Next Article in Journal
New 1,2,3-Triazoles from (R)-Carvone: Synthesis, DFT Mechanistic Study and In Vitro Cytotoxic Evaluation
Previous Article in Journal
Alpha-Linolenic Acid Mediates Diverse Drought Responses in Maize (Zea mays L.) at Seedling and Flowering Stages
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Visible-Light-Induced, Graphene Oxide-Promoted C3-Chalcogenylation of Indoles Strategy under Transition-Metal-Free Conditions

1
Department of Chemistry and Chemical Engineering, Gannan Normal University, Ganzhou 341000, China
2
Key Laboratory of Prevention and Treatment of Cardiovascular and Cerebrovascular Diseases of Ministry of Education, School of Pharmaceutical Science of Gannan Medical University, Ganzhou 341000, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(3), 772; https://doi.org/10.3390/molecules27030772
Submission received: 14 December 2021 / Revised: 16 January 2022 / Accepted: 19 January 2022 / Published: 25 January 2022

Abstract

:
An efficient and general method for the synthesis of 3-sulfenylindoles and 3-selenylindoles employing visible-light irradiation with graphene oxide as a promoter at room temperature has been achieved. The reaction features are high yields, simple operation, metal-free and iodine-free conditions, an easy-to-handle oxidant, and gram-scalable synthesis. This simple protocol allows one to access a wide range of 3-arylthioindoles, 3-arylselenylindoles, and even 3-thiocyanatoindoles with good to excellent yields.

1. Introduction

Organosulfur and organoselenium compounds, which possess broad biological and pharmaceutical activities, have been widely employed as important scaffolds for medicinal chemistry (Figure 1) [1,2,3,4,5,6]. Among them, 3-sulfenylindoles and 3-selenylindoles represent important classes of sulfur and selenium-containing compounds having more greater therapeutic values in the treatment of cancer [7,8,9,10,11,12], HIV [13,14,15], tubulin assembly inhibition [16,17], and bacterial diseases [18,19,20,21,22]. In this regard, numerous methods for the straightforward construction of C-S and C-Se bonds have been developed for the synthesis of 3-sulfenylindoles and 3-selenylindoles. Among these various approaches, the most commonly used methods involved the direct sulfenylation and selenation of the indole moieties with various electrophilic sulfur and selenium reagents [23,24,25,26,27,28,29,30,31,32,33,34,35].
However, these strategies suffer from limitations, such as the need for stoichiometric or super stoichiometric amounts of catalysts, strong acidic or oxidizing reagents, harsh reaction conditions, the complex synthetic process of activated sulfur or selenium reagents, and limited substrate scopes [36,37,38,39,40,41]. Most importantly, these reactions employ arylsulfur or arylselenium reagents such as benzenesulfonyl chlorides [42,43,44,45], N-(thiophenyl)succinimide [46,47], S-phenyl benzenesulfonothioate [48,49], disulfides [50,51,52], benzene-sulfonhydrazide [53,54,55,56], N-phenylselenophthalimide [57], N-phenylselenosuccinimide [58], and diselenides [59,60,61,62,63], generation of stoichiometric byproducts still cannot be avoided under the conditions used. Therefore, the development of green and sustainable synthetic methods is highly desirable under mild conditions so as to avoid the use of external oxidants, transition metal catalysts, or harsh reaction conditions.
In recent years, graphene oxide (GO) [64,65,66,67], which is a readily available and inexpensive material, has historically functioned primarily as a precursor to reduced graphene oxide (rGO) or chemically modified graphene (CMG) materials [68,69], and has generated tremendous excitement due to its potential applications in plastic electronics, solar cells, optical materials, and biosensors [70,71]. In addition, photo-induced organic transformations have emerged as an attractive and suitable approach in recent years [72,73,74,75,76,77,78,79,80,81]. Although GO has been reported as a photocatalyst for hydrogen production from water under UV irradiation [82], the potential application of GO in synthetic photochemistry is still rare [83].
More recently, Wu et al developed a procedure of GO-mediated thiolation of indoles with thiols in water (Scheme 1) [84]. This methodology provided an atom economical and transition-metal and iodine free procedure for the direct synthesis of 3-sulfenylindoles. Subsequently, Kumar and Rathore reported a benign oxidant, photocatalyst and transition-metal-free visible light induced methodology for the construction of carbon-chalcogen (S, Se, Te) bond that enables the 3-chalcogenyl indole (Scheme 1) [29]. However, most of these methods suffer from some drawbacks such as low atom efficiency and limited substrate scope. Recently, we reported a new and efficient method for the C3-chalcogenylation of indolines employing visible-light irradiation and graphene oxide as a promoter at room temperature [85]. However, the reaction substrates are expensive and difficult to obtain for this synthesis method. In continuation of our work on indole chemistry [86,87,88,89,90,91,92] and GO-promoted C-H functionalisation of indoles [93], herein, we wish to report the combination of GO and blue LEDs, which works in synergy to efficiently promote the organo chalcogenylation (S and Se) of indoles in DCE under air atmosphere by using commercially available substrates. The highlight of this work is that GO not only acts as an oxidant, but as a photocatalyst as well.

2. Results and Discussion

The GO material used in this investigation was prepared by Hummers oxidation of graphite and subsequent exfoliation, as reported [94,95]. The obtained GO material was characterized by X-ray powder diffraction (XRD), transmission electron microscopy (TEM), visible Raman spectroscopy, and atomic force microscopy (AFM) [96] (see the Supplementary Materials).
To commence our investigation, the reaction of indole 4a with 4-methylbenzenethiol 5a was performed using 40 wt % GO as a promoter under irradiation with sunlight in open air (Table 1). The reaction proceeded and produced the desired coupling product 6a with a 28% yield (entry 1). Different light sources, such as CWF bulb (22 W, λmax = 365 ± 10 nm), green LED (1.0 W, λmax = 530 ± 10 nm), and blue LED (3.0 W, λmax = 425 ± 15 nm), were tested. Blue LED was more effective than other light sources, indicating the higher activity of GO in the presence of high-intensity blue light (entries 2–4). The reaction in the absence of a light source either failed to take place at room temperature (entry 5), or only a trace amount of the target product was formed (entry 17). The solvent also plays an important role in this transformation. DCE (1,2-dichloroethane) was more effective than the other tested solvents, such as THF, DMSO, toluene, DMF, and 1,4-dioxane (entries 6–11). Subsequent efforts were directed toward optimizing the GO loadings (entries 12–16). Whereas 50 wt % GO afforded 87% of the target product, decreasing the loading to 20 wt % GO was found to be sufficient to drive the cross-coupling reaction to quantitative conversion. No product was detected without GO. On the basis of our screening experiments, the best reaction condition is using 50 wt % GO in DCE and irradiation with blue LED in open air at 25 °C for 12 h, which afforded the desired product 6a in high yield (87%, entry 14).
With the best experimental conditions for the synthesis of 6a in hand, we first evaluated the efficiency of different substituted indoles 4 while keeping 4-methylbenzenethiol 5a constant. Under the optimized conditions, the desired products 6aa-6ma could be efficiently obtained in good to excellent yields (Table 2). Various substituted indoles 4, i.e., electrondonating (EDG, R = Me, OMe, OBn) and electron-withdrawing (EWG, R = Cl, I, CN, CO2CH3) groups successfully afforded the corresponding 3-sulfenylindoles and had no significant effect on the reactivity and the regioselectivity of reactions. In general, the EDG were better than the EWG. Furthermore, the introduction of various groups at the N-1, C-2, -3, -4, -5, -6, or -7 position of the indoles all proceeded with 5a under standard reaction conditions. Exceptions to this are 4-methylindole and methyl-4-indolecarboxylate, showing moderate sulfenylation yields (6ha and 6ia), probably due to the steric hindrance effect (entries 8 and 9). Interestingly, introducing a methyl group at the C-3-position of the indole afforded the 2-sulfenylindole product 6ma in 84% yield.
Next, a diverse array of arylthiols were employed as substrates to explore the scope of this reaction (Table 3). These substrates also showed high reactivity in this transformation. All reactions proceeded smoothly when the thiophenol was bearing, regardless of electron-donating groups (Me and OMe) or electron-withdrawing groups (Cl, Br, and NO2) on the phenyl ring; the 3-sulfenylindoles were obtained in good to excellent yields.
The success in using aryl thiols encouraged us to examine the reaction of indole 4a with various heterocyclic thiols including benzo[d]thiazole-2-thiol, 1-methyl-1H-imidazole-2-thiol, 1,3,4-thiadiazole-2-thiol, 5-methyl-1,3,4-thiadiazole-2-thiol, 1-methyl-1H-tetrazole-5-thiol, and the results are summarized in Scheme 2. In general, the desired products were formed in moderate to excellent yields under the standard reaction conditions.
Organothiocyanates are valuable synthetic intermediates which can be easily transformed into an array of organosulfur molecules [97,98,99]. Under the optimized conditions, we sustained our studies by treating indoles or 1H-pyrrolo[2,3-b]pyridine with KSCN under the standard reaction conditions, and the corresponding thiocyanated product 7af were obtained with 43–85% yields (Scheme 2). The results have shown that electronegativities of substituents play a major role in governing the reactivity of the substrates. Electron-donating substitutents show better results than electron-withdrawing substitutents in this transformation.
The developed protocol can also be applied for the preparation of 3-selenyl-indoles using various indole derivatives 4 and diphenyl diselenide 8. In general, the desired products 9 were formed in good to excellent yields in 8 h (Scheme 3), which was more efficient than the generation of 3-sulfenylindoles with regard to the yields and reaction times.
In order to demonstrate the effectiveness of this new strategy, a gram scale reaction was performed under the standard conditions. 10 mmol indole 4a and 12 mmol 4-methylbenzenethiol 5a were subjected to the reaction in the presence of GO (468 mg, 40 wt %) in 50 mL DCE at room temperature. After 12 h, the desired product 6a was obtained in 84% yield, which demonstrated the practical application of this protocol to prepare 3-sulfenylindoles on a gram-scale (Scheme 4). To our delight, when the amount of GO was reduced to 40 wt %, the yield was not affected to any observable extent.
To gain some insight into the mechanism of this reaction, some control experiments were conducted as shown in Scheme 5. Because the visible-light-induced, GO-promoted cross-coupling reaction was performed under open air, the role of O2 in this reaction was explored. Initially, When the optimal reaction was performed under an oxygen atmosphere instead of open air, there was no effect on the yield, but a faster conversion of the starting material to the reaction product was observed, indicating that O2 could be involved in the reaction pathway. Similarly, when the reaction was carried out under an argon atmosphere, no major effect was observed, indicating that the reaction follows a different route in an argon environment.
Then, radical trapping experiments were conducted by adding butylated hydroxytoluene (BHT) or 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) into the standard conditions of 4a and 5a. Experimental results show that these reaction were completely inhibited, indicating the involvement of radical species in the transformation.
On the basis of our control experiments and several other reports from the literature [29,85,100,101,102], we proposed two plausible mechanisms for this reaction in argon and in oxygen environments as shown in Scheme 6. Graphene oxide might act as a radical initiator [29]. Under an argon atmosphere (path A), promoted by the functional groups on the surface of GO, 5-methylbenzenethiol transformed into phenylthiophenol radical 10. Next, the thioyl radical 10 interacted with 4a to produce the radical intermediate 11. After that, 11 was oxidized to the intermediate 12. Finally, deprotonation of intermediate 12 led to the formation of product 6a. GO probably plays a crucial role during the process of oxidation and deprotonation.
In 2012, Loh et al suggested that the edge sites with unpaired electrons in GO constitute the active catalytic sites and afford enhanced kinetics for the trapping and activation of molecular oxygen by a sequence of electron transport and reduction to superoxide radical [103,104]. Thus, in the case of an oxygen atmosphere (path B), the anion radical of O2 (O2•−), which is produced through a SET from unpaired electrons in GO, would abstract a proton from 12, which would generate the desired product 6a and perhydroxyl radical (HO2). The transfer of H from 5a to HO2 would generate 10 and H2O2.

3. Materials and Methods

3.1. General Information

Unless otherwise specified, commercial reagents and solvents were used without further purification. Commercially available chemicals were purchased from Leyan (Shanghai, China) and used without any further purification. 1H and 13C NMR spectra were recorded on a Bruker spectrometer at 400 and 100 MHz, respectively. The chemical shifts were given in parts per million relative to CDCl3 (7.26 ppm for 1H) and CDCl3 (77.0 ppm for 13C. Peak multiplicities were reported as follows: s, singlet; d, doublet; t, triplet; m, multiplet; br. s, broad singlet and J, coupling constant (Hz). Mass spectra were recorded with Bruker Dalton Esquire 3000 plus LC-MS apparatus. Elemental analyses are expressed as percentage values. HRFABMS spectra were recorded on a FTMS apparatus. Silica gel (300–400 mesh) was used for flash column chromatography, eluting (unless otherwise stated) with an ethyl acetate/petroleum ether (PE) (60–90 °C) mixture.

3.2. General Procedure of the Products 6

In a 10 mL Schlenk tube, indole (0.3 mmol), GO (17.6 mg), and thiol (0.36 mmol) were stirred in DCE (1 mL) for 12 h at room temperature under an air atmosphere irradiated by blue LEDs. The reaction mixture was concentrated under reduced pressure. The residue was purified by flash chromatography on silica gel (eluent: EtOAc/PE = 1:10) to yield the corresponding product 6.
3-(p-Tolylthio)-1H-indole (6aa). Yellow amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.37 (s, 1H), 7.63 (d, J = 7.9 Hz, 1H), 7.46 (d, J = 2.6 Hz, 1H), 7.43 (d, J = 8.1 Hz, 1H), 7.26 (dt, J = 1.0, 8.1 Hz, 1H, Ar-H), 7.18 (t, J = 7.1 Hz, 1H), 7.05 (d, J = 8.3 Hz, 2H), 6.99 (d, J = 8.3 Hz, 2H), 2.26 (s, 3H). 13C NMR (101 MHz, CDCl3): δ 136.5, 135.5, 134.7, 130.4, 129.5, 129.1, 126.3, 123.0, 120.8, 119.7, 111.5, 103.6, 20.8. MS (ESI): 240 (M + H+, 100). These assignments matched with those previously published [27].
5-Iodo-3-(p-tolylthio)-1H-indole (6ba). Brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.51 (s, 1H, NH), 7.99 (d, J = 1.5 Hz, 1H, Ar-H), 7.53 (dd, J = 8.5, 1.5 Hz, 1H, Ar-H), 7.43 (d, J = 1.5 Hz, 1H, Ar-H), 7.21 (d, J = 8.5 Hz, 1H, Ar-H), 7.06–7.01 (m, 4H, Ar-H), 2.29 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 135.6, 135.1, 134.9, 131.7, 131.4, 131.3, 129.6, 128.4, 126.3, 113.6, 102.8, 84.6, 20.9. MS (ESI): 366 (M + H+, 100). These assignments matched with those previously published [105].
5-Methyl-3-(p-tolylthio)-1H-indole (6ca). Yellow amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.19 (s, 1H, NH), 7.55 (d, J = 0.7 Hz, 1H, Ar-H), 7.40 (d, J = 2.6 Hz, 1H, Ar-H), 7.34 (d, J = 8.2 Hz, 1H, Ar-H), 7.21–7.13 (m, 3H, Ar-H), 7.08 (d, J = 8.2 Hz, 2H, Ar-H), 2.52 (s, 3H, CH3), 2.36 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 135.9, 134.9, 134.7, 130.94, 130.4, 129.7, 129.5, 126.2, 124.7, 119.2, 111.5, 102.4, 21.6, 21.0. MS (ESI): 254 (M + H+, 100). These assignments matched with those previously published [105].
3-(p-Tolylthio)-1H-indole-5-carbonitrile (6da). Yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 12.20 (s, 1H, NH), 7.99 (d, J = 1.6 Hz, 1H, Ar-H), 7.81 (s, 1H, Ar-H), 7.66 (d, J = 8.4 Hz, 1H, Ar-H), 7.53 (d, J = 8.4 Hz, 1H, Ar-H), 7.04 (d, J = 8.0 Hz, 2H, Ar-H), 6.99 (d, J = 8.0 Hz, 2H, Ar-H), 2.20 (s, 3H, CH3). 13C NMR (101 MHz, DMSO-d6): δ 139.1, 135.3, 135.2, 134.9, 130.1, 129.0, 126.9, 125.4, 124.1, 120.7, 114.3, 102.8, 102.5, 20.9. MS (ESI): 265 (M + H+, 100). These assignments matched with those previously published [106].
6-Methoxy-3-(p-tolylthio)-1H-indole (6ea). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.33 (s, 1H, NH), 7.51 (d, J = 8.6 Hz, 1H, Ar-H), 7.34 (d, J = 2.2 Hz, 1H, Ar-H), 7.08 (d, J = 8.2 Hz, 2H, Ar-H), 7.02 (d, J = 8.2 Hz, 2H, Ar-H), 6.90 (d, J = 2.2 Hz, 1H, Ar-H), 6.86 (dd, J = 8.6, 2.2 Hz, 1H, Ar-H), 3.87 (s, 3H, OCH3), 2.29 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 157.2, 137.3, 135.6, 134.7, 129.5, 129.3, 126.3, 123.3, 120.3, 110.8, 103.4, 95.2, 55.7, 20.9. MS (ESI): 270 (M + H+, 100). These assignments matched with those previously published [106].
7-(Benzyloxy)-3-(p-tolylthio)-1H-indole (6ga). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.71 (s, 1H, NH), 7.52 (d, J = 7.1 Hz, 2H, Ar-H), 7.49–7.40 (m, 4H, Ar-H), 7.27 (d, J = 8.4 Hz, 1H, Ar-H), 7.09 (t, J = 7.8 Hz, 1H, Ar-H), 7.07 (d, J = 8.0 Hz, 2H, Ar-H), 7.01 (d, J = 8.0 Hz, 2H, Ar-H), 6.81 (d, J = 7.8 Hz, 1H, Ar-H), 5.24 (s, 2H, OCH2), 2.28 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 145.6, 136.9, 135.7, 134.6, 130.8, 130.1, 129.5, 128.7, 128.3, 128.0, 127.2, 126.3, 121.2, 112.6, 104.0, 103.7, 70.4, 20.9. MS (ESI): 346 (M + H+, 100). Anal calcd for C22H19NOS: C, 76.49; H, 5.54; N, 4.05; S, 9.28. Found C, 76.35; H, 5.47; N, 4.33; S, 8.95.
4-Methyl-3-(p-tolylthio)-1H-indole (6ha). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3) δ 8.41 (s, 1H, NH), 7.43 (d, J = 2.6 Hz, 1H, Ar-H), 7.29 (d, J = 8.1 Hz, 1H, Ar-H), 7.17 (t, J = 8.1 Hz, 1H, Ar-H), 7.06–6.98 (m, 4H, Ar-H), 6.92 (d, J = 7.1 Hz, 1H), 2.70 (s, 3H, CH3), 2.29 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 137.9, 137.0, 134.3, 132.2, 131.7, 129.6, 127.0, 125.5, 123.1, 122.4, 109.4, 102.9, 20.9, 18.7. MS (ESI): 254 (M + H+, 100). These assignments matched with those previously published [106].
Methyl 3-(p-tolylthio)-1H-indole-4-carboxylate (6ia). Brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 9.19 (s, 1H, NH), 7.51 (d, J = 2.6 Hz, 1H, Ar-H), 7.49 (d, J = 1.0 Hz, 1H, Ar-H), 7.38 (d, J = 2.6 Hz, 1H, Ar-H), 7.24 (t, J = 7.8 Hz, 1H, Ar-H), 6.98 (s, 4H, Ar-H), 3.68 (s, 3H), 2.30 (d, J = 37.7 Hz, 3H). 13C NMR (101 MHz, CDCl3): δ 169.6, 137.6, 136.4, 134.6, 133.4, 129.4, 126.2, 125.4, 125.3, 122.1, 122.0, 115.1, 103.1, 51.9, 20.8. MS (ESI): 298 (M + H+, 100). Anal calcd for C17H15NO2S: C, 68.66; H, 5.08; N, 4.71; S, 10.78. Found C, 68.80; H, 5.26; N, 4.64; S, 10.57.
2-Methyl-3-(p-tolylthio)-1H-indole (6ja). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.26 (s, 1H, NH), 7.58 (d, J = 7.8 Hz, 1H, Ar-H), 7.36 (dt, J = 1.0, 7.8 Hz, 1H, Ar-H), 7.22 (dt, J = 1,0, 7.8 Hz, 1H, Ar-H), 7.15 (dt, J = 1.0, 7.8 Hz, 1H, Ar-H), 6.99 (s, 4H, Ar-H), 2.54 (s, 3H, CH3), 2.27 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 141.0, 135.7, 135.5, 134.4, 130.4, 129.5, 125.8, 122.1, 120.7, 119.0, 110.7, 99.9, 20.9, 12.2. MS (ESI): 254 (M + H+, 100). These assignments matched with those previously published [106].
1-Methyl-3-(p-tolylthio)-1H-indole (6ka). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 7.75 (d, J = 8.2 Hz, 1H, Ar-H), 7.45 (d, J = 8.2 Hz, 1H, Ar-H), 7.39 (dt, J = 1.0, 7.0 Hz, 1H, Ar-H), 7.37 (s, 1H, Ar-H), 7.28 (dt, J = 1.0, 7.0 Hz, 1H, Ar-H), 7.15 (d, J = 8.2 Hz, 2H, Ar-H), 7.07 (d, J = 8.2 Hz, 1H, Ar-H), 3.86 (s, 3H, NCH3), 2.35 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 137.6, 136.1, 134.9, 134.6, 123.0, 129.6, 126.3, 126.2, 122.6, 120.5, 119.8, 109.8, 101.3, 33.1, 21.0. MS (ESI): 254 (M + H+, 100). These assignments matched with those previously published [105].
2,5-Dimethyl-3-(p-tolylthio)-1H-indole (6la). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.03 (s, 1H, NH), 7.46 (s, 1H, Ar-H), 7.25 (d, J = 8.2 Hz, 1H, Ar-H), 7.10 (dd, J = 8.2, 1.2 Hz, 1H, Ar-H), 7.46 (s, 4H, Ar-H), 2.51 (s, 3H, CH3), 2.49 (s, 3H, CH3), 2.34 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 141.3, 136.0, 134.3, 133.8, 130.7, 130.1, 129.6, 125.7, 123.7, 118.7, 110.5, 99.0, 21.5, 20.9, 12.1. MS (ESI): 268 (M + H+, 100). These assignments matched with those previously published [106].
3-Methyl-2-(p-tolylthio)-1H-indole (6ma). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 7.96 (s, 1H, NH), 7.71 (d, J = 7.9 Hz, 1H, Ar-H), 7.33 (d, J = 3.6 Hz, 2H, Ar-H), 7.26 (m, 1H, Ar-H), 7.12 (d, J = 8.3 Hz, 2H, Ar-H), 7.09 (d, J = 8.3 Hz, 2H, Ar-H), 2.51 (s, 3H, CH3), 2.38 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 136.9, 135.9, 133.5, 130.0, 128.6, 127.1, 123.4, 122.3, 119.7, 119.5, 119.4, 111.0, 21.0, 9.6. MS (ESI): 254 (M + H+, 100). These assignments matched with those previously published [106].
3-((4-Chlorophenyl)thio)-1H-indole (6ab). Light yellow amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.47 (s, 1H, NH), 7.58 (d, J = 8.0 Hz, 1H, Ar-H), 7.49 (d, J = 2.6 Hz, 1H, Ar-H), 7.45 (d, J = 8.2 Hz, 1H, Ar-H), 7.29 (dt, J = 1.0, 8.0 Hz, 1H, Ar-H), 7.18 (dt, J = 1.0, 8.0 Hz, 1H, Ar-H), 7.12 (d, J = 8.7 Hz, 2H, Ar-H), 7.02 (d, J = 8.7 Hz, 2H, Ar-H). 13C NMR (101 MHz, CDCl3): δ 137.8, 136.5, 130.6, 130.5, 128.7, 128.6, 127.1, 123.1, 121.0, 119.4, 111.6, 102.4. MS (ESI): 260 (M + H+, 30), 262 (M + H+, 100). These assignments matched with those previously published [105].
3-((4-Bromophenyl)thio)-1H-indole (6ac). Brown amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.76 (s, 1H, NH), 7.80 (d, J = 2.7 Hz, 1H, Ar-H), 7.51 (d, J = 8.1 Hz, 1H, Ar-H), 7.41–7.35 (m, 3H, Ar-H), 7.20 (dt, J = 1.1, 8.1 Hz, 1H, Ar-H), 7.08 (dt, J = 1.1, 8.1 Hz, 1H, Ar-H), 6.96 (dt, J = 2.7, 8.6 Hz, 2H, Ar-H). 13C NMR (101 MHz, DMSO-d6): δ 139.5, 137.3, 133.1, 132.1, 128.9, 127.7, 122.7, 120.7, 118.6, 118.0, 112.9, 99.1. MS (ESI): 304 (M + H+, 100), 306 (M + H+, 100). These assignments matched with those previously published [105].
3-((4-Methoxyphenyl)thio)-1H-indole (6ad). Brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.38 (s, 1H, NH), 7.63 (d, J = 8.0 Hz, 1H, Ar-H), 7.46 (d, J = 2.6 Hz, 1H, Ar-H), 7.41 (d, J = 8.0 Hz, 1H, Ar-H), 7.25 (dt, J = 1.0, 8.0 Hz, 1H, Ar-H), 7.17 (dt, J = 1.0, 8.0 Hz, 1H, Ar-H), 7.13 (d, J = 8.9 Hz, 2H, Ar-H), 6.74 (d, J = 8.9 Hz, 2H, Ar-H), 3.73 (s, 3H, OCH3). 13C NMR (101 MHz, CDCl3): δ 157.8, 136.5, 123.0, 129.5, 129.0, 128.6, 122.9, 120.8, 119.7, 114.5, 111.5, 104.7, 55.3. MS (ESI): 256 (M + H+, 100). These assignments matched with those previously published [27].
3-((4-Ethylphenyl)thio)-1H-indole (6ae). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.34 (s, 1H, NH), 7.71 (d, J = 8.0 Hz, 1H, Ar-H), 7.45 (d, J = 2.3 Hz, 1H, Ar-H), 7.44 (d, J = 8.0 Hz, 1H, Ar-H), 7.32 (dt, J = 1.0, 8.0 Hz, 1H, Ar-H), 7.23 (t, J = 7.5 Hz, 1H, Ar-H), 7.13 (d, J = 8.2 Hz, 2H, Ar-H), 7.06 (d, J = 8.2 Hz, 2H, Ar-H), 2.61 (q, J = 7.6 Hz, 2H, CH2), 1.23 (t, J = 7.6 Hz, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 141.3, 136.5, 135.9, 130.7, 129.2, 128.5, 126.3, 123.1, 120.9, 119.7, 111.8, 103.2, 28.4, 15.7. MS (ESI): 254 (M + H+, 100). Anal calcd for C16H15NS: C, 75.85; H, 5.97; N, 5.53; S, 12.65. Found C, 75.59; H, 5.63; N, 5.71; S, 12.32.
3-((2,4-Dimethylphenyl)thio)-1H-indole (6af). Tawny amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.28 (s, 1H, NH), 7.71 (d, J = 8.0 Hz, 1H, Ar-H), 7.45 (d, J = 8.0 Hz, 1H, Ar-H), 7.41 (d, J = 2.6 Hz, 1H, Ar-H), 7.36 (t, J = 7.7 Hz, 1H, Ar-H), 7.27 (t, J = 7.7 Hz, 1H, Ar-H), 7.09 (s, 1H, Ar-H), 6.83 (d, J = 8.0 Hz, 1H, Ar-H), 6.78 (d, J = 8.0 Hz, 1H, Ar-H), 2.59 (s, 3H, CH3), 2.34 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 136.6, 134.7, 134.6, 134.4, 130.9, 130.5, 129.3, 127.1, 126.1, 123.0, 120.8, 119.7, 111.6, 103.0, 20.7, 19.9. MS (ESI): 254 (M + H+, 100). Anal calcd for C16H15NS: C, 75.85; H, 5.97; N, 5.53; S, 12.65. Found C, 76.04; H, 5.83; N, 5.66; S, 12.35.
3-((4-Nitrophenyl)thio)-1H-indole (6ag). Reddish brown amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.86 (s, 1H, NH), 8.02 (d, J = 9.0 Hz, 2H, Ar-H), 7.55 (t, J = 7.7 Hz, 1H, Ar-H), 7.54 (d, J = 2.6 Hz, 1H, Ar-H), 7.53 (d, J = 8.1 Hz, 1H, Ar-H), 7.34 (t, J = 8.1 Hz, 1H, Ar-H), 7.22 (t, J = 7.7 Hz, 1H, Ar-H), 7.15 (d, J = 9.0 Hz, 2H, Ar-H). 13C NMR (101 MHz, CDCl3): δ 150.0, 144.9, 136.7, 131.4, 128.5, 125.2, 123.9, 123.5, 121.4, 119.2, 112.1, 100.1. MS (ESI): 271 (M + H+, 100). These assignments matched with those previously published [107].
3-(Naphthalen-2-ylthio)-1H-indole (6ah). Reddish brown amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.74 (s, 1H, NH), 7.83 (d, J = 2.5 Hz, 1H, Ar-H), 7.77 (d, J = 8.6 Hz, 1H, Ar-H), 7.73 (d, J = 8.7 Hz, 1H, Ar-H), 7.61 (d, J = 7.5 Hz, 1H, Ar-H), 7.50 (d, J = 2.6 Hz, 1H, Ar-H), 7.48 (s, 1H, Ar-H), 7.40–7.33 (m, 3H, Ar-H), 7.20 (dd, J = 2.6 Hz, 1H, Ar-H), 7.17 (t, J = 2.6 Hz, 1H, Ar-H), 7.03 (t, J = 7.5 Hz, 1H, Ar-H). 13C NMR (101 MHz, DMSO-d6): δ 137.3, 137.2, 133.7, 133.0, 131.3, 129.1, 128.8, 128.1, 127.1, 127.0, 125.7, 124.9, 123.3, 122.6, 120.6, 118.8, 112.9, 99.8. MS (ESI): 276 (M + H+, 100). These assignments matched with those previously published [106].
5-Methoxy-3-((4-methoxyphenyl)thio)-1H-indole (6ai). Red amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.42 (s, 1H, NH), 7.40 (d, J = 2.6 Hz, 1H, Ar-H), 7.28 (d, J = 8.8 Hz, 1H, Ar-H), 7.17 (dt, J = 2.6, 8.8 Hz, 2H, Ar-H), 7.13 (d, J = 2.4 Hz, 1H, Ar-H), 6.95 (dd, J = 8.8, 2.4 Hz, 1H, Ar-H), 6.80 (dt, J = 2.6, 8.8 Hz, 2H, Ar-H), 3.84 (s, 3H, OCH3), 3.76 (s, 3H, OCH3). 13C NMR (101 MHz, CDCl3): δ 157.8, 155.0, 131.5, 131.0, 129.9, 129.8, 128.3, 114.6, 113.4, 112.5, 103.7, 104.0, 55.9, 55.4. MS (ESI): 286 (M + H+, 100). These assignments matched with those previously published [108].
3-(p-Tolylthio)-1H-pyrrolo[3,2-b]pyridine (6aj). Light yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.86 (s, 1H, NH), 8.36 (d, J = 4.0 Hz, 1H, Ar-H), 8.00 (d, J = 2.6 Hz, 1H, Ar-H), 7.86 (d, J = 8.1 Hz, 1H, Ar-H), 7.19 (dd, J = 8.1, 4.5 Hz, 1H, Ar-H), 7.00 (d, J = 7.2 Hz, 2H, Ar-H), 6.96 (d, J = 7.2 Hz, 2H, Ar-H), 2.19 (s, 3H, CH3). 13C NMR (101 MHz, DMSO-d6): δ 146.0, 143.8, 136.2, 135.9, 134.5, 129.8, 129.7, 126.5, 120.0, 117.7, 101.6, 20.9. MS (ESI): 241 (M + H+, 100). Anal calcd for C14H12N2S: C, 69.97; H, 5.03; N, 11.66; S, 13.34. Found C, 70.21; H, 5.37; N, 11.31; S, 13.15.
2-((1H-indol-3-yl)thio)benzo[d]thiazole (6ak). Brown amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 12.03 (s, 1H, NH), 8.04 (d, J = 2.8 Hz, 1H, Ar-H), 7.82 (dd, J = 2.8, 1.8 Hz, 1H, Ar-H), 7.80 (dd, J = 2.1, 1.0 Hz, 1H, Ar-H), 7.57 (d, J = 7.8 Hz, 1H, Ar-H), 7.56 (d, J = 7.8 Hz, 1H, Ar-H), 7.41 (dt, J = 1.2, 8.4 Hz, 1H, Ar-H), 7.30–7.23 (m, 2H, Ar-H), 7.15 (dt, J = 1.0, 7.1 Hz, 1H, Ar-H). 13C NMR (101 MHz, DMSO-d6): δ 173.8, 154.6, 137.2, 135.4, 134.4, 128.4, 126.6, 124.4, 123.1, 122.1, 121.6, 121.3, 118.5, 113.1, 97.7. MS (ESI): 283 (M + H+, 100). These assignments matched with those previously published [101].
3-((1-Methyl-1H-imidazol-2-yl)thio)-1H-indole (6al). Yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.51 (s, 1H, NH), 7.69 (s, 1H, Ar-H), 7.60 (d, J = 7.6 Hz, 1H, Ar-H), 7.39 (d, J = 7.6 Hz, 1H, Ar-H), 7.16 (s, 1H, Ar-H), 7.10 (t, J = 7.1 Hz, 1H, Ar-H), 7.04 (t, J = 7.1 Hz, 1H, Ar-H), 6.85 (s, 1H, Ar-H), 3.66 (s, 3H, CH3). 13C NMR (101 MHz, DMSO-d6): δ 140.0, 136.7, 131.0, 128.9, 128.5, 124.0, 122.4, 120.3, 119.1, 112.5, 100.6, 34.0. MS (ESI): 230 (M + H+, 100). Anal calcd for C12H11N3S: C, 62.86; H, 4.84; N, 18.33; S, 13.98. Found C, 63.10; H, 5.07; N, 18.05; S, 13.61.
2-((1H-indol-3-yl)thio)-1,3,4-thiadiazole (6am). Yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.98 (s, 1H, NH), 9.29 (s, 1H, Ar-H), 8.00 (d, J = 2.8 Hz, 1H, Ar-H), 7.56 (d, J = 8.0 Hz, 1H, Ar-H), 7.53 (d, J = 7.9 Hz, 1H, Ar-H), 7.25 (dt, J = 1.0, 8.0 Hz, 1H, Ar-H), 7.16 (d, J = 7.9 Hz, 1H, Ar-H). 13C NMR (101 MHz, DMSO-d6): δ 173.0, 154.3, 137.2, 133.6, 127.8, 123.2, 121.3, 118.4, 113.2, 98.6. MS (ESI): 234 (M + H+, 100). Anal calcd for C10H7N3S2: C, 51.48; H, 3.02; N, 18.01; S, 27.48. Found C, 51.83; H, 3.39; N, 17.85; S, 27.17.
2-((1H-indol-3-yl)thio)-5-methyl-1,3,4-thiadiazole (6an). Yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.95 (s, 1H, NH), 7.96 (d, J = 2.7 Hz, 1H, Ar-H), 7.55 (d, J = 2.7 Hz, 1H, Ar-H), 7.53 (d, J = 2.7 Hz, 1H, Ar-H), 7.24 (d, J = 7.5 Hz, 1H, Ar-H), 7.16 (d, J = 7.5 Hz, 1H, Ar-H), 2.50 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 172.4, 165.7, 137.1, 133.5, 128.0, 123.1, 121.3, 118.4, 113.1, 98.8, 15.6. MS (ESI): 248 (M + H+, 100). Anal calcd for C11H9N3S2: C, 53.42; H, 3.67; N, 16.99; S, 25.92. Found C, 53.76; H, 3.92; N, 16.84; S, 25.59.
3-((1-Methyl-1H-tetrazol-5-yl)thio)-1H-indole (6ao). Pink amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.85 (s, 1H, NH), 7.94 (d, J = 2.7 Hz, 1H, Ar-H), 7.54 (d, J = 8.7 Hz, 1H, Ar-H), 7.51 (d, J = 8.7 Hz, 1H, Ar-H), 7.22 (d, J = 7.2 Hz, 1H, Ar-H), 7.13 (d, J = 7.2 Hz, 1H, Ar-H), 4.03 (s, 3H, NCH3). 13C NMR (101 MHz, DMSO-d6): δ 153.8, 136.9, 133.5, 128.8, 122.9, 121.0, 118.6, 112.9, 94.5, 34.5. MS (ESI): 232 (M + H+, 100). Anal calcd for C10H9N5S: C, 51.93; H, 3.92; N, 30.28; S, 13.86. Found C, 52.20; H, 4.28; N, 29.91; S, 13.93.
3-Thiocyanato-1H-indole (7a). White amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.76 (s, 1H, NH), 7.83 (dd, J = 5.9, 3.1 Hz, 1H, Ar-H), 7.52 (d, J = 2.8 Hz, 1H, Ar-H), 7.45 (dt, J = 5.9, 3.1 Hz, 1H, Ar-H), 7.35 (t, J = 3.1 Hz, 1H, Ar-H), 7.33 (t, J = 3.1 Hz, 1H, Ar-H). 13C NMR (101 MHz, CDCl3): δ 136.0, 131.0, 127.7, 123.9, 121.9, 118.8, 112.1, 111.9, 92.3. MS (ESI): 175 (M + H+, 100). These assignments matched with those previously published [109].
5-Chloro-3-thiocyanato-1-H-indole (7b). Yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 12.20 (s, 1H, NH), 8.05 (s, 1H, Ar-H), 7.66 (s, 1H, Ar-H), 7.55 (d, J = 7.7 Hz, 1H, Ar-H), 7.26 (d, J = 7.7 Hz, 1H, Ar-H). 13C NMR (101 MHz, DMSO-d6): δ 135.3, 135.2, 129.1, 126.4, 123.5, 117.4, 115.0, 112.5, 89.9. MS (ESI): 209 (M + H+, 100), 211 (M + H+, 32). These assignments matched with those previously published [110].
5-Methyl-3-thiocyanato-1-H-indole (7c). White amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 12.00 (s, 1H, NH), 7.94 (d, J = 2.9 Hz, 1H, Ar-H), 7.35 (d, J = 8.2 Hz, 1H, Ar-H), 7.12 (t, J = 8.2 Hz, 1H, Ar-H), 6.93 (d, J = 7.1 Hz, 1H, Ar-H), 2.85 (s, 3H, CH3). 13C NMR (101 MHz, DMSO-d6): δ 137.2, 134.6, 129.9, 125.7, 123.4, 123.0, 114.3, 111.3, 89.8, 19.2. MS (ESI): 189 (M + H+, 100). These assignments matched with those previously published [111].
5-Methoxy-3-thiocyanato-1H-indole (7d). White amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.55 (s, 1H, NH), 7.67 (d, J = 8.7 Hz, 1H, Ar-H), 7.40 (d, J = 2.7 Hz, 1H, Ar-H), 6.97 (dd, J = 8.7, 2.1 Hz, 1H, Ar-H), 6.90 (d, J = 2.1 Hz, 1H, Ar-H), 3.86 (s, 3H, OCH3). 13C NMR (101 MHz, CDCl3): δ 157.7, 136.9, 129.8, 121.8, 119.5, 112.1, 111.9, 95.2, 92.3, 55.7. MS (ESI): 205 (M + H+, 100). These assignments matched with those previously published [109].
3-Thiocyanato-1-H-indole-5-carbonitrile (7e). Yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 12.52 (s, 1H, NH), 8.23 (d, J = 2.6 Hz, 1H, Ar-H), 8.21 (s, 1H, Ar-H), 7.71 (d, J = 8.5 Hz, 1H, Ar-H), 7.64 (dd, J = 8.6, 2.6 Hz, 1H, Ar-H). 13C NMR (101 MHz, DMSO-d6): δ 143.4, 141.1, 132.5, 130.9, 128.6, 125.1, 119.5, 117.2, 108.7, 96.8. MS (ESI): 200 (M + H+, 100). These assignments matched with those previously published [110].
5-Nitro-3-thiocyanato-1-H-indole (7f). Yellow amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 12.63 (s, 1H, NH), 8.51 (d, J = 1.8 Hz, 1H, Ar-H), 8.27 (s, 1H, Ar-H), 8.12 (dd, J = 9.0, 1.8 Hz, 1H, Ar-H), 7.70 (d, J = 9.0 Hz, 1H, Ar-H). 13C NMR (101 MHz, DMSO-d6): δ 142.6, 140.0, 137.5, 127.3, 118.6, 114.8, 114.1, 112.4, 93.6. MS (ESI): 220 (M + H+, 100). These assignments matched with those previously published [110].
4-Methyl-3-thiocyanato-1-H-indole (7g). White amorphous solid. 1H NMR (400 MHz, DMSO-d6): δ 11.91 (s, 1H, NH), 7.92 (d, J = 2.9 Hz, 1H, Ar-H), 7.46 (s, 1H, Ar-H), 7.43 (d, J = 8.3 Hz, 1H, Ar-H), 7.10 (dd, J = 8.3, 1.1 Hz, 1H, Ar-H), 2.45 (s, 3H, CH3). 13C NMR (101 MHz, DMSO-d6): δ 135.1, 133.5, 130.5, 128.2, 125.0, 117.7, 113.0, 112.8, 89.0, 21.6. MS (ESI): 189 (M + H+, 100). These assignments matched with those previously published [110].
3-Thiocyanato-1H-pyrrolo[2,3-b]pyridine (7h). White amorphous solid. 1H NMR (400 MHz, DMSO): δ 12.62 (s, 1H, NH), 8.40 (d, J = 4.5 Hz, 1H, Ar-H), 8.18 (s, 1H, Ar-H), 8.13 (d, J = 7.8 Hz, 1H, Ar-H), 7.31 (dd, J = 7.8, 4.7 Hz, 1H, Ar-H). 13C NMR (101 MHz, DMSO): δ 148.8, 145.0, 134.4, 127.0, 120.3, 117.8, 112.6, 89.5. MS (ESI): 176 (M + H+, 100). These assignments matched with those previously published [112].
3-(Phenylselanyl)-1H-indole (9a). Yellow amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.43 (s, 1H, NH), 7.69 (d, J = 7.9 Hz, 1H, Ar-H), 7.49 (d, J = 2.5 Hz, 1H, Ar-H), 7.46 (d, J = 8.2 Hz, 1H, Ar-H), 7.33–7.27 (m, 3H, Ar-H), 7.24–7.12 (m, 4H, Ar-H). 13C NMR (101 MHz, CDCl3): δ 136.4, 133.9, 131.3, 130.0, 129.0, 128.7, 125.6, 123.0, 120.9, 120.4, 111.4, 98.2. MS (ESI): 274 (M + H+, 100). These assignments matched with those previously published [30].
2-Methyl-3-(phenylselanyl)-1H-indole (9b). Yellow amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.20 (s, 1H, NH), 7.64 (d, J = 7.7 Hz, 1H, Ar-H), 7.36 (d, J = 7.9 Hz, 1H, Ar-H), 7.26 (dd, J = 3.5, 1.4 Hz, 1H, Ar-H), 7.23 (dd, J = 3.5, 1.4 Hz, 2H, Ar-H), 7.20 (d, J = 4.1 Hz, 1H, Ar-H), 7.19–7.12 (m, 3H, Ar-H), 2.56 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 141.0, 135.8, 134.0, 131.3, 129.0, 128.4, 125.5, 122.2, 120.7, 119.8, 110.6, 96.2, 13.2. MS (ESI): 288 (M + H+, 100). These assignments matched with those previously published [113].
5-Methyl-3-(phenylselanyl)-1H-indole (9c). Yellow amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.30 (s, 1H, NH), 7.51 (d, J = 0.5 Hz, 1H, Ar-H), 7.44 (d, J = 2.5 Hz, 1H, Ar-H), 7.35 (d, J = 8.3 Hz, 1H, Ar-H), 7.31–7.28 (m, 2H, Ar-H), 7.22–7.14 (m, 4H, Ar-H), 2.49 (s, 3H, CH3). 13C NMR (101 MHz, CDCl3): δ 134.7, 134.1, 131.5, 130.4, 130.3, 129.0, 128.6, 125.6, 124.7, 119.9, 111.1, 97.4, 21.5. MS (ESI): 288 (M + H+, 100). These assignments matched with those previously published [32].
5-Methoxy-3-(phenylselanyl)-1H-indole (9d). Yellow amorphous solid. 1H NMR (400 MHz, CDCl3): δ 8.42 (s, 1H, NH), 7.44 (d, J = 2.5 Hz, 1H, Ar-H), 7.33 (d, J = 8.8 Hz, 1H, Ar-H), 7.30 (dd, J = 8.2, 1.5 Hz, 1H, Ar-H), 7.29 (s, 1H, Ar-H), 7.21–7.13 (m, 4H, Ar-H), 6.97 (dd, J = 8.8, 2.5 Hz, 1H, Ar-H), 3.85 (s, 3H, OCH3). 13C NMR (101 MHz, CDCl3): δ 155.1, 134.0, 132.0, 131.4, 130.8, 129.1, 128.5, 125.6, 113.5, 112.4, 101.6, 97.5, 55.9. MS (ESI): 304 (M + H+, 100). These assignments matched with those previously published [113].

4. Conclusions

In summary, we have developed a practical GO-promoted and transition metal-free light induced methodology for the construction of a carbon-chalcogen (S and Se) bond that provides 3-chalcogenyl indoles in good to excellent yields under open air. The key features of this simple and robust protocol are: (1) metal-free and iodine-free conditions; (2) easy-to-handle oxidant; (3) open to the air; (4) atom-economic; (5) performed on a gram-scale; (6) regioselective; and (7) applicable to different sources of organochalcogenides with substituted indoles for this transformation. Moreover, very few methods report the combination of GO and light which works in synergy to efficiently promote the organic reactions [83].

Supplementary Materials

The following are available online, Figure S1: (A) SEM image of graphite. (B) SEM image of GO, Figure S2: TEM image of graphite, Figure S3: TEM image of GO, Figure S4: AFM image of GO, Figure S5: Raman image of GO I D /I G ratio = 0.87, Figure S6: XRD image of GO.

Author Contributions

Conceptualization, L.L. and Q.H.; methodology, L.L. and Q.H.; investigation, H.L. and H.H.; writing original draft preparation, X.P.; writing review and editing, L.L.; NMR research, X.P.; supervision, L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (grant number 21462002), Jiangxi Province Office of Education Support Program (grant numbers GJJ190755, GJJ190788), Practice and Innovation Training Program for College Students in Jiangxi Province (grant number 201910413013), Graduate Innovation Project of Gannan Medical University (grant number YC2021-X014), and Fundamental Research Funds for Gannan Medical University (grant number QD202023) for financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Material.

Conflicts of Interest

The authors declare that they have no conflict of interest.

Sample Availability

Samples of the compounds 4, 5, and 8 are available from the authors.

Electronic Supplementary Information (ESI) Available

Experimental procedures and characterization data, characterization of GO, and 1H and 13C NMR spectra of compounds 6, 7 and 9.

References

  1. Mellah, M.; Voituriez, A.; Schulz, E. Chiral sulfur ligands for asymmetric catalysis. Chem. Rev. 2007, 107, 5133–5209. [Google Scholar] [CrossRef] [PubMed]
  2. Feng, M.; Tang, B.; Liang, S.; Jiang, X. Sulfur containing scaffolds in drugs: Synthesis and application in medicinal chemistry. Curr. Top. Med. Chem. 2016, 16, 1200–1216. [Google Scholar] [CrossRef] [PubMed]
  3. Jarrett, J.T. The biosynthesis of thiol- and thioether-containing cofactors and secondary metabolites catalyzed by radical S-adenosylmethionine enzymes. J. Biol. Chem. 2015, 290, 3972–3979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ostrovidov, S.; Franck, P.; Joseph, D.; Martarello, L.; Kirsch, G.; Belleville, F.; Nabet, P.; Dousset, B. Screening of new antioxidant molecules using flow cytometry. J. Med. Chem. 2000, 43, 1762–1769. [Google Scholar] [CrossRef] [PubMed]
  5. Ninomiya, M.; Garud, D.R.; Koketsu, M. Biologically significant selenium-containing heterocycles. Coord. Chem. Rev. 2011, 255, 2968–2990. [Google Scholar] [CrossRef]
  6. Nogueira, C.W.; Rocha, J.B.T. Toxicology and pharmacology of selenium: Emphasis on synthetic organoselenium compounds. Arch. Toxicol. 2011, 85, 1313–1359. [Google Scholar] [CrossRef]
  7. Guan, Q.; Han, C.M.; Zuo, D.Y.; Zhai, M.A.; Li, Z.Q.; Zhang, Q.; Zhai, Y.P.; Jiang, X.W.; Bao, K.; Wu, Y.L.; et al. Synthesis and evaluation of benzimidazole carbamates bearing indole moieties for antiproliferative and antitubulin activities. Eur. J. Med. Chem. 2014, 87, 306–315. [Google Scholar] [CrossRef]
  8. La Regina, G.; Edler, M.C.; Brancale, A.; Kandil, S.; Coluccia, A.; Piscitelli, F.; Hamel, E.; De Martino, G.; Matesanz, R.; Díaz, J.F.; et al. Arylthioindole inhibitors of tubulin polymerization. 3. Biological evaluation, structure-activity relationships and molecular modeling studies. J. Med. Chem. 2007, 50, 2865–2874. [Google Scholar] [CrossRef]
  9. Avis, I.; Martinez, A.; Tauler, J.; Zudaire, E.; Mayburd, A.; Abu-Ghazaleh, R.; Ondrey, F.; Mulshine, J.L. Inhibitors of the arachidonic acid pathway and peroxisome proliferator-activated receptor ligands have superadditive effects on lung cancer growth inhibition. Cancer Res. 2005, 65, 4181–4190. [Google Scholar] [CrossRef] [Green Version]
  10. Pang, Y.; An, B.; Lou, L.; Zhang, J.; Yan, J.; Huang, L.; Li, X.; Yin, S. Design, synthesis, and biological evaluation of novel selenium-containing iso combretastatins and phenstatins as antitumor agents. J. Med. Chem. 2017, 60, 7300–7314. [Google Scholar] [CrossRef]
  11. Plano, D.; Karelia, D.N.; Pandey, M.K.; Spallholz, J.E.; Amin, S.; Sharma, A.K. Design, synthesis, and biological evaluation of novel selenium (Se-NSAID) molecules as anticancer agents. J. Med. Chem. 2016, 59, 1946–1959. [Google Scholar] [CrossRef] [PubMed]
  12. Nedel, F.; Campos, V.F.; Alves, D.; McBride, A.J.A.; Dellagostin, O.A.; Collares, T.; Savegnago, L.; Seixas, F.K. Substituted diaryl diselenides: Cytotoxic and apoptotic effect in human colon adenocarcinoma cells. Life Sci. 2012, 91, 345–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. La Regina, G.; Coluccia, A.; Brancale, A.; Piscitelli, F.; Gatti, V.; Maga, G.; Samuele, A.; Pannecouque, C.; Schols, D.; Balzarini, J.; et al. Indolylarylsulfones as HIV-1 non-nucleoside reverse transcriptase inhibitors: New cyclic substituents at indole-2-carboxamide. J. Med. Chem. 2011, 54, 1587–1598. [Google Scholar] [CrossRef] [PubMed]
  14. Sancineto, L.; Mariotti, A.; Bagnoli, L.; Marini, F.; Desantis, J.; Iraci, N.; Santi, C.; Pannecouque, C.; Tabarrini, O. Design and synthesis of diselenobisbenzamides (DISeBAs) as nucleocapsid protein 7 (NCp7) inhibitors with anti-HIV activity. J. Med. Chem. 2015, 58, 9601–9614. [Google Scholar] [CrossRef]
  15. Ragno, R.; Coluccia, A.; La Regina, G.; De Martino, G.; Piscitelli, F.; Lavecchia, A.; Novellino, E.; Bergamini, A.; Ciaprini, C.; Sinistro, A.; et al. Design, molecular modeling, synthesis, and anti-HIV-1 activity of new indolyl aryl sulfones. Novel derivatives of the indole-2-carboxamide. J. Med. Chem. 2006, 49, 3172–3184. [Google Scholar] [CrossRef]
  16. De Martino, G.; Edler, M.C.; La Regina, G.; Coluccia, A.; Barbera, M.C.; Barrow, D.; Nicholson, R.I.; Chiosis, G.; Brancale, A.; Hamel, E.; et al. New arylthioindoles:  Potent inhibitors of tubulin polymerization. 2. Structure-activity relationships and molecular modeling studies. J. Med. Chem. 2006, 49, 947–954. [Google Scholar] [CrossRef]
  17. Funk, C.D. Leukotriene modifiers as potential therapeutics for cardiovascular disease. Nat. Rev. Drug Discov. 2005, 4, 664–672. [Google Scholar] [CrossRef]
  18. Zhang, M.Z.; Chen, Q.; Yang, G.F. A review on recent developments of indole-containing antiviral agents. Eur. J. Med. Chem. 2015, 89, 421–441. [Google Scholar] [CrossRef]
  19. Nuth, M.; Guan, H.C.; Zhukovskaya, N.; Saw, Y.L.; Ricciardi, R.P. Design of potent poxvirus inhibitors of the heterodimeric processivity factor required for viral replication. J. Med. Chem. 2013, 56, 3235–3246. [Google Scholar] [CrossRef]
  20. Wen, Z.Y.; Xu, J.W.; Wang, Z.W.; Qi, H.; Xu, Q.L.; Bai, Z.S.; Zhang, Q.; Bao, K.; Wu, Y.L.; Zhang, W.G. 3-(3,4,5-Trimethoxyphenylselenyl)-1H-indoles and their selenoxides as combretastatin A-4 analogs: Microwave-assisted synthesis and biological evaluation. Eur. J. Med. Chem. 2015, 90, 184–194. [Google Scholar] [CrossRef]
  21. Sahu, P.K.; Umme, T.; Yu, J.; Nayak, A.; Kim, G.; Noh, M.; Lee, J.-Y.; Kim, D.-D.; Jeong, L.S. Selenoacyclovir and selenoganciclovir: Discovery of a new template for antiviral agents. J. Med. Chem. 2015, 58, 8734–8738. [Google Scholar] [CrossRef]
  22. Chen, H.X.; Olatunji, O.J.; Zhou, Y.F. Anti-oxidative, anti-secretory and anti-inflammatory activities of the extract from the root bark of Lycium chinense (Cortex Lycii) against gastric ulcer in mice. J. Nat. Med. 2016, 70, 610–619. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, F.L.; Tian, S.K. Iodine-catalyzed regioselective sulfenylation of indoles with sulfonyl hydrazides. Angew. Chem. Int. Ed. 2013, 52, 4929–4932. [Google Scholar] [CrossRef] [PubMed]
  24. Thurow, S.; Penteado, F.; Perin, G.; Alves, D.; Santi, C.; Monti, B.; Schiesser, C.H.; Barcellos, T.; Lenardão, E.J. Selenium dioxide-promoted selective synthesis of mono- and bis-sulfenylindoles. Org. Chem. Front. 2018, 5, 1983–1991. [Google Scholar] [CrossRef]
  25. Azeredo, J.B.; Godoi, M.; Martins, G.M.; Silveira, C.C.; Braga, A.L. A solvent- and metal-free synthesis of 3-chacogenyl-indoles employing DMSO/I2 as an eco-friendly catalytic oxidation system. J. Org. Chem. 2014, 79, 4125–4130. [Google Scholar] [CrossRef]
  26. Maeda, Y.; Koyabu, M.; Nishimura, T.; Uemura, S. Vanadium-catalyzed sulfenylation of indoles and 2-naphthols with thiols under molecular oxygen. J. Org. Chem. 2004, 69, 7688–7693. [Google Scholar] [CrossRef]
  27. Prasad, C.D.; Kumar, S.; Sattar, M.; Adhikary, A.; Kumar, S. Metal free sulfenylation and bis-sulfenylation of indoles: Persulfate mediated synthesis. Org. Biomol. Chem. 2013, 11, 8036–8040. [Google Scholar] [CrossRef]
  28. Rafique, J.; Saba, S.; Franco, M.S.; Bettanin, L.; Schneider, A.R.; Silva, L.T.; Braga, A.L. Direct, meta-free C(sp2)-H chalcogenation of indoles and imidazopyridines with dichalcogenides catalysed by KIO3. Chem. Eur. J. 2018, 24, 4173–4180. [Google Scholar] [CrossRef]
  29. Rathore, V.; Kumar, S. Visible-light-induced metal and reagent-free oxidative coupling of sp2 C–H bonds with organo-dichalcogenides: Synthesis of 3-organochalcogenyl indoles. Green Chem. 2019, 21, 2670–2676. [Google Scholar] [CrossRef]
  30. Cao, Y.; Liu, J.; Liu, F.M.; Jiang, L.Q.; Yi, W.B. Copper-catalyzed direct and odorless selenylation with a sodium selenite-based reagent. Org. Chem. Front. 2019, 6, 825–829. [Google Scholar] [CrossRef]
  31. Luo, D.P.; Wu, G.; Yang, H.; Liu, M.C.; Gao, W.X.; Huang, X.B.; Chen, J.X.; Wu, H.Y. Copper-catalyzed three-component reaction for regioselective aryl- and heteroarylselenation of indoles using selenium powder. J. Org. Chem. 2016, 81, 4485–4493. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, X.; Wang, C.G.; Jiang, H.; Sun, L.H. Convenient synthesis of selenyl-indoles via iodide ion-catalyzed electrochemical C-H selenation. Chem. Commun. 2018, 54, 8781–8784. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, Q.B.; Ban, Y.L.; Yuan, P.F.; Peng, S.J.; Fang, J.G.; Wu, L.Z.; Liu, Q. Visible-light-mediated aerobic selenation of (hetero)arenes with diselenides. Green Chem. 2017, 19, 5559–5563. [Google Scholar] [CrossRef]
  34. Vieira, B.M.; Thurow, S.; da Costa, M.; Casaril, A.M.; Domingues, M.; Schumacher, R.F.; Perin, G.; Alves, D.; Savegnago, L.; Lenard, E.J. Ultrasound-Assisted Synthesis and antioxidant activity of 3-selanyl-1 H-indole and 3-selanylimidazo[1,2-a]pyridine derivatives. Asian J. Org. Chem. 2017, 6, 1635–1646. [Google Scholar] [CrossRef]
  35. Luz, E.Q.; Seckler, D.; Araújo, J.S.; Angst, L.; Lima, D.B.; Rios, E.A.M.; Ribeiro, R.R.; Rampon, D.S. Fe(III)-catalyzed direct C3 chalcogenylation of indole: The effect of iodide ions. Tetrahedron 2019, 75, 1258–1266. [Google Scholar] [CrossRef]
  36. Silveira, C.C.; Mendes, S.R.; Wolf, L.; Martins, G.M. The use of anhydrous CeCl3 as a catalyst for the synthesis of 3-sulfenyl indoles. Tetrahedron Lett. 2010, 51, 2014–2016. [Google Scholar] [CrossRef]
  37. Wu, G.; Liu, Q.; Shen, Y.; Wu, W.; Wu, L. Regioselective thiocyanation of aromatic and heteroaromatic compounds using ammonium thiocyanate and oxone. Tetrahedron Lett. 2005, 46, 5831–5834. [Google Scholar] [CrossRef]
  38. Ranjit, S.; Lee, R.; Heryadi, D.; Shen, C.; Wu, J.; Zhang, P.; Huang, K.-W.; Liu, X. Copper-mediated C-H activation/C-S cross-coupling of heterocycles with thiols. J. Org. Chem. 2011, 76, 8999–9007. [Google Scholar] [CrossRef]
  39. Tudge, M.; Tamiya, M.; Savarin, C.; Humphrey, G.R. Development of a novel, highly efficient halide-catalyzed sulfenylation of indoles. Org. Lett. 2006, 8, 565–568. [Google Scholar] [CrossRef]
  40. Khazaei, A.; Zolfigol, M.A.; Mokhlesi, M.; Panah, F.D.; Sajjadifar, S. Simple and highly efficient catalytic thiocyanation of aromatic compounds in aqueous media. Helv. Chim. Acta 2012, 95, 106–114. [Google Scholar] [CrossRef]
  41. Silveira, C.C.; Mendes, S.R.; Wolf, L.; Martins, G.M.; von Mühlen, L. Efficient synthesis of 3-selanyl- and 3-sulfanylindoles employing trichloroisocyanuric acid and dichalcogenides. Tetrahedron 2012, 68, 10464–10469. [Google Scholar] [CrossRef]
  42. Rahaman, R.; Devi, N.; Bhagawati, J.R.; Barman, P. Microwave-assisted regioselective sulfenylation of indoles under solvent- and metal-free conditions. RSC Adv. 2016, 6, 18929–18935. [Google Scholar] [CrossRef]
  43. Chen, M.; Huang, Z.T.; Zheng, Q.Y. Visible light-induced 3-sulfenylation of N-methylindoles with arylsulfonyl chlorides. Chem. Commun. 2012, 48, 11686–11688. [Google Scholar] [CrossRef] [PubMed]
  44. Wu, Q.; Zhao, D.; Qin, X.; Lan, J.; You, J. Synthesis of di(hetero)aryl sulfides by directly using arylsulfonyl chlorides as a sulfur source. Chem. Commun. 2011, 47, 9188–9190. [Google Scholar] [CrossRef]
  45. Tocco, G.; Begala, M.; Esposito, F.; Caboni, P.; Cannas, V.; Tramontano, E. ZnO-mediated regioselective C-arylsulfonylation of indoles: A facile solvent-free synthesis of 2- and 3-sulfonylindoles and preliminary evaluation of their activity against drug-resistant mutant HIV-1 reverse transcriptases (RTs). Tetrahedron Lett. 2013, 54, 6237–6241. [Google Scholar] [CrossRef]
  46. Nalbandian, C.J.; Miller, E.M.; Toenjes, S.T.; Gustafson, J.L. A conjugate Lewis base-Brønsted acid catalyst for the sulfenylation of nitrogen containing heterocycles under mild conditions. Chem. Commun. 2017, 53, 1494–1497. [Google Scholar] [CrossRef]
  47. Hostier, T.; Ferey, V.; Ricci, G.; Pardo, D.G.; Cossy, J. TFA-promoted direct C-H sulfenylation at the C2 position of non-protected indoles. Chem. Commun. 2015, 51, 13898–13901. [Google Scholar] [CrossRef]
  48. Wang, F.X.; Zhou, S.D.; Wang, C.; Tian, S.K. N-Hydroxy sulfonamides as new sulfenylating agents for the functionalization of aromatic compounds. Org. Biomol. Chem. 2017, 15, 5284–5288. [Google Scholar] [CrossRef]
  49. Ravi, C.; Joshi, A.; Adimurthy, S. C3 sulfenylation of N-heteroarenes in water under catalyst-free conditions. Eur. J. Org. Chem. 2017, 2017, 3646–3651. [Google Scholar] [CrossRef]
  50. Sang, P.; Chen, Z.; Zou, J. K2CO3 promoted direct sulfenylation of indoles: A facile approach towards 3-sulfenylindoles. Green Chem. 2013, 15, 2096–2100. [Google Scholar] [CrossRef]
  51. Shi, Q.; Li, P.; Zhang, Y.; Wang, L. Visible light-induced tandem oxidative cyclization of 2-alkynylanilines with disulfides (diselenides) to 3-sulfenyl- and 3-selenylindoles under transition metal-free and photocatalyst-free conditions. Org. Chem. Front. 2017, 4, 1322–1330. [Google Scholar] [CrossRef]
  52. Yi, S.; Li, M.; Mo, W.; Hu, X.; Hu, B.; Sun, N.; Jin, L.; Shen, Z. Metal-free, iodine-catalyzed regioselective sulfenylation of indoles with thiols. Tetrahedron Lett. 2016, 57, 1912–1916. [Google Scholar] [CrossRef]
  53. Yang, Y.; Zhang, S.; Tang, L.; Hu, Y.; Zha, Z.; Wang, Z. Catalyst-free thiolation of indoles with sulfonyl hydrazides for the synthesis of 3-sulfenylindoles in water. Green Chem. 2016, 18, 2609–2613. [Google Scholar] [CrossRef]
  54. Qiu, J.K.; Hao, W.J.; Wang, D.C.; Wei, P.; Sun, J.; Jiang, B.; Tu, S.J. Selective sulfonylation and diazotization of indoles. Chem. Commun. 2014, 50, 14782–14785. [Google Scholar] [CrossRef] [PubMed]
  55. Li, X.; Xu, Y.; Wu, W.; Jiang, C.; Qi, C.; Jiang, H. Copper-catalyzed aerobic oxidative N-S bond functionalization for C-S Bond formation: Regio- and stereoselective synthesis of sulfones and thioethers. Chem.-Eur. J. 2014, 20, 7911–7915. [Google Scholar] [CrossRef]
  56. Nookaraju, U.; Begari, E.; Yetra, R.R.; Kumar, P. CeCl3⋅7H2O-NaI promoted regioselective sulfenylation of indoles with sulfonylhydrazides. Chem. Select 2016, 1, 81–85. [Google Scholar] [CrossRef]
  57. Nicolaou, K.C.; Claremon, D.A.; Barnette, W.E.; Seitz, S.P. N-Phenylselenophthalimide (N-PSP) and N-phenylselenosuccinimide (N-PSS). Two versatile carriers of the phenylseleno group. Oxyselenation of olefins and a selenium-based macrolide synthesis. J. Am. Chem. Soc. 1979, 101, 3704–3706. [Google Scholar] [CrossRef]
  58. Zhao, X.; Yu, Z.; Xu, T.; Wu, P. Novel Brønsted acid catalyzed three-component alkylations of indoles with N-phenylselenophthalimide and styrenes. Org. Lett. 2007, 9, 5263–5266. [Google Scholar] [CrossRef]
  59. Mohan, B.; Yoon, C.; Jang, S.; Park, K.H. Copper nanoparticles catalyzed Se(Te)-Se(Te) bond activation: A straightforward route towards unsymmetrical organochalcogenides from boronic acids. ChemCatChem 2015, 7, 405–412. [Google Scholar] [CrossRef]
  60. Saba, S.; Rafique, J.; Braga, A.L. Synthesis of unsymmetrical diorganyl chalcogenides under greener conditions: Use of an iodine/DMSO system, solvent- and metal-free approach. Adv. Synth. Catal. 2015, 357, 1446–1452. [Google Scholar] [CrossRef]
  61. Chatterjee, T.; Ranu, B.C. Solvent-controlled halo-selective selenylation of aryl halides catalyzed by Cu(II) supported on Al2O3. A general protocol for the synthesis of unsymmetrical organo mono- and bis-selenides. J. Org. Chem. 2013, 78, 7145–7153. [Google Scholar] [CrossRef] [PubMed]
  62. Becht, J.-M.; Le Drian, C. Formation of carbon-sulfur and carbon-selenium bonds by palladium-catalyzed decarboxylative cross-couplings of hindered 2,6-dialkoxybenzoic acids. J. Org. Chem. 2011, 76, 6327–6330. [Google Scholar] [CrossRef] [PubMed]
  63. Reddy, V.P.; Kumar, A.V.; Swapna, K.; Rao, K.R. Copper oxide nanoparticle-catalyzed coupling of diaryl diselenide with aryl halides under ligand-free conditions. Org. Lett. 2009, 11, 951–953. [Google Scholar] [CrossRef] [PubMed]
  64. Peng, X.J.; Xu, X.Y.; Liu, Q.; Liu, L.X. Graphene oxide and its derivatives: Their synthesis and use in organic synthesis. Curr. Org. Chem. 2019, 23, 188–204. [Google Scholar] [CrossRef]
  65. Chua, C.K.; Pumera, M. Carbocatalysis: The state of “metal-free” catalysis. Chem. Eur. J. 2015, 21, 12550–12562. [Google Scholar] [CrossRef] [PubMed]
  66. Navalon, S.; Dhakshinamoorthy, A.; Alvaro, M.; Garcia, H. Carbocatalysis by graphene-based materials. Chem. Rev. 2014, 114, 6179–6212. [Google Scholar] [CrossRef]
  67. Dreyer, D.R.; Todd, A.D.; Bielawski, C.W. Harnessing the chemistry of graphene oxide. Chem. Soc. Rev. 2014, 43, 5288–5301. [Google Scholar] [CrossRef] [PubMed]
  68. Su, C.-L.; Loh, K.-P. Carbocatalysts: Graphene oxide and its derivatives. Acc. Chem. Res. 2013, 46, 2275–2285. [Google Scholar] [CrossRef]
  69. Coraux, J.; Marty, L.; Bendiab, N.; Bouchiat, V. Functional hybrid systems based on large-area high-quality graphene. Acc. Chem. Res. 2013, 46, 2193–2201. [Google Scholar] [CrossRef]
  70. Dreyer, D.R.; Bielawski, C.W. Carbocatalysis: Heterogeneous carbons finding utility in synthetic chemistry. Chem. Sci. 2011, 2, 1233–1240. [Google Scholar] [CrossRef]
  71. Dreyer, D.R.; Park, S.; Bielawski, C.W.; Ruoff, R.S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228–240. [Google Scholar] [CrossRef] [PubMed]
  72. Qin, Y.; Zhu, L.; Luo, S. Organocatalysis in inert C-H bond functionalization. Chem. Rev. 2017, 117, 9433–9520. [Google Scholar] [CrossRef] [PubMed]
  73. Romero, N.A.; Nicewicz, D.A. Organic photoredox catalysis. Chem. Rev. 2016, 116, 10075–10166. [Google Scholar] [CrossRef] [PubMed]
  74. Skubi, K.L.; Blum, T.R.; Yoon, T.P. Dual catalysis strategies in photochemical synthesis. Chem. Rev. 2016, 116, 10035–10074. [Google Scholar] [CrossRef]
  75. Prier, C.K.; Rankic, D.A.; MacMillan, D.W.C. Visible light photoredox catalysis with transition metal complexes: Applications in organic Synthesis. Chem. Rev. 2013, 113, 5322–5363. [Google Scholar] [CrossRef] [Green Version]
  76. Hoffmann, N. Photochemical reactions as key steps in organic synthesis. Chem. Rev. 2008, 108, 1052–1103. [Google Scholar] [CrossRef]
  77. Colmenares, J.C.; Varma, R.S.; Nair, V. Selective photocatalysis of lignin-inspired chemicals by integrating hybrid nanocatalysis in microfluidic reactors. Chem. Soc. Rev. 2017, 46, 6675–6686. [Google Scholar] [CrossRef]
  78. Narayanam, J.M.R.; Stephenson, C.R.J. Visible light photoredox catalysis: Applications in organic synthesis. Chem. Soc. Rev. 2011, 40, 102–113. [Google Scholar] [CrossRef]
  79. Ravelli, D.; Dondi, D.; Fagnoni, M.; Albini, A. Photocatalysis. A multi-faceted concept for green chemistry. Chem. Soc. Rev. 2009, 38, 1999–2011. [Google Scholar] [CrossRef]
  80. Yoon, T.P.; Ischay, M.A.; Du, J. Visible light photocatalysis as a greener approach to photochemical synthesis. Nat. Chem. 2010, 2, 527–532. [Google Scholar] [CrossRef]
  81. Zeitler, K. Photoredox catalysis with visible light. Angew. Chem. Int. Ed. 2009, 48, 9785–9789. [Google Scholar] [CrossRef]
  82. Yeh, T.F.; Syu, J.M.; Cheng, C.; Chang, T.H.; Teng, H. Graphite oxide as a photocatalyst for hydrogen production from water. Adv. Funct. Mater. 2010, 20, 2255–2262. [Google Scholar] [CrossRef]
  83. Pan, Y.H.; Wang, S.; Kee, C.W.; Dubuisson, E.; Yang, Y.Y.; Loh, K.P.; Tan, C.H. Graphene oxide and Rose Bengal: Oxidative C–H functionalisation of tertiary amines using visible light. Green Chem. 2011, 13, 3341–3344. [Google Scholar] [CrossRef]
  84. Chen, M.; Luo, Y.; Zhang, C.; Guo, L.; Wang, Q.T.; Wu, Y. Graphene oxide mediated thiolation of indoles in water: A green and sustainable approach to synthesize 3-sulfenylindoles. Org. Chem. Front. 2019, 6, 116–120. [Google Scholar] [CrossRef]
  85. Liu, C.P.; Peng, X.J.; Hu, D.; Shi, F.; Huang, P.P.; Luo, J.J.; Liu, Q.; Liu, L.X. Direct C3 chalcogenylation of indolines using a graphene oxide-promoted and visible-light-induced synergistic effect. N. J. Chem. 2020, 44, 17245–17251. [Google Scholar] [CrossRef]
  86. Liao, H.W.; Peng, X.J.; Hu, D.; Xu, X.Y.; Huang, P.P.; Liu, Q.; Liu, L.X. CoCl2-promoted TEMPO oxidative homocoupling of indoles: Access to tryptanthrin derivatives. Org. Biomol. Chem. 2018, 16, 5699–5706. [Google Scholar] [CrossRef]
  87. Huang, P.P.; Peng, X.J.; Hu, D.; Liao, H.W.; Tang, S.B.; Liu, L.X. Regioselective synthesis of 2,3’-biindoles mediated by an NBS-induced homo-coupling of indoles. Org. Biomol. Chem. 2017, 15, 9622–9629. [Google Scholar] [CrossRef]
  88. Yin, B.; Huang, P.P.; Lu, Y.B.; Liu, L.X. TEMPO-catalyzed oxidative homocoupling route to 3,2’-biindolin-2-ones via an indolin-3-one intermediate. RSC Adv. 2017, 7, 606–610. [Google Scholar] [CrossRef] [Green Version]
  89. Deng, Z.F.; Peng, X.J.; Huang, P.P.; Jiang, L.L.; Ye, D.N.; Liu, L.X. A multifunctionalized strategy of indoles to C2-quaternary indolin-3-ones via a TEMPO/Pd-catalyzed cascade process. Org. Biomol. Chem. 2017, 15, 442–448. [Google Scholar] [CrossRef]
  90. Jiang, L.L.; Peng, X.J.; Huang, P.P.; Chen, Z.W.; Liu, L.X. TEMPO-catalyzed oxidative dimerization and cyanation of indoles for the synthesis of 2-(1H-indol-3-yl)-3-oxoindoline-2-carbonitriles. Tetrahedron 2017, 73, 1389–1396. [Google Scholar] [CrossRef]
  91. Lin, F.; Chen, Y.; Wang, B.S.; Qin, W.B.; Liu, L.X. Silver-catalyzed TEMPO oxidative homocoupling of indoles for the synthesis of 3,3’-biindolin-2-ones. RSC Adv. 2015, 5, 37018–37022. [Google Scholar] [CrossRef]
  92. Qin, W.B.; Zhu, J.Y.; Kong, Y.B.; Bao, Y.H.; Chen, Z.W.; Liu, L.X. Metal-free (Boc)2O-mediated C4-selective direct indolation of pyridines using TEMPO. Org. Biomol. Chem. 2014, 12, 4252–4259. [Google Scholar] [CrossRef] [PubMed]
  93. Peng, X.J.; Zeng, Y.; Liu, Q.; Liu, L.X.; Wang, H.S. Graphene oxide as a green carbon material for cross-coupling of indoles with ethers via oxidation and the Friedel-Crafts reaction. Org. Chem. Front. 2019, 6, 3615–3619. [Google Scholar] [CrossRef]
  94. Hummers, W.S.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  95. Primo, A.; Neatu, F.; Florea, M.; Parvulescu, V.; García, H. Graphenes in the absence of metals as carbocatalysts for selective acetylene hydrogenation and alkene hydrogenation. Nat. Commun. 2014, 5, 5291–5299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Su, C.L.; Tandiana, R.; Balapanuru, J.; Tang, W.; Pareek, K.; Nai, C.T.; Hayashi, T.; Loh, K.P. Tandem catalysis of amines using porous graphene oxide. J. Am. Chem. Soc. 2015, 137, 685–690. [Google Scholar] [CrossRef] [PubMed]
  97. Bayarmagnai, B.; Matheis, C.; Jouvin, K.; Goossen, L.J. Synthesis of difluoromethyl thioethers from difluoromethyl trimethylsilane and organothiocyanates generated in situ. Angew. Chem. Int. Ed. 2015, 54, 5753–5756. [Google Scholar] [CrossRef]
  98. Zhang, Z.H.; Liebeskind, L.S. Palladium-catalyzed, copper(I)-mediated coupling of boronic acids and benzylthiocyanate. A cyanide-free cyanation of boronic acids. Org. Lett. 2006, 8, 4331–4333. [Google Scholar] [CrossRef]
  99. Exner, B.; Bayarmagnai, B.; Jia, F.; Goossen, L.J. Iron-catalyzed decarboxylation of trifluoroacetate and its application to the synthesis of trifluoromethyl thioethers. Chem.-Eur. J. 2015, 21, 17220–17223. [Google Scholar] [CrossRef]
  100. Saba, S.; Rafique, J.; Franco, M.S.; Schneider, A.R.; Espíndola, L.; Silva, D.O.; Braga, A.L. Rose Bengal catalysed photo-induced selenylation of indoles, imidazoles and arenes: A metal free approach. Org. Biomol. Chem. 2018, 16, 880–885. [Google Scholar] [CrossRef]
  101. Saima, E.D.; Lavekar, A.G.; Sinha, A.K. Cooperative catalysis by bovine serum albumin-iodine towards cascade oxidative coupling-C(sp2)-H sulfenylation of indoles/hydroxyaryls with thiophenols on water. Org. Biomol. Chem. 2016, 14, 6111–6118. [Google Scholar] [CrossRef] [PubMed]
  102. Guo, W.; Tan, W.; Zhao, M.; Tao, K.; Zheng, L.Y.; Wu, Y.; Chen, D.; Fan, X.L. Photocatalytic direct C–S bond formation: Facile access to 3-sulfenylindoles via metal-free C-3 sulfenylation of indoles with thiophenols. RSC Adv. 2017, 7, 37739–37742. [Google Scholar] [CrossRef] [Green Version]
  103. Su, C.L.; Acik, M.; Takai, K.; Lu, J.; Hao, S.J.; Zheng, Y.; Wu, P.P.; Bao, Q.L.; Enoki, T.; Chabal, Y.J.; et al. Probing the catalytic activity of porous graphene oxide and the origin of this behaviour. Nat. Commun. 2012, 3, 1298–1306. [Google Scholar] [CrossRef] [PubMed]
  104. Lv, G.Q.; Wang, H.L.; Yang, Y.G.; Deng, T.S.; Chen, C.M.; Zhu, Y.L.; Hou, X.L. Graphene oxide: A convenient metal-free carbocatalyst for facilitating aerobic oxidation of 5-hydroxymethylfurfural into 2,5-diformylfuran. ACS Catal. 2015, 5, 5636–5646. [Google Scholar] [CrossRef]
  105. Kumaraswamy, G.; Rajua, R.; Narayanarao, V. Metal- and base-free syntheses of aryl/alkylthioindoles by the iodine-induced reductive coupling of aryl/alkyl sulfonyl chlorides with indoles. RSC Adv. 2015, 5, 22718–22723. [Google Scholar] [CrossRef]
  106. Yang, X.Q.; Bao, Y.H.; Dai, Z.H.; Zhou, Q.F.; Yang, F.L. Catalyst-free sulfenylation of indoles with sulfinic esters in ethanol. Green Chem. 2018, 20, 3727–3731. [Google Scholar] [CrossRef]
  107. Fang, X.L.; Tang, R.Y.; Zhong, P.; Li, J.H. Iron-catalyzed sulfenylation of indoles with disulfides promoted by a catalytic amount of iodine. Synthesis 2009, 24, 4183–4189. [Google Scholar] [CrossRef]
  108. Ackermann, L.; Dell’Acqua, M.; Fenner, S.; Vicente, R.; Sandmann, R. Metal-free direct arylations of indoles and pyrroles with diaryliodonium salts. Org. Lett. 2011, 13, 2358–2360. [Google Scholar] [CrossRef]
  109. Zhang, X.; Wang, C.G.; Jiang, H.; Sun, L.H. A low-cost electrochemical thio- and selenocyanation strategy for electron-rich arenes under catalyst- and oxidant-free conditions. RSC Adv. 2018, 8, 22042–22045. [Google Scholar] [CrossRef] [Green Version]
  110. Li, C.Q.; Long, P.L.; Wu, H.P.; Yin, H.Q.; Chen, F.X. N-Thiocyanato-dibenzenesulfonimide: A new electrophilic thiocyanating reagent with enhanced reactivity. Org. Biomol. Chem. 2019, 17, 7131–7134. [Google Scholar] [CrossRef] [PubMed]
  111. Jiang, H.F.; Yu, W.T.; Tang, X.D.; Li, J.X.; Wu, W.Q. Copper-catalyzed aerobic oxidative regioselective thiocyanation of aromatics and heteroaromatics. J. Org. Chem. 2017, 82, 9312–9320. [Google Scholar] [CrossRef]
  112. Wu, D.; Qiu, J.S.; Karmaker, P.G.; Yin, H.Q.; Chen, F.X. N-Thiocyanatosaccharin: A “sweet” electrophilic thiocyanation reagent and the synthetic applications. J. Org. Chem. 2018, 83, 1576–1583. [Google Scholar] [CrossRef]
  113. Yu, Y.Z.; Zhou, Y.; Song, Z.Q.; Liang, G. An efficient t-BuOK promoted C3-chalcogenylation of indoles with dichalcogenides. Org. Biomol. Chem. 2018, 16, 4958–4962. [Google Scholar] [CrossRef]
Figure 1. Selected examples of biologically active 3-selanyl- and 3-sulfanylindole compounds.
Figure 1. Selected examples of biologically active 3-selanyl- and 3-sulfanylindole compounds.
Molecules 27 00772 g001
Scheme 1. C3 chalcogenylation of indoles.
Scheme 1. C3 chalcogenylation of indoles.
Molecules 27 00772 sch001
Scheme 2. Heterocyclic thiols and potassium thiocyanate used as C3 sulfenylation of indoles.
Scheme 2. Heterocyclic thiols and potassium thiocyanate used as C3 sulfenylation of indoles.
Molecules 27 00772 sch002
Scheme 3. Synthesis of 3-selenylindoles. Reaction conditions: 4 (0.3 mmol), 8 (0.36 mmol), GO (50 wt %), and DCE (1 mL), for 8 h at rt under open air. Isolated yield.
Scheme 3. Synthesis of 3-selenylindoles. Reaction conditions: 4 (0.3 mmol), 8 (0.36 mmol), GO (50 wt %), and DCE (1 mL), for 8 h at rt under open air. Isolated yield.
Molecules 27 00772 sch003
Scheme 4. Scale-up reaction between 4a and 5a.
Scheme 4. Scale-up reaction between 4a and 5a.
Molecules 27 00772 sch004
Scheme 5. Control experiments.
Scheme 5. Control experiments.
Molecules 27 00772 sch005
Scheme 6. The proposed mechanism for the reaction.
Scheme 6. The proposed mechanism for the reaction.
Molecules 27 00772 sch006
Table 1. Optimization of the reaction conditions a.
Table 1. Optimization of the reaction conditions a.
Molecules 27 00772 i001
EntryGO (wt %) bLight SourceSolventYield (%) c
140SunlightCH3CN28
24022 W CWF bulbCH3CN22
3401W Green LEDCH3CN7
4403W blue LEDCH3CN61
540No lightCH3CN0
6403W blue LEDTHF5
7403W blue LEDDMSO7
8403W blue LEDToluene34
9403W blue LEDDCE78
10403W blue LEDDMF0
11403W blue LED1,4-Dioxane0
12203W blue LEDDCE67
13303W blue LEDDCE72
14503W blue LEDDCE87
15603W blue LEDDCE85
1603W blue LEDDCE0
1750No lightDCE<5
a Reaction conditions: 4a (0.3 mmol), 5a (0.36 mmol), and solvent (1 mL), for 12 h at rt under open air. b With respect to the substrate 4a. c Isolated yield.
Table 2. Scope of indoles a.
Table 2. Scope of indoles a.
Molecules 27 00772 i002
EntryR1R2R3ProductYield (%) b
1HHH6aa87
25-IHH6ba83
35-CH3HH6ca89
45-CNHH6da78
56-OCH3HH6ea90
67-ClHH6fa80
77-OBnHH6ga83
84-CH3HH6ha71
94-CO2CH3HH6ia67
10HH2-CH36ja86
11HCH3H6ka86
125-CH3H2-CH36la82
13 cHH3-CH36ma84
a Reaction conditions: 4 (0.3 mmol), 5a (0.36 mmol), GO (50 wt %) with respect to the substrate 4a, and DCE (1 mL), for 12 h at rt under open air. b Isolated yield. c 3-Methyl-2-(p-tolylthio)-1H- indole (6ma) was obtained.
Table 3. Scope of thiols a.
Table 3. Scope of thiols a.
Molecules 27 00772 i003
EntryRProductYield (%) b
14-ClPh6ab86
24-BrPh6ac88
34-OCH3Ph6ad83
44-C2H5Ph6ae80
52,4-dimethylphenyl6af82
64-NO2Ph6ag91
7Naphthalen-2-yl6ah78
8 c4-OCH3Ph6ai78
9 d4-CH36aj76
a Reaction conditions: 4a (0.3 mmol), 5 (0.36 mmol), GO (50 wt %) with respect to the substrate 4a, and DCE (1 mL), for 12 h at rt under open air. b Isolated yield. c 5-Methoxy-3-((4-methoxyphenyl)thio)-1H-indole (6ai) was obtained. d 3-(p-Tolylthio)-1H-pyrrolo[3,2-b]pyridine (6aj) was obtained.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, Q.; Peng, X.; Li, H.; He, H.; Liu, L. Visible-Light-Induced, Graphene Oxide-Promoted C3-Chalcogenylation of Indoles Strategy under Transition-Metal-Free Conditions. Molecules 2022, 27, 772. https://doi.org/10.3390/molecules27030772

AMA Style

Huang Q, Peng X, Li H, He H, Liu L. Visible-Light-Induced, Graphene Oxide-Promoted C3-Chalcogenylation of Indoles Strategy under Transition-Metal-Free Conditions. Molecules. 2022; 27(3):772. https://doi.org/10.3390/molecules27030772

Chicago/Turabian Style

Huang, Qing, Xiangjun Peng, Hong Li, Haiping He, and Liangxian Liu. 2022. "Visible-Light-Induced, Graphene Oxide-Promoted C3-Chalcogenylation of Indoles Strategy under Transition-Metal-Free Conditions" Molecules 27, no. 3: 772. https://doi.org/10.3390/molecules27030772

Article Metrics

Back to TopTop