K.L.; Scharf, D.H.; Litomska, A.; Hertweck, C. Enzymatic carbon-sulfur bond formation in natural product biosynthesis. Chem. Rev. 2017, 117, 5521–5577. [Google Scholar] [CrossRef]
  12. Park, J.K.; Lee, S. Sulfoxide and sulfone synthesis via electrochemical oxidation of sulfides. J. Org. Chem. 2021, 86, 13790–13799. [Google Scholar] [CrossRef]
  13. Forchetta, M.; Sabuzi, F.; Stella, L.; Conte, V.; Galloni, P. KuQuinone as a highly stable and reusable organic photocatalyst in selective oxidation of thioethers to sulfoxides. J. Org. Chem. 2022, 87, 14016–14025. [Google Scholar] [CrossRef]
  14. Cao, Y.; Huang, Y.; Blakemore, P.R. Synthesis of thioalkynes by desilylative sonogashira cross-coupling of aryl iodides and 1-methylthio-2-(trimethylsilyl)ethyne. Eur. J. Org. Chem. 2022, 2022, e202200498. [Google Scholar] [CrossRef]
  15. Cao, W.; Chen, P.; Wang, L.; Wen, H.; Liu, Y.; Wang, W.S.; Tang, Y. A highly regio- and stereoselective syntheses of α-halo enamides, vinyl thioethers, and vinyl ethers with aqueous hydrogen halide in two-phase systems. Org. Lett. 2018, 20, 4507–4511. [Google Scholar] [CrossRef]
  16. Clarke, A.K.; Parkin, A.; Taylor, R.J.K.; Unsworth, W.P.; Rossi-Ashton, J.A. Photocatalytic deoxygenation of sulfoxides using visible light: Mechanistic investigations and synthetic applications. ACS Catal. 2020, 10, 5814–5820. [Google Scholar] [CrossRef]
  17. Sakai, N.; Shimada, R.; Ogiwara, Y. Indium-catalyzed deoxygenation of sulfoxides with hydrosilanes. Asian J. Org. Chem. 2021, 10, 845–850. [Google Scholar] [CrossRef]
  18. Antoniak, D.; Paluba, B.; Basak, T.; Blaziak, K.; Barbasiewicz, M. Alkylation of nitroarenes via vicarious nucleophilic substitution-experimental and DFT mechanistic studies. Chem. Eur. J. 2022, 28, e202201153. [Google Scholar] [CrossRef] [PubMed]
  19. Wu, J.; Wang, Z.W.; Chen, X.-Y.; Wu, Y.C.; Wang, D.M.; Peng, Q.; Wang, P. Para-selective borylation of monosubstituted benzenes using a transient mediator. Sci. China Chem. 2020, 63, 336–340. [Google Scholar] [CrossRef]
  20. Gensch, T.; Klauck, F.J.R.; Glorius, F. Cobalt-catalyzed C-H thiolation through dehydrogenative cross-coupling. Angew. Chem. Int. Ed. 2016, 55, 11287–11291. [Google Scholar] [CrossRef] [PubMed]
  21. Ahmed, S.; Bhat, M.Y.; Hussain, F.; Ahmed, Q.N. BF3-Et2O Promoted heteronucleophilic addition reactions for the synthesis of unsymmetrical gem-diarylmethyl thioethers. Org. Lett. 2023, 25, 5017–5021. [Google Scholar] [CrossRef]
  22. Li, X.M.; Zhang, B.B.; Zhang, J.R.; Wang, X.; Zhang, D.K.; Du, Y.F.; Zhao, K. Synthesis of 3-methylthioindoles via intramolecular cyclization of 2-alkynylanilines mediated by DMSO/DMSO-d6 and SOCl2. Chin. J. Chem. 2021, 39, 1211–1224. [Google Scholar] [CrossRef]
  23. Lin, Z.G.; Huang, L.B.; Yuan, G.Q. Electrosynthesis of sulfonamides from DMSO and amines under mild conditions. Chem. Commun. 2021, 57, 3579–3582. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, T.L.; Li, H.; Nie, Z.W.; Su, M.-D.; Luo, W.-P.; Liu, Q.; Guo, C.-C. [3+1+1+1] Annulation to the Pyridine structure in quinoline molecules based on DMSO as a nonadjacent dual-methine synthon: Simple synthesis of 3-arylquinolines from arylaldehydes, arylamines, and DMSO. J. Org. Chem. 2022, 87, 2797–2808. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, H.L.; Wang, W.J.; Wang, B.D.; Tan, H.; Jiao, N.; Song, S. Electrophilic amidomethylation of arenes with DMSO/MeCN reagents. Org. Chem. Front. 2022, 9, 2430–2437. [Google Scholar] [CrossRef]
  26. Liu, H.; He, G.-C.; Zhao, C.-Y.; Zhang, X.-X.; Ji, D.-W.; Hu, Y.-C.; Chen, Q.-A. Redox-divergent construction of (dihydro)thiophenes with DMSO. Angew. Chem. Int. Ed. 2021, 133, 24486–24493. [Google Scholar] [CrossRef]
  27. Mukherjee, N.; Chatterjee, T. Iodine-catalyzed methylthiolative annulation of 2-alkynyl biaryls with DMSO: A metal-free approach to 9-sulfenylphen-anthrenes. J. Org. Chem. 2021, 86, 7881–7890. [Google Scholar] [CrossRef]
  28. Xia, Z.-H.; Gao, Z.-H.; Dai, L.; Ye, S. Visible-light-promoted oxo-difluoroalkylation of alkenes with DMSO as the oxidant. J. Org. Chem. 2019, 84, 7388–7394. [Google Scholar] [CrossRef]
  29. Kornfeind, J.; Iyer, P.S.; Keller, T.M.; Fleming, F.F. Oxidative DMSO cyclization cascade to bicyclic hydroxyketonitriles. J. Org. Chem. 2022, 87, 6097–6104. [Google Scholar] [CrossRef]
  30. Luo, F.; Pan, C.D.; Li, L.P.; Chen, F.; Cheng, J. Copper-mediated methylthiolation of aryl halides with DMSO. Chem. Commun. 2011, 47, 5304–5306. [Google Scholar] [CrossRef]
  31. Ghosh, K.; Ranjit, S.; Mal, D. A convenient method for the synthesis of aryl methyl sulfides via Cu(I)-mediated methylthiolation of haloarenes with DMSO. Tetrahedron Lett. 2015, 56, 5199–5202. [Google Scholar] [CrossRef]
  32. Cao, L.; Luo, S.-H.; Wu, H.-Q.; Chen, L.-Q.; Jiang, K.; Hao, Z.-F.; Wang, Z.-Y. Copper(I)-catalyzed alkyl- and arylsulfenylation of 3,4-dihalo-2(5H)-furanones (X=Br, Cl) with sulfoxides under mild conditions. Adv. Synth. Catal. 2017, 359, 2961–2971. [Google Scholar] [CrossRef]
  33. Kaiser, D.; Klose, I.; Oost, R.; Neuhaus, J.; Maulide, N. Bond-forming and-breaking reactions at sulfur (IV): Sulfoxides, sulfonium salts, sulfur ylides, and sulfinate salts. Chem. Rev. 2019, 119, 8701–8780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Saito, F. Sulfoxide reagent for one-pot, three-component syntheses of sulfoxides and sulfinamides. Angew. Chem. Int. Ed. 2022, 61, e202213872. [Google Scholar] [CrossRef]
  35. Wang, N.Z.; Saidhareddy, P.; Jiang, X.F. Construction of sulfur-containing moieties in the total synthesis of natural products. Nat. Prod. Rep. 2020, 37, 246–275. [Google Scholar] [CrossRef]
  36. Fu, Z.-H.; Tian, H.-D.; Ni, S.-F.; Wright, J.S.; Li, M.; Wen, L.-R.; Zhang, L.-B. Scalable selective electrochemical oxidation of sulfides to sulfoxides. Green Chem. 2022, 24, 4772–4777. [Google Scholar] [CrossRef]
  37. Wang, B.-W.; Jiang, K.; Li, J.-X.; Luo, S.-H.; Wang, Z.-Y.; Jiang, H.-F. 1,1-diphenyl-vinylsulfide as a functional AIEgen derived from the aggregation-caused-quenching molecule 1,1-diphenylethene through simple thioetherification. Angew. Chem. Int. Ed. 2020, 59, 2338–2343. [Google Scholar] [CrossRef]
  38. Cao, L.; Luo, S.-H.; Jiang, K.; Hao, Z.-F.; Wang, B.-W.; Pang, C.-M.; Wang, Z.-Y. Disproportionate coupling reaction of sodium sulfinates mediated by BF3·OEt2: An approach to symmetrical/unsymmetrical thiosulfonates. Org. Lett. 2018, 20, 4754–4758. [Google Scholar] [CrossRef]
  39. Zhou, Y.-J.; Yang, K.; Fang, Y.-G.; Luo, S.-H.; Chen, Q.; Yu, S.-W.; Wang, Z.-Y. A NaHCO3 promoted three-component cyclization: Easy access to benzodisulfide heterocycles. Asian J. Org. Chem. 2022, 11, e202200170. [Google Scholar] [CrossRef]
  40. Shi, J.; Tang, X.-D.; Wu, Y.-C.; Fang, J.-F.; Cao, L.; Chen, X.-Y.; Wang, Z.-Y. A radical coupling reaction of DMSO with sodium arylsulfinates in air: Mild utilization of DMSO as C1 resource for the synthesis of arylsulfonyl dibromomethane. RSC Adv. 2016, 6, 25651–25655. [Google Scholar] [CrossRef]
  41. Cao, L.; Li, J.-X.; Wu, H.-Q.; Jiang, K.; Hao, Z.-F.; Luo, S.-H.; Wang, Z.-Y. Metal-free sulfonylation of 3,4-dihalo-2(5H)-furanones (X=Cl, Br) with sodium sulfinates under air atmosphere in aqueous media via a radical pathway. ACS Sustain. Chem. Eng. 2018, 6, 4147–4153. [Google Scholar] [CrossRef]
  42. Wu, H.-Q.; Yang, K.; Luo, S.-H.; Wu, X.-Y.; Wang, N.; Chen, S.-H.; Wang, Z.-Y. C4-Selective synthesis of vinyl thiocyanates and selenocyanates through 3,4-dihalo-2(5H)-furanones. Eur. J. Org. Chem. 2019, 2019, 4572–4580. [Google Scholar] [CrossRef]
  43. Wu, H.-Q.; Yang, K.; Chen, X.-Y.; Arulkumar, M.; Wang, N.; Chen, S.-H.; Wang, Z.-Y. A 3,4-dihalo-2(5H)-furanone initiated ring-opening reaction of DABCO in the absence of a metal catalyst and additive and its application in a one-pot two-step reaction. Green Chem. 2019, 21, 3782–3788. [Google Scholar] [CrossRef]
  44. Yu, S.L.; Hong, C.; Liu, Z.; Zhang, H. Cobalt-catalyzed vinylic C-H addition to formaldehyde: Synthesis of butenolides from acrylic acids and HCHO. Org. Lett. 2021, 23, 8359–8364. [Google Scholar] [CrossRef]
  45. Irie, T.; Asami, T.; Sawa, A.; Uno, Y.; Hanada, M.; Taniyama, C.; Funakoshi, Y.; Masai, H.; Sawa, M. Discovery of novel furanone derivatives as potent Cdc7 kinase inhibitors. Eur. J. Med. Chem. 2017, 130, 406–418. [Google Scholar] [CrossRef]
  46. Zeiler, M.J.; Connors, G.M.; Durling, G.M.; Oliver, A.G.; Marquez, L.; Melander, R.J.; Quave, C.L.; Melander, C. Synthesis, stereochemical confirmation, and derivatization of 12(S), 16ϵ-dihydroxycleroda-3,13-dien-15,16-olide, a clerodane diterpene that sensitizes methicillin-resistant staphylococcus aureus to β-lactam antibiotics. Angew. Chem. Int. Ed. 2022, 134, e202117458. [Google Scholar] [CrossRef]
  47. Byczek-Wyrostek, A.; Kitel, R.; Rumak, K.; Skonieczna, M.; Kasprzycka, A.; Walczak, K. Simple 2(5H)-furanone derivatives with selective cytotoxicity towards non-small cell lung cancer cell line A549–synthesis, structure-activity relationship and biological evaluation. Eur. J. Med. Chem. 2018, 150, 687–697. [Google Scholar] [CrossRef] [PubMed]
  48. Bailly, F.; Queffelec, C.; Mbemba, G.; Mouscadet, J.-F.; Pommery, N.; Pommery, J.; Henichart, J.-P.; Cotelle, P. Synthesis and biological activities of a series of 4, 5-diaryl-3-hydroxy-2 (5H)-furanones. Eur. J. Med. Chem. 2008, 43, 1222–1229. [Google Scholar] [CrossRef] [PubMed]
  49. Wei, M.-X.; Feng, L.; Li, X.-Q.; Zhou, X.-Z.; Shao, Z.-H. Synthesis of new chiral 2,5-disubstituted 1,3,4-thiadiazoles possessing γ-butenolide moiety and preliminary evaluation of in vitro anticancer activity. Eur. J. Med. Chem. 2009, 44, 3340–3344. [Google Scholar] [CrossRef] [PubMed]
  50. Yang, K.; Yang, J.-Q.; Luo, S.-H.; Mei, W.-J.; Lin, J.-Y.; Zhan, J.-Q.; Wang, Z.-Y. Synthesis of N-2(5H)-furanonyl sulfonyl hydrazone derivatives and their biological evaluation in vitro and in vivo activity against MCF-7 breast cancer cells. Bioorg. Chem. 2021, 107, 104518. [Google Scholar] [CrossRef] [PubMed]
  51. Zhou, Y.-J.; Fang, Y.-G.; Yang, K.; Yu, S.-W.; Chen, Z.-J.; Wang, B.-C.; Zhan, H.-Y.; Wang, Z.-Y. Multicomponent selective thioetherification of KSAc: Easy access to symmetrical/unsymmetrical 4-alkylthio-3-halo-2(5H)-furanones. Asian J. Org. Chem. 2023, 12, e202300038. [Google Scholar] [CrossRef]
  52. CCDC 2270564 (for 3a) Contains the Supplementary Crystallographic Data for This Paper. These Data Can Be Obtained Free of Charge from The Cambridge Crystallographic Data Centre. Available online: www.ccdc.cam.ac.uk/data_request/cif (accessed on 17 June 2023).
  53. Murai, K.; Matsushita, T.; Nakamura, A.; Fukushima, S.; Shimura, M.; Fujioka, H. Asymmetric bromolactonization catalyzed by a C3-symmetric chiral trisimidazoline. Angew. Chem. Int. Ed. 2010, 49, 9174–9177. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, L.; Zhai, L.L.; Chen, J.Y.; Gong, Y.L.; Wang, P.; Li, H.L.; She, X.G. Catalyst-free 1,2-dibromination of alkenes using 1,3-dibromo-5,5-dimethylhydantoin (DBDMH) as a bromine source. J. Org. Chem. 2022, 87, 3177–3183. [Google Scholar] [CrossRef] [PubMed]
  55. Zhao, J.-Y.; Liao, Q.; Zhang, J.-W.; Lin, G.-Q.; Wang, Z.T.; Gao, D.D.; Li, Q.-H.; Tian, P. Thiourea-promoted cascade dehalogenation-cyclization of cyclohexadienone-containing 1,6-enynes. Adv. Synth. Catal. 2023, 365, 482–489. [Google Scholar] [CrossRef]
  56. Khazaei, A.; Zolfigol, M.A.; Rostami, A. 1,3-dibromo-5,5-dimethylhydantoin [DBDMH] as an efficient and selective agent for the oxidation of thiols to disulfides in solution or under solvent-free conditions. Synthesis 2004, 2004, 2959–2961. [Google Scholar] [CrossRef]
  57. Xu, S.H.; Wu, P.; Zhang, W. 1,3-Dibromo-5,5-dimethylhydantoin (DBH) mediated one-pot syntheses of α-bromo/amino ketones from alkenes in water. Org. Biomol. Chem. 2016, 14, 11389–11395. [Google Scholar] [CrossRef] [PubMed]
  58. Hua, J.W.; Xu, J.Q.; Xu, J.; Zhou, B.C.; Zhang, D.; Yang, Z.; Fang, Z.; Guo, K. Oxidative thioesterification of alkenes mediated by 1,3-dibromo-5,5-dimethylhydantoin and DMSO for the synthesis of α-ketothioesters. Eur. J. Org. Chem. 2019, 2019, 4056–4060. [Google Scholar] [CrossRef]
  59. Fu, D.; Dong, J.; Du, H.G.; Xu, J.X. Methanesulfinylation of benzyl halides with dimethyl sulfoxide. J. Org. Chem. 2019, 85, 2752–2758. [Google Scholar] [CrossRef]
  60. Fu, D.; Dong, J.; Wang, J.Y.; Xu, J.X. Direct 1,2-oxosulfenylation of acetylenic sulfones with DMSO. Asian J. Org. Chem. 2021, 10, 1756–1764. [Google Scholar] [CrossRef]
Figure 1. Some natural products and drugs containing the structure of the methyl sulfur group.
Figure 1. Some natural products and drugs containing the structure of the methyl sulfur group.
Molecules 28 05635 g001
Scheme 1. The synthesis of thioethers using DMSO.
Scheme 1. The synthesis of thioethers using DMSO.
Molecules 28 05635 sch001
Scheme 2. Green synthesis from 3,4-dihalo-2(5H)-furanone and various sulfur-containing reagents [32,40,41,43].
Scheme 2. Green synthesis from 3,4-dihalo-2(5H)-furanone and various sulfur-containing reagents [32,40,41,43].
Molecules 28 05635 sch002
Figure 2. The HR-MS spectrum of compound 3c.
Figure 2. The HR-MS spectrum of compound 3c.
Molecules 28 05635 g002
Scheme 3. Control experiments.
Scheme 3. Control experiments.
Molecules 28 05635 sch003
Scheme 4. A plausible reaction pathway.
Scheme 4. A plausible reaction pathway.
Molecules 28 05635 sch004
Scheme 5. Gram-scale reaction.
Scheme 5. Gram-scale reaction.
Molecules 28 05635 sch005
Scheme 6. Synthesis route of intermediate 1.
Scheme 6. Synthesis route of intermediate 1.
Molecules 28 05635 sch006
Scheme 7. Synthesis route of target products 3a3t.
Scheme 7. Synthesis route of target products 3a3t.
Molecules 28 05635 sch007
Table 1. Optimization of reaction conditions [a].
Table 1. Optimization of reaction conditions [a].
Molecules 28 05635 i001
EntryActivator (Equiv.)Temp. (°C)Time (h)Yield of 3a (%) [b]
1NBS (1.0)80565
2NCS (1.0)80543
3DBDMH (1.0)80587
4DBDMH (0.5)80578
5DBDMH (1.5)80593
6DBDMH (2.0)80590
7DBDMH (1.5)60585
8DBDMH (1.5)100587
9DBDMH (1.5)80889
10DBDMH (1.5)80357
[a] Reaction conditions: 1a (0.4 mmol) and 2 (2 mL) were added and stirred for 5 h. [b] Isolated yield.
Table 2. Substrate scope of various 5-substituted-3,4-dihalo-2(5H)-furanone intermediate 1 [a,b] [32].
Table 2. Substrate scope of various 5-substituted-3,4-dihalo-2(5H)-furanone intermediate 1 [a,b] [32].
Molecules 28 05635 i002
Molecules 28 05635 i003
[a] Reaction conditions: 1 (0.4 mmol), DBDMH (0.6 mmol), 2 (2 mL) were added and stirred at 80 °C for 5 h. [b] Isolated yield.
Table 3. Crystal data and structure refinement for 3a.
Table 3. Crystal data and structure refinement for 3a.
Compound3a
Empirical formulaC6H7BrO3S
Formula weight239.09
Temperature (K)297
Wavelength (Å)0.71073
Crystal systemMonoclinic
Space groupP21
Unit cell dimensions (Å, °)a = 4.1083 (8), b = 7.6804 (19), c = 13.739 (3)
α = 90, β = 90.079 (17), γ = 90
Volume (Å3)433.22 (16)
Z2
Density (calculated) (g/cm3)1.833
Absorption coefficient (mm−1)4.941
F(000)236.0
Theta range for data collection3.981 to 22.995
Index ranges−4 ≤ h ≤ 4, −9 ≤ k ≤ 9, 0 ≤ l ≤ 16
Reflections collected1464
Independent reflections1464 [Rsigma = 0.1075]
Completeness to theta = 1.78°99.6%
Absorption correctionMultiscan
Max. and min. transmission1.000 and 0.101
Refinement methodLeast squares minimization
Data/restraints/parameters1464/1/103
Goodness-of-fit on F21.040
Final R indices [I > 2 sigma (I)]R1 = 0.0545, wR2 = 0.0955
R indices (all data)R1 = 0.0707, wR2 = 0.1048
Largest diff. peak and hole0.58 and −0.84 e.Å−3
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

Zhou, Y.-J.; Fang, Y.-G.; Yang, K.; Lin, J.-Y.; Li, H.-Q.; Chen, Z.-J.; Wang, Z.-Y. DBDMH-Promoted Methylthiolation in DMSO: A Metal-Free Protocol to Methyl Sulfur Compounds with Multifunctional Groups. Molecules 2023, 28, 5635. https://doi.org/10.3390/molecules28155635

AMA Style

Zhou Y-J, Fang Y-G, Yang K, Lin J-Y, Li H-Q, Chen Z-J, Wang Z-Y. DBDMH-Promoted Methylthiolation in DMSO: A Metal-Free Protocol to Methyl Sulfur Compounds with Multifunctional Groups. Molecules. 2023; 28(15):5635. https://doi.org/10.3390/molecules28155635

Chicago/Turabian Style

Zhou, Yong-Jun, Yong-Gan Fang, Kai Yang, Jian-Yun Lin, Huan-Qing Li, Zu-Jia Chen, and Zhao-Yang Wang. 2023. "DBDMH-Promoted Methylthiolation in DMSO: A Metal-Free Protocol to Methyl Sulfur Compounds with Multifunctional Groups" Molecules 28, no. 15: 5635. https://doi.org/10.3390/molecules28155635

Article Metrics

Back to TopTop