Next Article in Journal
Targeting Epigenetic Changes Mediated by Members of the SMYD Family of Lysine Methyltransferases
Next Article in Special Issue
One-Pot Synthesis of Isoxazole-Fused Tricyclic Quinazoline Alkaloid Derivatives via Intramolecular Cycloaddition of Propargyl-Substituted Methyl Azaarenes under Metal-Free Conditions
Previous Article in Journal
Synthesis of New Amino-Functionalized Porphyrins:Preliminary Study of Their Organophotocatalytic Activity
Previous Article in Special Issue
A Facile Synthesis of 2-Oxazolines via Dehydrative Cyclization Promoted by Triflic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis and Application Dichalcogenides as Radical Reagents with Photochemical Technology

1
Department of Chemistry, Changzhi University, Changzhi 046011, China
2
College of Chemistry and Chemical Engineering, Yantai University, Yantai 264005, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(4), 1998; https://doi.org/10.3390/molecules28041998
Submission received: 1 February 2023 / Revised: 11 February 2023 / Accepted: 13 February 2023 / Published: 20 February 2023
(This article belongs to the Special Issue Synthesis and Modification of Nitrogen Heterocyclic Compounds)

Abstract

:
Dichalcogenides (disulfides and diselenides), as reactants for organic transformations, are important and widely used because of their potential to react with nucleophiles, electrophilic reagents, and radical precursors. In recent years, in combination with photochemical technology, the application of dichalcogenides as stable radical reagents has opened up a new route to the synthesis of various sulfur- and selenium-containing compounds. In this paper, synthetic strategies for disulfides and diselenides and their applications with photochemical technology are reviewed: (i) Cyclization of dichalcogenides with alkenes and alkynes; (ii) direct selenylation/sulfuration of C−H/C−C/C−N bonds; (iii) visible-light-enabled seleno- and sulfur-bifunctionalization of alkenes/alkynes; and (iv) Direct construction of the C(sp)–S bond. In addition, the scopes, limitations, and mechanisms of some reactions are also described.

1. Introduction

Organosulfur and organoselenium compounds play an increasingly important role in organic synthesis and material science as well as in pharmaceutical applications [1]. In particular, C–S bonds exist in numerous natural products and biological and pharmaceutical compounds, such as nelfinavir, esomeprazole, and cephalosporin-type antibiotics, most of which have demonstrated antibacterial, antiviral, antitumor, and anti-inflammatory activities [2]. Furthermore, various organoselenium compounds have been identified as therapeutic agents, with anticancer, antiviral, and anti-Alzheimer’s properties [3]. They have also been explored as fluorescent probes, catalysts in organic synthesis, and functional organic materials [4]. Thus, the efficient construction of C–S and C–Se bonds has generated considerable interest in organic synthesis in the past decades.
The disulfide bond is an important functional group, existing in a large number of chemical and biological molecules, such as oxidized glutathione (GSSG), some proteins (thrombospondin-1 and protein disulfide isomerase [PDI]), and even some natural drugs (mitomycin disulfides and leinamycin). Disulfide bonds play a vital role in organisms and are generally obtained by covalently crosslinking two cysteine residues. GSSG and reduced glutathione (GSH) can maintain a balance of oxidation and reduction in cells, while disulfide bonds predominantly maintain the stability of protein conformation to ensure biological activity [5]. Compared with sulfur, Se has a larger diameter and a weaker negative charge [6]. The diselenide bond (Se–Se) is more redox active than the S–S bond in the microenvironment of cancer cells because its bond energy is lower than that of the disulfide bond (Se – Se = 172 kJ/mol; S – S = 268 kJ/mol) [7]. In addition, Sun et al. [8] studied the bond angles of sulfur bonds and selenium bonds and found that S–S (94.6°/97.9°) and Se–Se (89.9°/93.3°) had bond angles closest to 90°. In comparison, the diselenide bond has sufficient structural flexibility to establish the most favorable conformation during self-assembly.
As a green and sustainable energy source, visible-light photocatalysis has received increasing attention in synthetic chemistry. Indeed, lots of breakthroughs have been made through the effective fusion of radical chemistry and photocatalysis strategies [9]. In recent years, many visible-light-enabled direct reactions, including C−H selenylation/sulfuration and cyclization of diselenides/disulfides with alkenes and alkynes, among others, have been explored and successfully applied to the synthesis of various sulfur and selenium-containing compounds, showing outstanding synthetic value and potential applications. Although there are several other reviews on dichalcogenides, none of these reviews mainly focused on the application of dichalcogenides as radical reagents. Especially, the disulfide bond (S–S) and the diselenide bond (Se–Se) are very important and amazing. Along with the development of photochemical technology, in this review, we describe a recently established visible-light-induced synthesis strategy for sulfur and selenium-containing compounds with dichalcogenide (disulfides and diselenides) radical precursors. In addition, the reaction scope, application, and mechanism for some of these methods are also discussed. Finally, a number of issues and challenges are explored, such as the need for more catalytic approaches that should be explored with electrochemical strategies and a migration cyclization strategy involving disulfide and diselenide as radical reagents that should also be established.

2. Synthesis of Dichalcogenides (Disulfides and Diselenides)

Disulfides and diselenides have interesting properties and are also widely used as versatile free radical reagents in many organic transformations. For these reasons, chemists have made considerable efforts to design and develop new and effective methods for the construction of disulfides and diselenides. However, a literature review indicates that synthetic methods for disulfides and diselenides are limited. Here, we briefly summarize some well-known synthetic strategies based on the precursors used to generate disulfides and diselenides. Each of these reactions will be discussed in summary in the following introduction.
Lithium reagent method (Figure 1a): In 1983, Cava [10] improved the lithium reagent method by using N,N-dialkyl-thiocarbamoyl chloride, avoiding the by-product of diaryl-monoselenide in the original method. The yield using this approach was improved from 75% to 90%, but this method is carried out at a low temperature (−78 °C), while the preparation of the reagents required for the synthesis of diselenide is also complex.
Selenol oxidation method (Figure 1b): In 1988, Krief [11] discovered that diselenides can be prepared rapidly and in large quantities by using H2O2 or the corresponding selenic acid as the oxidant. Recently, Chinese scholars have found that diselenides can also be efficiently prepared using trichlorocyanuric acid [12].
Selenium transfer reagent method (Figure 1c): In 1995, Zhou et al. [13] first reported the use of a selenium transfer reagent to synthesize diselenide. Under mild conditions, the reaction of selenoamides with various alkyl halogenated hydrocarbons can produce alkyl diselenides in high yields. The disadvantage of this method is that selenoamides are not easy to prepare.
Catalytic cross-coupling method (Figure 1d): In 2010, Oscai’s group [14] first reported the preparation of diselenide compounds via cross-coupling catalyzed by copper oxide nanoparticles. Compared with traditional methods, nanocatalysts have been effectively processed, with their high surface area and special morphology, making them effective as catalysts in organic synthesis. This method is applicable not only to the preparation of alkyl and aryl diselenides but also to heterocyclic diselenides, with high universality and generally high yields.
Copper-catalyzed one-pot method (Figure 1e): In 2013, Zhang et al. [15] developed a method using cuprous chloride as the catalyst, acetylacetone as a ligand, and an aromatic halide as the raw material, which was directly reacted with selenium to synthesize diaryl selenide compounds.
Sulfur reagent transfer method (Figure 1f): In 2003, Kaushik et al. [16] also reported a new and effective sulfur transfer reagent (C6H5CH2N(Et)3)6-Mo7S24) to react with halogenated hydrocarbons to obtain disulfides. The reaction conditions are mild, and the yield is satisfactory.
Oxidative coupling method (Figure 1g): In 2005, Joshaghani [17] reported a new method for the mass production of disulfides. This method uses tert-Butyl hydroperoxide (TBHP) as an oxidant to directly transform thiols to disulfides, which is fast and efficient.
Microwave irradiation method (Figure 1h): In 2007, Lenardao’s group [18] discovered that symmetric disulfides were rapidly synthesized from mercaptans under solvent-free conditions using a solid Al2O3/KF system as a catalyst under microwave irradiation at room temperature. This improved method is generally aimed at the conversion of liquid mercaptans to disulfides, and the yields range from moderate to good. In addition, the catalyst system could be used twice without any decrease in activity.
Reductive coupling method (Figure 1i): In 2009, Yao [19] reported that triphenylphosphine was used for the reductive coupling of benzenesulfonyl chloride at room temperature to easily obtain disulfides.
Nanocatalysis method (Figure 1j): In 2013, Saidi [20] found that using nano iron oxide as a catalyst and hydrogen peroxide as an oxidant at room temperature can efficiently oxidize mercaptan to disulfide ether. Nano iron oxide is an efficient and recyclable catalyst, and hydrogen peroxide is a green oxidant. The advantage of this method is that the catalyst is stable and reliable, the reaction conditions are mild, and the selectivity is high.
Oxidation of thiols (Figure 1k): In 2015, Noël et al. [21] used eosin Y dye to oxidize thiols to disulfide derivatives under visible light. Compared with heavy metal catalysts, using organic dyes as photosensitizers is less toxic and less expensive. This method is suitable for aliphatic thiols, aromatic thiols, and heterocyclic thiols, with yields ranging from 86% to 99%. The advantage of this method is that the reaction time can be reduced from hours to minutes. Moreover, drugs such as oxytocin and disulfiram can be synthesized using this method.
Conversion from Na2S2 (Figure 1l): This method is used to prepare symmetrical disulfide derivatives by the reaction of alkyl bromide with Na2S2. In 2017, Wang and colleagues [22] reported a method for the synthesis of symmetric disulfides. This method uses a bromine-containing dienamine reacted with Na2S2 to afford the corresponding symmetrical disulfides in 54% yield.
Recently, Chen’s group [23] reported a class of N-acyl hypersulfonamides containing S–N bonds based on copper-catalyzed amination of azobene-mediated thiols. This class of S–N compounds was shown to be a good thiol sulfide reagent for the preparation of various disulfides (Figure 1m).

3. Application of Dichalcogenides under Visible-Light Irradiation

3.1. Cyclization of Diselenides/Disulfides with Alkenes and Alkynes

3.1.1. Cyclization of Diselenides with Alkenes and Alkynes

The ingenious design of novel radical addition cyclizations has emerged as an increasingly powerful tactic to construct complex molecules in an atom- and step-economical manner. Over the past decade, visible-light-driven photoredox catalysis has stimulated a resurgence of interest in the exploration of radical reactions.
In 2013, Ragains and coworkers reported a visible-light-promoted selenocyclization of alkenes at room temperature (Scheme 1) [24]. In this reaction, bench-stable PhSeSePh was combined with CBr4 under the irradiation of a 5W blue light-emitting diode (LED), resulting in the in situ generation of reactive PhSeBr. This reaction showed a broad substrate scope, generating O-heterocycles in high yields along with N-heterocycles in moderate to high yields. Notably, in CH2Cl2 as a solvent, diphenyl ditelluride provided successful tellurofunctionalization products in 53–75% yields. To further demonstrate the application of this method, the amaryllidaceae alkaloid, γ-lycorane, was synthesized. Free radical validation experiment and density functional theory (DFT) calculation suggested visible-light irradiation promoted the phenylselenyl radical abstraction of bromine from CBr4 to generate phenylselenyl bromide in situ. The detailed mechanism of these reactions is still under investigation.
In 2017, an efficient photocatalyst-free approach for the preparation of selenium-substituted spiro [4,5] trienones based on visible-light-induced selenium radical cyclization of N-aryl alkynamides was described by Baidya and coworkers (Scheme 2) [25]. Diverse N-aryl alkynamides and diaryl diselenides bearing electron-donating as well as electron-withdrawing groups in an aryl ring achieved the products under an oxygen atmosphere at room temperature. In addition, good yields were achieved in gram-scale reactions. A spiro-ring-opening strategy was realized to give fully substituted acryl amides 5.
To probe the reaction mechanism, control experiments were performed. The reaction yields dropped significantly in the presence of radical scavengers such as butylated hydroxytoluene (BHT) and 2,2,6,6-tetramethylpiperidinyl 1-oxide (TEMPO), and a plausible reaction mechanism was proposed, as shown in Scheme 2. First, under visible-light irradiation, the diselenide is homolytically cleaved to produce a selenium radical 4a, which subsequently attacks the triple bond position of compound 3 to obtain radical intermediate 7. The final radical 7 is obtained as intermediate 8 by intramolecular cyclization. The intermediate 8 undergoes an oxidative de-aromatization reaction in an oxygen environment to obtain the desired product 5.
In 2019, Xu and coworkers reported an Se radical-triggered multi-component tandem cyclization of alkyne-tethered cyclohexadienones 11 and diaryl diselenides 12 under the irradiation of 25 W white LEDs at 40 °C (Scheme 3) [26]. The cascade cyclization starts with the addition of selenyl radical 12a to the alkyne, generating a vinylic radical 14, and proceeds in a 5-exo-trig fashion to 15, with the formation of the C−C bond. The intermediate 15 captures another selenyl radical, 12a, yielding bis-selenyl chromenones 15a. In the presence of water and a base, the final step of nucleophilic substitution affords compound 13. Therefore, in the absence of a base, the authors obtained the bis-selenyl derivatives 15a.
In the same year, Kim et al. developed a practical method to synthesize selenated cyclopentanone derivatives 17 by photooxidation-catalyzed selenation through a ring expansion reaction with diselenides and alkenyl cyclobutanol derivatives (Scheme 4) [27]. A simple method was provided for the preparation of selenated cyclopentanone derivatives. The reaction occurred in acetonitrile solvent and was irradiated with blue light in air for 11 h. Ru(bpy)3Cl2·6H2O was used as a photocatalyst, with a yield of up to 94%. The authors performed a series of relevant control experiments to verify the mechanism of this reaction. The report described how the formation of the product was significantly inhibited by the addition of the radical scavenger TEMPO under standard conditions, demonstrating that the transformation involved a radical pathway. The authors proposed the reaction mechanism shown in Scheme 4. For example, the photocatalyst Ru(bpy)3Cl2·6H2O is converted to the excited state [Ru(bpy)3Cl2·6H2O]* under visible-light irradiation. Afterwards, the selenium radical 19 reacts with 1-(1-phenylvinyl)cyclobutanol 16 to give the carbon radical 21, which is oxidized by oxygen to the cationic intermediate 22 (path a). By 1,2-alkyl migration of intermediate 21, the alcohol rearranges to form cyclopentanone 17. Alternatively, radical 19 can be oxidized by oxygen to the selenium cation 20, which reacts with 1-(1-phenylethenyl)cyclobutanol 16 to give the corresponding cation 22 (path b).
In 2020, the Yang group disclosed a visible-light-induced, catalyst-free, and metal-free method for the construction of 4-sulfo/seleno-α,α-difluoro-γ-lactams via radical-initiated tandem cyclization (Scheme 5) [28]. In this process, 23 reacts smoothly with diphenyl disulfides under irradiation with 30 W blue LEDs, with K2HPO4 as a base to afford the desired products at room temperature. It is worth mentioning that the addition of Ru(bpy)3PF6 or Ir(ppy)2(dtbbpy)PF6 resulted in relatively lower yields of 39%. Control experiments demonstrated that visible light was vital for this transformation. Interestingly, for this photoinduced transformation, the 5-exo-trig cyclization dominated the process, and no 6-endo-trig cyclization product was isolated. This protocol exhibits good functional group tolerance and affords a variety of 4-thio- and 4-seleno-substituted 3,3-difluoro-γ-lactams in moderate to good yields. To gain insight into the mechanisms underlying this reaction, a series of control experiments were conducted. The addition of TEMPO suppressed the reaction and generated the addition compounds TEMPO−23 and TEMPO−18. The results of emission quenching experiments indicated that it is impossible to form the corresponding electron–donor–acceptor complexes for substrates 23 and 18. Based on these results, a feasible mechanism for this reaction was proposed (Scheme 5). Initially, the radicals Ph−X and Ph−XBr are generated under visible-light irradiation. The consumption of Ph–XBr by the base also promotes this transformation. The radical Ph–X then abstracts the bromine atom from substrate 23 to offer the radical intermediate 25, which undergoes rapid 5-exo-trig cyclization to form the radical 26. Ultimately, the radical 26 and Ph−X easily form the target 24 via radical–radical cross-coupling. Another radical pathway could not be completely excluded, as indicated by the combination of products detected.
In 2020, Reuter’s group [29] reported the visible-light-assisted dearomatic carbon selenide iodine cyclization of aromatic homologues (Scheme 6). At different temperatures, alkynones can deliver selenized cyclohexadienones and spiro diepoxides. The reaction mixture was exposed to blue LED illumination at 18 °C, using acetonitrile as solvent. Starting from 1,4-diphenylbut-3-yn-2-one 27 and diphenyl diselenides as the selenium source, selenospiro-cyclohexadienones were obtained using oxygen in up to 88% yield. In addition, experimental studies on free radical control confirmed that the process proceeds through the free radical pathway. The mechanism was rationally described based on control experiments and competition experiments, as shown in Scheme 6. Namely, the addition of aryl selenium radical 4a produced under blue LED irradiation to the alkyne bond of 27 produces vinyl radicals 29, which subsequently undergo intramolecular cyclization to give intermediate 30. In an oxygen environment, intermediate 30 undergoes peroxidation and is converted to the peroxy intermediate 31, which finally provides spirocyclized cyclic alkenones 28 together with an OH radical. Two OH radicals combine to deliver hydrogen peroxide followed by in situ generations of benzeneperoxyseleninic acid 33 in the presence of diaryl diselenide 4, which functions as an epoxidizing agent. The spiro-cyclized cyclic enone 28 undergoes double epoxidation with the in situ-generated epoxidizing agent 33, facilitating the epoxidation of the adjacent enones to the di-epoxide 32.
In 2020, Xu and coworkers further developed the visible-light-induced selenocyclization reaction of indolyl-ynones 30 with diselenides at room temperature under an air atmosphere (Scheme 7) [30]. Diverse 3-selenospiroindolenines bearing various functional groups were obtained in moderate to good yields. Similarly, a phenylselenyl free radical is generated from diphenyl diselenide under visible-light irradiation. The desired product is then obtained through the radical addition/oxidation/deprotonation pathway.
Based on free-radical verification experiments and previous literature reviews, Scheme 7 proposes and describes two possible pathways. The irradiation of diphenyl diselenide 18 produces phenylselenyl free radical 19. In path a, the addition of indolyl-ynones 34 to the phenylselenyl radical 19 affords an alkenyl radical 36, followed by cyclizing with the indole at its 3-position to form a spirocyclic radical intermediate 37. Oxidation of 37 to 38 in air then undergoes dehydrogenation in base conditions to afford the desired product 35. In path b, the reaction proceeds by oxidation of the phenylselenyl radical 19 to PhSe+ in air, which reacts with the alkyne group of indolyl-ynone 34, leading to the formation of seleniumion. Then, cyclization at the 3-position of indole gives the desired product 35.
Very recently, Wang disclosed a regio- and chemoselective radical cascade cyclization of 1,6-enynes 39 and areneselenosulfonates 40 under 34 W blue LED irradiation in the air without any photocatalysts (Scheme 8) [31]. Numerous substrates 39 were examined, and the corresponding cyclized products 41 were obtained in good to excellent yields. This reaction also proceeded smoothly using diaryldiselenides 40 with 1,6-enynes and produced the desired products with moderate to good yields. However, this method was not applicable when the chain length was increased from one to two or three. The internal alkene and free amine in enyne were also not tolerant of this transformation. This protocol offers an effective approach to building selenium-substituted pyrrolidine derivatives via multiple chemical bond constructions in a 5-exo-dig fashion, including one C–S bond, one C–Se bond, and one C–C bond.
Notably, the cyclization reaction does not require photocatalysts or other additives. The bond energy of the Se–Se bond allows the production of photoinduced phase Se radicals and is further used in a wide range of reactions. In 2020, Tran and colleagues reported the synthesis of diselenyl quinoxalate 44 from o-diisocyanate arenes 42 and diaryl or dialkyl diselenides 43 with moderate to good yields (Scheme 9) [32].
The addition of the radical inhibitor TEMPO under the standard reaction suggested that the transformation involved a radical process. Control experiments also showed that diselenides were activated through a photocatalytic energy transfer pathway. Based on the results of relevant control experiments and previous references, two possible mechanisms have been depicted, as shown in Scheme 9. The first mechanism involves a photoinduced aza-Bergman-type cyclization followed by trapping of the 11 species (path A). In the second mechanism (path B), the R2Se radical species 43a attacks the isocyanide group to form the imidoyl radical intermediate 42a, followed by intramolecular addition to the vicinal isocyanide group. Finally, intermediate 45 reacts with another equivalent of diorganoyl diselenide, R2Se 43a, affording the bis-selenylquinoxalines 44.
In 2021, Zhu and colleagues [33] described a visible-light-mediated diaryl selenide cyclization of 1,6-enylene 46 with diselenides ether radical cyclization (Scheme 10). It is proved that terminal alkyne and inner alkyne derived from 1,6-enylene are suitable for this synthesis, and the cyclized product 47 with a good yield is obtained. Through validation experiments based on free radicals, the mechanism of radical cyclization is proposed. First, diaryl diethers 18 produce aryl selenium radicals 19 under visible-light irradiation. Then, a selenium free radical 19 is added to the carbon–carbon triple bond of alkyne 46 to form a carbon radical intermediate 48. Then the intermediate vinyl free radical 48 undergoes an intramolecular 5-external triangular cyclization reaction to generate a tertiary carbon-alkyl radical 49. Finally, the diselenium compound is captured by radical 49 to obtain the target product 47 and the aryl selenyl radical 19.

3.1.2. Cyclization of Disulfides with Alkenes and Alkynes

In 2017, Wang’s group [34] developed a simple and effective scheme for the preparation of indoles using H2O2 and visible light to drive the reaction of 2-alkynyl aniline with disulfides. The corresponding products were synthesized at room temperature with high yields and without additional conditions. In the same year, another method was published for the preparation of benzothiophenes, which was achieved by visible-light induction of o-alkynyl anilines and disulfides without the addition of metals or photocatalysts (Scheme 11). The reaction offers good yields, broad substrate expansion, and simple operating conditions.
Benzofuran-like scaffolds are usually found in natural products and display biological and pharmaceutical activity. In 2021, Li’s group [35] developed a simple and effective method for the visible-light-induced tandem cyclization of 1,6-benzenes with disulfides for the synthesis of functionalized benzofurans (Scheme 12).
The reaction proceeded smoothly under metal-free conditions in room-temperature air, providing the desired products with broad functional group tolerance and good yields. On the basis of experiments verified by radicals, the authors propose a plausible mechanism, as shown in Scheme 12. On the one hand, PC (Mes-Acr-Me-ClO4) is transformed into the excited state PC* under visible-light irradiation, which is subsequently oxidized by oxygen to PC+ while generating superoxide radical anions [36]. In the presence of N-methyl-2-pyrrolidone (NMP), PC+ again becomes PC in the catalytic cycle. On the other hand, the superoxide radical anion is added to the alkyne 56 to provide the alkyne radical anion 59, which reacts with diphenyl disulfide 57 to produce the peroxyalkynyl sulfide intermediate 62 and release the sulfur group (PhS). The resulting 62 then undergoes a [1,5]-hydrogen transfer to give intermediate 60, which then loses the hydroxyl anion to yield the desired product 58. Moreover, the formed phenylthio group (PhS) can be added to another substrate molecule 56 to give vinyl 61, which reacts with the superoxide radical anion to produce intermediate 62. Intermediate 62 repeats the [1,5]-hydrogen transfer and elimination of the hydroxyl anion (HO-) to give the corresponding product 58.

3.2. Direct Selenylation/Sulfuration of C−H/C−C/C−N Bonds

3.2.1. Direct Selenylation/Sulfuration of C−H Bonds

Coumarins form a well-known class of naturally occurring heteroarenes, which exhibit a diverse range of biological and medicinal activity, including anticancer, antibacterial, and anticoagulant properties [37]. Their use in materials science is also well established [38]. In 2018, Yang and colleagues [39] reported the photoinduced C(sp2)−H selenium functionalization of 4-amino-substituted coumarins 63 in the presence of ammonium persulfate (Scheme 13). Under optimal conditions, various 4-amino-substituted coumarin derivatives were found to be converted to the desired products with diselenides in moderate to good yields; unfortunately, 2H-chromen-2-one is not suitable for this visible-light-induced selenization reaction.
Although the mechanism of the visible-light-induced selenization pathway remains unclear, Scheme 13 proposes two possible routes based on radical verification experiments and previous literature reports [40,41,42]. First, (NH4)2S2O8 is irradiated with visible light to give the reactive radical anion SO4•−. Then, a single-electron transfer process occurs between 63 and SO4•− to give the corresponding radical intermediate 63a. In path 1, intermediate 63b reacts with diselenide 4 to form intermediate 65b. In path 2, intermediate 63c reacts with diselenide 4 to form the C−3 selenide product 65c. A single-electron transfer process occurs between 66 and SO4•− to give the corresponding radical cation 69.
The use of inexpensive and readily available organic dyes as photocatalysts has attracted extensive attention in the synthesis community [43]. In 2018, Braga and colleagues reported a more environmentally friendly, photoinduced synthesis of selenium aromatic compounds 71 (Scheme 14). Under visible-light irradiation, using rose bengal (RB) as the photocatalyst and air as the terminal oxidant, the indole was selenized with a half molar equivalent of diselenide to obtain a selenium-containing indole compound in good yield. The reaction has broad functional group tolerance and is also applicable to a variety of heterocycles, including indazole, imidazopyridine, imidazolyl, and imidazothiazole, as well as aniline, anisole, β-naphthylamine, and β-naphthol, for example. Validation experiments based on free radicals allowed the authors to propose a possible reaction mechanism. First, RB changes from the ground state to the excited state RB* under visible-light irradiation, and 70 undergoes single-electron transfer to form radical cation 72. 72 obtains resonance formula 72a through resonance and reacts with diselenide to form intermediate 73 in the presence of O2. Next, intermediate 73 loses a proton at C−3, yielding the corresponding 3-selenoindole 71. At the same time, the remaining selenium radicals are oxidized back to diselenide 12a, entering the second cycle.
In 2019, Lemir’s group published a method for synthesizing 3-selenoindoles from diselenides and indoles or electron-rich olefins as raw materials and ethanol as a green solvent (Scheme 15) [44]. By this simple and environmentally friendly method, several 3-selenoindoles and many asymmetric diselenides can be obtained in high yields. In contrast, indoles with electron-withdrawing groups do not react.
Recently, α-selenoketones have attracted attention due to their versatile applications in organic synthesis as well as their biological activity in the treatment of disorders such as depression and anxiety [45]. Most of the α-selenofunctionalization protocols include the in situ generation of nucleophilic selenium reagents or previous formation of an electrophilic species from diselenides. In 2020, Schneider [46] reported an efficient photoinduced α-selenide reaction of ketones without metal additives or photosensitizers, providing a green scheme for the synthesis of various α-selenide ketones using light energy (Scheme 16). In this experiment, α-selenide ketones were prepared using equal amounts of diselenides and alkyl ketones 76 under compact florescent lamp ultraviolet A (CFL UVA) irradiation at 26 W for 6 h with 20 mol% pyrrolidine as the organocatalyst and CH3CN as solvent. Subsequently, the universality of the substrate was explored, and it was found that when the diorganodiselenide and the cyclic and acyclic ketones had different substituents, the reaction could proceed smoothly to create a series of α-selenide products with moderate to excellent yields. These findings demonstrate the generality of the method.
The addition of the radical inhibitor TEMPO to the standard reaction suggests that the transformation involves a radical process. Control experiments also showed that diselenide is activated through a photocatalytic energy transfer pathway. Based on the results of relevant control experiments and previous references, a possible mechanism was depicted, as shown in Scheme 16. First, 79 is formed in situ by the condensation of pyrrolidine 77 with cyclohexanone 76. Then, diselenide is cleaved to selenium radicals by Se−Se bonding under visible-light irradiation. The selenium radical undergoes addition to the enamine 79 to form the selenated intermediate 80. Subsequently, the species 80 can be oxidized by atmospheric O2 to the imino cation 81, providing the desired product 78 and regenerating the organocatalyst 77.
In 2022, Choudhury et al. [47] reported a visible-light-mediated C(sp2)−H selenation reaction of aminopyrazoles (Scheme 17). The authors chose 5-amino-3-methyl-1H-phenylpyrazoles 83 and diphenyl diselenides as model substrates and demonstrated experimentally that the reaction does not require additional metal oxidants; only 12 h of visible-light irradiation in nitrogen at room temperature is required to obtain the desired target products. The reaction also exhibits broad substrate versatility with moderate to good yields and is also suitable for gram-scale synthesis. Finally, the reaction provides important pharmaceutical heterocyclic compounds such as aminopyrazoles, isoxazoles, isothiazoles, and selenium uridines pyrimidines.
Organosulfur compounds are widely used in medicine, pesticides, new materials, and other fields. Therefore, it is very important to develop new methods for the formation of carbon–sulfur bonds. A reaction strategy for the generation of sulfur radicals from disulfides catalyzed by visible-light has been rapidly developed, and the direct sulfurization of carbon–hydrogen bonds has been widely used. In 2016, Wang’s group [48] developed a simple and effective method for the preparation of α-aryl thioethers by direct thiolysis at α-C(sp3)−H of ethers under visible-light induction using acridine red as a novel photocatalyst (Scheme 18). For this reaction, diphenyldisulfide 88 and tetrahydrofuran 89 were selected as model substrates, acetone as the solvent, and TBHP as the oxidant. When the reaction is irradiated with green light for 12 h at room temperature, the target product 90 can be obtained with a good yield.
According to the free radical experiment results, the authors propose a mechanism, as shown in Scheme 18. First, the AR in the ground state becomes AR* in the excited state under green light irradiation, which is highly reductive. Subsequent interaction with the oxide tBuOOH (TBHP) gives a hydroxyl radical HO and a tert-butoxy radical tBuO, while AR* is transformed into the ground state AR. Subsequently, HO or tBuO plucks a hydrogen from tetrahydrofuran (THF) to give the key alkoxy radical intermediate 91. Intermediate 91 reacts with PhSSPh 88a to give the desired product 2-(phenylthio)-tetrahydrofuran 90a and a new radical PhS.
In 2019, the Kumar group disclosed an attractive mild visible-light-induced strategy for the organo-chalcogenylation (S, Se) of indole in an oxygen atmosphere using commercially available substrates in acetone (Scheme 19) [49]. Further, cyclic voltammetry, UV-visible analysis, and electron paramagnetic resonance (EPR) spectroscopic studies have been carried out to establish the mechanism of the light-induced reaction.
Although allyl sulfides are widely used, the direct-catalyzed allyl C(sp3)−H sulfidation reaction remains unclear. In 2020, Hong et al. [50] reported a direct photooxidation-catalyzed allyl C(sp3)−H sulfidation reaction by visible-light (Scheme 20). The authors initially optimized the reaction conditions using disulfides 97 and 2,3-dimethyl-2-butene 98 as model substrates. The results showed that high yields of allyl sulfides 99 were obtained with 2.0 mol% (CF3ppy)3 as the catalyst, sodium hydroxide as the oxidant, and dimethylacetamide (DMA) as the solvent under 40 W blue LEDs irradiation for 24 h. In addition, the authors explored the universality of the substrate, and discovered that these reaction conditions were unfortunately not applicable to dialkyl disulfides, with only the corresponding thiols obtained. However, the authors proposed a feasible mechanism for the reaction. First, the sulfur group produced by the interaction of disulfides with an iridium catalyst under light may induce a single-electron transfer of allyl C(sp3)−H to obtain allyl radicals, which then undergo radical coupling with the disulfide. In addition, a redox process may occur in the presence of a photocatalyst in which the allyl radical is oxidized to a cationic intermediate, which is ionically coupled to the thioaryl anion to obtain the target product.
Recently, in 2020, the Glorius group published a photocatalytic dithioether reaction for the simple hydrogen methane-thionation of olefins without gaseous and toxic methane-thionate (Scheme 21) [51]. Fortunately, when the reaction uses 100 as a substrate, the reaction requires only the substitution of the thiol coupling with dimethyl disulfide and illumination with a blue light in DMA solvent for 16 h to give the desired product in a very good yield [52,53]. In addition, the reaction is also feasible for tribromoimidazole and 5-chloro-2-aminothiazole, demonstrating the suitability of the methanethiol scheme for a variety of structures.

3.2.2. Direct Selenylation of C−C Bonds

Selenium amino acids are important components for the synthesis of selenoproteins. However, there are few methods for the preparation of selenium amino acids. In 2016, Fu’s group [54] reported a method for the preparation of α-selenamine-based compounds using N-Bis(Boc)-Asp(oPht)-oMe 103 and diphenyl diselenides 104 as a model substrate (Scheme 22). More importantly, this method enables the synthesis of a series of bioactive and chiral selenoamino acid derivatives. A rational mechanism for the synthesis of chiral α-selenoamino acids is proposed in Scheme 22. First, 103, Ru(bpy)32+ is irradiated by visible-light to the excited state [Ru(bpy)32+]*, which is reduced by diethyl 2,6-dimethyl-1,4-dihydropyridine-3,5-dicarboxylate(HE) or N,N-diisopropylethylamine (DIPEA) to give Ru(bpy)3+, where HE or DIPEA became 106 or 106a. The subsequent interaction of Ru(bpy)3+ with 103 gives the radical anion 103a and the regenerated catalyst Ru(bpy)32+, after which 103a is eliminated to give the benzodicarboximide anion 107 and carbon dioxide to give the radical 103b. Finally, 104a and 106 are subjected to free radical coupling to obtain the target product 105. 104a interacts with 106 and 107 to give 108, 109, and 110.
C−H bond activation is one of the most challenging topics in organic synthesis and has received extensive attention from chemists. Finding environmentally friendly catalyst systems for carbon–hydrogen bond activation to construct carbon–selenium bonds has important theoretical and practical value. In 2019, Zhou et al. described the photocatalytic vulcanization, selenization, and boration of cyclosterone oxime esters by the cleavage of C−C bonds initiated by imidamines (Scheme 23) [55]. In this method, using fac-Ir(ppy)3, diphenyl diselenides 112 and irradiation with blue LEDs several functionalized organoselenides 113 were prepared. Based on the experimental results and previous reports [56,57,58], a plausible mechanism is proposed whereby Ir(III) is converted to the strongly reduced excited state Ir(III)* under visible-light irradiation, reducing 111a to the imine group 114. The radical 114 is cleaved by the ring-opening C−C bond to generate the highly reactive cyanoalkyl radical 115, which is captured by disulfides, diselenides, and diborides, respectively, forming an sp3 C−S bond and C−Se bond. The single electron transfer(SET) process of RX reduction to Ir(IV) produces cation 112b, which regenerates Ir.

3.2.3. Direct Selenylation/Sulfuration of C−N Bonds

As a common chemical bond, C−N bonds widely exist in organic molecules, and the activation and breaking of C−N bonds play an important role in organic reactions and life chemistry. In 2020, Zhao [59] reported the substitution of azosulfone groups in heterocycles by organoselenide groups driven by blue LEDs. The method is also applicable to the preparation of selenide aromatics and organotelluride products (Scheme 24). The addition of the radical inhibitor TEMPO under standard reactions suggests that the transformation involves a radical process. The authors propose an initial mechanism whereby the (hetero)aryl substrate 116 becomes excited upon irradiation with visible light, followed by the decomposition of the azosulfone group to form the aryl radicals 119. Aryl radicals 119 with diselenides 117 affords the corresponding selenium-containing compound 118 and the organoselenides RY* group.
In the past few years, pyridinium salts have been widely used in organic synthesis as radical precursors due to their ability to reduce single-electron transfer [60]. These salts typically form new C−C, C−N, C−O, and C−B bonds via transition metal catalysis, photocatalysis, and Lewis base activation. However, the formation of carbon–chalcogen bonds by cleavage of pyridinium salts has not been reported to date [61,62,63]. In 2021, Zhao et al. [64] reported a method for the synthesis of asymmetric selenides from alkylpyridinium salts and diselenides under visible-light catalysis (Scheme 25). Radical verification experiments show that the reaction is a radical pathway. First, the authors chose benzyl Katritzky salt 120 and diselenide 12 as model substrates. The selenide product 121 was obtained by reacting with DIPEA as the oxidant and eosin Y as the photocatalyst in an acetonitrile solution at room temperature for 20 h. The reaction is germane to a wide range of applications. First, the excited state of eosin Y (EY*) is reduced by single-electron transfer 120 to yield a dihydropyridine radical 122, which subsequently decomposes into a benzyl radical 123 and an aromatized pyridine derivative 124. Single-electron reduction to EY+ from DIPEA concurrently regenerates the ground-state eosin Y. Next, benzyl radical 123 reacts with diselenide 12 to give product 121.
In recent years, aryl hydrazides have attracted much attention as electrophilic substitutes in cross-coupling reactions. Aromatics can construct various chemical bonds, including C−C, [65] C−N, [66] C−S, [67] C−Se, [68] and C−P [69] bonds. In particular, aryl hydrazines can act as aryl reactants by releasing nitrogen gas. In 2019, Yu’s group [70] reported the testing of the substrates phenylhydrazine 125 and diphenyl sulfide 126 using the organic dye Na2-eosin Y as the catalyst, hydrogen peroxide as the oxidant, and dimethyl sulfoxide (DMSO) as the optimal reaction conditions to yield the expected sulfide product 127 (Scheme 26). Under optimal conditions, they explored a range of visible-light-promoted hydrazine sulfides and showed that substrates with strong electron-absorbing groups were not converted to the corresponding sulfides with high efficiency. This may be due to the low reactivity of the corresponding radical intermediates.
When the sulfidation reaction was carried out with TEMPO or BHT, only a trace amount of 127 was detected based on gas chromatology (GC) analysis. Based on these experimental results and related reports [71], a plausible mechanism is proposed. First, the photocatalyst Na2-eosin Y is irradiated to the excited state PC* and then converted to PC- by single-electron transfer, and PC- is oxidized by oxygen (from H2O2 and air) to provide the state photocatalyst and O2-·. At the same time, phenylhydrazine 120 is oxidized to generate the radical cation 12 and the radical cation 128 is deprotonated to provide radical 129. The radical 129 line is then deprotonated to convert the intermediate 131. Subsequently, nitrogen is eliminated from 131 to form the phenyl radical 132. On the other hand, the benzene sulfide radical 126a is generated by homolytic cleavage of 126 under visible-light irradiation. Finally, the target product, diphenyl sulfide 127a, is obtained by the radical coupling of 132 and 126b.

3.3. Visible-Light-Enabled Seleno-and Sulfur-Bifunctionalization of Alkenes/Alkynes

Alkenes/alkynes are simple and abundant bulk commodities, and the vicinal difunctionalization of these feedstocks is of great importance in rapidly increasing molecular complexity with a variety of functional groups.
In 2013, Ogawa et al. [72] used a diselenide-Ph2P(O)H hybridization system to achieve highly regioselective hydroselenation of deactivated terminal alkynes under visible-light irradiation to give vinyl selenide 135. Based on free radical validation experiments, a reasonable mechanism was proposed, as shown in Scheme 27. First, visible-light irradiation triggers the cleavage of the Se−Se single bond in the diselenides, generating the corresponding selenyl radicals. Next, the attack of the selenyl radicals on the terminal carbon of the olefin leads to the production of the vinyl intermediate. The vinyl radicals are captured by selenol to produce hydroselenate products, while the selenium radicals are regenerated. In addition, Ph2P(O)SeR generated as a by-product does not undergo addition reactions with the alkynes. This is in contrast to the addition reaction between the unoxidized Ph2PSePh and the alkynes.
In 2019, Chen et al. [73] reported a multi-component cascade reaction of diselenides, alkynes, and sulfur dioxide in the presence of visible light. A novel β-sulfonylvinylsilane was prepared by this cascade reaction (Scheme 28). In addition, the conversion reactions of β-sulfonylvinylsilanes to 1,4-oxopyrimidine-4,4-dioxides and sulfonylacetylene derivatives were also investigated. The reaction was carried out in a 1,4-diazabicyclo(2,2,2)octane (DABCO)-(SO2)2 solution with CH3CN as the solvent and irradiated under blue light for 6 h. The substituents with halogen, methyl, methoxy, nitro, methoxycarbonyl, and amino groups were detected, respectively. The alkynes and diselenides of alkynes and diselenides showed experimentally that all reactions proceeded smoothly, but the conditions were not applicable to alkylalkynes. The formation of the product was significantly inhibited by the addition of the radical scavenger TEMPO under standard conditions, demonstrating that the transformation involved a radical pathway.
According to previous literature reports [74,75], a feasible reaction mechanism is proposed in Scheme 28. First, under blue light, the diselenide is excited to form selenium radicals. The addition of 136 gives the vinyl radical intermediate 138, which is then trapped by sulfur dioxide. A sulfonyl radical intermediate 139 is then formed. Subsequent attack on 136 produces the vinyl intermediate 140, which is captured by disulfide. Finally, product 137 is generated and provides recovery of the selenium radical [76].
Difunctionalization of alkenes enables rapid construction of complex molecules with broad applications in organic synthesis. In 2020, Yuan’s group [77] reported a mild and effective strategy for the synthesis of acyloxy selenides via visible-light-induced bifunctionalization of styrene with diaryl diselenides and carboxylic acid dibasic compounds (Scheme 29). The work was carried out under white LEDs (20 W) with 4-methylstyrenes 141, diphenyl diselenides, and acetic acids 142 as substrates, CH3CN as a solvent, and 1.0 mol% RB as the required catalyst. The reaction was carried out for 14 h at room temperature to give the target product 143 with a 58% yield.
Controlled experimental studies confirmed that this process was carried out through a radical chain propagation process. On the basis of controlled and competitive experiments, a reasonable mechanism was depicted, as shown in Scheme 29 [78]. Initially, the rose bengal (RB) produces an excited state under the irradiation of blue LED light. Subsequently, the excited state RB* acquires electrons from diphenyl diselenide, providing RB•− and phenylselenide 19, which reacts with 4-methylstyrenes 141 to form 144, and another phenylselenide reacts with acetic acid 142 to form 145 and PhSeH, which is then oxidized by oxygen to PhSeSePh. Finally, the two radicals 144 and 145 react to give the target product 143.
β-Oxysulfides are a familiar organic molecule, widely used in organic synthesis precursors and functional compounds [79,80]. This unit is also present in biologically active pharmaceuticals and natural compounds [81]. However, there is no literature on the preparation of β-alkoxysulfides using diols as substrates. In 2019, Wang et al. reported a photochemical method for the preparation of bifunctional β-oxysulfides without photocatalysts and oxidants (Scheme 30). In this paper, styrene 146 and diphenyl disulfide 47 were used as model substrates to start the study. The desired products, 1-phenyl-2-(phenylthio)ethane-1-ol 147, were obtained in the presence of dichloromethane (CH2Cl2), water, CBr4, and blue LEDs in the system [82]. Subsequently, the prevalence of the reaction substrates, olefins, and thioethers was explored, and it was found that most of the expected products could be obtained. However, when aliphatic olefins or aliphatic disulfides were used as substrates, none of the expected products were observed.
To elaborate on the possible reaction mechanism, the authors performed a controlled experiment where the difunctionalization was almost suppressed when TEMPO was added to the reaction, which indicates that free radical processes are involved in the reaction. A potential mechanism is outlined in Scheme 30. First, under the irradiation of blue light, the disulfide 126a is homogenized to generate the sulfur radical 126b, and then 126b reacts with CBr4 to obtain the tribromomethyl radical, before the hydrogen is extracted from H2O or HOCH2CH2OH to obtain the corresponding hydroxyl group free radicals (OH) or alkoxy groups (OCH2CH2OH). An addition reaction of 126b with styrene 146 yields the intermediate 148. Finally, intermediate 148 with OH or OCH2CH2OH yields the desired products 147aa or 147ab.

4. Construction of the C(sp)-S Bond

The carbon–heteroatom bond formation is used as a bridge for linking organic molecules to access complicated compounds with biological and pharmaceutical activity [83]. Although many synthetic strategies have been established for the C(sp3)–S and C(sp2)–S bond-forming formation, there are still important challenges for the construction of C(sp)–S bonds, particularly for the achievement of C(sp)–S bond formation [84]. In 2021, Wang et al. reported a strategy for the preparation of alkynyl sulfides using visible-light irradiation (Scheme 31) [85]. First, 2,2′-diaminodiphenyl disulfides 149 and phenylethynyl bromides 150 were chosen as model substrates, irradiated under blue light, and stirred in CH2Cl2 under an N2 atmosphere for 12 h to obtain the desired coupling products 151. The functional group tolerance of this visible-light-promoted C(sp)−S cross-coupling was investigated under optimized reaction conditions using various brominated alkynes and found to behave well for various brominated alkynes with different aromatic rings and good functional group compatibility.
To clearly illustrate the reaction and establish that the coupling reaction is initiated by disulfide 149 or bromoalkynes 150, control experiments were performed. The coupling reaction is stopped when the radical scavenger TEMPO is added to the reaction system; this significant inhibition means that the reaction may involve a free radical process. Based on the results of controlled experiments and the relevant literature, the transformation mechanism was proposed in Scheme 31. First, the intermediate product o-NH2C6H4S-(152) was produced by visible-light-induced homolysis of 2,2’-diaminodiphenyl disulfide (149). Then, 152 was added to brominated alkynes 150 to form an intermediate product 153, which was converted to product 151a by an elimination reaction.

5. Conclusions

In this review, the recent progress in the synthesis and application of disulfide and diselenide as radical reagents with photochemical technology is reviewed, and the reaction scopes, limitations, and mechanisms of some of these reactions are also discussed. From the above discussion, it can be seen that the visible-light-promoted disulfide and diselenide radical reactions provide a stable and robust platform for the efficient construction of various selenium-/sulfur-containing heterocyclic compounds or bifunctional products. The successful examples described in this review convincingly demonstrate the high potential of disulfide and diselenide for drug discovery and applications. Among these synthesis methods, photochemistry and synergistic metal/photoredox catalysis techniques provide a more friendly and sustainable alternative for the application of disulfide and diselenide and illustrate the development prospects of thioether as free radicals.
Although significant progress has been made in this field over the last few decades, several issues and challenges remain in fully developing the potential applications of disulfide and diselenide, which should be further exploited and improved. More catalytic means should be explored in the application of disulfide and diselenide as free radical reagents, especially in terms of electrochemical synthesis methods, because electrochemical strategies have become increasingly powerful tools for the synthesis of organic compounds in recent years. Furthermore, the migration cyclization strategies involving disulfide and diselenide as radical reagents should be established to construct more complex polycyclic compounds, since polycyclic frameworks play an important role in medicinal chemistry and organic materials. In addition, the harmful consequences of metal catalysts and peroxides used in the reaction process cannot be ignored. It is our hope that this paper will encourage more researchers to contribute to this emerging field by developing more environmentally friendly catalytic systems.

Author Contributions

Writing—original draft preparation, C.W., Y.Z., T.Y. and F.L.; writing—review and editing, K.S., X.W.; funding acquisition, K.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Qingchuang Technology Support Program of University in Shandong Province (2021KJ066).

Data Availability Statement

The data presented in this study are available upon request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, X.; Meng, J.P.; Zhao, D.Y.; Tang, S.; Sun, K. Synthesis and applications of thiosulfonates and selenosulfonates as free-radical reagents. Chin. Chem. Lett. 2023, 34, 107736. [Google Scholar] [CrossRef]
  2. Wang, X.; Zhang, Y.; Sun, K.; Meng, J.P.; Zhang, B. Study on the photoelectric technology in the synthesis of selenium-containing heterocycles. Chin. J. Org. Chem. 2021, 41, 4588–4609. [Google Scholar] [CrossRef]
  3. Chen, Z.; Lai, H.Q.; Hou, L.Y.; Chen, T.F. Rational design and action mechanisms of chemically innovative organoselenium in cancer therapy. Chem. Commun. 2020, 56, 179–196. [Google Scholar] [CrossRef] [PubMed]
  4. Ortgies, S.; Breder, A. Oxidative alkene functionalizations via selenium-π-acid catalysis. ACS Catal. 2017, 7, 5828–5840. [Google Scholar] [CrossRef]
  5. Lomaestro, B.M.; Malone, M. Glutathione in health and disease: Pharmacotherapeutic issues. Ann. Pharmacother. 1995, 29, 1263–1273. [Google Scholar] [CrossRef]
  6. Ma, N.; Li, Y.; Ren, H.F.; Xu, H.P.; Li, Z.B.; Zhang, X. Selenium-containing block copolymers and their oxidation-responsive aggregates. Polym. Chem. 2010, 1, 1609–1614. [Google Scholar] [CrossRef]
  7. Ji, S.; Cao, W.; Yu, Y.; Xu, H. Dynamic Diselenide Bonds: Exchange Reaction Induced by Visible Light without Catalysis. Angew. Chem. Int. Ed. 2014, 53, 6781–6785. [Google Scholar] [CrossRef]
  8. Polshettiwar, V.; Nivsarkar, M.; Acharya, J. A new reagent for the efficient synthesis of disulfides from alkyl halides. Tetrahedron Lett. 2003, 44, 887–889. [Google Scholar] [CrossRef]
  9. Sun, B.; Luo, C.; Zhang, X.; Guo, M.; Sun, M.; Yu, H.; Chen, Q.; Yang, W.; Wang, M.; Zuo, S. Probing the impact of sulfur/selenium/carbon linkages on prodrug nanoassemblies for cancer therapy. Nat. Commun. 2019, 10, 3211–3219. [Google Scholar] [CrossRef] [Green Version]
  10. Jen, K.Y.; Cava, M.P. Improved synthesis of aromatic diselenides. J. Org. Chem. 1983, 48, 1449–1451. [Google Scholar] [CrossRef]
  11. Krief, A.; Mahieu, A.F.D.; Dumont, W.; Trabelsi, M. Efficient syntheses of diselenides from selenols. Synthesis 1988, 2, 131–133. [Google Scholar] [CrossRef]
  12. Zhong, P.; Guo, M.P. Preparation of diselenides by the oxidation of selenols using trichloroisocyanuric acid. Synth. Commun. 2001, 31, 1507–1510. [Google Scholar] [CrossRef]
  13. De, R.M.; Rong, Z.H.; Qiang, F.W.; Zhou, X.J. First example of selenium transfer reaction of primary selenoamides and selenourea. Novel synthesis of dialkyl diselenides from alkyl halides. J. Organomet. Chem. 1995, 485, 19–24. [Google Scholar] [CrossRef]
  14. Singh, D.; Deobald, A.M.; Camargo, L.R.S.; Tabarelli, G.; Rodrigues, O.E.D.; Braga, A.L. An efficient one-pot synthesis of symmetrical diselenides or ditellurides from halides with CuO nanopowder/Seo or Teo/base. Org. Lett. 2010, 12, 3288–3291. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, S.; Ran, L.; Luo, J.; Xiong, F.; Jing, L.; Qin, D. Synthesis of diaryl selenyl ethers by copper-catalyzed one-pot method. Chem. Res. Appl. 2013, 25, 1632–1638. [Google Scholar]
  16. Chen, X.; Cui, M.F.; Zhang, J. Phase transfer catalytic synthesis of bicyclohexyl disulfide. Fine Chem. 2005, 22, 874–877. [Google Scholar]
  17. Joshaghani, M.; Khosropour, A.; Jafary, H. Mild and highly efficient preparation of symmetrical disulfides and diselenides using bipyridinum hydrobromide perbromide as a new oxidative reagent. Phosphorus Sulfur Silicon 2005, 180, 117–123. [Google Scholar] [CrossRef]
  18. Lenardao, E.J.; Lara, R.G.; Silva, M.S. Clean and fast oxidative transformation of thiolsto disulfides under solvent-free conditions. Tetrahedron Lett. 2007, 39, 7668–7670. [Google Scholar] [CrossRef]
  19. Kabalka, G.W.; Reddy, M.S.; Yao, M.L. Synthesis of diaryl disulfides via the reductive coupling of arylsulfonyl chlorides. Tetrahedron Lett. 2009, 50, 7340–7342. [Google Scholar] [CrossRef]
  20. Rajabi, F.; Kakeshpool, T.; Saidi, M.R. Supported iron oxide nanoparticles: Recoverable and efficient catalyst for oxidative S−S coupling of thiolsto disulfides. Catal. Commun. 2013, 40, 13–17. [Google Scholar] [CrossRef]
  21. Talla, A.; Driessen, B.; Straathof, N.J.W.; Milroy, L.G.; Brunsveld, L.; Noel, T. Metal-free photocatalytic aerobic oxidation of thiols to disulfides in batch and continuous-flow. Adv. Synth. Catal. 2015, 357, 2180–2186. [Google Scholar] [CrossRef]
  22. Yu, B.; Zheng, Y.; Yuan, Z. Toward direct protein S-persulfidation: A prodrug approach that directly delivers hydrogen persulfide. J. Am. Chem. Soc. 2018, 140, 30–33. [Google Scholar] [CrossRef] [PubMed]
  23. Bai, Z.; Zhu, S.; Hu, Y.; Yang, P.; Chu, X.; He, G.; Wang, H.; Chen, G. Synthesis of N-acyl sulfenamides via copper catalysis and their use as S-sulfenylating reagents of thiols. Nat. Commun. 2022, 13, 6445. [Google Scholar] [CrossRef] [PubMed]
  24. Conner, E.S.; Crocker, K.E.; Fernando, R.G.; Fronczek, F.R.; Stanley, G.G.; Ragains, J.R. Visible-light-promoted selenofunctionalization of alkenes. Org. Lett. 2013, 15, 5558–5561. [Google Scholar] [CrossRef] [PubMed]
  25. Sahoo, H.; Mandal, A.; Dana, S.; Baidyaa, M. Visible-light-induced synthetic approach for selenylative spirocyclization of N-aryl alkynamides with molecular oxygen as oxidant. Adv. Synth. Catal. 2018, 360, 1099–1103. [Google Scholar] [CrossRef]
  26. Ma, X.L.; Wang, Q.; Feng, X.Y.; Mo, Z.Y.; Pan, Y.M.; Chen, Y.Y.; Xin, M.; Xu, Y.L. Metal-free visible-light induced cyclization/substitution cascade reaction of alkyne-tethered cyclohexadienones and diselenides: Access to 5-hydroxy-3-selenyl-4a,8a-dihydro-2H-chromen-6(5H)-ones. Green Chem. 2019, 21, 3547–3551. [Google Scholar] [CrossRef]
  27. Jung, H.I.; Kim, D.Y. Synthesis of β-selenylated cyclopentanones via photoredox-catalyzed selenylation ring-expansion cascades of alkenyl cyclobutanols. Synlett 2019, 30, 1361–1365. [Google Scholar] [CrossRef] [Green Version]
  28. Ye, Z.P.; Xia, P.; Liu, F.; Hu, Y.Z.; Song, D.; Xiao, J.A.; Huang, P.; Xiang, H.Y.; Chen, X.Q.; Yang, H. Visible-light-induced, catalyst-free radical cross-coupling cyclization of N-allylbromodifluoroacetamides with disulfides or diselenides. J. Org. Chem. 2020, 85, 5670–5682. [Google Scholar] [CrossRef]
  29. Sahoo, S.R.; Das, B.; Sarkar, D.; Henkel, F.; Reuter, H. Visible-light assisted selenylative intramolecular dearomative carbo-spirocyclisation (IDCS) of homologated-ynones. Eur. J. Org. Chem. 2020, 7, 891–896. [Google Scholar] [CrossRef]
  30. Zhou, X.J.; Liu, H.Y.; Mo, Z.Y.; Ma, X.L.; Chen, Y.Y.; Tang, H.T.; Pan, Y.M.; Xu, Y.L. Visible-light-promoted selenylative spirocyclization of indolyl-ynones toward the formation of 3-selenospiroindolenine anticancer agents. Chem. Asian J. 2020, 15, 1536–1539. [Google Scholar] [CrossRef]
  31. Mutra, M.R.; Kudale, V.S.; Li, J.; Tsaia, W.H.; Wang, J.J. Alkene versus alkyne reactivity in unactivated 1,6-enynes: Regio- and chemoselective radical cyclization with chalcogens under metal- and oxidant-free conditions. Green Chem. 2020, 22, 2288–2300. [Google Scholar] [CrossRef]
  32. Tran, C.C.; Kawaguchi, S.; Sato, F.; Nomoto, A.; Ogawa, A. Photoinduced cyclizations of o-diisocyanoarenes with organic diselenides and thiols that afford chalcogenated quinoxalines. J. Org. Chem. 2020, 85, 7258–7266. [Google Scholar] [CrossRef] [PubMed]
  33. Hou, H.; Sun, Y.; Pan, Y.J.; Yu, H.G.; Han, Y.; Shi, Y.C.; Yan, C.G.; Zhu, S.Q. Visible-light mediated diaryl selenylative cyclization of 1,6-Enynes. J. Org. Chem. 2021, 86, 1273–1280. [Google Scholar] [CrossRef] [PubMed]
  34. Shi, Q.; Li, P.; Zhang, Y.; Wang, L. Visible-light-induced a 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]
  35. Li, X.; Xu, Z.; Wang, L.; Wang, F.; Yang, J.; Li, P. A facile synthesis of functionalized benzofurans via visible-light-induced tandem cyclization of 1,6-enynes with disulfides. ChemPhotoChem 2021, 5, 142–148. [Google Scholar] [CrossRef]
  36. Romero, N.A.; Nicewicz, D.A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116, 10075–10166. [Google Scholar] [CrossRef]
  37. Li, H.; Yao, Y.; Li, L. Coumarins as potential antidiabetic agents. J. Pharmacol. 2017, 69, 1253–1264. [Google Scholar] [CrossRef] [Green Version]
  38. Cuevas, J.M.; Seoane-Rivero, R.; Navarro, R.; Marcos-Fernandez, A. Coumarins into polyurethanes for smart and functional materials. Polymers 2020, 12, 630. [Google Scholar] [CrossRef] [Green Version]
  39. Yang, D.; Li, G.; Xing, C.; Cui, W.; Li, K.; Wei, W. Metal- and photocatalyst-free visible-light-promoted regioselective selenylation of coumarin derivatives via oxidation-induced C−H functionalization. Org. Chem. Front. 2018, 5, 2974–2979. [Google Scholar] [CrossRef]
  40. Jin, J.; MacMillan, D.W.C. Direct α-arylation of ethers through the combination of photoredox-mediated C−H functionalization and the minisci reaction. Angew. Chem. Int. Ed. 2015, 54, 1565–1569. [Google Scholar] [CrossRef] [Green Version]
  41. Devari, S.; Shah, B.A. Visible-light-promoted C−H functionalization of ethers and electron-deficient arenes. Chem. Commun. 2016, 52, 1490–1493. [Google Scholar] [CrossRef] [PubMed]
  42. Zhao, Y.; Huang, B.; Yang, C.; Xia, W. Visible-light-promoted direct amination of phenols via oxidative cross-dehydrogenative coupling reaction. Org. Lett. 2016, 18, 3326–3329. [Google Scholar] [CrossRef] [PubMed]
  43. Saba, S.; Rafifique, 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] [PubMed]
  44. Lemir, I.D.; Castro-Godoy, W.D.; Heredia, A.A.; Schmidt, L.C.; Arguello, J.E. Metal- and photocatalyst-free synthesis of 3-selenylindoles and asymmetric diaryl selenides promoted by visible-light. RSC Adv. 2019, 9, 22685–22694. [Google Scholar] [CrossRef] [Green Version]
  45. Sousa, F.S.S.; Birmann, P.T.; Baldinotti, R.; Fronza, M.G.; Balaguez, R.; Alves, D.; Bruning, C.A.; Savegnago, L. α-(phenylselanyl) acetophenone mitigates reserpine-induced pain-depression dyad: Behavioral, biochemical and molecular docking evidences. Brain Res. Bull. 2018, 142, 129–137. [Google Scholar] [CrossRef]
  46. Dalberto, B.T.; Schneider, P.H. Photoinduced metal-free α-selenylation of ketones. RSC Adv. 2020, 10, 10502–10509. [Google Scholar] [CrossRef] [Green Version]
  47. Ali, D.; Parvin, T.; Choudhury, L.H. Visible-light-mediated C(sp2)−H selenylation of amino pyrazole and amino uracils in the presence of rose bengal as an organophotocatalyst. J. Org. Chem. 2022, 87, 1230–1239. [Google Scholar] [CrossRef]
  48. Zhu, X.; Xie, X.; Li, P.; Guo, J.; Wang, L. Visible-light-induced direct thiolation at α-C(sp3)−H of ethers with disulfides using acridine red as photocatalyst. Org. Lett. 2016, 18, 1546–1549. [Google Scholar] [CrossRef]
  49. Rathore, V.; Kuma, S. Visible-light-induced metal and reagent-free oxidative coupling of C(sp2)−H bonds with organo-dichalcogenides: Synthesis of 3-organochalcogenyl indoles. Green Chem. 2019, 21, 2670–2676. [Google Scholar] [CrossRef]
  50. Kim, J.; Kang, B.; Hong, S.H. Direct allylic C(sp3)−H thiolation with disulfides via visible-light photoredox catalysis. ACS Catal. 2020, 10, 6013–6022. [Google Scholar] [CrossRef]
  51. Sandfort, F.; Knecht, T.; Pinkert, T.; Daniliuc, C.G.; Glorius, F. Site-selective thiolation of (multi)halogenated heteroarenes. J. Am. Chem. Soc. 2020, 142, 6913–6919. [Google Scholar] [CrossRef]
  52. Tyson, E.L.; Ament, M.S.; Yoon, T.P. Transition metal photoredox catalysis of radical thiol-ene reactions. J. Org. Chem. 2013, 78, 2046–2050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Czyz, M.L.; Weragoda, G.K.; Monaghan, R.; Connell, T.U.; Brzozowski, M.A.; Scully, D.; Burton, J.; Lupton, D.W.; Polyzos, A.A. Visible-light photocatalytic thiolation of aryl, heteroaryl and vinyl iodides. Org. Biomol. Chem. 2018, 16, 1543–1551. [Google Scholar] [CrossRef] [PubMed]
  54. Jiang, M.; Yang, H.J.; Fu, H. Visible-light photoredox synthesis of chiral α-selenoamino acids. Org. Lett. 2016, 18, 1968–1971. [Google Scholar] [CrossRef] [PubMed]
  55. Anand, D.; He, Y.; Li, L.; Zhou, L. A photocatalytic sp3 C−S, C−Se and C−B bond formation through C−C bond cleavage of cycloketone oxime esters. Org. Biomol. Chem. 2019, 17, 533–540. [Google Scholar] [CrossRef]
  56. He, B.Q.; Yu, X.Y.; Wang, P.Z.; Chen, J.R.; Xiao, W.J. A photoredox catalyzed iminyl radical-triggered C−C bond cleavage/addition/Kornblum oxidation cascade of oxime esters and styrenes: Synthesis of ketonitriles. Chem. Commun. 2018, 54, 12262–12265. [Google Scholar] [CrossRef]
  57. Li, L.; Chen, H.; Mei, M.; Zhou, L. Visible-light promoted γ-cyanoalkyl radical generation: Three-component cyanopropylation etherification of unactivated alkenes. Chem. Commun. 2017, 53, 11544–11547. [Google Scholar] [CrossRef]
  58. Zhang, J.J.; Duan, X.H.; Wu, Y.; Yang, J.C.; Guo, L.N. Transition-metal free C−C bond cleavage/borylation of cycloketone oxime esters. Chem. Sci. 2018, 10, 161–166. [Google Scholar] [CrossRef] [Green Version]
  59. Liu, J.J.; Tian, M.M.; Li, Y.X.; Shan, X.W.; Li, A.K.; Lu, K.; Fagnoni, M.; Protti, S.; Zhao, X. Metal-free synthesis of unsymmetrical aryl selenides and tellurides via visible-light-driven activation of arylazo sulfones. Eur. J. Org. Chem. 2020, 47, 7358–7367. [Google Scholar] [CrossRef]
  60. He, F.S.; Ye, S.Q.; Wu, J. Recent advances in pyridinium salts as radical reservoirs in organic synthesis. ACS Catal. 2019, 9, 8943–8960. [Google Scholar] [CrossRef]
  61. Basch, C.H.; Liao, J.N.; Xu, J.Y.; Piane, J.J.; Watson, M.P. Harnessing alkyl amines as electrophiles for nickel-catalyzed cross couplings via C−N Bond Activation. J. Am. Chem. Soc. 2017, 139, 5313–5316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Klauck, F.J.R.; Jame, M.J.; Glorius, F. Deaminative strategy for the visible-light-mediated generation of alkyl radicals. Angew. Chem. Int. Ed. 2017, 56, 12336–12339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Hu, J.P.; Wang, G.Q.; Li, S.H.; Shi, Z.Z. Selective C−N borylation of alkyl amines promoted by lewis base. Angew. Chem. Int. Ed. 2018, 57, 15227–15232. [Google Scholar] [CrossRef] [PubMed]
  64. Ji, L.S.; Qiao, J.M.; Li, A.K.; Jiang, Z.Y.; Lu, K.; Zhao, X. Metal-free synthesis of unsymmetrical selenides from pyridinium salts and diselenides catalysed by visible-light. Tetrahedron Lett. 2021, 72, 153071. [Google Scholar] [CrossRef]
  65. Ding, Y.; Zhang, W.; Li, H.; Meng, Y.; Zhang, T.; Chen, Q.Y.; Zhu, C. Metal-free synthesis of ketones by visible-light induced aerobic oxidative radical addition of aryl hydrazines to alkenes. Green Chem. 2017, 19, 2941–2944. [Google Scholar] [CrossRef]
  66. Ding, Y.; Li, H.; Meng, Y.; Zhang, T.; Li, J.; Chen, Q.Y.; Zhu, C. Direct synthesis of hydrazones by visible-light mediated aerobic oxidative cleavage of the C[double bond, length as m-dash]C bond. Org. Chem. Front. 2017, 4, 1611–1614. [Google Scholar] [CrossRef]
  67. Taniguchi, T.; Naka, T.; Imoto, M.; Takeda, M.; Nakai, T.; Mihara, M.; Mizuno, T.; Nomoto, A.; Ogawa, A. Transition-metal-free and oxidant-free cross-coupling of arylhydrazines with disulfides: Base-promoted synthesis of unsymmetrical aryl sulfides. J. Org. Chem. 2017, 82, 6647–6655. [Google Scholar] [CrossRef]
  68. Taniguchi, T.; Murata, A.; Takeda, M.; Mizuno, T.; Nomoto, A.; Ogawa, A. Atom-economical synthesis of unsymmetrical diaryl selenides from arylhydrazines and diaryl diselenides. Eur. J. Org. Chem. 2017, 2017, 4928–4934. [Google Scholar] [CrossRef]
  69. Hosseini-Sarvari, M.; Jafari, F.; Mohajeri, A.; Hassani, N. Cu2O/TiO2 nanoparticles as visible-light photocatalysts concerning C(sp2)−P bond formation. Catal. Sci. Technol. 2018, 8, 4044–4051. [Google Scholar] [CrossRef]
  70. Li, R.; Shi, T.; Chen, X.L.; Lv, Q.Y.; Zhang, Y.L.; Peng, Y.Y.; Qu, L.B.; Yu, B. Visible-light-promoted organic dye-catalyzed sulfidation and phosphorylation of arylhydrazines toward aromatic sulfides and diarylphosphoryl hydrazides. N. J. Chem. 2019, 43, 13642–13646. [Google Scholar] [CrossRef]
  71. Hu, H.; Chen, X.; Sun, K.; Wang, J.; Liu, Y.; Liu, H.; Yu, B.; Sun, Y.; Qu, L.; Zhao, Y. Silver-catalyzed decarboxylative cascade radical cyclization of tert-carboxylic acids and o-(allyloxy)arylaldehydes towards chroman-4-one derivatives. Org. Chem. Front. 2018, 5, 2925–2929. [Google Scholar] [CrossRef]
  72. Kobiki, Y.; Kawaguchi, S.; Ogawa, A. Highly regioselective hydroselenation of inactivated terminal alkynes using diselenide-Ph2P(O)H mixed systems under visible-light irradiation. Tetrahedron Lett. 2013, 54, 5453–5456. [Google Scholar] [CrossRef]
  73. Chen, H.; Ding, R.; Tang, H.; Pan, Y.; Xu, Y.; Chen, Y. Simultaneous construction of C−Se and C−S bonds via the visible-light-mediated multicomponent cascade reaction of diselenides, alkynes, and SO2. Chem. Asian J. 2019, 14, 3264–3268. [Google Scholar] [CrossRef] [PubMed]
  74. Ogawa, A.; Doi, M.; Ogawa, I.; Hirao, T. Highly Selective Three-Component Coupling of Ethyl Propiolate, Alkenes, and Diphenyl Diselenide under Visible-Light Irradiation. Angew. Chem. Int. Ed. 1999, 38, 2027–2029. [Google Scholar] [CrossRef]
  75. Ogawa, A.; Ogawa, I.; Sonoda, N. A Novel Three-Component Coupling of Alkynes, Vinylcyclopropanes, and Diphenyl Diselenide under Visible-Light Irradiation. J. Org. Chem. 2000, 65, 7682–7685. [Google Scholar] [CrossRef]
  76. Tsuchii, K.; Doi, M.; Hirao, T.; Ogawa, A. Highly selective sequential addition and cyclization reactions involving diphenyl diselenide, an alkyne, and alkenes under visible-light irradiation. Angew. Chem. Int. Ed. 2003, 42, 3490–3493. [Google Scholar] [CrossRef]
  77. Chen, J.; Chen, R.; Mei, L.; Yan, S.; Wu, Y.; Li, Q.; Yuan, B. Visible-light-induced difunctionalization of styrenes: An efficient and green protocol for the synthesis of β-acyloxyselenides. Asian J. Org. Chem. 2020, 9, 181–184. [Google Scholar] [CrossRef]
  78. Srivastava, V.; Singh, P.K.; Singh, P.P. Visible-light photoredox catalysed amidation of carboxylic acids with amines. Tetrahedron Lett. 2019, 60, 40–43. [Google Scholar] [CrossRef]
  79. Luly, J.R.; Yi, N.; Soderquist, J.; Stein, H.; Cohen, J.; Perun, T.J.; Plattner, J.J. New inhibitors of human renin that contain novel Leu-Val replacements. J. Med. Chem. 1987, 30, 1609–1616. [Google Scholar] [CrossRef]
  80. Alvarez-Ibarra, C.; Cuervo-Rodrìguez, R.; Fernìnedz-Monreal, M.C.; Ruiz, M.P. Synthesis, configurational assignment and conformational analysis of beta-hydroxy sulfoxides, bioisosteres of oxisuran metabolites, and their o-methyl derivatives. J. Org. Chem. 1994, 59, 7284–7291. [Google Scholar] [CrossRef]
  81. Corey, E.J.; Clark, D.A.; Goto, G.; Marfat, A.; Mioskowski, C.; Samuelsson, B.; Hammarstrom, S. Stereospecific total synthesis of a "slow reacting substance" of anaphylaxis, leukotriene C-1. J. Am. Chem. Soc. 1980, 102, 1436–1439. [Google Scholar] [CrossRef]
  82. Ruan, H.; Meng, L.G.; Zhu, L.; Wang, L. Visible-light-induced hydroxysulfurization and alkoxysulfurization of styrenes in the absence of photocatalyst: Synthesis of β-Hydroxysulfides and β-alkoxysulfides. Adv. Synth. Catal. 2019, 361, 3217–3222. [Google Scholar] [CrossRef]
  83. Wang, H.; Gao, X.; Lv, Z.; Abdelilah, T.; Lei, A. Recent advances in oxidative R1-H/R2-H cross-coupling with hydrogen evolution via photo-electrochemistry. Chem. Rev. 2019, 119, 6769–6787. [Google Scholar] [CrossRef]
  84. 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] [PubMed]
  85. Ye, R.; Ruan, H.; Xu, H.; Li, Z.; Meng, L.G.; Wang, L. Amino-assisted synthesis of alkynylthioethers via a visible-light-induced C(sp)–SII coupling between bromoalkynes and 2,2′-diaminodiaryldisulfides. Org. Chem. Front. 2021, 8, 5345–5351. [Google Scholar] [CrossRef]
Figure 1. Summary of recent synthetic strategies for disulfides and diselenides.
Figure 1. Summary of recent synthetic strategies for disulfides and diselenides.
Molecules 28 01998 g001
Scheme 1. Visible-light-promoted synthesis of β-selenyl O-heterocycles and N-heterocycles.
Scheme 1. Visible-light-promoted synthesis of β-selenyl O-heterocycles and N-heterocycles.
Molecules 28 01998 sch001
Scheme 2. Visible-light-promoted synthesis of spiro [4,5] trienones.
Scheme 2. Visible-light-promoted synthesis of spiro [4,5] trienones.
Molecules 28 01998 sch002
Scheme 3. Visible-light-induced cascade reaction of alkyne-tethered cyclohexadienones and diselenides.
Scheme 3. Visible-light-induced cascade reaction of alkyne-tethered cyclohexadienones and diselenides.
Molecules 28 01998 sch003
Scheme 4. Ru (II) catalyzed, photochemical synthesis of β-organoselenyl cyclopentanones.
Scheme 4. Ru (II) catalyzed, photochemical synthesis of β-organoselenyl cyclopentanones.
Molecules 28 01998 sch004
Scheme 5. Visible-light-induced cross-coupling cyclization.
Scheme 5. Visible-light-induced cross-coupling cyclization.
Molecules 28 01998 sch005
Scheme 6. Spirocyclization of homologated-ynones.
Scheme 6. Spirocyclization of homologated-ynones.
Molecules 28 01998 sch006
Scheme 7. Visible-light-promoted synthesis of 3-selenospiroindolenines.
Scheme 7. Visible-light-promoted synthesis of 3-selenospiroindolenines.
Molecules 28 01998 sch007
Scheme 8. Visible-light-promoted synthesis of seleno-containing pyrrolidines.
Scheme 8. Visible-light-promoted synthesis of seleno-containing pyrrolidines.
Molecules 28 01998 sch008
Scheme 9. Photoinduced cyclizations of o-diisocyanoarenes with diselenides.
Scheme 9. Photoinduced cyclizations of o-diisocyanoarenes with diselenides.
Molecules 28 01998 sch009
Scheme 10. Photoinduced regio- and chemoselective radical cyclization of unactivated 1,6-enynes with selenides.
Scheme 10. Photoinduced regio- and chemoselective radical cyclization of unactivated 1,6-enynes with selenides.
Molecules 28 01998 sch010
Scheme 11. Synthesis of benzo heterocycles.
Scheme 11. Synthesis of benzo heterocycles.
Molecules 28 01998 sch011
Scheme 12. Synthesis of benzofurans.
Scheme 12. Synthesis of benzofurans.
Molecules 28 01998 sch012
Scheme 13. (NH4)S2O8 mediated photoinduced selenylation coumarins.
Scheme 13. (NH4)S2O8 mediated photoinduced selenylation coumarins.
Molecules 28 01998 sch013
Scheme 14. RB catalyzed photochemical selenylation of indoles and (hetero)arenes.
Scheme 14. RB catalyzed photochemical selenylation of indoles and (hetero)arenes.
Molecules 28 01998 sch014
Scheme 15. Blue LEDs induced C(sp2)−H bond selenylation of indoles and activated arenes.
Scheme 15. Blue LEDs induced C(sp2)−H bond selenylation of indoles and activated arenes.
Molecules 28 01998 sch015
Scheme 16. Photoinduced metal-free α-selenylation of ketones.
Scheme 16. Photoinduced metal-free α-selenylation of ketones.
Molecules 28 01998 sch016
Scheme 17. Visible-light-mediated C(sp2)−H selenylation.
Scheme 17. Visible-light-mediated C(sp2)−H selenylation.
Molecules 28 01998 sch017
Scheme 18. Visible-light-induced direct thiolation of ether with disulfides at α-C(sp3)−H.
Scheme 18. Visible-light-induced direct thiolation of ether with disulfides at α-C(sp3)−H.
Molecules 28 01998 sch018
Scheme 19. Light mediated C(sp2)−H bond chalcogenation (S, Se,) of indoles.
Scheme 19. Light mediated C(sp2)−H bond chalcogenation (S, Se,) of indoles.
Molecules 28 01998 sch019
Scheme 20. Direct thiolation of allyl C(sp3)−H.
Scheme 20. Direct thiolation of allyl C(sp3)−H.
Molecules 28 01998 sch020
Scheme 21. Site-selective thiolation of (multi)halogenated heteroarenes.
Scheme 21. Site-selective thiolation of (multi)halogenated heteroarenes.
Molecules 28 01998 sch021
Scheme 22. Visible-light photoredox synthesis of chiral α-selenoamino acids.
Scheme 22. Visible-light photoredox synthesis of chiral α-selenoamino acids.
Molecules 28 01998 sch022
Scheme 23. Photochemical C(sp3)−Se bond formation through C−C bond cleavage of cyclobutanone oxime esters.
Scheme 23. Photochemical C(sp3)−Se bond formation through C−C bond cleavage of cyclobutanone oxime esters.
Molecules 28 01998 sch023
Scheme 24. Blue LEDs mediated seleno-substitutions of azosulfone group.
Scheme 24. Blue LEDs mediated seleno-substitutions of azosulfone group.
Molecules 28 01998 sch024
Scheme 25. Visible-light-catalyzed synthesis of asymmetric selenides from pyridinium salts.
Scheme 25. Visible-light-catalyzed synthesis of asymmetric selenides from pyridinium salts.
Molecules 28 01998 sch025
Scheme 26. Sulfurization of aryl hydrazines.
Scheme 26. Sulfurization of aryl hydrazines.
Molecules 28 01998 sch026
Scheme 27. Photochemical regioselective hydroselenylation of terminal alkynes.
Scheme 27. Photochemical regioselective hydroselenylation of terminal alkynes.
Molecules 28 01998 sch027
Scheme 28. Visible-light-mediated cascading reactions construct C−Se and C−S bonds simultaneously.
Scheme 28. Visible-light-mediated cascading reactions construct C−Se and C−S bonds simultaneously.
Molecules 28 01998 sch028
Scheme 29. Visible-light-induced difunctionalization of styrenes with a binary system of diaryl diselenides and carboxylic acids.
Scheme 29. Visible-light-induced difunctionalization of styrenes with a binary system of diaryl diselenides and carboxylic acids.
Molecules 28 01998 sch029
Scheme 30. Synthesis of β-hydroxysulfide and β-alkoxysulfide.
Scheme 30. Synthesis of β-hydroxysulfide and β-alkoxysulfide.
Molecules 28 01998 sch030
Scheme 31. Synthesis of alkynyl sulfides.
Scheme 31. Synthesis of alkynyl sulfides.
Molecules 28 01998 sch031
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

Wang, C.; Zhang, Y.; Sun, K.; Yu, T.; Liu, F.; Wang, X. Synthesis and Application Dichalcogenides as Radical Reagents with Photochemical Technology. Molecules 2023, 28, 1998. https://doi.org/10.3390/molecules28041998

AMA Style

Wang C, Zhang Y, Sun K, Yu T, Liu F, Wang X. Synthesis and Application Dichalcogenides as Radical Reagents with Photochemical Technology. Molecules. 2023; 28(4):1998. https://doi.org/10.3390/molecules28041998

Chicago/Turabian Style

Wang, Cairong, Yan Zhang, Kai Sun, Tingting Yu, Fei Liu, and Xin Wang. 2023. "Synthesis and Application Dichalcogenides as Radical Reagents with Photochemical Technology" Molecules 28, no. 4: 1998. https://doi.org/10.3390/molecules28041998

Article Metrics

Back to TopTop