Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Jan 17;123(4):1364–1416. doi: 10.1021/acs.chemrev.2c00591

Recent Progress in Cyclic Aryliodonium Chemistry: Syntheses and Applications

Xiaopeng Peng †,, Abdur Rahim §, Weijie Peng , Feng Jiang , Zhenhua Gu §,*, Shijun Wen ‡,*
PMCID: PMC9951228  PMID: 36649301

Abstract

graphic file with name cr2c00591_0061.jpg

Hypervalent aryliodoumiums are intensively investigated as arylating agents. They are excellent surrogates to aryl halides, and moreover they exhibit better reactivity, which allows the corresponding arylation reactions to be performed under mild conditions. In the past decades, acyclic aryliodoniums are widely explored as arylation agents. However, the unmet need for acyclic aryliodoniums is the improvement of their notoriously low reaction economy because the coproduced aryl iodides during the arylation are often wasted. Cyclic aryliodoniums have their intrinsic advantage in terms of reaction economy, and they have started to receive considerable attention due to their valuable synthetic applications to initiate cascade reactions, which can enable the construction of complex structures, including polycycles with potential pharmaceutical and functional properties. Here, we are summarizing the recent advances made in the research field of cyclic aryliodoniums, including the nascent design of aryliodonium species and their synthetic applications. First, the general preparation of typical diphenyl iodoniums is described, followed by the construction of heterocyclic iodoniums and monoaryl iodoniums. Then, the initiated arylations coupled with subsequent domino reactions are summarized to construct polycycles. Meanwhile, the advances in cyclic aryliodoniums for building biaryls including axial atropisomers are discussed in a systematic manner. Finally, a very recent advance of cyclic aryliodoniums employed as halogen-bonding organocatalysts is described.

1. Introduction

Hypervalent iodine(III) compounds, due to their environmentally benign and highly electron-deficient nature, have been widely studied and utilized as powerful chemicals.16 Hypervalent iodine compounds are reported to display similar structural properties and reactivity as transition metals, such as Ag(I), Hg(II), and Tl(III), while they are usually low toxic, bench stable, and easily handled. Thus, they may serve as efficient alternative chemicals to avoid toxic heavy metal oxidants and expensive organometallic catalysts to initiate organic transformations.711

Over the past few decades, diverse hypervalent iodine(III) compounds (λ3-iodanes) have been discovered and prepared, and their representative structures are shown in Scheme 1.1214 These trivalent iodine scaffolds are often shown in a trigonal bipyramidal geometry with two hybridized lone pairs of the iodine in equatorial positions.5,12,15,16 However, increasing evidence implies that these two lone pairs are almost pure s-type and pure p-type (Scheme 1A).1721 Both s-type and p-type lone pairs are more stabilized for five-membered diphenyl trivalent iodine center (iodonium) compared to acyclic diphenyl iodonium. The lowering of the energies of the lone pairs may result from the compressed C–I–C angle in the cyclic ring system.20 The aryliodoniums are found to have an approximately T-shaped structure, as shown in Scheme 1A. Compared to a regular covalent bond, the I–Z bond (Z = halogen, O, N, C) is longer, weaker, and more polarized to be ionic bond alike.5,22,23 Such unique structural features may explain the high reactivity of aryliodonium compounds. Additionally, the newly formed aryl iodides have higher stability than aryliodoniums to become another thermodynamic driving force for aryliodonium-mediated oxidative transformations.24 Now, hypervalent aryliodoniums have attracted more and more research interest due to their synthetic merits as arylating agents to react with a variety of nucleophiles.12,16

Scheme 1. Main Classes of Trivalent Iodine Compounds.

Scheme 1

Since the discovery of iodonium salts in the 1890s,12,25 aryliodoniums as important organic synthons have been successfully applied to generate various structural scaffolds.2633 Meanwhile, diaryliodoniums display some biological activity.3440 For example, they are reported to act as NADPH oxidase inhibitors due to their potential production of aryl radicals to form covalent bonds with NADPH oxidase to block the enzymatic function.41 Structurally, aryliodonium salts can be classified into two categories: acyclic and cyclic aryliodoniums. Because of their high reactivity, air/moisture stability, and ready availability, acyclic aryliodoniums have been extensively utilized as efficient arylation reagents in organic synthesis.27,4253 For example, cross-coupling reactions involving acyclic aryliodoniums with carbonyl moieties,54 alkynes,55 alkenes,56,57 and hetero atoms to construct various functionalized molecules5860 are realized. Compared to acyclic aryliodoniums, cyclic counterparts are less explored, although they could be employed for further transformations to set up a cascade reaction to achieve reaction economy. Normally, most cyclic aryliodoniums, especially five-membered ones, are less reactive than acyclic aryliodoniums. The rigid geometry of cyclic aryliodoniums is claimed to hinder the Y-shape conformation which has higher energy status to allow the reductive elimination, while the dominant and lower energy T-shape conformation prevents the reductive elimination to complete the arylation process (Scheme 2A).61 Acyclic diaryliodoniums may undergo “T–Y” isomerization easily to be vulnerable to the following reductive elimination. For a similar reason, the five-membered cyclic aryliodoniums are less reactive than big-sized cyclic counterparts that tend to have the property of acyclic ones (Scheme 2B). In a more recently published paper, Legault and co-workers have computationally calculated the free energy to activate the reductive elimination for cyclic and acyclic aryliodoniums.62 For example, activation barriers for five- and six-membered diphenyliodoniums are 139 and 123 kJ/mol, respectively, when bromide is used as the nucleophile. The acyclic diphenyleneiodonium has less free energy barrier with 112 kJ/mol (Scheme 2A,B).

Scheme 2. Conformation-Dependent Reductive Elimination and Arylation with Aryliodoniums.

Scheme 2

Although acyclic aryliodoniums have made significant advancements as arylating agents to assemble different compounds, the concern associated with their arylation process is the iodoarenes that are produced along with the desired arylated products and are often discarded as waste, which is problematic and should be addressed in the context of atom economy (Scheme 2C).1,63 Compared to acyclic aryliodoniums, cyclic aryliodoniums are more atom-economical because the generated iodoarene remains as one part of the arylated products, which have further elaborating potential (Scheme 2D).31,64,65

Owing to the highly electron-deficient property and unique ring system, cyclic aryliodoniums can be naturally considered to build various polycyclic scaffolds.32,6671 Such cyclic aryliodonium associated transformations have received increasing interest and made significant advancements. So far, there are a number of reviews on hypervalent iodine(III).1,2,5,12,15,7275 However, acyclic aryliodoniums and their applications are more focused,12,72,73 for example, Olofsson et al. reviewed synthetic developments and applications of acyclic diaryliodonium salts.12 Han and co-workers recently provided a systematic summary of structures and synthetic strategies of acyclic aryliodoniums.72 In addition, the authors also briefly illustrate the preparation methods of cyclic aryliodoniums in a small manner. More recently, Fu et al. reviewed the progress of synthetic methods toward diarylioniums and applications, in which cyclic aryliodoniums are briefly mentioned.73 Due to the emergence of cyclic iodoniums in the past decade, some reviews have been published on cyclic aryliodoniums. Goswami et al. have summarized the advancements of cyclic diaryliodoniums made before 2017 and are focused on their applications for construction of polycycle compounds.74 Jiang et al. have reviewed the progress of various atom-economical applications of cyclic diaryliodoniums to synthesize functionalized molecules.75 There are other reviews on hypervalent iodine(III) published before 2017.1,2,5,15

In view of the fact that cyclic aryliodonium chemistry is a rapidly evolving field and the vast achievements that have been made in the last 5 years, it is of high importance to summarize the new progress in this fast-growing research area from different perspectives. Therefore, our current review intends to provide a comprehensive and in-depth illustration of the recent progress (2017–2022) in new synthetic routes to cyclic aryliodoniums and their further applications to construct diverse arylation products, including our recent work from the Wen group and the Gu group. Moreover, the very recent development of cyclic aryliodoniums as organocatalysts, for example, halogen-bonding donors, is also included. In the past several decades, benziodoxol(on)es as shown in Scheme 1B, another unique type of cyclic aryliodoniums have been investigated extensively as functional-group transfer reagents for alkynylation,7679 azidation,80,81 and trifluoromethylation (Togni’s reagents).8287 The research focused on benziodoxol(on)es as functional-group transfer reagents has been reviewed by Togni,15 Nachtsheim,24 and more recently by Patel and co-workers.88 Therefore, this type of cyclic aryliodonium will not be discussed in further detail here. Furthermore, this review aims to inspire chemical researchers to take the benefits of utilizing hypervalent iodines as environmentally reagents to accomplish the construction of more complicated compounds and to realize more diverse reaction transformations.

2. Advance in Design and Synthesis of Cyclic Aryliodoniums

2.1. Synthesis of Cyclic Diaryliodoniums

The past decade has witnessed significant advancements in the chemistry of cyclic aryliodoniums, and they have been frequently utilized in the construction of complex molecules through effective multiple functionalizations, including fluorenes,89 carbazoles,26 and phenanthrenes.90 Cyclic diaryliodoniums are more recently investigated for their enantioselective ring-opening reactions to prepare chiral biaryl atropisomers that can provide novel chiral ligands for organic synthesis. The preparation of cyclic arylidooniums is well established, in which the conventional synthetic routes to cyclic aryliodoniums rely on stepwise procedures,91 including the Suzuki-coupling reaction to provide 2-amino substituted biaryls 3 and the following Sandmeyer-type iodination to construct 2-iodobiaryls 4. The final intramolecular oxidative cyclization assembles the cyclic diaryliodoniums 5 (Scheme 3).

Scheme 3. Conventional Approach for the Synthesis of Cyclic Diaryliodoniums.

Scheme 3

Although the conventional strategies have enabled the facile production of diverse types of cyclic diaryliodoniums, the procedures are reagents-wasting and time-consuming and require multiple steps. Recently, a number of effective synthetic routes are reported with a wide scope in a one-pot Friedel–Crafts alkylation/oxidation sequence or through green electrochemical strategies.

In 2021, the groups of Moran and He have respectively reported their electrochemical preparation of cyclic five-membered diaryliodoniums 5 (Scheme 4), which is an environmentally friendly approach.49,92 In Moran’s method, cyclic aryliodoniums 5 are obtained in good to excellent yields via anodic oxidation of 2-iodobiaryls 4 by constant current electrolysis in a mixture of HFIP/MeCN/TfOH, without any electrolyte salts. A variety of cyclic diaryliodoniums 5 with different substitutions are prepared. Moreover, the practicality and ready scalability of this method are validated by large-scale experiments. However, the diversity of iodoniums is limited to diphenyl iodoniums, while the synthesis of heterocyclic diaryliodonium 5o fails. He’s approach not only prepares cyclic diaryliodoniums (5p5r) and benziodoxolones 6a6b but also acyclic aryliodine(III) reagents 6c, including iodosylarenes. These environmentally friendly strategies are complementary to conventional methods to prepare cyclic diaryliodoniums and avoid the excessive use of toxic oxidant chemical reagents. It should be noted that substituted 2-iodobiphenyls 4 are not usually commercially available and have to be synthesized by multistep reactions, the same as the procedure shown in Scheme 3.

Scheme 4. Electrochemical Synthesis of Cyclic Aryliodoniums.

Scheme 4

Six-membered aryliodoniums are not easily accessed like the aforementioned synthesis of five-membered aryliodoniums. The group of Nachtsheim93 has reported a two-step procedure for the construction of six-membered cyclic diphenyliodoniums 9 (Scheme 5). In their strategy, strong Brønsted acids including HBF4 promote the effective Friedel–Crafts alkylation with various substituted o-iodobenzylalcohols 7, giving diphenylmethane scaffolds, the key intermediates that are followed by an oxidative cyclization to provide a variety of cyclic aryliodoniums 9 in moderate to excellent yields. Notably, both primary and secondary benzyl alcohols 7 are suitable substrates. The synthetic issue is the limited commercial availability of o-iodobenzylalcohols.

Scheme 5. Concise Synthetic Route to Cyclic Six-Membered Diphenyliodoniums.

Scheme 5

Heteroatom-bridged cyclic aryliodoniums 12 are underrepresented compared to carbon-bridged counterparts 9, likely due to the low availability of efficient synthetic protocols. Recently, the group of Nachtsheim has further developed a one-pot procedure of O- and N-bridged cyclic diaryliodoniums 12 by treating substituted phenols or anilines 10 with in situ generated arynes which derive from (trimethylsilyl)aryl triflates 11 (Scheme 6).94 In this procedure, the Selectfluor is found as the most effective oxidant, and it could selectively oxidize int1 to int2. The electron-poor substrates (10b) and slightly electron-rich compounds (10c) are compatible in this approach. The more electron-rich and aryl substituted derivatives (10d, 10h) fail to provide the expected O-bridged iodoniums, revealing that the electronic property of substituents and extended π-systems have a significant effect. Similarly, the compounds containing electron-poor substituent (10j) work smoothly, and electron-rich substituent bearing substrates (10k) fail in the synthesis of N-bridged diaryliodoniums. However, the phenyl-substituted aniline is successful to afford 12l in good yield. X-ray analysis have confirmed the structures of compounds 12a and 12i.

Scheme 6. Synthetic Approach to Cyclic O- and N-Bridged Six-Membered diphenyliodoniums.

Scheme 6

2.2. Design and Synthesis of Cyclic Heterocyclic Iodoniums

Although the synthetic routes to cyclic aryliodoniums are well established, the structure and properties of such cyclic moiety have been centered on simple carbocyclic scaffolds, and their heterocycle-flanked analogues remain underexplored.95 Because heterocyclic scaffolds have attracted considerable interest in the academic research field and pharmaceutical industry, heterocyclic aryliodoniums possess a great potential to build high-valued heterocyclic frameworks. However, such heterocyclic iodoniums (HCIs) are less reported. In 2018, the Wen group developed an efficient strategy to synthesize a series of novel HCIs 14 containing heterocyclic motifs, including flavone (14ad, 14i), isoquinoline (14eh), and dibenzo-oxepin (14j, 14k) (Scheme 7).96 It is worth mentioning that the HCIs are not easily synthesized from 13′, in which the iodide is substituted on the benzene ring. Furthermore, the structures of HCIs 14 are unambiguously clarified by X-ray crystallographic analysis.

Scheme 7. Design and Synthesis of HCIs.

Scheme 7

Particularly, due to the advantages in construction of heterocyclic compounds, N-containing HCIs have received the attention of other groups. In a more recent report from the group of Nachtsheim,97 HCIs 16 flanked by pyrazole moieties are designed and synthesized from 2-pyrazole-phenyl iodides 15 using mCPBA as an oxidant (Scheme 8A). The range of substrates is extended to both electron-withdrawing (15d, 15f) and electron-donating substituents (15c, 15e) or halogen substituents (15g). Postnikov and co-workers have developed another approach to construct a wide range of imidazole-substituted cyclic iodoniums 19 in the presence of Oxone (Scheme 8B).98 In this two-step strategy, acyclic iodonium salts 17 undergo internal migratory N-arylation to provide 1-phenyl-5-iodoimidazoles 18. Finally, the oxidative cyclization generates the desired cyclic HCIs 19. Various functional groups containing electron-donating groups (18c) on aromatic rings proceed steadily. Likewise, the halogen-substituted precursors (18a,b, 18d) work well and afford the target products in moderate yields. However, the precursors bearing thiophene moiety fail the cyclization under the standard condition. In addition, X-ray crystallography studies show that fluorine-substituted iodonium salt 19a crystallizes in a dimeric structural pattern in which there are short contacts between the iodine center and two surrounding sulfate anions. The weak bonding of counterion to iodine(III) may reduce the overall electronic density around the iodine, which invites another counterion to complex, leading to the dimer formation.21 Similiarly, bromide-substituted iodonium salt 19d shows a dimeric nature and moreover forms a more complicated structure due to the incorporation of water and additional π-stacking.

Scheme 8. Design and Synthesis of Pyrazole and Imidazole Substituted Cyclic HCIs.

Scheme 8

In 2018, the same group of Postnikov reported a facile synthesis of pseudocyclic benzimidazolyl-substituted aryliodoniums 21 through the condensation of [hydroxy(tosyloxy)iodo]arenes 20 and arenes.99 The treatment of a base further converts pseudo aryliodoniums 21 into novel five-membered iodine–nitrogen heterocycles 22, which represent novel nitrogen-containing analogues of neutral diaryliodoniums (Scheme 9). The starting iodonium tosylates 20 with either electron-donating (20ac) or electron-withdrawing groups (20df) on aromatic rings are well tolerated. The following X-ray structural analysis of 21d confirms the presence of a long nitrogen-to-iodine bond.

Scheme 9. Design and Synthesis of Pseudo and Neutral HCIs.

Scheme 9

As aforementioned, the growing interest in the investigation of N-heterocycle based single cyclic HCIs have been showed in recent years. However, the bis-N-heterocycle-stabilized λ3-iodanes (BNHIs) are rarely explored.100 Recently, the group of Nachtsheim has reported their synthetic approach to the uniquely designed BNHIs (24),101 as depicted in Scheme 10. The heterocycles can be pyrazoles, triazoles, and benzoimidazoles. However, the synthesis of benzoxazole-stabilized iodanes (24i,j) fail in the similar approach. Furthermore, the X-ray analysis elucidates the solid-state structure and further indicates the strong intramolecular interactions between the pyrazole nitrogen atoms and the hypervalent iodine center. These unique cyclic iodoniums can be used as promising oxidants to achieve sulfoxidations102,103 and phenol dearomatizations.104

Scheme 10. Design and Synthesis of BNHIs.

Scheme 10

2.3. Design and Synthesis of Cyclic Vinyl(aryl)iodoniums

Cyclic vinyl(aryl)iodoniums are a class of cyclic monoaryl iodoniums that have found applications in organic syntheses and even as radical initiators.105107 Because cyclic vinyl(aryl)iodoniums are not relatively easily prepared compared to cyclic diaryliodoniums, much less attention has been paid. In 2019, Moran and co-workers reported a straightforward strategy to prepare five-membered cyclic vinyl(aryl)iodoniums 26 by oxidizing β-iodostyrenes 25,108 as described in Scheme 11A. With their reported procedure, substituted (E)-β-iodostyrenes bearing electron-donating groups (25b,c) work well to provide the desired iodoniums in moderate to good yields (57–64%). However, the presence of strong electron withdrawing groups, such as trifluoromethyl in 25d, lead to poor yield (12%). The substrates with o- or p-methoxy substituents (25e,f) are not tolerated in this system. Accordingly, the X-ray crystal analysis confirms that the cyclic vinyl(aryl)iodonium 26a exists in a dimer with two triflate anions loosely binding to the two iodine(III) centers, which is in line with the previous reports.109

Scheme 11. Synthesis of Cyclic Vinyl(aryl)iodoniums.

Scheme 11

The Wen group also recently reported a strategy to prepare vinyl(aryl)iodoniums 28 with more functionality, as shown in Scheme 11B.110 In this work, four types of vinyl(aryl)iodoniums with diverse substituents R2 in the vinyl side are obtained in good yields using the conventional procedure for the synthesis of cyclic aryliodoniums. Notably, vinyl(aryl)iodoniums 28 are stable in either powder or solution at ambient or elevated temperature. However, these special aryliodoniums will undergo a reductive ring opening in the presence of trimethylamine, and the result invites further study.

2.4. Design and Synthesis of Polycyclic Aryliodoniums

More recently, hypervalent iodine compounds containing multiple iodine(III) centers have drawn interest for applications as material precursors,111 supramolecular materials,112 and bidentate Lewis acids.113,114 Stang and co-workers have reported a series of macrocyclic aryliodoniums in which multiple iodine(III) atoms bridge the biaryl together, for example, compounds 29 and 30(115,116) (Scheme 12A). These macrocyclic aryliodoniums are more like acyclic. The hypervalent iodines embedded in polycyclic aromatic hydrocarbons are still rare, likely due to potential high positive charges and ring strain.

Scheme 12. (A) Representative Macrocyclic Aryliodoniums; (B,C) Syntheses of Cyclic Di-iodonium and Tri-iodonium.

Scheme 12

The molecular modelling of 30 is reproduced with permission from ref (116). Copyright 2003 American Chemical Society. The X-ray structure of 31-I is reproduced with permission from ref (117). Copyright 2022 Royal Society of Chemistry

In 2022, the group of Protasiewicz reported successful the synthesis of a sterically crowded tetrakis-di(acetoxy)iodoarenes 31-I from the bulky 1,4-di-iodobenzene 31. Subsequently, 31-I could facilitate transformations to a bulky bicyclic di-iodonium salt 32 via ready cyclization under mild conditions, probably because the crowdedness of complex 31-I may promote the reaction (Scheme 12B).117 Single crystal X-ray diffraction studies characterize the solid structure of hypervalent iodines 31-I and 32 and further reveal varying degrees of intramolecular I–O interactions. Notably, owing to the steric properties, the cyclic di-iodonium 32 possesses an excellent solubility compared to other hypervalent iodines.

Xu and co-workers have reported the synthesis of triphenylenotriiodonium, namely iodine-doped sumanene 34, which represents the first example of iodine-doped buckybowls.118 The synthesis of iodine-doped sumanene 34 is illustrated in Scheme 12C. First, 1,5,9-triaminotriphenylene 33 as the starting material undergoes a classic Sandmeyer reaction procedure to give 1,5,9-triiodotriphenylene 33-I in modest yield. Subsequently, treatment of 33-I with mCPBA and triflic acid in CH2Cl2 leads to the final compound 34, while other strong acids such as H2SO4, HNTf2, and CF3CO2H could not afford the tricyclic products. The triiodonium compound could be prepared on a gram scale. This polycyclic aryliodonium possesses a planar geometry as predicted by DFT calculation.

2.5. Design and Synthesis of Chiral Cyclic Aryliodoniums

Chiral hypervalent iodine reagents have attracted attention in asymmetric reactions,119 either being directly employed as chiral oxidants or participating in an array of chiral transformations.120,121 The chiral hypervalent iodine-catalyzed process is considered as a fundamental tool to obtain important enantio-enriched molecules and synthetic intermediates with wide applications in pharmaceutical and medicinal chemistry.122 Chiral hypervalent iodine chemistry has been steadily increasing interest in recent years, especially the acyclic aryliodoniums as potential chiral reagents in asymmetric synthesis over the last decades.123 Among them, binaphthyl-based chiral iodoniums 36 have interesting properties as a result of the naphthyl moieties restricting the axial rotation (Scheme 13A).124126 Several approaches to binaphthyl-based chiral acyclic aryliodoniums are disclosed in the past years. Recently, Wirth et al. have prepared novel binaphthyl-based chiral acyclic iodonium reagents 40 from λ3-iodanes and BINOL (R)-analogues 38,125 as depicted in Scheme 13B. The cyclic aryliodoniums are less developed in the past years. Ochiai and co-workers first reported the synthesis and characterization of chiral cyclic iodonium 37, however its chirality may be difficult to maintain due to high ring strain (Scheme 13A).127 More recently, Yoshida and co-worker128 have designed and prepared another series of chiral binaphthyl-based iodoniums 43, as shown in Scheme 13C. These newly chiral cyclic aryliodoniums 43 are synthesized via a three-step procedure with excellent yield and high ee values. Furthermore, the binaphthyl-based iodonium salt 43 can be efficiently utilized as halogen-bonding donors (discussed later).

Scheme 13. Synthesis of New Chiral Iodine(III) Compounds.

Scheme 13

Taken together, while cyclic aryliodoniums become increasingly important to deliver essential functionalization compounds, the novel design and efficient synthesis of a broad range of cyclic aryliodoniums are highly demanded. Up to date, numerous synthetic approaches have been developed to prepare diverse cyclic aryliodonium species in easily handling, mild, and environmentally friendly manners. In recent years, the emergence of nascent diaryliodonium salts has been witnessed more frequently. In addition, polycyclic aryliodoniums with multiple iodine centers are interesting species to discover, however, their development is scarcely reported, likely due to rigid and steric features. Furthermore, it is worth mentioning that chiral cyclic diaryliodonium mediated catalysis possess a promising future in asymmetric induction and regiocontrol procedures, such as asymmetric oxidative dearomatization reactions. However, the development of chiral cyclic aryliodoniums based on different structures and types of chirality is just fledging, and their synthesis still remains challenging. Collectively, the emergence of novel and intriguing cyclic aryliodoniums will allow more seminal reaction transformations to broaden synthetic applications of aryliodoniums. Therefore, it is desirable to discover new strategies to design and synthesize these special cyclic aryliodonium species.

3. Synthetic Applications of Cyclic Aryliodoniums

Like acyclic aryliodoniums, cyclic aryliodoniums can serve as effective arylating reagents toward a wide range of nucleophiles to generate arylated compounds under the catalysis of metal catalysts or metal-free conditions. Moreover, the iodoarenes that remain in the products after the first arylation with cyclic aryliodoniums would undergo a second potential arylation to set up cascade reactions for further transformations.

In the past decade, a number of synthetic transformations and applications utilizing cyclic aryliodoniums have been reported.65,90,129137 The domino arylation procedures are utilized to access more complicated polycycles through the construction of C–C and C–heteroatom bonds. In the past five years, significant advancements have been also made in enantioselective ring-opening reactions of cyclic aryliodoniums with nucleophiles to construct axially chiral atropisomers, which represent an important motif in natural products,138143 biologically active compounds,144162 and chiral ligands.163181 Based on the currently available literature, we summarize a recent progress on the use of cyclic aryl iodoniums as electrophilic aryl transfer reagents for the bond formations of C–N, C–O, C–C, and C–other atom (including S, Se, Si, Te) to construct polycyclic compounds and facilitate the synthesis of axially chiral atropisomers. The synthetically appealing protocols for the arylation of various nucleophiles with cyclic aryliodoniums have been well developed. Selected arylations and their associated approaches to synthesize diverse compounds are summarized in Scheme 14, mediated by metal or metal-free systems.

Scheme 14. Selected Arylation Approaches to Construct Diverse Compounds.

Scheme 14

3.1. Construction of Polycyclic Arenes via Arylation/Cyclization

3.1.1. N-Arylation/Cyclization to Assemble Polycyclic Arenes

Nitrogen-containing nucleophiles are perfect arylation acceptors to react with cyclic aryliodoniums. As a result, the C–N bond is formed during the first arylation. The following intramolecular second N-arylation accomplishes the cyclization to build nitrogen containing polycycles.26,132 The Wen group has recently developed a procedure to access a variety of N-carbonyl acridanes 46 via a dual Buchwald-coupling cascade process by employing six-membered cyclic diaryliodoniums 44 and readily available nitriles 45 (Scheme 15).182 It is worth mentioning that iodine-bearing diarylmethane amide might be involved in this transformation. Various nitriles 45 bearing aryl (45af), heterocyclic groups (45g,h) and alkyl (45i,j) are compatible and successfully assemble the corresponding N-carbonyl acridanes in modest to good yields. Importantly, acridane 46k potently inhibits tubulin polymerization with an IC50 value of 1.6 μM, comparable to the positive control drug CA-4 (1.4 μM) and better than colchicine (7.0 μM), as shown in Scheme 15A–C. It is noteworthy that five-membered cyclic diaryliodoniums could not engage such transformation, implying the relatively poor reactivity of five-membered species, which is line with Legault’s energy calculation.62 Furthermore, the study also indicates the great potential of cyclic diaryliodoniums in building drug-like compounds with polycyclic scaffolds.

Scheme 15. Dual N-Arylation of Nitriles with Cyclic Diaryliodoniums to Construct Acridanes and the Biological Activity of Acridane 46k; (A–C) The Inhibition of Tubulin Polymerization by Compound 46k, CA-4, and Colchicine.

Scheme 15

In 2019, Zhang and co-workers reported a practical protocol for the construction of diversely functionalized tribenzo[b,d,f]azepines 49 via palladium-catalyzed cascade N-arylation and π-extended decarboxylative annulation between 2-aminobenzoic acids 47 and cyclic diaryliodoniums 48 under high temperature,30 as shown in Scheme 16. Regardless of the electronic properties of substituents in the 2-aminobenzoic acids, the desired products are generated in fair yields. Notably, the heterocyclic acids, including 2-aminopyridinyl carboxylic acids 47c, are also effective in this transformation. In addition, symmetric cyclic diaryliodoniums substituted with various groups work smoothly to provide corresponding products (49df), while strong electron-withdrawing groups (e.g., −CF3, 48f) were deleterious for the yields. Furthermore, the mechanistic study indicates that the carboxylic acid group in the benzoic acids may play a key role in the initiation process.

Scheme 16. Cascade Decarboxylative Annulation Involving Cyclic Diaryliodoniums.

Scheme 16

Indole-containing scaffolds represent an important class of bioactive natural products,183187 drugs,188193 and functional materials.194,195 The Wen group has reported the first example of palladium-catalyzed assembly of benzocarbazoles 52 from simple indoles and cyclic diaryliodoniums 50 via dual C–H activation of indoles at positions C2 and C3 (Scheme 17A).196 However, the double annulation at positions N1 and C2 products of indoles cannot be obtained with this protocol. Zhang et al. have realized the preparation of various N1,C2-indole-fused compounds 54 (e.g., phenanthridines) by coupling indole-2-carboxylic acids 53 with cyclic diaryliodoniums 50 in the presence of palladium species,70 as depicted in Scheme 17B. In this efficient decarboxylative annulation procedure, highly regioselective N1/C2 annulated products can be prepared in modest to excellent yields, complementary to the previous report of C2/C3 annulation.196 The mechanism studies confirm that the carboxylic acid as a traceless directing group plays a key role in determining the reaction regioselectivity. Electron-donating (53a,d) or electron-withdrawing groups (53b,c) and heterocyclic substituents of indoles (53e,f) proceed smoothly to generate the corresponding cyclized products 54.

Scheme 17. Pd-Catalyzed Regioselective Annulation with Indoles.

Scheme 17

As mentioned before, the Wen group has reported a synthetic method of cyclic vinyl monoaryl iodoniums.110 These unique monoaryl iodoniums 55 can react with various nucleophiles 56, including anilines, phenols, aromatic acids, and indoles to initiate the first arylation under metal-free conditions (Scheme 18). It is worth mentioning that the C–C double bond unexpectedly migrates to the adjacent methyl (57-I) and the carbene species is conceptualized as a key intermediate from mechanistic studies. The presence of the acetylene group and boronic acid in the nucleophiles is well tolerated. In addition, the simple indole gives the C3-arylated product. Other alkylamines except benzylamine are not tolerated in this transformation, probably due to their strong basicity. Furthermore, the following transformation of the arylated products 57-I using diversity-oriented synthesis strategy finally builds a chemical drug library of diverse heterocyclic fragments 57 that are in demand in the drug discovery field.197215

Scheme 18. Metal-Free Arylation with Cyclic Vinyl Monoaryliodoniums to Construct Polycyclic Arenes via Further Transformations.

Scheme 18

3.1.2. O-Arylation/Cyclization to Assemble Polycyclic Arenes

Structurally complex oxygen-incorporated polycyclic frameworks are important motifs in biologically active molecules216223 and organic materials,224227 and the efficient synthesis of oxygen-incorporated polycyclic arenes has received considerable attention from the research community.

The group of Zhang has presented sequential O-arylation and alkynylation of cyclic diaryliodoniums 58 with anthranilic acids and alkynes on the basis of copper catalysis (Scheme 19).69 Through this three-component approach, a number of alkyne substituted biaryls 59 are efficiently produced. Notably, a range of commercially available anthranilic acids (e.g., alkyl, alkenyl, and aryl halides) are well tolerated in this cascade and give the desired biaryls 59af in modest to excellent yields (55–90%). The heterocyclic acids also undergo this transformation successfully under the optimal conditions (59f). Furthermore, azidation of 59 with trimethylsilyl azide and the subsequent intramolecular [3 + 2] Click cycloaddition provides diverse novel ten-membered poly(hetero)aryl containing compounds 60, which are essential motifs in G-quadruplex binders228231 and telomerase inhibitors.232234

Scheme 19. O-Arylation/C-Alkynylation of Anthranilic Acids to Build Large Cyclic Molecules.

Scheme 19

The Wen group has pioneered the synthesis of diverse oxygen-incorporated polycycles 62 and 63 for the first time through dual O-arylation of both HCIs 61 and typical medium-ring sized cyclic diaryliodoniums 61′ with environmentally benign water as the oxygen source (Scheme 20).96,235 Cheap copper acetate is employed as an effective catalyst in these procedures. The obtained heteropolycycles 62 incorporate diverse fragments that may be essential in biologically active derivatives236245 and natural products.237,241,246249 Particularly, the pyridine-fused HCIs (61ho) and sulfur-containing 61q also react with water smoothly to provide desired products in moderate yield. Further elaboration of 62h is performed to access more complex heterocyclic compounds (64, 65), demonstrating the practical applicability of these HCIs. Similarly, five- and six-membered diphenyliodoniums 61′ are compatible with this transformation to provide expected products 63, while seven-membered and bigger-sized aryliodoniums fail (Scheme 20). Liu et al. has also reported a similar strategy to construct diverse dibenzofuran derivatives 63 by reacting five-membered cyclic diaryliodoniums in water.68 In their publication, several organic semiconducting material scaffolds have been facilitated with moderate yields by further transformations.

Scheme 20. Oxygenation of HCIs or Diphenyliodoniums with Water to Construct Oxygen-Bridged Polycycles.

Scheme 20

3.1.3. C-Arylation/Cyclization to Assemble Polycyclic Arenes

Polycyclic aromatic hydrocarbons (PAHs) have attracted much attention due to their novel optoelectronic, self-assembly, and charge-transport properties resulting from their unique π-conjugated scaffold.250256 Therefore, synthetic strategies to produce PAHs have been extensively explored. In 2015, the Wen group disclosed palladium-catalyzed insertion of internal alkynes into cyclic diaryliodoniums via [4 + 2] benzannulation to generate various phenanthrenes,90 one of the vital molecular structural features in functional materials.257 Recently, a new synthetic approach to dibenzopleiadiene-embedded PAHs 68 through a palladium-catalyzed cascade intermolecular decarbonylation and intramolecular decarboxylative annulation between commercially available naphthalene anhydride 66 and cyclic diaryliodoniums 67 has been described by Leowanawat et al. (Scheme 21).258 The protocol also works well with HCIs and provides heterocycle-containing PAHs, for example, 68g, smoothly. An unsymmetrical and hindered cyclic diaryliodonium successfully transformed to the corresponding dibenzopleiadiene 68h with relatively poor yield, and its crystal structure is determined by X-ray analysis. The molecular stacking and shape-index Hirshfeld surface study reveal that the π–π interactions could form between the benzene rings in the naphthyl group with moderate π–π distance of 3.67 Å and the π–π interactions between two benzene rings in the biphenyl motifs is not likely formed due to the relatively long π–π distances of 4.64 Å, as shown in Scheme 21. Although this synthetic strategy enables the assembly of various dibenzopleiadiene-embedded PAHs, the high reaction temperature (160 °C) may limit its practical applications. Thus, more convenient and practical methods for the construction of functional PAHs are demanded.

Scheme 21. Annulative Decarbonylation/Decarboxylation with Cyclic Diaryliodoniums to Construct Polyaromatics; (A,B) Crystal Structure and Molecular Packing of 68h.

Scheme 21

In 2018, Zhang’s group reported the transformation for the synthesis of triphenylenes 71 by using o-chlorobenzoic acids 70 and cyclic diaryliodoniums 69. The carboxylic acid promotes the C-arylation process under relatively mild reaction conditions,259 as detailed in Scheme 22. A range of commercially available o-chloro aromatic acids with different substituents (70ag) are compatible under the optimized reaction conditions and give the desired functionalized triphenylenes (71ag) in moderate to excellent yields. However, the heterocyclic substrates, such as sulfur-containing acid 70h, give relatively low yields, likely due to the catalyst poison by the sulfur. Later, such site-selective decarboxylative annulation of 69 is also applicable to 2-chloropyridinyl acids 72. As a result, diverse polycyclic dibenzo[f,h]quinolones 73 are furnished in modest to good yields.260 This transformation is highly applicable because o-chloro benzoic acids 70 and pyridine carboxylic acids 72 are readily commercially available. Unfortunately, unsymmetrical cyclic diaryliodoniums give an inseparable mixture of products (73j and 73k) with a ratio (1:1, 1.7:1).

Scheme 22. Site-Selective Decarboxylative Annulation to Construct Polycycles.

Scheme 22

In 2019, a palladium catalyzed annulation approach of inactivated simple arenes 75 with cyclic diaryliodoniums 74 to prepare triphenylenes frameworks was reported by Lee’s group (Scheme 23A).261 In this efficient annulation process, two C–C bonds are formed via the coupling of arenes 75 and cyclic diaryliodoniums 74 involving double C–H activation. Notably, arenes with either electron-donating or -withdrawing groups (75ag) and indole (75h) are suitable under the optimized conditions. However, the reaction is sensitive to the steric effect and provides the desired products with slightly low yields (76g). Using similar strategy, helical carbohelicenes 78, via Pd(II)-catalyzed regioselective C8–H/C7–H arylation of 1-naphthamides 77 with cyclic diaryliodoniums 74 have been accomplished by You and co-workers (Scheme 23B).67 The sterically hindered tert-butylaminocarbonyl is employed as a directing group for the C–H activation. By this transformation, various helical carbohelicenes 78 are assembled by a one-pot process in good to excellent yields. A wide range of 1-naphthamides and cyclic diaryliodoniums are applicable in this transformation. The X-ray analysis has determined the crystal structure of 78a. Furthermore, the potential photophysical property of the resulting carbohelicenes 78 is also investigated. The results show that single or double carbohelicenes emit violet to deep-blue fluorescence, while [4,5,4]-fused triple carbohelicenes display sky-blue emissions. This founding has the potential to facilitate the development of novel thermally activated delayed fluorescence materials by further applications. Notwithstanding, the harsh reaction conditions to construct carbohelicene may restrict its practicability.

Scheme 23. Regioselective Dual C–H Arylations of Cyclic Diaryliodoniums to Construct Benzenoid Triphenylenes and Carbohelicenes.

Scheme 23

The above transition-metal-catalyzed annulation between cyclic diaryliodoniums and arenes could construct π-extended arenes efficiently. However, the use of transition heavy metals poses some drawbacks, such as high toxicity and requirement of additional purification. Hence, the more economical and convenient approaches to delivering diverse PAHs are still desirable. More recently, the alkylamine-mediated free radical intramolecular annulation to produce various PAHs 80 under environmentally friendly reaction conditions is reported by the Wen group (Scheme 24).262 Notably, tert-butylamine is employed as an organocatalyst to initiate this synthetic operation for the construction of iodine-containing PAHs 80. Based on this protocol, both phenyl directly linked to cyclic diaryliodoniums (79ac) and phenyl installed via methylene to cyclic diaryliodoniums (79df) proceed smoothly to afford six- or seven-membered PAHs, respectively. For the reaction mechanism, a plausible pathway involving radical species is proposed, in which the tert-butylamine plays an important role.

Scheme 24. Modular Metal-Free Catalytic Radical Annulation of Cyclic Diaryliodoniums to Construct π-Extended PAHs.

Scheme 24

Fluorene is a common core structure of functional materials263270 and biologically active pharmaceuticals.264,268,271279 Recently, the Wen group has presented a palladium-catalyzed approach for synthesis of substituted fluorenes 83 via sequence Mizoroki–Heck/reductive Heck domino reactions between cyclic diaryliodoniums 81 and terminal alkenes 82 in the presence of Et3N/HCOONa as base/reducing agents (Scheme 25).280 Mechanistically, palladium species coordinate with cyclic diaryliodoniums 81, and subsequent insertion into alkenes followed by β-H elimination process generates a key intermediate int11. Then, a further cyclization to yield complex int12. Subsequently, int12 undergoes anion exchange with formate to give intermediate int13 and finally yields the compound 83. In this procedure, commercially available aryl and alkyl terminal alkenes are compatible, and the acrylates and their mimics also perform well with good yields. Strikingly, the more inert alkyl alkenes including 82h also provide the expected products in reasonable yields. Pleasingly, polycyclic heteroarene compound 83j could be conveniently generated with modest yield by utilizing HCI 82j. With this procedure, nearly 30 novel fluorenes are efficiently constructed, extensively expanding the structural diversity of the fluorene-based compound library. The practicality of this strategy is also demonstrated to construct the anthracenes from the six-membered cyclic diaryliodoniums and various commercially available terminal alkenes.281 Importantly, the screening of the new synthetic anthracenes successfully leads to the discovery of several novel compounds with potent antiproliferative activities.

Scheme 25. Tandem Modular Mizoroki–Heck/Reductive Heck Reactions to Construct Fluorenes.

Scheme 25

In 2017, the Wen group disclosed an approach to construct structurally diversified fluorenes with an all-carbon quaternary center (86ah), which is inaccessible from the aforementioned method as shown in Scheme 25. In this approach (Scheme 26), dual C–C arylation of enolizable methylene species takes place in the presence of palladium catalyst.282 With this new finding, various novel spirofluorenes containing barburate acid (86g) or indane motifs (86h), which are featured in medicinal drugs and functional materials are obtained. Notwithstanding, this strategy is limited to the enolizable compounds bearing a dual electron-withdrawing group. In 2020, You et al. reported another approach to synthesize spirofluorenyl naphthalenones 88 from easily available naphthols 87 through direct spiroannulation with cyclic diaryliodoniums 84 under mild reaction conditions.66 In this spiroannulation procedure, dearomatization is involved, and consequently the hydroxyl group in naphthols 87 is converted into a carbonyl group. It is worth noting that either naphthols 87 or cyclic diaryliodoniums 84 with various halogen substitutions (e.g., F, Cl, CF3) are well tolerated and deliver the corresponding spirofluorenyl naphthalenones smoothly. Furthermore, these halogen-bearing products could undergo further transformations to build thermally activated delayed fluorescent materials.

Scheme 26. Pd-Catalyzed Spiroannulation of Cyclic Diaryliodoniums with Naphthols.

Scheme 26

3.1.4. Other Atom Arylation (Halo/S/Se/Si/Te/B) /Cyclization

Heteroatoms-containing molecules or heterocyclic compounds are widely present in pharmaceutical drugs203,240,283289 and organic materials.286,290,291 Due to their unique chemical properties, the development of new methodologies to access structurally diversified heteroatom-containing scaffolds is of high importance.210,292300 The advances in synthesis of such heterocyclic compounds from cyclic aryliodoniums have been made.

As aforementioned, Postnikov et al. have successfully constructed diverse imidazole-flanked cyclic iodoniums.98 Later, they further develop a metal-free heterocyclization reaction of elemental sulfur (S8) with newly produced HCIs 89. As a result, a series of benzo[5,1-b]imidazothiazoles 90 are obtained in modest to high yields (Scheme 27A).98 Their work illustrates how HCIs can be efficiently utilized to access complex building blocks. Considering the essential functions of selenium or silicon-containing frameworks in drugs301309 or optoelectronic materials,310315 concise synthetic routes to these scaffolds are receiving much attention.316318 In 2017, Xu and co-workers disclosed copper-catalyzed dual arylations between cyclic diaryliodoniums 91 and Se, potassium thioacetate (AcSK), or dichlorodimethylsilane for construction of triselenasumanene, chalcogenasumanenes, and silicon-containing heterosumanene 92, which may pose considerable synthetic challenges of preparation by conventional approaches (Scheme 27B).118 It is noteworthy that this protocol would inspire the study of iodine-doping to the aromatic frameworks for the synthesis of novel curved and planar π-extended materials, such as nanoloops, buckybowls, and nanographenes.

Scheme 27. Synthesis of Cyclic Se, S, and Si-Containing Molecules from Cyclic Diaryliodoniums.

Scheme 27

Tellurium-containing motifs have a wide application in the range from metallurgy to electronics and the chemical field,319326 such as solar panels,327329 thermoelectric materials,330332 semiconductor devices,333335 alloys,336339 and chemical raw materials.340,341 An attempt has been reinvested in the development of synthetic procedures for tellurium-containing compounds by Xu and co-workers in 2019.342 They can construct various π-extended tellurium-containing arenes 94, including tritellurasumanene 94l, through efficient intermolecular and intramolecular coupling process between tellurium powder and cyclic diaryliodoniums 93 (Scheme 28). Importantly, these unique compounds demonstrate excellent thermal stabilities of the π-extended tellurium-containing aromatic compounds, thereby providing the potential for further applications.

Scheme 28. Synthesis of Tellurium-Containing Compounds from Cyclic Diaryliodoniums.

Scheme 28

The C1-carborane anions [CHB11H11] have novel conjugation behavior between σ and π aromaticity and are bestowed with unique properties and functions, for example, being used as ionic liquid crystals.343345 In 2018, Uchiyama et al. developed a one-pot annulation approach to provide biaryl-fused C1-carborane anion derivatives 97 using cyclic diaryliodoniums 95 and C1-carborane anion 96, as described in Scheme 29.346 In this transformation, without prefunctionalization of the C1-carborane anion, the C–B bond is built under palladium catalysis after the initial formation of C–C bond through C-arylation. However, the substrate scope is narrow to symmetrical cyclic diaryliodoniums. Furthermore, the DFT calculation studies demonstrate that the B–H bond deprotonation/metalation may undergo in a concerted manner.

Scheme 29. One-Pot Procedure for Synthesis of Biaryl-Fused C1-Carborane Anion Derivatives from Cyclic Diaryliodoniums.

Scheme 29

The X-ray structure of 97e is reproduced with permission from ref (346). Copyright 2018 Royal Society of Chemistry.

In the past several years, cyclic aryliodoniums as dual-arylation reagents have found wide applications in the synthesis of various polyfunctionalized compounds and complex polycyclic skeletons. Numerous effective transformations have been discovered to broaden the synthetic application of cyclic aryliodoniums. However, conventional transition-metals are often required and used, causing considerable toxicity and cost inevitably. Meanwhile, the scope is usually limited to symmetrical cyclic aryliodoniums in most cases, while it remains a challenge to achieve the regioselectivity for unsymmetrical counterparts. Probably, preinstallation of directing groups (DGs) in unsymmetrical cyclic diaryliodoniums may serve as an efficient strategy to obtain the expected regioselectivity. Collectively, more efficient and practical strategies to use cyclic aryliodoniums for the construction of novel functionalized compounds are highly desirable.

3.2. Construction of Biaryl Compounds via Single Arylation

3.2.1. Construction of Racemic Biaryl Compounds

With carboxylic acids as nucleophile, Zhang and co-workers have described a general and convenient approach to prepare various valuable biaryl esters 100 from cyclic diaryliodoniums 98 (Scheme 30A).28 This procedure tolerates a wide range of substrates (e.g., benzoic, aliphatic, and heterocyclic carboxylic acids) with good yields (70–97%). Moreover, in this system, excellent regioselective arylation of carboxylic acids can be obtained without observation of the arylation at the hydroxyl/amino positions (100ac). Compared to Olofsson’s protocol,347 this approach gives better yields with amino and hydroxyl groups contained. Furthermore, this protocol provides rapid access to a diverse range of functionalized biaryl esters 100 with an iodo-substituent, which can be readily transformed to diversified scaffolds.

Scheme 30. Copper-Catalyzed Arylation of Carboxylic Acids (A) and Phosphinic or Phosphoric Acids (B) with Cyclic Diaryliodoniums.

Scheme 30

Wong et al. have developed a facile synthesis of valuable 2′-iodo substituted biaryl phosphinic/phosphoric acid esters 102 through copper-catalyzed single arylation of P(O)–OH bonds with cyclic diaryliodoniums 98 (Scheme 30B).348 Various functional P(O)–OH compounds 101 and a range of substituted cyclic diaryliodoniums 98 are compatible under the optimized reaction conditions. This new synthetic method possesses high potential for further applications, such as the construction of biologically active compounds and chiral phosphine ligands.

Recently, Zhang et al. have reported a practical Pd-catalyzed C(sp3)-H biarylation approach to construct various functionalized biaryls 105 by employing 2-methylbenzaldehydes 104 and cyclic diaryliodoniums 103 with tert-leucine as a transient directing group,349 as depicted in Scheme 31. The tert-leucine has a significant influence on reaction efficiency and selectivity, and good to excellent yields are obtained with various 2-methylbenzaldehydes 104 and substituted cyclic diaryliodoniums 103. Mechanistically, the reaction undergoes a C–H activation process to yield Pd(IV) intermediate from Pd(II) species int15, which is followed by oxidative addition with cyclic diaryliodoniums to give intermediate int16. To validate the practicality of this approach, various functionalized biaryls 105 bearing iodide are further transformed into a wide range of compounds with potential applications in pharmaceutical motifs, such as PAH products and imidazole derivatives.

Scheme 31. Pd-Catalyzed C(sp3)–H Arylation via Transient Directing Group Strategy.

Scheme 31

The 2, 2′-dihalobiaryl scaffolds represent an essential precursor for the construction of heterofluorenes and other extended π-conjugated complex compounds. Yoshikai and co-worker have explored novel iodinative ring-opening reactions of cyclic diaryliodoniums 106, resulting in various 2,2′-diiodobiaryls 107 in modest to excellent yields,111 as shown in Scheme 32. A plausible catalytic cycle involving oxidative addition and reductive elimination process is proposed. However, the halogenation of aryliodoniums is limited to produce diiodobiaryl. More recently, Zhang and co-workers have broadened the halogenation of cyclic diaryliodoniums 106 to bromine and chloride.91 Diverse symmetric and unsymmetric cyclic diaryliodoniums (106a-i) regardless of chemical properties of substituted groups proceed smoothly and give the corresponding 2,2′-dihalobiaryl derivatives (108a-i) with excellent yields.

Scheme 32. Cu-Catalyzed Halogenation of Cyclic Diaryliodoniums.

Scheme 32

3.2.2. Construction of Chiral Biaryl Atropisomers

Axially chiral compounds, especially the axially chiral biaryl atropisomers, have obtained considerable attention from organic chemists and medicinal chemists due to their widespread presence in biologically active molecules such as Mastigophorene A,350 Korupensamine A,351 and Steganacin (Scheme 33).352 In addition, axially chiral biaryl atropisomers find wide-ranging utilities in functional materials,156,169,353355 for example, fluorescent sensor356 and rotary molecular motor,357 or they are used as privileged chiral ligands (e.g., BINAP, BINOL)358364 and catalysts for asymmetric synthesis.365,366 Owing to the importance of these scaffolds, great progress has been achieved in the synthesis of axially chiral biaryl compounds over the past decades. Because the single arylation with cyclic aryliodoniums can generate biaryls, the facile and practical protocol for the construction of chiral biaryl atropisomers from cyclic aryliodoniums is promising and appealing. A number of elegant methods have been reported for the construction of these biaryl atropisomeric scaffolds.

Scheme 33. Representative Molecules Containing Axially Chiral Biaryl Atropisomers.

Scheme 33

Due to their essential role in both medicinal and synthetic chemistry, several approaches for the construction of biaryl atropisomers have been developed, including metal-catalyzed asymmetric cross-coupling367369 and de novo aromatic ring formation.148,370 For example, the Hayashi group has disclosed a nickel-catalyzed ring opening of dinaphthothiophene to afford axially chiral 1,10-binaphthyls.371 They then employ cyclic dinaphthaleneiodonium to perform ring opening catalyzed by palladium species to obtain 2-iodo-2′-functionalized-1,1′-binaphthyls.372 However, these biaryls from cyclic aryliodoniums are racemic. Recently, the Gu group has developed efficient methods to construct diverse functionalized axially chiral biaryl atropisomers from cyclic diaryliodoniums.32,58,373378 The cheap copper-catalyzed atroposelective ring-opening procedures involve amination, oxygenation, and thiolation. The two ortho-substituents installed adjacent to cyclic diaryliodoniums may exert torsional strain on the biaryl linkage to restrict its rotation. Thus, such hindered substituents and the chiral ligands of copper further stabilize axial configuration of enantio-enriched atropisomer upon ring opening (Scheme 34). Finally, cyclic aryliodoniums react with various nucleophiles and result in excellent enantioselectivity and yields.

Scheme 34. Diverse Atroposelective Ring-Opening of Cyclic Aryliodoniums via Single Arylation.

Scheme 34

3.2.2.1. Construction of Biaryl Atropisomers via Single N-Arylation

Nitrogen-containing atropisomers are widely utilized in chiral ligand skeletons.156,379383 In 2018, the Gu group reported a highly efficient ring-opening amination process of cyclic aryliodoniums 109 toward the amines to deliver optically pure 2-iodo-[1,10-biphenyl]-2-amines derivatives 110 (Scheme 35).378 In this approach, enantioselective amination of cyclic diaryliodoniums (e.g., N-arylation of amine) is designed to assemble enantioenriched biaryls by employing chiral Cu-bis(oxazolinyl)pyridine complex (catalysts 1, 2) as the catalysts. Mechanistically, computational investigation indicates that the two conformers of the cyclic diaryliodoniums had a low rotational barrier and may rapidly interconvert. Consequently, the active catalysts and cyclic diaryliodoniums would give intermediate int19 and int20. However, because of steric repulsion between the methyls in cyclic diaryliodonium and the benzyl groups of bis(oxazolinyl)pyridine, int20 is sterically disfavored and quickly converts to int19. Finally, a ring-opening reaction and subsequent reductive elimination undergo to establish axial chiral products. Both anilines and benzylamines are successfully utilized for the construction of expected atropisomers with reasonable yields and notable ee values ranging from 97% to 99%. Further studies validate that various cyclic aryliodoniums with monosubstituted or disubstituted methyl group also generate the corresponding atropisomers in excellent efficiency and enantioselectivity. However, one of the limitations in this method is that the hindered anilines and aliphatic amines are not suitable.

Scheme 35. Atroposelective Ring-Opening of Cyclic Aryliodoniums to Construct Atropisomers with Amines.

Scheme 35

The abbreviated reaction pathway is reproduced with permission from ref (378). Copyright 2018 Elsevier.

More recently, Gu and co-workers have found that the combination of CuCl and anionic chiral cobalt(III) as cocatalysts (catalyst 3) is able to construct nitrogen-containing atropisomer skeleton 111 from bulky aromatic amines, such as bulky 2,6-dimethylaniline.58 To understand the reaction mechanism, several control experiments have been performed. The results validate that the copper catalyst is essential to this transformation. Later, catalyst 2 can be further employed to realize asymmetric ring-opening reaction of cyclic aryliodoniums with both benzylic and aliphatic amines in high yields and enantioselectivity.377 Importantly, this procedure can well tolerate different substituted aliphatic amines and transform them smoothly to axial atropoenriched biaryl products 112, which are difficult to be prepared by conventional asymmetric synthetic approaches. Very recently, Wang et al. have explored an efficient enantioselective copper-catalyzed ring-opening reaction of cyclic diaryliodoniums with imides to afford a series of axially chiral 2-imidobiaryl compounds, which could be readily derivatized to various types of aromatic amines.384

In 2019, the Gu group reported a novel chiral nitrogen-containing ligand [3,5-di(tert-butyl)phenyl bis(oxazoline) (ligand 1) with Cu(OTf)2 for atroposelective amination of cyclic aryliodoniums 113 with O-alkylhydroxylamines 114 to construct 2-hydroxyamino-2′-iodobiaryls 115 with up to 99% ee values (Scheme 36).137 The benzenesulfonamide moiety with diverse substitutions (114ac) in the para-position proceeds well to give the corresponding 2-hydroxyamino-2′-iodobiaryls (115ac) with excellent enantioselectivity (99%). Different substituents on the oxygen or nitrogen atoms (114dh) are well tolerated to form the amination products (115dh) with high enantioselectivity up to 98% ee. Likewise, substituted cyclic aryliodoniums (113ik) are suitable under the current conditions.

Scheme 36. Cu-Catalyzed Asymmetric N-Arylation with Alkylhydroxylamine.

Scheme 36

The Gu group has recently reported a novel Cu/bisoxazoline-catalytic atroposelective ring-opening amination of five-membered cyclic aryliodoniums 116 with oximes 117 to assemble atropisomeric nitrones 118 in good optical purity in the presence of cesium carbonate (Scheme 37).385 A metal-catalytic cycle is proposed. Initially, ligand 2 coordinates copper catalyst interactions with cyclic aryliodoniums to afford a chiral copper complex int21, and then the intermediate int22 is generated through oxidative addition of cyclic aryliodoniums with Cu(I) species. Subsequently, int22 undergoes ligand exchange with oximes to produce intermediate int23, followed by reductive elimination to provide the enantioenriched nitrones and regenerate the Cu (I) species. To validate the utility of this methodology, the produced nitrones are efficiently transformed to axially chiral anilines 119 and N-heterocyclic compounds 120.

Scheme 37. Cu-Catalyzed Asymmetric Ring-Opening Reaction of Cyclic Aryliodoniums and Oximes.

Scheme 37

Gu and co-workers also explored an additional copper-catalyzed enantioselective ring-opening system of cyclic aryliodoniums 121 with 1,2,3-triazoles 122 as nitrogen nucleophiles to afford various axially chiral triazole-containing molecules 123 (Scheme 38).386 This transformation leads to the formation of a C–N bond with excellent site selectivity for three different nitrogen atoms of 1,2,3-triazoles 122. In this protocol, a broad range of 1,2,3-triazoles 122 could proceed smoothly to prepare products 123 with excellent enantioselectivity and stereoselectivity. Additionally, various substituted groups on five-membered cyclic aryliodoniums (121ln), including alkyl and halogen, are suitable for this transformation and deliver functionalized 1,2,3-triazole derivatives (123ln) successfully. Furthermore, it is worth mentioning that these axially chiral triazole-containing scaffolds could serve as powerful substrates for the construction of different chiral phosphine ligands (124, 125).

Scheme 38. Cu-Catalyzed Asymmetric Ring-Opening Reaction of Cyclic Aryliodoniums and 1,2,3-Triazoles.

Scheme 38

3.2.2.2. Construction of Biaryl Atropisomers via Halogenation

The biological activities and the unique structures of axially chiral 2,2′-diiodobiaryl derivatives have attracted continuous attention in the pharmaceutical and medicinal chemistry fields.387,388 Particularly, there was ample evidence that the axially chiral 2,2′-diiodobiaryl scaffolds possess the potential to yield diverse privileged C2-symmetic chiral ligands.389 In addition, the axially chiral 2,2′-diiodobiaryls are also employed as hypervalent iodine precursors or application in other asymmetric transformations.390

Recently, the He group has employed Cu/PyBox and simple tetrabutylammonium halides 127 to construct axially chiral 2,2′-dihalobiaryls 128 in good to excellent yields and ee values (Scheme 39).391 The use of CuI with the chiral bisoxazoline ligand (ligand 3) in solvent hexafluoroisopropanol (HFIP) is essential to this transformation. A comprehensive scope of substituted symmetrical or unsymmetrical cyclic aryliodoniums 126, such as halogen substituents (−F, −Cl, −Br) are proved to be suitable and could access the desired products in high yields and excellent enantioselectivity. The potential utility of this strategy is also investigated. The axially chiral 2,2′-dihalobiaryl product 128a can be further converted to a number of privileged C2-symmetric chiral ligands or catalysts by diversity-oriented transformations. For example, BAMOL (1,1′-biaryl-2, 2′-dimethanol) derivative 129 and BIPHEMP 130 are obtained conveniently, and 129 exhibits excellent activity in the Diels–Alder reaction.

Scheme 39. Copper-Catalyzed Asymmetric Halogenation of Cyclic Aryliodoniums for Synthesis of Axially Chiral 2,2′-Dihalobiaryls.

Scheme 39

3.2.2.3. Construction of Biaryl Atropisomers via Single O-Arylation

Chiral phosphine-containing ligands possess distinct values and exhibit exceptional utilities in either organocatalysis or transition-metal mediated asymmetric catalysis.392,393 In 2019, the Gu group realized an asymmetric ring-opening/oxidative phosphorylation procedure, which could readily provide various atropisomeric phosphinate products 133 via a simple Cu-catalyzed ring-opening oxygenation (e.g., O-arylation) between cyclic aryliodoniums 131 and diarylphosphine oxides 132 in the presence of TEMPO and the addition of bis(oxazoline) ligand (ligand 2) (Scheme 40).394 Based on 18O-labeled control experiments and DFT studies, a plausible catalytic cycle involving C–O formation rather than C–P formation is proposed. Briefly, the axially chiral Cu(III) intermediate is generated through oxidative addition of cyclic aryliodoniums with Cu(I) species and subsequently coordinates with diarylphosphine oxide to give intermediate int24 that undergoes oxidation to generate intermediate int25 and reductive elimination to provide atropisomeric phosphinate product 133. This versatile protocol permits easy chemical elaborations of different diphenylphosphine oxide (132af) into atropo-enantioenriched phosphine-containing scaffolds (133af) with excellent ee values (99%). The range of coupling partners is successfully extended to electron-withdrawing 132e, electron-donating substituents (132ac) or halogen substituents 132d installed at diarylphosphine oxides. Additionally, butyl(phenyl)phosphine oxide 132h also displays good compatibility, and their asymmetric ring-opening/oxidative phosphorylation is conducted to assemble desired diastereomeric products 133h with excellent enantioselective control (99%). The aryliodoniums with different substituted groups (131im) are efficiently transformed to the expected products (133im). The dicyclohexylphosphine oxide derivatives fail the transformation likely due to their inert property. In the following transformations, atropisomeric 2′-hydroxy diaryl phosphine oxides 134 could be readily produced via tert-butyllithium-mediated intramolecular P-transfer of the atropisomeric phosphine oxides 133. Further computational studies indicate that the phosphine oxide transfer procedure involves a concerted C–P bond formation and P–O bond cleavage process.

Scheme 40. Copper-Catalyzed Asymmetric Phosphorylation of Cyclic Aryliodoniums and Subsequent Transformation.

Scheme 40

The abbreviated reaction pathway is reproduced with permission from ref (394). Copyright 2003 American Chemical Society.

In a recent work by Zhang’s group,395 a facile and efficient copper/chiral Box ligand is employed in an enantioselective acyloxylation for the construction of various axially chiral acyloxylated 2-iodobiaryl derivatives 137 in excellent yields and enantioselectivity. In this transformation, cyclic aryliodoniums 135 and ubiquitous aliphatic or (hetero)aromatic carboxylic acids 136 are utilized (Scheme 41). A broad range of benzoic acid (136ah) or alkyl carboxylic derivatives (136mp) bearing different substituents are effective and give the corresponding products in excellent yields and ee values. In addition, heteroaromatic carboxylic acids (136j,k), such as 2-furan/thiophene carboxylic acids also proceed smoothly to produce the corresponding chiral biaryls. However, 2-quinolinecarboxylic acid 136l was not suitable for the reaction to generate 137l, likely due to the strong coordinating effect of the quinoline nitrogen. Furthermore, various substituted cyclic aryliodoniums (135qs) perform well in this system, and the reactions proceed steadily to give the corresponding axially chiral acyloxylated 2-iodobiaryls (137qs). Remarkably, this efficient approach has been applied for the late-stage modification of several marketed drug molecules to produce novel drug-like atropisomers, including indomethacin 138, febuxostat 139, and diclofenac 140. Finally, the value of this strategy is further demonstrated by the subsequent diversity-oriented transformations of the biaryl acyloxylated products to a wide range of functionalized axially chiral biaryls, for example, chiral bidentate pyridine and phosphine ligands or functional molecules.

Scheme 41. Copper-Catalyzed Enantioselective Acyloxylation of Cyclic Aryliodoniums with Carboxylic Acids.

Scheme 41

α,β-Unsaturated carboxylic acids 142 are also employed to realize the enantioselective acyloxylation of cyclic aryliodoniums 141 using copper/(Ph)-bis(oxazoline) as the chiral catalyst system to generate oxygenated axially chiral biaryls 143, disclosed by the Gu group (Scheme 42).376 Pleasingly, various substituted α,β-unsaturated carboxylic acids 142 including aromatic carboxylic acids (142eg) are compatible, and corresponding products (143eg) have been provided in high yields and excellent ee values. Likewise, a variety of cyclic aryliodoniums (141hn) with different groups, e.g., alkyl, halogen. and aryl, are proceeding well to furnish the desired products (143hn) with high yield (>94%) and excellent stereocontrol (>98% ee).

Scheme 42. Enantioselective Acyloxylation of Cyclic Aryliodoniums with Unsaturated Carboxylic Acids and Further Transformation to Bridged Biaryl Atropisomers.

Scheme 42

The construction of strained seven-membered rings containing a bridged biaryl atropisomer is challenging.396403 Based on the previous achievements in atropisomer construction, a facile approach to construct lactone-bridged biaryl atropisomers 144 from acyoxylated biaryls 143 has been developed.376 The approach involves a palladium-catalyzed intramolecular diastereoselective Mizoroki–Heck cyclization (Scheme 42). A comprehensive range of acyloxylated biaryls 143 are well tolerated, and the corresponding lactone-bridged biaryl atropisomers 144 are obtained in moderate to good yields and excellent ee values (>96% ee).

After the successful pursuit of enantioselective ring-opening single arylation of cyclic aryliodoniums, Gu et al. have reported a boronic acid-promoted Cu/Ph-PyBox catalyzed enantioselective ring-opening alkoxygenation of cyclic aryliodoniums 145 with diols 146 (Scheme 43).375 In this transformation, various functional groups on glycol derivatives (146ad) are accommodated. By changing alkyl groups to aromatic groups in 1,2-diols, the alkoxylated biaryls (147c,d) are acquired, albeit in slightly low enantioselectivity. A variety of symmetrical or unsymmetrical cyclic aryliodoniums are well tolerated under optimized reaction conditions (147e–h). At the meanwhile, ketohydroxylated compounds are also obtained (147ik). Additionally, 1,4-diols are proceeding smoothly to provide corresponding products (147lo) with attractive yields (>72%) and enantioselectivities (>85% ee). Furthermore, the product 147a could be readily converted to valuable phosphine ligands (148, 149). These applications demonstrate the utilities of the axially alkoxylated biaryls. Finally, a potential catalytic cycle involving a key ion pair intermediate (I-147-7) is proposed. First, the oxidative addition of I-147-7 with copper complex I-147-1 affords intermediate I-147-2, followed by transferring oxygen of I-147-7 to the Cu(III) center of I-147-2. The produced intermediate I-147-3 could undergo reductive elimination to regenerate copper catalyst I-147-1 and provide I-147-4. Finally, intermediate I-147-4 exchange with diols to release the boronic acid and provide the major enantiomer product 147.

Scheme 43. Cyclic Aryliodoniums with Diols to Generate Oxygenated Axially Biaryls.

Scheme 43

3.2.2.4. Construction of Biaryl Atropisomers via Single S-Arylation

Sulfur-containing compounds are the important moiety in drugs and natural products.404406 Due to their biological activities and pharmaceutical values, efficient synthetic approaches are attracting attention in pharmaceutical chemistry407412 and material science fields.413420 There have appeared several concise and efficient methods to synthesize sulfur-containing compounds through direct thiolation of cyclic aryliodoniums with elemental sulfur (e.g., S-arylation) recently.71,130,421,422

An efficient transformation of six-membered cyclic aryliodoniums 150 with thioacid salts 151 to prepare chiral diarylmethanes 152 via copper/[cyclopropyl bis(oxazoline)]-catalyzed enantioselective ring-opening desymmetrizing reaction was disclosed by the Gu group in 2018 (Scheme 44).32 Based on the previous computational studies and crystal structure of cyclic aryliodoniums, a plausible catalytic mechanism of this transformation is proposed to rationalize the stereochemical control. Initially, Cu/bis(oxazoline) complex int28 as the active catalyst initiates the first step. Intermediate int29 is generated through the oxidative addition of cyclic aryliodoniums with copper species. Potassium thioates coordinates with Cu(III) center of int29 to afford intermediate int30, which would produce the final products 152 by subsequent reductive elimination. In this strategy, potassium thioates 151 are utilized as the sulfur sources, and the corresponding functionalized chiral diarylmethanes 152 are efficiently achieved with good yields and enantiomeric excesses (>91% ee). Moreover, various cyclic aryliodoniums with different substituents (150dn) are compatible to give the desired chiral diarylmethanes (152dn) exclusively. Interestingly, the steric effect of cyclic aryliodoniums rarely affects the stereo induction. This is the first report of chiral sulfur-containing compounds’ construction through direct S-arylation of cyclic aryliodoniums with sulfur reagents. Recently, Yang et al. have reported a copper-catalyzed enantioselective desymmetrizing ring-opening reaction of six-membered cyclic diaryliodoniums with sulphonamide via N-arylation to afford a series of N-monoarylsulfonamides in good to excellent yields and enantioselectivities at room temperature.423

Scheme 44. Copper-Catalyzed Enantioselective Ring-Opening/Thiolation of Six-Membered Aryliodoniums to Construct Chiral Diarylmethanes.

Scheme 44

A Cu/Ph-bis(oxazoline)-catalyzed asymmetric ring-opening/thiolation reaction of five-membered aryliodoniums 153 and potassium thioates 154 for the preparation of atropisomeric biaryls is also reported (Scheme 45).374 Mechanistically, a metal-catalyzed asymmetric ring-opening pathway is proposed. Briefly, the Cu(II)/bis(oxazoline) complex int31 undergoes oxidative addition with cyclic aryliodoniums to produce intermediate int32. The subsequent ligand exchanges with thioate anions to give intermediate int33, followed by reductive elimination to afford 155. Notably, various potassium benzothioates (154ai) bearing either electron-donating or withdrawing groups on the phenyl ring proceed steadily and afford the corresponding products (155ai) with excellent yields and highly enantioselectivity (96–99% ee). Additionally, the aliphatic thioates (154jl) also work well and deliver the expected products (155jl) with stereocontrol (>91% ee). Moreover, various cyclic aryliodoniums (153mp) are suitable. Furthermore, the utilities of these axially chiral biaryl thioethers are briefly demonstrated through the synthesis of a typical axially chiral P,S-ligand 157 from 155j via a two-step reaction. It is worth mentioning that iodine and sulfur-containing axially chiral biaryl ligands are difficult to be constructed by conventional approaches and have potential further applications in a variety of asymmetric synthesis process.

Scheme 45. Copper-Catalyzed Enantioselective Ring-Opening of Five-Membered Aryliodoniums to Construct Chiral Biaryls with Thioacids.

Scheme 45

More recently, Gu et al. have developed another facile and practical approach to synthesize atropisomeric biaryl trifluoromethylsulfane derivatives 160 in good enantioselectivity via a copper-catalyzed ring-opening reaction (Scheme 46).373 In the proposed mechanism, this reaction commences via trifluoromethylthiol intermediate int35, which is formed through copper/PyBox complex int34 interacting with trifluoromethanethiolate CsSCF3. Intermediate int35 undergoes atroposelective oxidative addition with cyclic aryliodoniums to deliver intermediate int36, followed by reductive elimination to give the final product 160 and regenerate complex int34. Through this strategy, symmetrical cyclic aryliodoniums 158 and cheap CsSCF3159 are utilized to accomplish the enantioselective biaryl trifluoromethylthiolation procedure. A broad range of atropisomeric biaryl trifluoromethylsulfanes 160 substituted with electron-donating (e.g., -methyl) and -withdrawing groups (e.g., -benzophenone) are systematically furnished in modest to good yields and high enantioselectivity. However, the regioselectivity is decreased apparently when unsymmetrical cyclic aryliodoniums bearing chloride and methyl (158i,j) are employed. In addition, the product 160a could be smoothly transformed to chiral biphenyls 161 and phosphine ligands 162, which may exhibit great potential application in asymmetric synthesis.

Scheme 46. Copper-Catalyzed Enantioselective Ring-Opening of Five-Membered Aryliodoniums to Construct Chiral Biaryls with Trifluoromethanethiolate.

Scheme 46

Recently, the group of Zhang has reported a chiral copper(II)-bisoxazoline catalyzed enantioselective ring-opening/thiolation of cyclic aryliodoniums 163 with readily available 2-mercaptobenzoxazole and 2-mercaptobenzothiazole derivatives 164. A series of axially chiral sulfur-containing biaryls 165 are efficiently prepared in excellent yields and enantioselectivities (Scheme 47).424 2-Mercaptobenzoxazole derivatives bearing electron-donating (e.g., -methoxy, 164b), electron-withdrawing (e.g., -nitro, 164a), or halogen (e.g., -F, Cl, Br, 164gi) substituents on the phenyl rings all provide the desired products in excellent yields (>96%) and ee values (>98%). In addition, diverse cyclic aryliodoniums (163df, 163p) are investigated, and all of the transformation proceeds smoothly. Furthermore, a gram-scale reaction is well performed. Meanwhile, the benzoxazole motif of 165a is readily transformed into different alkyls (e.g., 166, 167) without a decreased ee value.

Scheme 47. Enantioselective Synthesis of Axially Chiral Biaryls via Copper-Catalyzed Thiolation of Cyclic Aryliodoniums.

Scheme 47

3.3. Organocatalytic Application of Cyclic Aryliodoniums as Halogen-Bonding Donors

The significant synthetic application of cyclic aryliodoniums as arylation reagents has been achieved, and their other potential application remains to be explored. Metallic catalysts have exhibited high reactivity to catalyze various transformations.425431 However, they have some disadvantages, for example, possible metal contamination of products and pollution of environments.432,433 Organocatalysts possess some privileges such as excellent air and moisture stability and pose no contamination of products.434441 More recently, halogen bonding (XB), the noncovalent interaction of Lewis bases with an electron-deficient halogen atom, has attracted increased attention in the research field of chemistry.21,442445 Thereby, the discovery of efficient halogen-containing motifs as XB donors has been subjected to intense research.

Before 2016, most of XB donors were explicitly designed based on iodine(I) derivatives. Hypervalent iodine(III) species are underrepresented to be utilized as Lewis acids, likely due to uncertain Lewis acidity. In the past few years, quantification of iodine(III)-based Lewis acidity and their thermodynamic parameters for the binding to weak Lewis bases have been determined, suggesting that the hypervalent iodine(III) compounds possess great potential application as XB donors. Moreover, a recently published research convincingly demonstrates that cyclic aryliodoniums possess greater halogen-bond donating catalytic activity than acyclic aryliodoniums, likely due to the additional energy required for the rotation of the phenyl ring in acyclic aryliodoniums cation during the ligation of substrates.446 Both s-type and p-type lone pairs in five-membered diphenyl iodonium are more stabilized compared to acyclic diphenyl iodonium,20 which is crucial for the halogen bonding. In addition, Legault and co-workers demonstrate that cyclic compounds have higher Lewis acidity than their acyclic counterparts.47 Collectively, cyclic aryliodoniums may have numerous advantages for halogen-bonding catalysis.

3.3.1. Binding Models of Cyclic Aryliodoniums as XB Donors

Cyclic diaryliodoniums have two binding sites approximatively 90° from each other, and their counterion may bind to one site (Scheme 48A). Most Lewis bases do not have the geometrical requirements to bind both sites simultaneously. In the case of cyclic diaryliodoniums, substituents near the ortho binding positions can block a binding axis completely. Therefore, through sterically blocking one or two of the electrophilic axes, the Lewis acidity of cyclic aryliodoniums may be significantly reduced or even lost (Scheme 48B). It is important to note that the highly biaxial directional nature of halogen bonding results in a mostly square planar geometry at the iodonium center. Based on the planar geometry structure of cyclic diaryliodoniums, as shown in Scheme 48C,D, Huber and co-worker have designed and identified several double carbonyl-containing compounds as suitable biaxial ligands, including dimethyl isophthalate and diethyl-3,3′-(1,2-phenylene)dipropiolate, which could simultaneously occupy the two binding sites of cyclic diaryliodoniums.447

Scheme 48. Conceptual XB Modes of Cyclic Arylidoniums Bound to Substrates.

Scheme 48

Huber’s group has performed a systematic study on the biaxial coordination of cyclic aryliodoniums 168 with suitable bidentate substrates 169.447 The results show that only geometrically suitable diesters and diamides could interact with unhindered iodine(III) species to generate the desired biaxial complex at high affinity, as shown in Scheme 49. A significant drop of binding strength is observed while one or two ortho positions of the iodonium center are blocked (170a-1 vs 170b-1, and 170c-1). For the para-diesters and diamides, the DFT calculations confirm that only one carbonyl group binds to the iodine(III) center (170a-4 and 170a-5). The crystal structural analysis unambiguously confirms the biaxial binding between diesters and unhindered iodoiums.

Scheme 49. Binding Constants of Cyclic Aryliodoniums to Diesters and Diamides.

Scheme 49

The X-ray structure of 170a-1 is reproduced with permission from ref (447). Copyright 2020 American Chemical Society.

So far, several research groups have reported their studies on various cyclic aryliodoniums as XB donors to assess their halogen-bonding strength via examination of their Lewis acidity.448,449 Moreover, the cyclic aryliodoniums as the XB donor in several model reactions have been performed to evaluate their activity as organocatalysts.

3.3.2. Structural Variation of Cyclic Aryliodoniums as XB Donors

3.3.2.1. Monodentate Aryliodoniums

Liu et al. have employed acyclic aryliodoniums as Lewis acid catalysts for the multicomponent Mannich process directly.450 In 2018, Huber et al. pioneeringly investigated the monodentate cyclic aryliodoniums as noncovalent Lewis acids to activate the halide abstraction and their role in halogen-bonding reactions.451 They investigate a series of substituted monodentate aryliodoniums 171 as potential halogen-bonding donors (Scheme 50). The further reaction kinetics study indicates that tetrakis[3,5-bis(trifluoromethyl)phenyl]borate (BArF4) salts (171-1b) are more potent as Lewis acids, possibly due to the noncoordinating nature of the anion, compared to triflate salts (171-1a) and chlorine salts (171-1c). No expected product or lower activity is observed in the reactions involving Lewis acid 171-2a or methylated derivatives (171-3 and 171-4), probably due to the poor stability of the iodonium compounds or spatial blockage of the electrophilic axis, as shown in Scheme 48. More importantly, the authors have established two benchmark reactions, including solvolysis of benzhydryl chloride 172 and Diels–Alder reaction of cyclopentadiene 175 with methyl vinyl ketone 174, revealing the iodoniums 171-1b as the most effective, which is in line with the kinetics studies. It is worth mentioning that aryliodoniums are used in catalytic amount for these two benchmark reactions.

Scheme 50. Two Model Reactions with Different Cyclic Aryliodoniums.

Scheme 50

3.3.2.2. Bidentate Aryliodoniums

The monodentate aryliodoniums have already exhibited great potential for use in noncovalent organocatalysis, however, their catalytic activity is still moderate.451 The bidentate aryliodoniums species chelating dinuclear iodine(III) featuring two σ-holes at the iodine atom exhibit remarkable Lewis acidity,114,452 and they possess greater catalytic activity in the interaction with a nucleophile than monodentate aryliodoniums derivatives (Scheme 51).21,47,113

Scheme 51. Interaction of Monodentate and Bidentate Diaryliodoniums with Nucleophiles (σ-Holes at the Iodine Atoms Are Represented as Red Ovals).

Scheme 51

Yoshikai and co-workers have reported bidentate cyclic aryliodoniums 177 with double hypervalent iodine centers.111 Later, Huber and co-workers have developed a new bidentate aryliodoniums 178 with a different counteranion as XB donor. In order to validate the transformation efficacy of this new XB donor, a variety of benchmark reactions are conducted, including Michael addition reactions as well as a Diels–Alder reaction (Scheme 52).114 Notably, compared to the iodine(I)-based or other cyclic aryliodoniums, 178 with BArF4 anion is more efficacious. The further assay verifies the catalytic activity of bidentate cyclic aryliodoniums is also enhanced while BF4 anion is used. The comparison catalytic efficiency experiments with monodentate iodolium compounds confirms that the enhanced activity of iodine(III) is not merely the result of the presence of two iodine centers but very likely the consequence of bidentate binding. Furthermore, XB catalytic efficacy by the bidentate iodine(III) derivatives is confirmed by several comparison experiments, a solid-state structure, and DFT calculations on the likely transition state.

Scheme 52. Bidentate Cyclic Aryliodoniums as Organocatalysts in Michael Addition and Diels-Alder Reaction.

Scheme 52

Recently, a more electron-deficient bidentate cyclic aryliodoniums 189 based on 188 is used to mediate Diels–Alder reaction by Fernandez et al. (Scheme 53).113 Based on the control experiments in the absence of any catalyst or in the presence of iodine(I) reagents as catalysts, bidentate cyclic aryliodoniums accelerate cycloaddition progress, likely through lowering the reaction barrier. The quantitative analysis reveal a dramatic reduction of the steric repulsion between the diene and dienophile, which rationalize the privilege of iodine(III) catalysts in this transformation. Notably, the activity of aryliodoniums is enhanced significantly while increasing the electrophilicity with fluorine. This result implies that cyclic aryliodoniums could be further developed as highly active catalysts.

Scheme 53. Bidentate Cyclic Aryliodonium Catalyzed Diels–Alder Reaction.

Scheme 53

3.3.2.3. Heterocyclic Aryliodoniums

N-Heterocycles as stabilizing functional groups in cyclic aryliodoniums scaffold are underexplored. In 2020, Nachtsheim and co-workers developed the N-heterocycles-containing cyclic aryliodoniums 191 as XB donors (Scheme 54). Further, they also explore the efficacy of 191 in some benchmark reactions by using benzhydryl halides 190 for subjection to amidation.97 Notably, 191 exhibits higher transformation efficiency than the canonical cyclic diphenyliodonium 171-1a. However, unsubstituted iodolopyrazolium triflates are not suitable in this transformation.

Scheme 54. XB-Based Halide Abstraction.

Scheme 54

Very recently, Nachtsheim et al. have reported another series of N-heterocyclic iodoniums (i.e., iod(az)olium) to be used as effective XB donors.453 Especially, N-methylated dicationic heterocyclic iodoniums (193195) exhibit an outstanding performance in the performed benchmark reactions (Scheme 55). These special iodoniums display higher catalytic activity than previously reported monodentate organic iodine(I) and (III) XB donors. Moreover, a small amount of catalyst loadings can efficiently promote Michael addition and Diels–Alder reactions.

Scheme 55. N-Heterocyclic Iod(az)olium Catalyzed Reactions.

Scheme 55

3.3.3. Synthetic Applications of Cyclic Aryliodoniums as XB Donors

Recently, an approach utilizing cyclic diaryliodoniums 171-1a as a XB donor catalyst for [4 + 2] cycloaddition of 2-alkenylindole 201 with 2-vinylindoles 202 to give the tetrahydrocarbazoles 203 is reported by Takayoshi and co-workers (Scheme 56).454 Compared to cyclic diaryliodoniums with other counterions (e.g., Cl), the ones containing triflate anion (OTf) promote the reaction more efficiently and show the higher catalytic activity. Finally, the scope and generality of this strategy are examined, and tetrahydrocarbazoles are obtained with good yield and diastereoselectivity. However, N-benzyl 2-vinylindoles (203df) results in low diastereoselectivity.

Scheme 56. [4 + 2] Cycloadditions of 2-Alkenylindoles with Iodoniums as XB Donor.

Scheme 56

The utilization of hypervalent iodine(III)-based XB donors has found increasing application in organocatalysis, and the vast achievement have been obtained.449,455457 However, the activation of metal–ligand bonds by XB have so far been limited to a few reactions with elemental iodine or bromine. Recently, the Huber group has presented the activation of gold chlorine bond by using cyclic aryliodonium 171-1b as a XB donor (Scheme 57).442 Compared with the other Lewis acid such as preorganized bis(benzimidazolium),458 the gold(I)-catalyzed cycloisomerization of 1,6-enyne 204 to form cyclic products 205a and 205b is also efficient, while 171-1b act as both Lewis acid and metal activator (Scheme 57A). The rearrangement is believed to proceed through cyclopropyl gold carbene species, and the efficacy of 171-1b underlines that the XB activation is beyond the simple Lewis acidity. Another benchmark reaction is also performed to broaden the application of 171-1b (Scheme 57B). The propargylic amide 206 is converted to oxazoline 208 under mediation of 172-1b. Furthermore, Nachtsheim and co-workers have used a heterocyclic iod(az)olium salt 207 to achieve the same cyclization.453 Mechanistically, both aryliodonoiums 171-1b and 207 in these transformations are not only employed as XB donors, but also act as activators in metal–halogen bonds, which is confirmed by the 31P NMR investigations and DFT calculations.

Scheme 57. Activation of Au–Cl Bond by Cyclic Aryliodoniums.

Scheme 57

Cyclic aryliodoniums can be also employed as noncovalent organocatalysts to promote other reactions. Recently, a Knorr-type reaction of N-acetyl hydrazides 209 with acetyl acetone 210 to generate N-acyl pyrazoles 211 procedure is developed by Bolotin’s group,459 as depicted in Scheme 58. The different hypervalent iodine species are examined, and the five-membered cyclic aryliodonium 171-1a exhibit the highest activity. With the optimized condition, a series of aromatic and aliphatic acylhydrazides proceed smoothly to construct desired N-acyl pyrazoles in modest yields. DFT calculations and 1H NMR titration data indicate that the catalytic activity of the iodine(III) results from its binding with a ketone and the solvent (MeOH) also has a significant influence on the transformation.

Scheme 58. Cyclic Aryliodoniums as Efficient Organocatalysts for Knorr-Type Reactions.

Scheme 58

Owing to the remarkable catalytic activity, aryliodoniums species are used for the multicomponent reaction to approach a series of new scaffolds. More recently, as described in Scheme 59, the group of Bolotin has realized a Groebke–Blackburn–Bienaymé reaction of aldehydes 212, isocyanides 213, with aminopyridine 214, which leads to a series of imidazopyridines 215 under catalysis of cyclic diaryliodonium 171-1a.446 Various aldehydes (212ag) and isocyanides (213hk), proceed smoothly. Importantly, the further mechanistic studies reveal that the ortho-H atoms in the vicinal position to the iodine atom play an essential role in this transformation, including forming additional noncovalent bonds with the ligated substrate and increasing the maximum electrostatic potential on the σ-hole at the iodine atom.

Scheme 59. Cyclic Aryliodoniums for Synthesis of Imidazopyridines.

Scheme 59

Most recently, the development of new enantioselective reactions using chiral hypervalent iodine compounds has been reported. In 2022, the Yoshida group has used enantiomeric aryliodonium 218 as a XB donor to realize the construction of chiral N,S-acetals 219 (Scheme 60), the motifs present in natural products and pharmaceuticals. Both yields and enantioselectivities in most cases are excellent, facilitating the addition of thiol 217 to isatin-derived ketimine 216.128 Tetrafluorine borate renders XB donor activity stronger than chloride and bromide. Additionally, the further NMR experiments and DFT calculations demonstrate that the imine substrate forms a bidentate intermediate with 218 through both a hydrogen bond and a halogen bond. In this transformation, an excessive amount of nucleophile 217 (7 equiv to 216) remains an unmet need for further reaction condition optimization.

Scheme 60. Chiral XB-Based Dual Catalysed Asymmetric Additions of Bulky Thiols to Ketimines.

Scheme 60

The abbreviated reaction pathway is reproduced with permission from ref (128). Copyright 2022 Wiley-VCH.

The fast research progress in hypervalent iodine(III)-based XB donors has been witnessed in the past few years. Due to their high electron-withdrawing abilities and remarkable Lewis acidity, the cyclic aryliodoniums have been evaluated for their activities as organocatalysts in several designed reactions. As mentioned above, the halide abstraction and further activation of neutral substrates are effectively accomplished by these cyclic aryliodoniums. Due to their high halogen bonding strength and high directionality of cyclic aryliodonium as organocatalysts, new research progresses in crystal engineering, construction of biomolecular systems, and organic materials can be expected in the near future. The bidentate aryliodoniums and heteroaryl moieties bearing aryliodoniums are less developed although they have higher catalytic activity as Lewis acid in transformations. To date, there are few examples of their catalytic application to achieve enantioselective transformations with their Lewis acidity.128,460 Taken together, the bidentate aryliodoniums, heterocyclic aryliodoniums, and chiral hypervalent bromine(III) salts require further investigation in halogen-bonding donors and bifunctional asymmetric catalysis.

4. Conclusion and Perspective

In the past several years, the advancement in the cyclic aryliodonium synthetic chemistry field has been made. Cyclic aryliodoniums as powerful arylation reagents have been intensively investigated, leading to construction of various polycyclic arenes and axially biaryl scaffolds. Compared to the well explored acyclic aryliodoniums, cyclic ones are less reactive but more atom and reaction economical in terms of their involved transformations. In the early studies, the cyclic aryliodoniums are often limited to cyclic diphenyliodoniums, and they are simply employed as arylation reagents. Nowadays, more types of cyclic aryliodoniums including heterocyclic iodoniums and monoaryliodoniums are developed. Polycyclic cyclic aryliodoniums and chiral cyclic aryliodoniums still remains underexplored, although those novel iodoniums may be advantageous in building complex scaffolds and promoting asymmetrical reactions. Importantly, the synthetic application of cyclic aryliodoniums has been expanded to construction of complex polycyclic arenes and chiral biaryl atropisomers both of which are important building blocks in pharmaceuticals, functional materials, and chiral ligands. More recently, cyclic aryliodoniums are explored to act as halogen-bonding donors and subsequently employed as organocatalysts to mediate several types of reactions including Michael addition and Diels–Alder reaction. It will be undoubted that these research areas will continue to attract significant research activity in the future.

Although the chemistry of cyclic diaryliodoniums has been intensively explored, and significant progress has been made in the past decade, there are still some challenges and problems to be resolved. First, the burgeoning cyclic diaryliodoniums are needed to further explore, such as heterocyclic aryliodoniums and polycyclic aryliodoniums. Due to their increasing significance for the pharmaceutical and functional material areas, efficient synthesis of heterocyclic compounds have attracted a great attention. Heterocyclic aryliodoniums serve as a synthon platform to generate heterocycles. However, heteroatom-containing cyclic diaryliodoniums are not easy to acquire by conventional approaches. Polycyclic aryliodoniums provide convenience to obtain polycyclic scaffolds including polycyclic aromatic hydrocarbons and graphene analogues. Due to potential high positive charges and ring strain, the hypervalent iodines embedded in polycyclic aromatic hydrocarbons are still rarely reported. More innovative approaches will come in the near future to obtain heterocyclic and polycyclic aryliodoniums.

From the point of view of environmental and economical issues, time-consuming and stepwise preparation of cyclic aryliodoniums, and relatively limited substrate compatibility would hinder the further application of cyclic diaryliodoniums. Meanwhile, cyclic aryliodoniums involved in arylation reactions usually require transition metals, and few cases without metal are reported. It would be promising if more synthetic approaches are developed to utilize the metal-like chemical property of hypervalent iodine(III), for example, exploring potential free radical reactions and their Lewis acidity. Compared to symmetrical cyclic diaryliodoniums, unsymmetrical species often encounter a poor regioselectivity, which may be solved by installing a directing group in unsymmetrical cyclic aryliodoniums. The advancement of cyclic aryliodoniums to building biarylic atropisomers scaffolds has also been conspicuous. Along biaryl scaffolds, the design and synthesis of heterobiaryl and nonbiaryl atropisomers have gained a rapid progress. Discovery of novel heterobiaryl and nonbiaryl-based cyclic iodoniums to access heterobiaryl and nonbiaryl atropisomers still remains desirable. Finally, construction of chiral cyclic aryliodonium frameworks could expand their application as XB donors to initiate asymmetrical reactions, while most of the currently reported cyclic aryliodoniums are achiral. In the future, we believe that more variants of the atom-economical applications of cyclic aryliodoniums will be achieved, which will facilitate effective synthesis of more structurally complex functionalized compounds.

Acknowledgments

We are grateful to the grant support from National Natural Science Foundation of China (81872440, 21961003, 82260676), the start-up fund of Gannan Medical University (QD202144-2067), and the Jiangxi Provincial Natural Science Foundation (20212ACB206001).

Biographies

Xiaopeng Peng received his Bachelor’s degree in Pharmacy from Gannan Medical University in 2013. He joined the research group of Professor Shijun Wen to jointly cultivate his Master’s degree, and then pursued and obtained his Ph.D. at Southern Medical University in 2021. His work encompasses the development of novel methodologies of cyclic diaryliodoniums applications.

Abdur Rahim was born in Chuadanga, Bangladesh. He graduated in Chemistry, from University of Rajshahi, Bangladesh in 2004 and also received his M.S. degree from the same university under supervision of Prof. Azizul Islam in 2005. He also obtained his Master of Philosophy degree in Organic Chemistry, supervised by Prof. Mosharef in 2017, from the University of Chittagong, Bangladesh. He is currently a doctoral student in the research group of Prof. Gu at the University of Science and Technology of China (USTC).

Weijie Peng received his Bachelor’s degree in Jiangxi Medical College and then obtained his Ph.D. from Central South University in 2008. Research in his group mainly focuses on organic synthesis, drug discovery, bioprinting, and tissue engineering.

Feng Jiang received his B.S. and M.S. degrees from Nanchang University and Zhejiang University of Technology in 1999 and 2003, respectively. From 2007 to 2011, he undertook Ph.D. studies at Shanghai Jiao Tong University under the supervision of Professor Wanbin Zhang. He then joined Professor Aiping Lyu’s group at Hong Kong Baptist University to continue his postdoctoral research (2015–2017). Since 2018, he has worked as a professor in the School of Pharmaceutical Sciences at Gannan Medical University. Professor Jiang’s current research interests include pharmaceutical chemistry and organic synthesis.

Zhenhua Gu is a professor of Chemistry at the University of Science and Technology of China (USTC). He studied chemistry at Nanjing University and received his Ph.D. at the Shanghai Institute of Organic Chemistry in 2007 with Prof. Shengming Ma. After postdoctoral research at University of California, Berkeley, with Professor K. Peter C. Vollhardt and at the University of California Santa Barbara with Professor Armen Zakarian, he began his independent academic career at USTC in 2012. Research in his group mainly focuses on the development of new synthetic methods via transition metal catalysis.

Shijun Wen is a professor at Sun Yat-sen University Cancer Center (SYSUCC). He received a Bachelor’s degee in Chemistry at the School of Chemistry in Jilin University in 1999 and then moved to the Shanghai Institute of Organic Chemistry to pursue his Ph.D. degree with Prof. Zhujun Yao from 1999 to 2004. After postdoctoral research at the University of Southampton with Professor A. Ganesan and the University of Cambridge with Professor Chris Abell, he became a research faculty member in SYSUCC in 2010. His research interests are synthetic methodology to access complex drug-like chemicals from cyclic arylidoniums and medichinal chemistry study to design and synthesize small moles to specifically target epigenetic associated enzymes.

Special Issue

This paper is an additional review for Chem. Rev. 2020, volume 122, issue (20), , “Glycosciences”.

Author Contributions

CRediT: Xiaopeng Peng conceptualization, data curation, writing-original draft, writing-review & editing; Abdur Rahim writing-original draft, writing-review & editing; Weijie Peng writing-review & editing; Feng Jiang writing-review & editing; Zhenhua Gu supervision, writing-original draft, writing-review & editing; Shijun Wen conceptualization, data curation, funding acquisition, supervision, writing-original draft, writing-review & editing.

The authors declare no competing financial interest.

References

  1. Zhdankin V. V.; Stang P. J. Chemistry of Polyvalent Iodine. Chem. Rev. 2008, 108, 5299–5358. 10.1021/cr800332c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Zhdankin V. V.; Stang P. J. Recent Developments in the Chemistry of Polyvalent Iodine Compounds. Chem. Rev. 2002, 102, 2523–2584. 10.1021/cr010003+. [DOI] [PubMed] [Google Scholar]
  3. Zhdankin V. V. Organoiodine(V) Reagents in Organic Synthesis. J. Org. Chem. 2011, 76, 1185–1197. 10.1021/jo1024738. [DOI] [PubMed] [Google Scholar]
  4. Zhao R.; Shi L. Reactions between Diazo Compounds and Hypervalent Iodine(III) Reagents. Angew. Chem., Int. Ed. Engl. 2020, 59, 12282–12292. 10.1002/anie.202003081. [DOI] [PubMed] [Google Scholar]
  5. Yoshimura A.; Zhdankin V. V. Advances in Synthetic Applications of Hypervalent Iodine Compounds. Chem. Rev. 2016, 116, 3328–3435. 10.1021/acs.chemrev.5b00547. [DOI] [PubMed] [Google Scholar]
  6. Yang X. G.; Zheng K.; Zhang C. Electrophilic Hypervalent Trifluoromethylthio-Iodine(III) Reagent. Org. Lett. 2020, 22, 2026–2031. 10.1021/acs.orglett.0c00405. [DOI] [PubMed] [Google Scholar]
  7. Wirth T. Hypervalent Iodine Chemistry in Synthesis: Scope and New Directions. Angew. Chem., Int. Ed. Engl. 2005, 44, 3656–3665. 10.1002/anie.200500115. [DOI] [PubMed] [Google Scholar]
  8. Li Y.; Hari D.; Vita M.; Waser J. Cyclic Hypervalent Iodine Reagents for Atom-Transfer Reactions: Beyond Trifluoromethylation. Angew. Chem., Int. Ed. Engl. 2016, 55, 4436–4454. 10.1002/anie.201509073. [DOI] [PubMed] [Google Scholar]
  9. Stang P. J.; Zhdankin V. V. Organic Polyvalent Iodine Compounds. Chem. Rev. 1996, 96, 1123–1178. 10.1021/cr940424+. [DOI] [PubMed] [Google Scholar]
  10. Stang P. J. Polyvalent Iodine in Organic Chemistry. J. Org. Chem. 2003, 68, 2997–3008. 10.1021/jo030022e. [DOI] [PubMed] [Google Scholar]
  11. Moriarty R. M. Organohypervalent Iodine: Development, Applications, and Future Directions. J. Org. Chem. 2005, 70, 2893–2903. 10.1021/jo050117b. [DOI] [PubMed] [Google Scholar]
  12. Merritt E. A.; Olofsson B. Diaryliodonium Salts: a Journey from Obscurity to Fame. Angew. Chem., Int. Ed. Engl. 2009, 48, 9052–9070. 10.1002/anie.200904689. [DOI] [PubMed] [Google Scholar]
  13. Matsumoto A.; Lee H. J.; Maruoka K. Development of New Radical-mediated Selective Reactions Promoted by Hypervalent Iodine(III) Reagents. Chem. Rec. 2020, 21, 1342–1357. 10.1002/tcr.202000132. [DOI] [PubMed] [Google Scholar]
  14. Lopez-Gomez M.; Mediavilla J. D.; Omar M.; Jimenez-Alonso J. Infrequent Cause of Upper Digestive Hemorrhage: Systemic Mastocytosis. Rev. Esp. Enferm. Dig. 1991, 80, 144–145. [PubMed] [Google Scholar]
  15. Charpentier J.; Fruh N.; Togni A. Electrophilic Trifluoromethylation by Use of Hypervalent Iodine Reagents. Chem. Rev. 2015, 115, 650–682. 10.1021/cr500223h. [DOI] [PubMed] [Google Scholar]
  16. Li X.; Chen P.; Liu G. Recent Advances in Hypervalent Iodine(III)-Catalyzed Functionalization of Alkenes. Beilstein. J. Org. Chem. 2018, 14, 1813–1825. 10.3762/bjoc.14.154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Birchall T.; Myers R. D. Iodine-127 Mössbauer Study of Iodine(III) Interhalogen Cations and Related Compounds. J. Chem. Soc. Dalton. Trans. 1983, 5, 885–888. 10.1039/DT9830000885. [DOI] [Google Scholar]
  18. Ehrlich B. S.; Kaplan M. Iodine Mssbauer Studies of Chemical Bonding in Iodobenzene and Several Polyvalent Iodine Derivatives. J. Chem. Phys. 1971, 54, 612. 10.1063/1.1674886. [DOI] [Google Scholar]
  19. Jones C. 129I Mssbauer Studies of Iodine in the + 1 and + 3 Oxidation States. J. Chem. Phys. 1975, 62, 4343. 10.1063/1.430332. [DOI] [Google Scholar]
  20. Robidas R.; Reinhard D. L.; Huber S. M.; Legault C. Y. A Quantum-chemical Analysis on the Lewis Acidity of Diarylhalonium Ions. ChemPhysChem 2022, e202200634 10.1002/cphc.202200634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Robidas R.; Reinhard D. L.; Legault C. Y.; Huber S. M. Iodine(III)-Based Halogen Bond Donors: Properties and Applications. Chem. Rec. 2021, 21, 1912–1927. 10.1002/tcr.202100119. [DOI] [PubMed] [Google Scholar]
  22. Amey R. L.; Martin J. C. Synthesis and Reaction of Substituted Arylalkoxyiodinanes: Formation of Stable Bromoarylalkoxy and Aryldialkoxy Heterocyclic Derivatives of Tricoordinate Organoiodine(III). J. Org. Chem. 1979, 44, 1779–1784. 10.1021/jo01325a007. [DOI] [Google Scholar]
  23. Ochiai M.; Sueda T.; Miyamoto K.; Kiprof P.; Zhdankin V. V. trans Influences on Hypervalent Bonding of Aryl Lambda(3)-Iodanes: Their Stabilities and Isodesmic Reactions of Benziodoxolones and Benziodazolones. Angew. Chem., Int. Ed. Engl. 2006, 45, 8203–8206. 10.1002/anie.200603055. [DOI] [PubMed] [Google Scholar]
  24. Boelke A.; Finkbeiner P.; Nachtsheim B. J. Atom-Economical Group-Transfer Reactions with Hypervalent Iodine Compounds. Beilstein. J. Org. Chem. 2018, 14, 1263–1280. 10.3762/bjoc.14.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Hartmann C.; Meyer V. Ueber Eine Neue Klasse Jodhaltiger, Stickstofffreier Organischer Basen. Ber. Dstch. Chem, Ges. 1894, 27, 426–432. 10.1002/cber.18940270183. [DOI] [Google Scholar]
  26. Zhu D.; Liu Q.; Luo B.; Chen M.; Pi R.; Huang P.; Wen S. Synthesis of Carbazoles via One-Pot Copper-Catalyzed Amine Insertion into Cyclic Diphenyleneiodoniums as a Strategy to Generate a Drug-Like Chemical Library. Adv. Synth. & Catal. 2013, 355, 2172–2178. 10.1002/adsc.201300271. [DOI] [Google Scholar]
  27. Matsumoto M.; Wada K.; Urakawa K.; Ishikawa H. Diaryliodonium Salt-Mediated Intramolecular C-N Bond Formation Using Boron-Masking N-Hydroxyamides. Org. Lett. 2020, 22, 781–785. 10.1021/acs.orglett.9b04076. [DOI] [PubMed] [Google Scholar]
  28. Xie H.; Yang S.; Zhang C.; Ding M.; Liu M.; Guo J.; Zhang F. Copper-Catalyzed Selective Diphenylation of Carboxylic Acids with Cyclic Diaryliodonium Salts. J. Org. Chem. 2017, 82, 5250–5262. 10.1021/acs.joc.7b00513. [DOI] [PubMed] [Google Scholar]
  29. Peng X.; Sun Z.; Kuang P.; Li L.; Chen J.; Chen J. Copper-Catalyzed Selective Arylation of Nitriles with Cyclic Diaryl Iodonium Salts: Direct Access to Structurally Diversified Diarylmethane Amides with Potential Neuroprotective and Anticancer Activities. Org. Lett. 2020, 22, 5789–5795. 10.1021/acs.orglett.0c01829. [DOI] [PubMed] [Google Scholar]
  30. Hu T.; Ye Z.; Zhu K.; Xu K.; Wu Y.; Zhang F. Synthesis of Tribenzo[b,d,f]azepines via Cascade π-Extended Decarboxylative Annulation Involving Cyclic Diaryliodonium Salts. Org. Lett. 2020, 22, 505–509. 10.1021/acs.orglett.9b04269. [DOI] [PubMed] [Google Scholar]
  31. Liu Z.; Zhu D.; Luo B.; Zhang N.; Liu Q.; Hu Y.; Pi R.; Huang P.; Wen S. Mild Cu(I)-Catalyzed Cascade Reaction of Cyclic Diaryliodoniums, Sodium Azide, and Alkynes: Efficient Synthesis of Triazolophenanthridines. Org. Lett. 2014, 16, 5600–5603. 10.1021/ol502654a. [DOI] [PubMed] [Google Scholar]
  32. Li B.; Chao Z.; Li C.; Gu Z. Cu-Catalyzed Enantioselective Ring Opening of Cyclic Diaryliodoniums toward the Synthesis of Chiral Diarylmethanes. J. Am. Chem. Soc. 2018, 140, 9400–9403. 10.1021/jacs.8b05743. [DOI] [PubMed] [Google Scholar]
  33. Wang M.; Fan Q.; Jiang X. Nitrogen-Iodine Exchange of Diaryliodonium Salts: Access to Acridine and Carbazole. Org. Lett. 2018, 20, 216–219. 10.1021/acs.orglett.7b03564. [DOI] [PubMed] [Google Scholar]
  34. Zhang Y.; Deng B.; Li Z. Inhibition of NADPH Oxidase Increases Defense Enzyme Activities and Improves Maize seed Germination under Pb Stress. Ecotoxicol. Environ. Saf. 2018, 158, 187–192. 10.1016/j.ecoenv.2018.04.028. [DOI] [PubMed] [Google Scholar]
  35. Roy K. K.; Lu J.; Doroshow J. H. Effects of Iodonium Analogs on NADPH Oxidase 1 in Human Colon Cancer Cells. Antioxidants. 2021, 10, 1757. 10.3390/antiox10111757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Reis J.; Massari M.; Marchese S.; Ceccon M.; Aalbers F. S.; Corana F.; Valente S.; Mai A.; Magnani F.; Mattevi A. A Closer Look into NADPH Oxidase Inhibitors: Validation and Insight into Their Mechanism of Action. Redox. Biol. 2020, 32, 101466. 10.1016/j.redox.2020.101466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Pintard C.; Ben Khemis M.; Liu D.; Dang P. M.; Hurtado-Nedelec M.; El-Benna J. Apocynin Prevents GM-CSF-induced-ERK1/2 Activation and -Neutrophil Survival Independently of its Inhibitory Effect on the Phagocyte NADPH Oxidase NOX2. Biochem. Pharmacol. 2020, 177, 113950. 10.1016/j.bcp.2020.113950. [DOI] [PubMed] [Google Scholar]
  38. Konate M. M.; Antony S.; Doroshow J. H. Inhibiting the Activity of NADPH Oxidase in Cancer. Antioxid. Redox. Signal. 2020, 33, 435–454. 10.1089/ars.2020.8046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Kitamoto K.; Miura Y.; Karnan S.; Ota A.; Konishi H.; Hosokawa Y.; Sato K. Inhibition of NADPH Oxidase 2 Induces Apoptosis in Osteosarcoma: The Role of Reactive Oxygen Species in Cell Proliferation. Oncol. Lett. 2018, 15, 7955–7962. 10.3892/ol.2018.8291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Altenhofer S.; Radermacher K. A.; Kleikers P. W.; Wingler K.; Schmidt H. H. Evolution of NADPH Oxidase Inhibitors: Selectivity and Mechanisms for Target Engagement. Antioxid. Redox. Signal. 2015, 23, 406–427. 10.1089/ars.2013.5814. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Lu W.; Hu Y.; Chen G.; Chen Z.; Zhang H.; Wang F.; Feng L.; Pelicano H.; Wang H.; Keating M. J.; et al. Novel Role of NOX in Supporting Aerobic Glycolysis in Cancer Cells with Mitochondrial Dysfunction and as a Potential Target for Cancer Therapy. PLoS. Biol. 2012, 10, e1001326 10.1371/journal.pbio.1001326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Wu H.; Wang Q.; Zhu J. Copper-Catalyzed Enantioselective Arylative Desymmetrization of Prochiral Cyclopentenes with Diaryliodonium Salts. Angew. Chem., Int. Ed. Engl. 2018, 57, 2721–2725. 10.1002/anie.201713329. [DOI] [PubMed] [Google Scholar]
  43. Shibata K.; Takao K. I.; Ogura A. Diaryliodonium Salt-Based Synthesis of N-Alkoxyindolines and Further Insights into the Ishikawa Indole Synthesis. J. Org. Chem. 2021, 86, 10067–10087. 10.1021/acs.joc.1c00820. [DOI] [PubMed] [Google Scholar]
  44. Parida S. K.; Jaiswal S.; Singh P.; Murarka S. Multicomponent Synthesis of Biologically Relevant S-Aryl Dithiocarbamates Using Diaryliodonium Salts. Org. Lett. 2021, 23, 6401–6406. 10.1021/acs.orglett.1c02220. [DOI] [PubMed] [Google Scholar]
  45. Neerbye Berntsen L.; Nova A.; Wragg D. S.; Sandtorv A. H. Cu-catalyzed N-3-Arylation of Hydantoins Using Diaryliodonium Salts. Org. Lett. 2020, 22, 2687–2691. 10.1021/acs.orglett.0c00642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Meng H.; Wen L.; Xu Z.; Li Y.; Hao J.; Zhao Y. Nonafluoro- tert-butoxylation of Diaryliodonium Salts. Org. Lett. 2019, 21, 5206–5210. 10.1021/acs.orglett.9b01813. [DOI] [PubMed] [Google Scholar]
  47. Mayer R. J.; Ofial A. R.; Mayr H.; Legault C. Y. Lewis Acidity Scale of Diaryliodonium Ions toward Oxygen, Nitrogen, and Halogen Lewis Bases. J. Am. Chem. Soc. 2020, 142, 5221–5233. 10.1021/jacs.9b12998. [DOI] [PubMed] [Google Scholar]
  48. Lei G.; Zhang H.; Chen B.; Xu M.; Zhang G. Copper-catalyzed Enantioselective Arylalkynylation of Alkenes. Chem. Sci. 2020, 11, 1623–1628. 10.1039/C9SC04029C. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Elsherbini M.; Moran W. J. Scalable Electrochemical Synthesis of Diaryliodonium Salts. Org. Biomol. Chem. 2021, 19, 4706–4711. 10.1039/D1OB00457C. [DOI] [PubMed] [Google Scholar]
  50. Chen Y.; Gu Y.; Meng H.; Shao Q.; Xu Z.; Bao W.; Gu Y.; Xue X. S.; Zhao Y. Metal-Free C-H Functionalization via Diaryliodonium Salts with a Chemically Robust Dummy Ligand. Angew. Chem., Int. Ed. Engl. 2022, 61, e202201240 10.1002/anie.202201240. [DOI] [PubMed] [Google Scholar]
  51. Chen H.; Wang L.; Han J. Aryl Radical-induced Desulfonylative ipso-Substitution of Diaryliodonium Salts: an Efficient Route to Sterically Hindered Biarylamines. Chem. Commun. 2020, 56, 5697–5700. 10.1039/D0CC01766C. [DOI] [PubMed] [Google Scholar]
  52. Chen H.; Han J.; Wang L. Intramolecular Aryl Migration of Diaryliodonium Salts: Access to ortho-Iodo Diaryl Ethers. Angew. Chem., Int. Ed. Engl. 2018, 57, 12313–12317. 10.1002/anie.201806405. [DOI] [PubMed] [Google Scholar]
  53. An G.; Wang L.; Han J. Palladium Catalyzed Regioselective Cyclization of Arylcarboxylic Acids via Radical Intermediates with Diaryliodonium Salts. Org. Lett. 2021, 23, 8688–8693. 10.1021/acs.orglett.1c03016. [DOI] [PubMed] [Google Scholar]
  54. Altomonte S.; Telu S.; Lu S.; Pike V. W. Pd(0)-Mediated 11C-Carbonylation of Aryl(mesityl)iodonium Salts as a Route to [11C]Arylcarboxylic Acids and Derivatives. J. Org. Chem. 2017, 82, 11925–11932. 10.1021/acs.joc.7b01704. [DOI] [PubMed] [Google Scholar]
  55. Chen Z.; Wang W. F.; Yang H.; Wu X. F. Palladium-Catalyzed Four-Component Carbonylative Cyclization Reaction of Trifluoroacetimidoyl Chlorides, Propargyl Amines, and Diaryliodonium Salts: Access to Trifluoromethyl-Containing Trisubstituted Imidazoles. Org. Lett. 2020, 22, 1980–1984. 10.1021/acs.orglett.0c00328. [DOI] [PubMed] [Google Scholar]
  56. Yuan H.; Du Y.; Liu F.; Guo L.; Sun Q.; Feng L.; Gao H. Tandem Approach to NOBIN Analogues from Arylhydroxylamines and Diaryliodonium Salts via [3,3]-Sigmatropic Rearrangement. Chem. Commun. 2020, 56, 8226–8229. 10.1039/D0CC02919J. [DOI] [PubMed] [Google Scholar]
  57. Ye B.; Zhao J.; Zhao K.; McKenna J. M.; Toste F. D. Chiral Diaryliodonium Phosphate Enables Light Driven Diastereoselective α-C(sp3)-H Acetalization. J. Am. Chem. Soc. 2018, 140, 8350–8356. 10.1021/jacs.8b05962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Zhang X.; Zhao K.; Li N.; Yu J.; Gong L. Z.; Gu Z. Atroposelective Ring Opening of Cyclic Diaryliodonium Salts with Bulky Anilines Controlled by a Chiral Cobalt(III) Anion. Angew. Chem., Int. Ed. Engl. 2020, 59, 19899–19904. 10.1002/anie.202008431. [DOI] [PubMed] [Google Scholar]
  59. Zeitler S. M.; Chakma P.; Golder M. R. Diaryliodonium Salts Facilitate Metal-free Mechanoredox Free Radical Polymerizations. Chem. Sci. 2022, 13, 4131–4138. 10.1039/D2SC00313A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Zaheer M. K.; Gupta E.; Kant R.; Mohanan K. Metal-free α-Arylation of α-Fluoro-α-Nitroacetamides Employing Diaryliodonium Salts. Chem. Commun. 2020, 56, 153–156. 10.1039/C9CC07859B. [DOI] [PubMed] [Google Scholar]
  61. Grushin V. V. Cyclic Diaryliodonium Ions: Old Mysteries Solved and New Applications Envisaged. Chem. Soc. Rev. 2000, 29, 315–324. 10.1039/a909041j. [DOI] [Google Scholar]
  62. Robidas R.; Huber S. M.; Legault C. Y. Computational Investigation of Cyclic Substituted Iodine(III) Halogen Bond Donors. Arkivoc. 2022, 2021, 128–140. 10.24820/ark.5550190.p011.639. [DOI] [Google Scholar]
  63. Besson T.; Fruit C. Recent Advances in Transition-Metal-Free Late-Stage C-H and N-H Arylation of Heteroarenes Using Diaryliodonium Salts. Pharmaceuticals (Basel). 2021, 14, 661. 10.3390/ph14070661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Zhu D.; Chen M.; Li M.; Luo B.; Zhao Y.; Huang P.; Xue F.; Rapposelli S.; Pi R.; Wen S. Discovery of Novel N-Substituted Carbazoles as Neuroprotective Agents with Potent anti-Oxidative Activity. Eur. J. Med. Chem. 2013, 68, 81–88. 10.1016/j.ejmech.2013.07.029. [DOI] [PubMed] [Google Scholar]
  65. Zhu D.; Wu Y.; Wu B.; Luo B.; Ganesan A.; Wu F. H.; Pi R.; Huang P.; Wen S. Three-Component Pd/Cu-Catalyzed Cascade Reactions of Cyclic Iodoniums, Alkynes, and Boronic Acids: an Approach to Methylidenefluorenes. Org. Lett. 2014, 16, 2350–2353. 10.1021/ol5006714. [DOI] [PubMed] [Google Scholar]
  66. Liang W.; Yang Y.; Yang M.; Zhang M.; Li C.; Ran Y.; Lan J.; Bin Z.; You J. Dearomatizing [4 + 1] Spiroannulation of Naphthols: Discovery of Thermally Activated Delayed Fluorescent Materials. Angew. Chem., Int. Ed. Engl. 2021, 60, 3493–3497. 10.1002/anie.202012842. [DOI] [PubMed] [Google Scholar]
  67. Wang M.; Zhang M.; Luo Y.; Liu Z.; Yang C.; Lan J.; Wu D.; You J. Pd(II)-Catalyzed Regioselective Multiple C-H Arylations of 1-Naphthamides with Cyclic Diaryliodonium Salts: One-Step Access to [4]- and [5]Carbohelicenes. Org. Lett. 2020, 22, 135–139. 10.1021/acs.orglett.9b04046. [DOI] [PubMed] [Google Scholar]
  68. Li J.; Xu Q. N.; Wang Z. B.; Li Y.; Liu L. Synthesis of Dibenzofurans from Cyclic Diaryliodonium Triflates and Water via Oxygen-Iodine Exchange Approach. ACS Omega. 2018, 3, 12923–12929. 10.1021/acsomega.8b02345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Xie H.; Ding M.; Liu M.; Hu T.; Zhang F. Synthesis of Functionalized Biaryls and Poly(hetero)aryl Containing Medium-Sized Lactones with Cyclic Diaryliodonium Salts. Org. Lett. 2017, 19, 2600–2603. 10.1021/acs.orglett.7b00933. [DOI] [PubMed] [Google Scholar]
  70. Ye Z.; Li Y.; Xu K.; Chen N.; Zhang F. Cascade π-Extended Decarboxylative Annulation Involving Cyclic Diaryliodonium Salts: Site-Selective Synthesis of Phenanthridines and Benzocarbazoles via a Traceless Directing Group Strategy. Org. Lett. 2019, 21, 9869–9873. 10.1021/acs.orglett.9b03775. [DOI] [PubMed] [Google Scholar]
  71. Wang M.; Wei J.; Fan Q.; Jiang X. Cu(II)-Catalyzed Sulfide Construction: Both Aryl Groups Utilization of Intermolecular and Intramolecular Diaryliodonium Salt. Chem. Commun. 2017, 53, 2918–2921. 10.1039/C6CC09201B. [DOI] [PubMed] [Google Scholar]
  72. Wang Y.; An G.; Wang L.; Han J. Diaryliodonium Salts: Structures and Synthesis. Cur. Org. Chem. 2020, 24, 2070–2105. 10.2174/1385272824999200507124328. [DOI] [Google Scholar]
  73. Wang D.; Li Q.; Li M.; Du Z.; Y Fu. Recent Progress in Arylation Reactions with Diaryliodonium Salts. Curr. Org. Chem. 2021, 25, 1298–1320. 10.2174/1385272825666210512015410. [DOI] [Google Scholar]
  74. Chatterjee N.; Goswami A. Synthesis and Application of Cyclic Diaryliodonium Salts: A Platform for Bifunctionalization in a Single Step. Eur. J. Org. Chem. 2017, 2017, 3023–3032. 10.1002/ejoc.201601651. [DOI] [Google Scholar]
  75. Wang M.; Chen S.; Jiang X. Atom-Economical Applications of Diaryliodonium Salts. Chem. Asian. J. 2018, 13, 2195–2207. 10.1002/asia.201800609. [DOI] [PubMed] [Google Scholar]
  76. Meng B.; Shi Q.; Meng Y.; Chen J.; Cao W.; Wu X. Asymmetric Catalytic Alkynylation of Thiazolones and Azlactones for Synthesis of Quaternary α-amino Acid Precursors. Org. Biomol. Chem. 2021, 19, 5087–5092. 10.1039/D1OB00582K. [DOI] [PubMed] [Google Scholar]
  77. Luo X.; Wang P. Ynonylation of Acyl Radicals by Electroinduced Homolysis of 4-Acyl-1,4-dihydropyridines. Org. Lett. 2021, 23, 4960–4965. 10.1021/acs.orglett.1c01243. [DOI] [PubMed] [Google Scholar]
  78. Le Du E.; Duhail T.; Wodrich M. D.; Scopelliti R.; Fadaei-Tirani F.; Anselmi E.; Magnier E.; Waser J. Structure and Reactivity of N-Heterocyclic Alkynyl Hypervalent Iodine Reagents. Chem.—Eur. J. 2021, 27, 10979–10986. 10.1002/chem.202101475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Amos S. G. E.; Cavalli D.; Le Vaillant F.; Waser J. Direct Photoexcitation of Ethynylbenziodoxolones: An Alternative to Photocatalysis for Alkynylation Reactions. Angew. Chem., Int. Ed. Engl. 2021, 60, 23827–23834. 10.1002/anie.202110257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Bertho S.; Rey-Rodriguez R.; Colas C.; Retailleau P.; Gillaizeau I. Regio- and Stereoselective Iron-Catalyzed Oxyazidation of Enamides Using a Hypervalent Iodine Reagent. Chem.—Eur. J. 2017, 23, 17674–17677. 10.1002/chem.201704499. [DOI] [PubMed] [Google Scholar]
  81. Smyrnov V.; Muriel B.; Waser J. Synthesis of Quinolines via the Metal-free Visible-Light-Mediated Radical Azidation of Cyclopropenes. Org. Lett. 2021, 23, 5435–5439. 10.1021/acs.orglett.1c01775. [DOI] [PubMed] [Google Scholar]
  82. Wu S.; Li J.; He R.; Jia K.; Chen Y. Terminal Trifluoromethylation of Ketones via Selective C-C Cleavage of Cycloalkanols Enabled by Hypervalent Iodine Reagents. Org. Lett. 2021, 23, 9204–9209. 10.1021/acs.orglett.1c03526. [DOI] [PubMed] [Google Scholar]
  83. Shen H.; Liu Z.; Zhang P.; Tan X.; Zhang Z.; Li C. Trifluoromethylation of Alkyl Radicals in Aqueous Solution. J. Am. Chem. Soc. 2017, 139, 9843–9846. 10.1021/jacs.7b06044. [DOI] [PubMed] [Google Scholar]
  84. Murugan A.; Babu V. N.; Polu A.; Sabarinathan N.; Bakthadoss M.; Sharada D. S. Regioselective C3-H Trifluoromethylation of 2H-Indazole under Transition-Metal-Free Photoredox Catalysis. J. Org. Chem. 2019, 84, 7796–7803. 10.1021/acs.joc.9b00676. [DOI] [PubMed] [Google Scholar]
  85. Mestre J.; Lishchynskyi A.; Castillon S.; Boutureira O. Trifluoromethylation of Electron-Rich Alkenyl Iodides with Fluoroform-Derived ″Ligandless″ CuCF3. J. Org. Chem. 2018, 83, 8150–8160. 10.1021/acs.joc.8b00927. [DOI] [PubMed] [Google Scholar]
  86. Kalim J.; Duhail T.; Le T. N.; Vanthuyne N.; Anselmi E.; Togni A.; Magnier E. Merging Hypervalent Iodine and Sulfoximine Chemistry: a New Electrophilic Trifluoromethylation Reagent. Chem. Sci. 2019, 10, 10516–10523. 10.1039/C9SC04289J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Chen Y.; Ma G.; Gong H. Copper-Catalyzed Reductive Trifluoromethylation of Alkyl Iodides with Togni’s Reagent. Org. Lett. 2018, 20, 4677–4680. 10.1021/acs.orglett.8b02005. [DOI] [PubMed] [Google Scholar]
  88. Dahiya A.; Sahoo A. K.; Chakraborty N.; Das B.; Patel B. K. Updates on Hypervalent-Iodine Reagents: Metal-free Functionalisation of Alkenes, Alkynes and Heterocycles. Org. Biomol. Chem. 2022, 20, 2005–2027. 10.1039/D1OB02233D. [DOI] [PubMed] [Google Scholar]
  89. Peng X.; Luo H.; Wu F.; Zhu D.; Ganesan A.; Huang P.; Wen S. Synthesis of Fluorenes with an All-Carbon Quaternary Center via Palladium-Catalyzed Dual Arylation using Cyclic Diaryliodonium Triflates. Adv. Synth. & Catal. 2017, 359, 1152–1156. 10.1002/adsc.201601260. [DOI] [Google Scholar]
  90. Wu Y.; Wu F.; Zhu D.; Luo B.; Wang H.; Hu Y.; Wen S.; Huang P. Pd Catalyzed Insertion of Alkynes into Cyclic Diaryliodoniums: a Direct Access to multi-Substituted Phenanthrenes. Org. Biomol. Chem. 2015, 13, 10386–10391. 10.1039/C5OB01597A. [DOI] [PubMed] [Google Scholar]
  91. Zhu K.; Song Z.; Wang Y.; Zhang F. Synthesis of 2,2’-Dihalobiaryls via Cu-Catalyzed Halogenation of Cyclic Diaryliodonium Salts. Org. Lett. 2020, 22, 9356–9359. 10.1021/acs.orglett.0c03614. [DOI] [PubMed] [Google Scholar]
  92. Zu B.; Ke J.; Guo Y.; He C. Synthesis of Diverse Aryliodine(III) Reagents by Anodic Oxidation. Chin. J. Chem. 2021, 39, 627–632. 10.1002/cjoc.202000501. [DOI] [Google Scholar]
  93. Caspers L. D.; Spils J.; Damrath M.; Lork E.; Nachtsheim B. J. One-Pot Synthesis and Conformational Analysis of Six-Membered Cyclic Iodonium Salts. J. Org. Chem. 2020, 85, 9161–9178. 10.1021/acs.joc.0c01125. [DOI] [PubMed] [Google Scholar]
  94. Damrath M.; Caspers L. D.; Duvinage D.; Nachtsheim B. J. One-Pot Synthesis of Heteroatom-Bridged Cyclic Diaryliodonium Salts. Org. Lett. 2022, 24, 2562–2566. 10.1021/acs.orglett.2c00691. [DOI] [PubMed] [Google Scholar]
  95. Grelier G.; Darses B.; Dauban P. Hypervalent Organoiodine Compounds: from Reagents to Valuable Building Blocks in Synthesis. Beilstein. J. Org. Chem. 2018, 14, 1508–1528. 10.3762/bjoc.14.128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Zhu D.; Wu Z.; Luo B.; Du Y.; Liu P.; Chen Y.; Hu Y.; Huang P.; Wen S. Heterocyclic Iodoniums for the Assembly of Oxygen-Bridged Polycyclic Heteroarenes with Water as the Oxygen Source. Org. Lett. 2018, 20, 4815–4818. 10.1021/acs.orglett.8b01969. [DOI] [PubMed] [Google Scholar]
  97. Boelke A.; Kuczmera T. J.; Caspers L. D.; Lork E.; Nachtsheim B. J. Iodolopyrazolium Salts: Synthesis, Derivatizations, and Applications. Org. Lett. 2020, 22, 7261–7266. 10.1021/acs.orglett.0c02593. [DOI] [PubMed] [Google Scholar]
  98. Antonkin N. S.; Vlasenko Y. A.; Yoshimura A.; Smirnov V. I.; Borodina T. N.; Zhdankin V. V.; Yusubov M. S.; Shafir A.; Postnikov P. S. Preparation and Synthetic Applicability of Imidazole-Containing Cyclic Iodonium Salts. J. Org. Chem. 2021, 86, 7163–7178. 10.1021/acs.joc.1c00483. [DOI] [PubMed] [Google Scholar]
  99. Vlasenko Y. A.; Postnikov P. S.; Trusova M. E.; Shafir A.; Zhdankin V. V.; Yoshimura A.; Yusubov M. S. Synthesis of Five-Membered Iodine-Nitrogen Heterocycles from Benzimidazole-Based Iodonium Salts. J. Org. Chem. 2018, 83, 12056–12070. 10.1021/acs.joc.8b01995. [DOI] [PubMed] [Google Scholar]
  100. Boelke A.; Vlasenko Y. A.; Yusubov M. S.; Nachtsheim B. J.; Postnikov P. S. Thermal Stability of N-Heterocycle-Stabilized Iodanes - a Systematic Investigation. Beilstein. J. Org. Chem. 2019, 15, 2311–2318. 10.3762/bjoc.15.223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Boelke A.; Sadat S.; Lork E.; Nachtsheim B. J. Pseudocyclic bis-N-Heterocycle-Stabilized Iodanes - Synthesis, Characterization and Applications. Chem. Commun. 2021, 57, 7434–7437. 10.1039/D1CC03097C. [DOI] [PubMed] [Google Scholar]
  102. Qurban J.; Elsherbini M.; Alharbi H.; Wirth T. Synthesis, Characterisation, and Reactivity of Novel Pseudocyclic Hypervalent Iodine Reagents with Heteroaryl Carbonyl Substituents. Chem. Commun. 2019, 55, 7998–8000. 10.1039/C9CC03905H. [DOI] [PubMed] [Google Scholar]
  103. Mikhael M.; Adler S. A.; Wengryniuk S. E. Chemoselective Oxidation of Equatorial Alcohols with N-Ligated lambda(3)-Iodanes. Org. Lett. 2019, 21, 5889–5893. 10.1021/acs.orglett.9b02018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Wang X.; Gallardo-Donaire J.; Martin R. Mild ArI-Catalyzed C(sp2)-H or C(sp3)-H Functionalization/C-O Formation: an Intriguing Catalyst-Controlled Selectivity Switch. Angew. Chem., Int. Ed. Engl. 2014, 53, 11084–11087. 10.1002/anie.201407011. [DOI] [PubMed] [Google Scholar]
  105. Zawia E.; Moran W. J. Aqueous DMSO Mediated Conversion of (2-(Arylsulfonyl)vinyl)iodonium Salts to Aldehydes and Vinyl Chlorides. Molecules. 2016, 21, 1073. 10.3390/molecules21081073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Sheng J.; Wang Y.; Su X.; He R.; Chen C. Copper-Catalyzed [2 + 2+2] Modular Synthesis of Multisubstituted Pyridines: Alkenylation of Nitriles with Vinyliodonium Salts. Angew. Chem., Int. Ed. Engl. 2017, 56, 4824–4828. 10.1002/anie.201700696. [DOI] [PubMed] [Google Scholar]
  107. Liu C.; Wang Q. Arylation, Vinylation, and Alkynylation of Electron-Deficient (Hetero)arenes Using Iodonium Salts. Org. Lett. 2016, 18, 5118–5121. 10.1021/acs.orglett.6b02550. [DOI] [PubMed] [Google Scholar]
  108. Kepski K.; Rice C. R.; Moran W. J. Cyclic Vinyl(aryl)iodonium Salts: Synthesis and Reactivity. Org. Lett. 2019, 21, 6936–6939. 10.1021/acs.orglett.9b02540. [DOI] [PubMed] [Google Scholar]
  109. Beringer F. M.; Ganis P.; Avitabile G.; Jaffe H. Synthesis, Structure, and Reactions of a Benziodolium Cation. J. Org. Chem. 1972, 37, 879–886. 10.1021/jo00971a015. [DOI] [Google Scholar]
  110. Wang H.; Liang L.; Guo Z.; Peng H.; Qiao S.; Saha N.; Zhu D.; Zeng W.; Chen Y.; Huang P.; et al. Highly Reactive Cyclic Monoaryl Iodoniums Tuned as Carbene Generators Couple with Nucleophiles under Metal-Free Conditions. iScience. 2020, 23, 101307. 10.1016/j.isci.2020.101307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Wu B.; Yoshikai N. Conversion of 2-Iodobiaryls into 2,2’-Diiodobiaryls via Oxidation-Iodination Sequences: A Versatile Route to Ladder-Type Heterofluorenes. Angew. Chem., Int. Ed. Engl. 2015, 54, 8736–8739. 10.1002/anie.201503134. [DOI] [PubMed] [Google Scholar]
  112. Li G.; Rheingold A. L.; Protasiewicz J. D. Remote Substituents as Potential Control Elements for the Solid-State Structures of Hypervalent Iodine(III) Compounds. Inorg. Chem. 2021, 60, 7865–7875. 10.1021/acs.inorgchem.1c00339. [DOI] [PubMed] [Google Scholar]
  113. Portela S.; Cabrera-Trujillo J. J.; Fernandez I. Catalysis by Bidentate Iodine(III)-Based Halogen Donors: Surpassing the Activity of Strong Lewis Acids. J. Org. Chem. 2021, 86, 5317–5326. 10.1021/acs.joc.1c00534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Heinen F.; Reinhard D. L.; Engelage E.; Huber S. M. A Bidentate Iodine(III)-Based Halogen-Bond Donor as a Powerful Organocatalyst. Angew. Chem., Int. Ed. Engl. 2021, 60, 5069–5073. 10.1002/anie.202013172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Stang P. J.; Zhdankin V. V. Preparation and Characterization of a Macrocyclic Tetraaryltetraiodonium Compound, Cyclo(Ar4I4)4+.4X. A Unique, Charged, Cationic Molecular Box. J. Am. Chem. Soc. 1993, 115, 9808–9809. 10.1021/ja00074a061. [DOI] [Google Scholar]
  116. Radhakrishnan U.; Stang P. J. Synthesis and Characterization of Cationic Iodonium Macrocycles. J. Org. Chem. 2003, 68, 9209–9213. 10.1021/jo030246x. [DOI] [PubMed] [Google Scholar]
  117. Li G.; Smith R.; Gembicky M.; Rheingold A. L.; Protasiewicz J. D. Sterically Crowded 1,4-Diiodobenzene as a Precursor to Difunctional Hypervalent Iodine Compounds. Chem. Commun. 2022, 58, 1159–1162. 10.1039/D1CC06486J. [DOI] [PubMed] [Google Scholar]
  118. Tan Q.; Zhou D.; Zhang T.; Liu B.; Xu B. Iodine-Doped Sumanene and its Application for the Synthesis of Chalcogenasumanenes and Silasumanenes. Chem. Commun. 2017, 53, 10279–10282. 10.1039/C7CC05885C. [DOI] [PubMed] [Google Scholar]
  119. Parra A. Chiral Hypervalent Iodines: Active Players in Asymmetric Synthesis. Chem. Rev. 2019, 119, 12033–12088. 10.1021/acs.chemrev.9b00338. [DOI] [PubMed] [Google Scholar]
  120. Hui C.; Pu F.; Xu J. Metal-Catalyzed Asymmetric Michael Addition in Natural Product Synthesis. Chem.—Eur. J. 2017, 23, 4023–4036. 10.1002/chem.201604110. [DOI] [PubMed] [Google Scholar]
  121. Dong L.; Chen F. E. Asymmetric Catalysis in Direct Nitromethane-free Henry Reactions. RSC Adv. 2020, 10, 2313–2326. 10.1039/C9RA10263A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Pan Y.; Liu L.; Guan F.; Li E.; Jin J.; Li J.; Che Y.; Liu G. Characterization of a Prenyltransferase for Iso-A82775C Biosynthesis and Generation of New Congeners of Chloropestolides. ACS Chem. Biol. 2018, 13, 703–711. 10.1021/acschembio.7b01059. [DOI] [PubMed] [Google Scholar]
  123. Muniz K.; Barreiro L.; Romero R. M.; Martinez C. Catalytic Asymmetric Diamination of Styrenes. J. Am. Chem. Soc. 2017, 139, 4354–4357. 10.1021/jacs.7b01443. [DOI] [PubMed] [Google Scholar]
  124. Zhang H.; Wirth T. Oxidation of BINOLs by Hypervalent Iodine Reagents: Facile Synthesis of Xanthenes and Lactones. Chem.—Eur. J. 2022, 28, e202200181 10.1002/chem.202200181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Zhang H.; Cormanich R. A.; Wirth T. Chiral Ligands in Hypervalent Iodine Compounds: Synthesis and Structures of Binaphthyl-Based λ3 -Iodanes. Chem.—Eur. J. 2022, 28, e202103623 10.1002/chem.202103623. [DOI] [PubMed] [Google Scholar]
  126. Murai T.; Hayashi T.; Yamada K.; Maekawa Y.; Minoura M. Fluorinative Hydrolysis of Phosphorothioic Acid Esters with a Binaphthyl Group through Axis-to-Center Chirality Transfer Leading to the Formation of P-Chiral Phosphorothioic Monofluoridic Acid Salts. Chem. Commun. 2014, 50, 12473–12475. 10.1039/C4CC05198J. [DOI] [PubMed] [Google Scholar]
  127. Ochiai M.; Kitagawa Y.; Takayama N.; Takaoka Y.; Shiro M. Synthesis of Chiral Diaryliodonium Salts, 1,1’-Binaphthyl-2-yl(phenyl)iodonium Tetrafluoroborates: Asymmetric α-Phenylation of β-Keto Ester Enolates. J. Am. Chem. Soc. 1999, 121, 9233–9234. 10.1021/ja992236c. [DOI] [Google Scholar]
  128. Yoshida Y.; Fujimura T.; Mino T.; Sakamoto M. Chiral Binaphthyl-Based Iodonium Salt (Hypervalent Iodine(III)) as Hydrogen- and Halogen-Bonding Bifunctional Catalyst: Insight into Abnormal Counteranion Effect and Asymmetric Synthesis of N,S-Acetals. Adv. Synth. & Catal. 2022, 364, 1091–1098. 10.1002/adsc.202101380. [DOI] [Google Scholar]
  129. Zhu D.; Liu P.; Lu W.; Wang H.; Luo B.; Hu Y.; Huang P.; Wen S. Relayed Regioselective Alkynylation/Olefination of Unsymmetrical Cyclic Diaryliodonium Species Catalyzed by Cu and Pd: Affording Fluorescent Cytotoxic Benzoxazoles. Chem.—Eur. J. 2015, 21, 18915–18920. 10.1002/chem.201503791. [DOI] [PubMed] [Google Scholar]
  130. Wang M.; Chen S.; Jiang X. Construction of Functionalized Annulated Sulfone via SO2/I Exchange of Cyclic Diaryliodonium Salts. Org. Lett. 2017, 19, 4916–4919. 10.1021/acs.orglett.7b02388. [DOI] [PubMed] [Google Scholar]
  131. Vase K. H.; Holm A. H.; Norrman K.; Pedersen S. U.; Daasbjerg K. Covalent Grafting of Glassy Carbon Electrodes with Diaryliodonium Salts: New Aspects. Langmuir. 2007, 23, 3786–3793. 10.1021/la0629227. [DOI] [PubMed] [Google Scholar]
  132. Riedmuller S.; Nachtsheim B. J. Palladium-Catalyzed Synthesis of N-Arylated Carbazoles using Anilines and Cyclic Diaryliodonium Salts. Beilstein. J. Org. Chem. 2013, 9, 1202–1209. 10.3762/bjoc.9.136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Purkait N.; Kervefors G.; Linde E.; Olofsson B. Regiospecific N-Arylation of Aliphatic Amines under Mild and Metal-Free Reaction Conditions. Angew. Chem., Int. Ed. Engl. 2018, 57, 11427–11431. 10.1002/anie.201807001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Prakash M.; Muthusamy S.; Kesavan V. Copper(I) Bromide Catalyzed Arylation of Cyclic Enamides and Naphthyl-1-Acetamides using Diaryliodonium Salts. J. Org. Chem. 2014, 79, 7836–7843. 10.1021/jo501162s. [DOI] [PubMed] [Google Scholar]
  135. Pan J. L.; Chen T.; Zhang Z. Q.; Li Y. F.; Zhang X. M.; Zhang F. M. A Cu-Mediated One-Pot Michael Addition/α-Arylation Strategy using a Diaryliodonium Salt: a Direct and Efficient Approach to α-Aryl-β-Substituted Cyclic Ketone Scaffolds. Chem. Commun. 2016, 52, 2382–2385. 10.1039/C5CC09837H. [DOI] [PubMed] [Google Scholar]
  136. Lindstedt E.; Stridfeldt E.; Olofsson B. Mild Synthesis of Sterically Congested Alkyl Aryl Ethers. Org. Lett. 2016, 18, 4234–4237. 10.1021/acs.orglett.6b01975. [DOI] [PubMed] [Google Scholar]
  137. Li Q.; Zhang M.; Zhan S.; Gu Z. Copper-Catalyzed Enantioselective Ring-Opening of Cyclic Diaryliodoniums and O-Alkylhydroxylamines. Org. Lett. 2019, 21, 6374–6377. 10.1021/acs.orglett.9b02267. [DOI] [PubMed] [Google Scholar]
  138. Gao Z.; Qian J.; Yang H.; Hang X. C.; Zhang J.; Jiang G. Chiral Bronsted Acid-Catalyzed Dynamic Kinetic Resolution of Atropisomeric ortho-Formyl Naphthamides. Chem. Commun. 2020, 56, 7265–7268. 10.1039/D0CC02380A. [DOI] [PubMed] [Google Scholar]
  139. Fugard A. J.; Lahdenpera A. S. K.; Tan J. S. J.; Mekareeya A.; Paton R. S.; Smith M. D. Hydrogen-Bond-Enabled Dynamic Kinetic Resolution of Axially Chiral Amides Mediated by a Chiral Counterion. Angew. Chem., Int. Ed. Engl. 2019, 58, 2795–2798. 10.1002/anie.201814362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Faseke V. C.; Sparr C. Stereoselective Arene-Forming Aldol Condensation: Synthesis of Axially Chiral Aromatic Amides. Angew. Chem., Int. Ed. Engl. 2016, 55, 7261–7264. 10.1002/anie.201602689. [DOI] [PubMed] [Google Scholar]
  141. Di Iorio N.; Crotti S.; Bencivenni G. Organocatalytic Desymmetrization Reactions for the Synthesis of Axially Chiral Compounds. Chem. Rec. 2019, 19, 2095–2104. 10.1002/tcr.201800194. [DOI] [PubMed] [Google Scholar]
  142. Bao X.; Rodriguez J.; Bonne D. Enantioselective Synthesis of Atropisomers with Multiple Stereogenic Axes. Angew. Chem., Int. Ed. Engl. 2020, 59, 12623–12634. 10.1002/anie.202002518. [DOI] [PubMed] [Google Scholar]
  143. Asada T.; Hoshimoto Y.; Kawakita T.; Kinoshita T.; Ogoshi S. Axial Chirality around N-P Bonds Induced by Complexation between E(C6F5)3 (E = B, Al) and an N-Phosphine Oxide-Substituted Imidazolinylidene: A Key Intermediate in the Catalytic Phosphinoylation of CO2. J. Org. Chem. 2020, 85, 14333–14341. 10.1021/acs.joc.9b03210. [DOI] [PubMed] [Google Scholar]
  144. Moliterno M.; Cari R.; Puglisi A.; Antenucci A.; Sperandio C.; Moretti E.; Di Sabato A.; Salvio R.; Bella M. Quinine-Catalyzed Asymmetric Synthesis of 2,2’-Binaphthol-Type Biaryls under Mild Reaction Conditions. Angew. Chem., Int. Ed. Engl. 2016, 55, 6525–6529. 10.1002/anie.201601660. [DOI] [PubMed] [Google Scholar]
  145. Mancinelli M.; Perticarari S.; Prati L.; Mazzanti A. Conformational Analysis and Absolute Configuration of Axially Chiral 1-Aryl and 1,3-Bisaryl-xanthines. J. Org. Chem. 2017, 82, 6874–6885. 10.1021/acs.joc.7b01010. [DOI] [PubMed] [Google Scholar]
  146. Luo J.; Zhang T.; Wang L.; Liao G.; Yao Q. J.; Wu Y. J.; Zhan B. B.; Lan Y.; Lin X. F.; Shi B. F. Enantioselective Synthesis of Biaryl Atropisomers by Pd-Catalyzed C-H Olefination using Chiral Spiro Phosphoric Acid Ligands. Angew. Chem., Int. Ed. Engl. 2019, 58, 6708–6712. 10.1002/anie.201902126. [DOI] [PubMed] [Google Scholar]
  147. Liu Y.; Wu X.; Li S.; Xue L.; Shan C.; Zhao Z.; Yan H. Organocatalytic Atroposelective Intramolecular [4 + 2] Cycloaddition: Synthesis of Axially Chiral Heterobiaryls. Angew. Chem., Int. Ed. Engl. 2018, 57, 6491–6495. 10.1002/anie.201801824. [DOI] [PubMed] [Google Scholar]
  148. Link A.; Sparr C. Organocatalytic Atroposelective Aldol Condensation: Synthesis of Axially Chiral Biaryls by Arene Formation. Angew. Chem., Int. Ed. Engl. 2014, 53, 5458–5461. 10.1002/anie.201402441. [DOI] [PubMed] [Google Scholar]
  149. Liao G.; Li B.; Chen H. M.; Yao Q. J.; Xia Y. N.; Luo J.; Shi B. F. Pd-Catalyzed Atroposelective C-H Allylation through β-O Elimination: Diverse Synthesis of Axially Chiral Biaryls. Angew. Chem., Int. Ed. Engl. 2018, 57, 17151–17155. 10.1002/anie.201811256. [DOI] [PubMed] [Google Scholar]
  150. Li Z.; Wang X.; Cui Y. M.; Ma J. H.; Fang L. L.; Han L. L.; Yang Q.; Xu Z.; Xu L. W. Combined Dynamic Kinetic Resolution and C-H Functionalization for Facile Synthesis of Non-Biaryl-Atropisomer-Type Axially Chiral Organosilanes. Chem.—Eur. J. 2021, 27, 4336–4340. 10.1002/chem.202100237. [DOI] [PubMed] [Google Scholar]
  151. Li X.; Parkin S. R.; Lehmler H. J. Absolute Configuration of 2,2’,3,3′,6-Pentachlorinatedbiphenyl (PCB 84) Atropisomers. Environ. Sci. Pollut. Res. Int. 2018, 25, 16402–16410. 10.1007/s11356-017-9259-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Li T.-Z.; Liu S.-J.; Tan W.; Shi F. Catalytic Asymmetric Construction of Axially Chiral Indole-Based Frameworks: An Emerging Area. Chem.—Eur. J. 2020, 26, 15779–15792. 10.1002/chem.202001397. [DOI] [PubMed] [Google Scholar]
  153. Lavarda G.; Bhattacharjee N.; Brancato G.; Torres T.; Bottari G. Enabling Racemization of Axially Chiral Subphthalocyanine-Tetracyanobutadiene-Aniline Enantiomers by Triplet State Photogeneration. Angew. Chem., Int. Ed. Engl. 2020, 59, 21224–21229. 10.1002/anie.202010133. [DOI] [PubMed] [Google Scholar]
  154. LaPlante S. R.; Fader L. D.; Fandrick K. R.; Fandrick D. R.; Hucke O.; Kemper R.; Miller S. P. F.; Edwards P. J. Assessing Atropisomer Axial Chirality in Drug Discovery and Development. J. Med. Chem. 2011, 54, 7005–7022. 10.1021/jm200584g. [DOI] [PubMed] [Google Scholar]
  155. LaPlante S. R.; Edwards P. J.; Fader L. D.; Jakalian A.; Hucke O. Revealing Atropisomer Axial Chirality in Drug Discovery. ChemMedChem. 2011, 6, 505–513. 10.1002/cmdc.201000485. [DOI] [PubMed] [Google Scholar]
  156. Kitagawa O. Chiral Pd-Catalyzed Enantioselective Syntheses of Various N-C Axially Chiral Compounds and Their Synthetic Applications. Acc. Chem. Res. 2021, 54, 719–730. 10.1021/acs.accounts.0c00767. [DOI] [PubMed] [Google Scholar]
  157. Jin L.; Zhang P.; Li Y.; Yu X.; Shi B. F. Atroposelective Synthesis of Conjugated Diene-Based Axially Chiral Styrenes via Pd(II)-Catalyzed Thioether-Directed Alkenyl C-H Olefination. J. Am. Chem. Soc. 2021, 143, 12335–12344. 10.1021/jacs.1c06236. [DOI] [PubMed] [Google Scholar]
  158. Jiang F.; Chen K. W.; Wu P.; Zhang Y. C.; Jiao Y.; Shi F. A Strategy for Synthesizing Axially Chiral Naphthyl-Indoles: Catalytic Asymmetric Addition Reactions of Racemic Substrates. Angew. Chem., Int. Ed. Engl. 2019, 58, 15104–15110. 10.1002/anie.201908279. [DOI] [PubMed] [Google Scholar]
  159. Ikeura Y.; Ishichi Y.; Tanaka T.; Fujishima A.; Murabayashi M.; Kawada M.; Ishimaru T.; Kamo I.; Doi T.; Natsugari H. Axially Chiral N-Benzyl-N,7-Dimethyl-5-Phenyl-1,7-Naphthyridine-6-Carboxamide Derivatives as Tachykinin NK1 Receptor Antagonists: Determination of the Absolute Stereochemical Requirements. J. Med. Chem. 1998, 41, 4232–4239. 10.1021/jm980042m. [DOI] [PubMed] [Google Scholar]
  160. Iida A.; Matsuoka M.; Hasegawa H.; Vanthuyne N.; Farran D.; Roussel C.; Kitagawa O. N-C Axially Chiral Compounds with an ortho-Fluoro Substituent and Steric Discrimination between Hydrogen and Fluorine Atoms Based on a Diastereoselective Model Reaction. J. Org. Chem. 2019, 84, 3169–3175. 10.1021/acs.joc.8b03043. [DOI] [PubMed] [Google Scholar]
  161. Hasegawa F.; Kawamura K.; Tsuchikawa H.; Murata M. Stable C-N Axial Chirality in 1-Aryluracil Scaffold and Differences in In Vitro Metabolic Clearance between Atropisomers of PDE4 Inhibitor. Bioorg. Med. Chem. 2017, 25, 4506–4511. 10.1016/j.bmc.2017.06.042. [DOI] [PubMed] [Google Scholar]
  162. Glunz P. W. Recent Encounters with Atropisomerism in Drug Discovery. Bioorg. Med. Chem. Lett. 2018, 28, 53–60. 10.1016/j.bmcl.2017.11.050. [DOI] [PubMed] [Google Scholar]
  163. Zhu S.; Chen Y. H.; Wang Y. B.; Yu P.; Li S. Y.; Xiang S. H.; Wang J. Q.; Xiao J.; Tan B. Organocatalytic Atroposelective Construction of Axially Chiral Arylquinones. Nat. Commun. 2019, 10, 4268. 10.1038/s41467-019-12269-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Zhao Y.; Ku C. F.; Xu X. Y.; Tsang N. Y.; Zhu Y.; Zhao C. L.; Liu K. L.; Li C. C.; Rong L.; Zhang H. J. Stable Axially Chiral Isomers of Arylnaphthalene Lignan Glycosides with Antiviral Potential Discovered from Justicia procumbens. J. Org. Chem. 2021, 86, 5568–5583. 10.1021/acs.joc.1c00068. [DOI] [PubMed] [Google Scholar]
  165. Zhang J. W.; Xiang S. H.; Li S.; Tan B. Copper-Catalyzed Synthesis of Axially Chiral Biaryls with Diaryliodonium Salts as Arylation Reagents. Molecules. 2021, 26, 3223. 10.3390/molecules26113223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Zhang C. L.; Gao Y. Y.; Wang H. Y.; Zhou B. A.; Ye S. Enantioselective Synthesis of Axially Chiral Benzothiophene/Benzofuran-Fused Biaryls by N-Heterocyclic Carbene Catalyzed Arene Formation. Angew. Chem., Int. Ed. Engl. 2021, 60, 13918–13922. 10.1002/anie.202103415. [DOI] [PubMed] [Google Scholar]
  167. Yildirim A.; Konuklar F. A.; Catak S.; Van Speybroeck V.; Waroquier M.; Dogan I.; Aviyente V. Solvent-Catalyzed Ring-Chain-Ring Tautomerization in Axially Chiral Compounds. Chem.—Eur. J. 2012, 18, 12725–12732. 10.1002/chem.201200363. [DOI] [PubMed] [Google Scholar]
  168. Yao Q. J.; Xie P. P.; Wu Y. J.; Feng Y. L.; Teng M. Y.; Hong X.; Shi B. F. Enantioselective Synthesis of Atropisomeric Anilides via Pd(II)-Catalyzed Asymmetric C-H Olefination. J. Am. Chem. Soc. 2020, 142, 18266–18276. 10.1021/jacs.0c09400. [DOI] [PubMed] [Google Scholar]
  169. Yan S.; Xia W.; Li S.; Song Q.; Xiang S. H.; Tan B. Michael Reaction Inspired Atroposelective Construction of Axially Chiral Biaryls. J. Am. Chem. Soc. 2020, 142, 7322–7327. 10.1021/jacs.0c01963. [DOI] [PubMed] [Google Scholar]
  170. Xu K.; Li W.; Zhu S.; Zhu T. Atroposelective Arene Formation by Carbene-Catalyzed Formal [4 + 2] Cycloaddition. Angew. Chem., Int. Ed. Engl. 2019, 58, 17625–17630. 10.1002/anie.201910049. [DOI] [PubMed] [Google Scholar]
  171. Xia W.; An Q. J.; Xiang S. H.; Li S.; Wang Y. B.; Tan B. Chiral Phosphoric Acid Catalyzed Atroposelective C-H Amination of Arenes. Angew. Chem., Int. Ed. Engl. 2020, 59, 6775–6779. 10.1002/anie.202000585. [DOI] [PubMed] [Google Scholar]
  172. Wang Y. B.; Wu Q. H.; Zhou Z. P.; Xiang S. H.; Cui Y.; Yu P.; Tan B. Asymmetric Construction of Axially Chiral 2-Arylpyrroles by Chirality Transfer of Atropisomeric Alkenes. Angew. Chem., Int. Ed. Engl. 2019, 58, 13443–13447. 10.1002/anie.201907470. [DOI] [PubMed] [Google Scholar]
  173. Song R.; Xie Y.; Jin Z.; Chi Y. R. Carbene-Catalyzed Asymmetric Construction of Atropisomers. Angew. Chem., Int. Ed. Engl. 2021, 60, 26026–26037. 10.1002/anie.202108630. [DOI] [PubMed] [Google Scholar]
  174. Shen D.; Xu Y.; Shi S. L. A Bulky Chiral N-Heterocyclic Carbene Palladium Catalyst Enables Highly Enantioselective Suzuki-Miyaura Cross-Coupling Reactions for the Synthesis of Biaryl Atropisomers. J. Am. Chem. Soc. 2019, 141, 14938–14945. 10.1021/jacs.9b08578. [DOI] [PubMed] [Google Scholar]
  175. Shan G.; Flegel J.; Li H.; Merten C.; Ziegler S.; Antonchick A. P.; Waldmann H. C-H Bond Activation for the Synthesis of Heterocyclic Atropisomers Yields Hedgehog Pathway Inhibitors. Angew. Chem., Int. Ed. Engl. 2018, 57, 14250–14254. 10.1002/anie.201809680. [DOI] [PubMed] [Google Scholar]
  176. Shaaban S.; Li H.; Otte F.; Strohmann C.; Antonchick A. P.; Waldmann H. Enantioselective Synthesis of Five-Membered-Ring Atropisomers with a Chiral Rh(III) Complex. Org. Lett. 2020, 22, 9199–9202. 10.1021/acs.orglett.0c03355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Quinonero O.; Jean M.; Vanthuyne N.; Roussel C.; Bonne D.; Constantieux T.; Bressy C.; Bugaut X.; Rodriguez J. Combining Organocatalysis with Central-to-Axial Chirality Conversion: Atroposelective Hantzsch-Type Synthesis of 4-Arylpyridines. Angew. Chem., Int. Ed. Engl. 2016, 55, 1401–1405. 10.1002/anie.201509967. [DOI] [PubMed] [Google Scholar]
  178. Qiu H.; Shuai B.; Wang Y. Z.; Liu D.; Chen Y. G.; Gao P. S.; Ma H. X.; Chen S.; Mei T. S. Enantioselective Ni-Catalyzed Electrochemical Synthesis of Biaryl Atropisomers. J. Am. Chem. Soc. 2020, 142, 9872–9878. 10.1021/jacs.9b13117. [DOI] [PubMed] [Google Scholar]
  179. Qin T.; Skraba-Joiner S. L.; Khalil Z. G.; Johnson R. P.; Capon R. J.; Porco J. A. Jr. Atropselective Syntheses of (−) and (+) Rugulotrosin A Utilizing Point-To-Axial Chirality Transfer. Nat. Chem. 2015, 7, 234–240. 10.1038/nchem.2173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Peluso P.; Mamane V.; Aubert E.; Cossu S. Recent Trends and Applications in Liquid-phase Chromatography Enantioseparation of Atropisomers. Electrophoresis. 2017, 38, 1830–1850. 10.1002/elps.201600502. [DOI] [PubMed] [Google Scholar]
  181. Newton C. G.; Braconi E.; Kuziola J.; Wodrich M. D.; Cramer N. Axially Chiral Dibenzazepinones by a Palladium(0)-Catalyzed Atropo-enantioselective C-H Arylation. Angew. Chem., Int. Ed. Engl. 2018, 57, 11040–11044. 10.1002/anie.201806527. [DOI] [PubMed] [Google Scholar]
  182. Peng X.; Li L.; Ren Y.; Xue H.; Liu J.; Wen S.; Chen J. Synthesis of N-Carbonyl Acridanes as Highly Potent Inhibitors of Tubulin Polymerization via One-Pot Copper-Catalyzed Dual Arylation of Nitriles with Cyclic Diphenyl Iodoniums. Adv. Synth. & Catal. 2020, 362, 2030–2038. 10.1002/adsc.201901460. [DOI] [Google Scholar]
  183. Zheng C.; You S. L. Catalytic Asymmetric Dearomatization (CADA) Reaction-Enabled Total Synthesis of Indole-Based Natural Products. Nat. Prod. Rep. 2019, 36, 1589–1605. 10.1039/C8NP00098K. [DOI] [PubMed] [Google Scholar]
  184. Vaca J.; Salazar F.; Ortiz A.; Sansinenea E. Indole Alkaloid Derivatives as Building Blocks of Natural Products from Bacillus Thuringiensis and Bacillus Velezensis and Their Antibacterial and Antifungal Activity Study. J. Antibiot. (Tokyo). 2020, 73, 798–802. 10.1038/s41429-020-0333-2. [DOI] [PubMed] [Google Scholar]
  185. Patil P.; Thakur A.; Sharma A.; Flora S. J. S. Natural Products and Their Derivatives as Multifunctional Ligands against Alzheimer’s Disease. Drug. Dev. Res. 2020, 81, 165–183. 10.1002/ddr.21587. [DOI] [PubMed] [Google Scholar]
  186. Liu Y.; Cui Y.; Lu L.; Gong Y.; Han W.; Piao G. Natural Indole-Containing Alkaloids and Their Antibacterial Activities. Arch. Pharm. 2020, 353, 2000120 10.1002/ardp.202000120. [DOI] [PubMed] [Google Scholar]
  187. Gentry E. C.; Rono L. J.; Hale M. E.; Matsuura R.; Knowles R. R. Enantioselective Synthesis of Pyrroloindolines via Noncovalent Stabilization of Indole Radical Cations and Applications to the Synthesis of Alkaloid Natural Products. J. Am. Chem. Soc. 2018, 140, 3394–3402. 10.1021/jacs.7b13616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Wan Y.; Li Y.; Yan C.; Yan M.; Tang Z. Indole: A Privileged Scaffold for the Design of Anti-Cancer Agents. Eur. J. Med. Chem. 2019, 183, 111691. 10.1016/j.ejmech.2019.111691. [DOI] [PubMed] [Google Scholar]
  189. Surur A. S.; Huluka S. A.; Mitku M. L.; Asres K. Indole: The After Next Scaffold of Antiplasmodial Agents?. Drug. Des. Devel. Ther. 2020, 14, 4855–4867. 10.2147/DDDT.S278588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Song F.; Li Z.; Bian Y.; Huo X.; Fang J.; Shao L.; Zhou M. Indole/Isatin-Containing Hybrids as Potential Antibacterial Agents. Arch. Pharm. 2020, 353, 2000143 10.1002/ardp.202000143. [DOI] [PubMed] [Google Scholar]
  191. Rathi A. K.; Syed R.; Singh V.; Shin H.-S.; Patel R. V. Kinase Inhibitor Indole Derivatives as Anticancer Agents: A Patent Review. Recent. Pat. Anticancer. Drug. Discovery 2017, 12, 55–72. 10.2174/1574892811666161003112119. [DOI] [PubMed] [Google Scholar]
  192. Norwood V. M. t.; Huigens R. W. 3rd. Harnessing the Chemistry of the Indole Heterocycle to Drive Discoveries in Biology and Medicine. Chembiochem. 2019, 20, 2273–2297. 10.1002/cbic.201800768. [DOI] [PubMed] [Google Scholar]
  193. Jia Y.; Wen X.; Gong Y.; Wang X. Current Scenario of Indole Derivatives with Potential Anti-Drug-Resistant Cancer Activity. Eur. J. Med. Chem. 2020, 200, 112359. 10.1016/j.ejmech.2020.112359. [DOI] [PubMed] [Google Scholar]
  194. Yuan Z.; Lin C.; He Y.; Tao B.; Chen M.; Zhang J.; Liu P.; Cai K. Near-Infrared Light-Triggered Nitric-Oxide-Enhanced Photodynamic Therapy and Low-Temperature Photothermal Therapy for Biofilm Elimination. ACS Nano 2020, 14, 3546–3562. 10.1021/acsnano.9b09871. [DOI] [PubMed] [Google Scholar]
  195. Li C.; Iscen A.; Palmer L. C.; Schatz G. C.; Stupp S. I. Light-Driven Expansion of Spiropyran Hydrogels. J. Am. Chem. Soc. 2020, 142, 8447–8453. 10.1021/jacs.0c02201. [DOI] [PubMed] [Google Scholar]
  196. Wu Y.; Peng X.; Luo B.; Wu F.; Liu B.; Song F.; Huang P.; Wen S. Palladium Catalyzed Dual C-H Functionalization of Indoles with Cyclic Diaryliodoniums, An Approach to Ring Fused Carbazole Derivatives. Org. Biomol. Chem. 2014, 12, 9777–9780. 10.1039/C4OB02170C. [DOI] [PubMed] [Google Scholar]
  197. Yu L.; Liu M.; Chen F.; Xu Q. Heterocycles from Methylenecyclopropanes. Org. Biomol. Chem. 2015, 13, 8379–8392. 10.1039/C5OB00868A. [DOI] [PubMed] [Google Scholar]
  198. Vitaku E.; Smith D. T.; Njardarson J. T. Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem. 2014, 57, 10257–10274. 10.1021/jm501100b. [DOI] [PubMed] [Google Scholar]
  199. Tonk R. K.; Bawa S.; Kumar D. Therapeutic Potential of Cinnoline Core: A Comprehensive Review. Mini. Rev. Med. Chem. 2020, 20, 196–218. 10.2174/1389557519666191011095858. [DOI] [PubMed] [Google Scholar]
  200. Sharma P. C.; Bansal K. K.; Sharma A.; Sharma D.; Deep A. Thiazole-Containing Compounds as Therapeutic Targets for Cancer Therapy. Eur. J. Med. Chem. 2020, 188, 112016. 10.1016/j.ejmech.2019.112016. [DOI] [PubMed] [Google Scholar]
  201. Salvio R.; Placidi S.; Bella M. Benzazetidines and Related Compounds: Synthesis and Potential. Chem.—Eur. J. 2020, 26, 10157–10174. 10.1002/chem.201905668. [DOI] [PubMed] [Google Scholar]
  202. Panda S. S.; Jones R. A.; Bachawala P.; Mohapatra P. P. Spirooxindoles as Potential Pharmacophores. Mini. Rev. Med. Chem. 2017, 17, 1515–1536. 10.2174/1389557516666160624125108. [DOI] [PubMed] [Google Scholar]
  203. Oukoloff K.; Lucero B.; Francisco K. R.; Brunden K. R.; Ballatore C. 1,2,4-Triazolo[1,5-a]pyrimidines in Drug Design. Eur. J. Med. Chem. 2019, 165, 332–346. 10.1016/j.ejmech.2019.01.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Matiadis D.; Sagnou M. Pyrazoline Hybrids as Promising Anticancer Agents: An Up-to-Date Overview. Int. J. Mol. Sci. 2020, 21, 5507. 10.3390/ijms21155507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. Liang X.; Zhang L.; Li F.; Luan S.; He C.; Yin L.; Yin Z.; Zou Y.; Yue G.; Li L.; et al. Autophagy-Regulating N-Heterocycles Derivatives as Potential Anticancer Agents. Future. Med. Chem. 2020, 12, 223–242. 10.4155/fmc-2019-0294. [DOI] [PubMed] [Google Scholar]
  206. Li Petri G.; Spano V.; Spatola R.; Holl R.; Raimondi M. V.; Barraja P.; Montalbano A. Bioactive Pyrrole-Based Compounds with Target Selectivity. Eur. J. Med. Chem. 2020, 208, 112783. 10.1016/j.ejmech.2020.112783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Leng L.; Yang L.; Chen J.; Leng S.; Li H.; Li H.; Yuan X.; Zhou W.; Huang H. A Review on Pyrolysis of Protein-Rich Biomass: Nitrogen Transformation. Bioresour. Technol. 2020, 315, 123801. 10.1016/j.biortech.2020.123801. [DOI] [PubMed] [Google Scholar]
  208. Lang D. K.; Kaur R.; Arora R.; Saini B.; Arora S. Nitrogen-Containing Heterocycles as Anticancer Agents: An Overview. Anticancer. Agents. Med. Chem. 2020, 20, 2150–2168. 10.2174/1871520620666200705214917. [DOI] [PubMed] [Google Scholar]
  209. Kumar S.; Narasimhan B. Therapeutic Potential of Heterocyclic Pyrimidine Scaffolds. Chem. Cent. J. 2018, 12, 38. 10.1186/s13065-018-0406-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Hellwig P. S.; Peglow T. J.; Penteado F.; Bagnoli L.; Perin G.; Lenardao E. J. Recent Advances in the Synthesis of Selenophenes and Their Derivatives. Molecules. 2020, 25, 5907. 10.3390/molecules25245907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  211. Fesatidou M.; Petrou A.; Athina G. Heterocycle Compounds with Antimicrobial Activity. Curr. Pharm. Des. 2020, 26, 867–904. 10.2174/1381612826666200206093815. [DOI] [PubMed] [Google Scholar]
  212. Facchetti G.; Rimoldi I. Anticancer Platinum(II) Complexes Bearing N-Heterocycle Rings. Bioorg. Med. Chem. Lett. 2019, 29, 1257–1263. 10.1016/j.bmcl.2019.03.045. [DOI] [PubMed] [Google Scholar]
  213. Chatterjee T.; Shetti V. S.; Sharma R.; Ravikanth M. Heteroatom-Containing Porphyrin Analogues. Chem. Rev. 2017, 117, 3254–3328. 10.1021/acs.chemrev.6b00496. [DOI] [PubMed] [Google Scholar]
  214. Ali R.; Khan M. A.; Siddiqui N. Past, Present and Future of Antiepileptic Drug Therapy - Finding a Place for Heterocyclics. Mini. Rev. Med. Chem. 2015, 15, 1024–1050. 10.2174/138955751512150731113549. [DOI] [PubMed] [Google Scholar]
  215. Al-Harthy T.; Zoghaib W.; Abdel-Jalil R. Importance of Fluorine in Benzazole Compounds. Molecules. 2020, 25, 4677. 10.3390/molecules25204677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Shiraiwa M.; Selzle K.; Poschl U. Hazardous Components and Health Effects of Atmospheric Aerosol Particles: Reactive Oxygen Species, Soot, Polycyclic Aromatic Compounds and Allergenic Proteins. Free. Radic. Res. 2012, 46, 927–939. 10.3109/10715762.2012.663084. [DOI] [PubMed] [Google Scholar]
  217. Ni Z.; Zhang C.; Wang Z.; Zhao S.; Fan X.; Jia H. Performance and Potential Mechanism of Transformation of Polycyclic Aromatic Hydrocarbons (PAHs) on Various Iron Oxides. J. Hazard. Mater. 2021, 403, 123993. 10.1016/j.jhazmat.2020.123993. [DOI] [PubMed] [Google Scholar]
  218. Nezikova B.; Degrendele C.; Bandowe B. A. M.; Holubova Smejkalova A.; Kukucka P.; Martinik J.; Mayer L.; Prokes R.; Pribylova P.; Klanova J.; et al. Three Years of Atmospheric Concentrations of Nitrated and Oxygenated Polycyclic Aromatic Hydrocarbons and Oxygen Heterocycles at A Central European Background Site. Chemosphere. 2021, 269, 128738. 10.1016/j.chemosphere.2020.128738. [DOI] [PubMed] [Google Scholar]
  219. Li Q.; Zhang X.; Cao Y.; Shi C.; Tao P.; Zhao Q.; Yuan A. An Oxygen-Bridged Triarylamine Polycyclic Unit Based Tris-Cyclometalated Heteroleptic Iridium(III) Complex: Correlation between the Structure and Photophysical Properties. Dalton. Trans. 2019, 48, 4596–4601. 10.1039/C9DT00344D. [DOI] [PubMed] [Google Scholar]
  220. Duan W.; Meng X.; Sun Y.; Jia C. Association between Polycyclic Aromatic Hydrocarbons and Osteoporosis: Data from NHANES, 2005–2014. Arch. Osteoporos. 2018, 13, 112. 10.1007/s11657-018-0527-4. [DOI] [PubMed] [Google Scholar]
  221. Craig R. A. 2nd; Stoltz B. M. Polycyclic Furanobutenolide-Derived Cembranoid and Norcembranoid Natural Products: Biosynthetic Connections and Synthetic Efforts. Chem. Rev. 2017, 117, 7878–7909. 10.1021/acs.chemrev.7b00083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Hayakawa K.; Tang N.; Toriba A. Recent Analytical Methods for Atmospheric Polycyclic Aromatic Hydrocarbons and Their Derivatives. Biomed. Chromatogr. 2017, 31, e3862 10.1002/bmc.3862. [DOI] [PubMed] [Google Scholar]
  223. Fu P. P.; Xia Q.; Sun X.; Yu H. Phototoxicity and Environmental Transformation of Polycyclic Aromatic Hydrocarbons (PAHs)-Light-Induced Reactive Oxygen Species, Lipid Peroxidation, and DNA Damage. J. Environ. Sci. Health. C. Environ. Carcinog. Ecotoxicol. Rev. 2012, 30, 1–41. 10.1080/10590501.2012.653887. [DOI] [PubMed] [Google Scholar]
  224. Xiao Y.; Zhong J.; Zhong B.; Huang J.; Jiang L.; Jiang Y.; Yuan J.; Sun J.; Dai L.; Yang C.; et al. Exosomes as Potential Sources of Biomarkers in Colorectal Cancer. Cancer. Lett. 2020, 476, 13–22. 10.1016/j.canlet.2020.01.033. [DOI] [PubMed] [Google Scholar]
  225. Han F.; Wan S.; Sun Q.; Chen N.; Li H.; Zheng L.; Zhang N.; Huang Z.; Hong L.; Sun Q. Donor Plasma Mitochondrial DNA Is Correlated with Posttransplant Renal Allograft Function. Transplantation. 2019, 103, 2347–2358. 10.1097/TP.0000000000002598. [DOI] [PubMed] [Google Scholar]
  226. Xiao Z.; Yang S.; Yang Z.; Yang J.; Yip H. L.; Zhang F.; He F.; Wang T.; Wang J.; Yuan Y.; et al. Carbon-Oxygen-Bridged Ladder-Type Building Blocks for Highly Efficient Nonfullerene Acceptors. Adv. Mater. 2019, 31, 1804790 10.1002/adma.201804790. [DOI] [PubMed] [Google Scholar]
  227. Kadri T.; Rouissi T.; Kaur Brar S.; Cledon M.; Sarma S.; Verma M. Biodegradation of Polycyclic Aromatic Hydrocarbons (PAHs) by Fungal Enzymes: A Review. J. Environ. Sci. 2017, 51, 52–74. 10.1016/j.jes.2016.08.023. [DOI] [PubMed] [Google Scholar]
  228. Zhu B. C.; He J.; Liu W.; Xia X. Y.; Liu L. Y.; Liang B. B.; Yao H. G.; Liu B.; Ji L. N.; Mao Z. W. Selectivity and Targeting of G-Quadruplex Binders Activated by Adaptive Binding and Controlled by Chemical Kinetics. Angew. Chem., Int. Ed. Engl. 2021, 60, 15340–15343. 10.1002/anie.202104624. [DOI] [PubMed] [Google Scholar]
  229. Xu C. X.; Liu L. Y.; Lv B.; Zhao H. Y.; Cao Q.; Zhai T.; Mao Z. W. Two Novel Fan-Shaped Trinuclear Pt(ii) Complexes Act as G-Quadruplex Binders and Telomerase Inhibitors. Dalton. Trans. 2020, 49, 9322–9329. 10.1039/D0DT01767A. [DOI] [PubMed] [Google Scholar]
  230. Miglietta G.; Russo M.; Capranico G. G-Quadruplex-R-Loop Interactions and the Mechanism of Anticancer G-Quadruplex Binders. Nucleic. Acids. Res. 2020, 48, 11942–11957. 10.1093/nar/gkaa944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Karpenko A. Heterotopic Nervous Tissue in the Spinal Leptomeninges of Open Cranioschises. Cesk. Patol. 1974, 10, 214–219. [PubMed] [Google Scholar]
  232. Zhang H.; Lei Z.; Tian R.; Wang Z. Polyamidoamine Starburst Dendrimer-Activated Chromatography Paper-Based Assay for Sensitive Detection of Telomerase Activity. Talanta 2018, 178, 116–121. 10.1016/j.talanta.2017.09.034. [DOI] [PubMed] [Google Scholar]
  233. Shukla S.; Jeong H. C.; Sturgeon C. M.; Parker R.; Batista L. F. Z. Chemical Inhibition of PAPD5/7 Rescues Telomerase Function and Hematopoiesis in Dyskeratosis Congenita. Blood. Adv. 2020, 4, 2717–2722. 10.1182/bloodadvances.2020001848. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Seimiya H. Crossroads of Telomere Biology and Anticancer Drug Discovery. Cancer. Sci. 2020, 111, 3089–3099. 10.1111/cas.14540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Zhu D.; Li M.; Wu Z.; Du Y.; Luo B.; Huang P.; Wen S. Copper-Catalyzed One-Pot Synthesis of Dibenzofurans, Xanthenes, and Xanthones from Cyclic Diphenyl Iodoniums. Eur. J. Org. Chem. 2019, 2019, 4566–4571. 10.1002/ejoc.201900745. [DOI] [Google Scholar]
  236. Zhirnov V. V.; Velihina Y. S.; Mitiukhin O. P.; Brovarets V. S. Intrinsic Drug Potential of Oxazolo[5,4-d]pyrimidines and Oxazolo[4,5-d]pyrimidines. Chem. Biol. Drug. Des. 2021, 98, 561–581. 10.1111/cbdd.13911. [DOI] [PubMed] [Google Scholar]
  237. Yan Y.; Liu Q.; Zang X.; Yuan S.; Bat-Erdene U.; Nguyen C.; Gan J.; Zhou J.; Jacobsen S. E.; Tang Y. Resistance-Gene-Directed Discovery of A Natural-Product Herbicide with A New Mode of Action. Nature. 2018, 559, 415–418. 10.1038/s41586-018-0319-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Salehian F.; Nadri H.; Jalili-Baleh L.; Youseftabar-Miri L.; Abbas Bukhari S. N.; Foroumadi A.; Tuylu Kucukkilinc T.; Sharifzadeh M.; Khoobi M. A Review: Biologically Active 3,4-Heterocycle-Fused Coumarins. Eur. J. Med. Chem. 2021, 212, 113034. 10.1016/j.ejmech.2020.113034. [DOI] [PubMed] [Google Scholar]
  239. Plescia F.; Maggio B.; Daidone G.; Raffa D. 4-(3H)-Quinazolinones N-3 Substituted with A Five Membered Heterocycle: A Promising Scaffold Towards Bioactive Molecules. Eur. J. Med. Chem. 2021, 213, 113070. 10.1016/j.ejmech.2020.113070. [DOI] [PubMed] [Google Scholar]
  240. Nadar S.; Khan T. Pyrimidine: An Elite Heterocyclic Leitmotif in Drug Discovery-Synthesis and Biological Activity. Chem. Biol. Drug. Des. 2022, 100, 818–842. 10.1111/cbdd.14001. [DOI] [PubMed] [Google Scholar]
  241. Miao L.; Li L.; Huang Y.; Delcassian D.; Chahal J.; Han J.; Shi Y.; Sadtler K.; Gao W.; Lin J.; et al. Delivery of mRNA Vaccines with Heterocyclic Lipids Increases Anti-Tumor Efficacy by STING-Mediated Immune Cell Activation. Nat. Biotechnol. 2019, 37, 1174–1185. 10.1038/s41587-019-0247-3. [DOI] [PubMed] [Google Scholar]
  242. Mathew T.; Papp A. A.; Paknia F.; Fustero S.; Surya Prakash G. K. Benzodiazines: Recent Synthetic Advances. Chem. Soc. Rev. 2017, 46, 3060–3094. 10.1039/C7CS00082K. [DOI] [PubMed] [Google Scholar]
  243. Ma L.; Miao D.; Lee J. J.; Li T.; Chen Y.; Su G.; Zhao Y. Synthesis and Biological Evaluation of Heterocyclic Ring-Fused Dammarane-Type Ginsenoside Derivatives as Potential Anti-Tumor Agents. Bioorg. Chem. 2021, 116, 105365. 10.1016/j.bioorg.2021.105365. [DOI] [PubMed] [Google Scholar]
  244. Liu J.; Cremosnik G. S.; Otte F.; Pahl A.; Sievers S.; Strohmann C.; Waldmann H. Design, Synthesis, and Biological Evaluation of Chemically and Biologically Diverse Pyrroquinoline Pseudo Natural Products. Angew. Chem., Int. Ed. Engl. 2021, 60, 4648–4656. 10.1002/anie.202013731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  245. Forezi L. da S. M.; Lima C. G. S.; Amaral A. A. P.; Ferreira P. G.; de Souza M. C. B. V.; Cunha A. C.; da Silva F. de C.; Ferreira V. F. Bioactive 1,2,3-Triazoles: An Account on their Synthesis, Structural Diversity and Biological Applications. Chem. Rec. 2021, 21, 2782–2807. 10.1002/tcr.202000185. [DOI] [PubMed] [Google Scholar]
  246. Homer J. A.; Sperry J. Mushroom-Derived Indole Alkaloids. J. Nat. Prod. 2017, 80, 2178–2187. 10.1021/acs.jnatprod.7b00390. [DOI] [PubMed] [Google Scholar]
  247. Hill R. A.; Sutherland A. Hot Off the Press. Nat. Prod. Rep. 2018, 35, 1236–1240. 10.1039/C8NP90046A. [DOI] [PubMed] [Google Scholar]
  248. Furstner A. Lessons from Natural Product Total Synthesis: Macrocyclization and Postcyclization Strategies. Acc. Chem. Res. 2021, 54, 861–874. 10.1021/acs.accounts.0c00759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Baudoin O. Multiple Catalytic C-H Bond Functionalization for Natural Product Synthesis. Angew. Chem., Int. Ed. Engl. 2020, 59, 17798–17809. 10.1002/anie.202001224. [DOI] [PubMed] [Google Scholar]
  250. Zhao P.; Guo Y.; Luan X. Total Synthesis of Dalesconol A by Pd(0)/Norbornene-Catalyzed Three-Fold Domino Reaction and Pd(II)-Catalyzed Trihydroxylation. J. Am. Chem. Soc. 2021, 143, 21270–21274. 10.1021/jacs.1c12118. [DOI] [PubMed] [Google Scholar]
  251. Sahoo B. M.; Ravi Kumar B. V. V.; Banik B. K.; Borah P. Polyaromatic Hydrocarbons (PAHs): Structures, Synthesis and their Biological Profile. Curr. Org. Synth. 2020, 17, 625–640. 10.2174/1570179417666200713182441. [DOI] [PubMed] [Google Scholar]
  252. Peeters E.; Mackie C.; Candian A.; Tielens A. A Spectroscopic View on Cosmic PAH Emission. Acc. Chem. Res. 2021, 54, 1921–1933. 10.1021/acs.accounts.0c00747. [DOI] [PubMed] [Google Scholar]
  253. Liao Q.; Liu H.; Lu C.; Liu J.; Waigi M. G.; Ling W. Root Exudates Enhance the PAH Degradation and Degrading Gene Abundance in Soils. Sci. Total Environ. 2021, 764, 144436. 10.1016/j.scitotenv.2020.144436. [DOI] [PubMed] [Google Scholar]
  254. Gbeddy G.; Goonetilleke A.; Ayoko G. A.; Egodawatta P. Transformation and Degradation of Polycyclic Aromatic Hydrocarbons (PAHs) in Urban Road Surfaces: Influential Factors, Implications and Recommendations. Environ. Pollut. 2020, 257, 113510. 10.1016/j.envpol.2019.113510. [DOI] [PubMed] [Google Scholar]
  255. Faseke V. C.; Raps F. C.; Sparr C. Polyketide Cyclizations for the Synthesis of Polyaromatics. Angew. Chem., Int. Ed. Engl. 2020, 59, 6975–6983. 10.1002/anie.201911255. [DOI] [PubMed] [Google Scholar]
  256. Drwal E.; Rak A.; Gregoraszczuk E. L. Review: Polycyclic Aromatic Hydrocarbons (PAHs)-Action on Placental Function and Health Risks in Future Life of Newborns. Toxicology 2019, 411, 133–142. 10.1016/j.tox.2018.10.003. [DOI] [PubMed] [Google Scholar]
  257. Del Tito A.; Abdulla H. O.; Ravelli D.; Protti S.; Fagnoni M. Photocatalyzed Syntheses of Phenanthrenes and Their Aza-Analogues. A Review. Beilstein. J. Org. Chem. 2020, 16, 1476–1488. 10.3762/bjoc.16.123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  258. Kantarod K.; Worakul T.; Soorukram D.; Kuhakarn C.; Reutrakul V.; Surawatanawong P.; Wattanathana W.; Leowanawat P. Dibenzopleiadiene-Embeded Polyaromatics via [4 + 3] Annulative Decarbonylation/Decarboxylation. Org. Chem. Front. 2021, 8, 522–530. 10.1039/D0QO00942C. [DOI] [Google Scholar]
  259. Yang S.; Wang F.; Wu Y.; Hua W.; Zhang F. Synthesis of Functionalized Triphenylenes via a Traceless Directing Group Strategy. Org. Lett. 2018, 20, 1491–1495. 10.1021/acs.orglett.8b00071. [DOI] [PubMed] [Google Scholar]
  260. Yang S.; Hua W.; Wu Y.; Hu T.; Wang F.; Zhang X.; Zhang F. Site-Selective Synthesis of Functionalized Dibenzo[f,h]Quinolines and Their Derivatives Involving Cyclic Diaryliodonium Salts via A Decarboxylative Annulation Strategy. Chem. Commun. 2018, 54, 3239–3242. 10.1039/C8CC00300A. [DOI] [PubMed] [Google Scholar]
  261. Lee J. B.; Jeon M. H.; Seo J. K.; von Helden G.; Rohde J. U.; Zhao B. S.; Seo J.; Hong S. Y. Annulative π-Extension of Unactivated Benzene Derivatives through Nondirected C-H Arylation. Org. Lett. 2019, 21, 7004–7008. 10.1021/acs.orglett.9b02583. [DOI] [PubMed] [Google Scholar]
  262. Zhu D.; Peng H.; Sun Y.; Wu Z.; Wang Y.; Luo B.; Yu T.; Hu Y.; Huang P.; Wen S. Modular Metal-Free Catalytic Radical Annulation of Cyclic Diaryliodoniums to Access π-Extended Arenes. Green. Chem. 2021, 23, 1972–1977. 10.1039/D0GC04183A. [DOI] [Google Scholar]
  263. Zhang W.; Zhao K.; Guan F.; Yin J.; Feng Y.; Li J.; Li Y. Microstructure and Electrical Properties of Fluorene Polyester Based Nanocomposite Dielectrics. Polymers. 2021, 13, 3053. 10.3390/polym13183053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Yamamoto T.; Chang T. C.; Tanaka K. Epoc Group: Transformable Protecting Group with Gold(III)-Catalyzed Fluorene Formation. Chem. Sci. 2021, 12, 10703–10709. 10.1039/D1SC03125B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Wu M.; Li J.; Zhang R.; Tian X.; Han Z.; Lu X.; Guo K.; Liu Z.; Wang Z. Synthesis and Properties of Dithiafulvenyl Functionalized Spiro[fluorene-9,9’-xanthene] Molecules. Org. Lett. 2018, 20, 780–783. 10.1021/acs.orglett.7b03918. [DOI] [PubMed] [Google Scholar]
  266. Li G.; Zheng J.; Klimes K.; Zhu Z. Q.; Wu J.; Zhu H.; Li J. Novel Carbazole/Fluorene-Based Host Material for Stable and Efficient Phosphorescent OLEDs. ACS Appl. Mater. Interfaces. 2019, 11, 40320–40331. 10.1021/acsami.9b13245. [DOI] [PubMed] [Google Scholar]
  267. Lee D. Y.; Sivakumar G.; Manju; Misra R.; Seok S. I. Carbazole-Based Spiro[fluorene-9,9’-xanthene] as an Efficient Hole-Transporting Material for Perovskite Solar Cells. ACS Appl. Mater. Interfaces. 2020, 12, 28246–28252. 10.1021/acsami.0c06318. [DOI] [PubMed] [Google Scholar]
  268. Kardelis V.; Denk M. M.; Adronov A. Click-Functionalization of a Poly(Tetrazine-co-Fluorene)-Conjugated Polymer with a Series of trans-Cyclooctene Derivatives. Angew. Chem., Int. Ed. Engl. 2021, 60, 2980–2986. 10.1002/anie.202010795. [DOI] [PubMed] [Google Scholar]
  269. Dong X.; Zheng M.; Wan B.; Liu X.; Xu H.; Zha J. Low-Permittivity Copolymerized Polyimides with Fluorene Rigid Conjugated Structure. Materials (Basel). 2021, 14, 6266. 10.3390/ma14216266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Beygisangchin M.; Abdul Rashid S.; Shafie S.; Sadrolhosseini A. R. Polyaniline Synthesized by Different Dopants for Fluorene Detection via Photoluminescence Spectroscopy. Materials (Basel). 2021, 14, 7382. 10.3390/ma14237382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Kim L. J.; Xue M.; Li X.; Xu Z.; Paulson E.; Mercado B.; Nelson H. M.; Herzon S. B. Structure Revision of the Lomaiviticins. J. Am. Chem. Soc. 2021, 143, 6578–6585. 10.1021/jacs.1c01729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Kim H. S.; You Y.; Mun J.; Gadhe C. G.; Moon H.; Lee J. S.; Pae A. N.; Kohara M.; Keum G.; Kim B. M.; et al. Structure-Activity Relationships of Fluorene Compounds Inhibiting HCV Variants. Antiviral Res. 2020, 174, 104678. 10.1016/j.antiviral.2019.104678. [DOI] [PubMed] [Google Scholar]
  273. Kaneko M.; Li Z.; Burk M.; Colis L.; Herzon S. B. Synthesis and Biological Evaluation of (2S,2’S)-Lomaiviticin A. J. Am. Chem. Soc. 2021, 143, 1126–1132. 10.1021/jacs.0c11960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  274. Ju H.; Murugan N. A.; Hou L.; Li P.; Guizzo L.; Zhang Y.; Bertagnin C.; Kong X.; Kang D.; Jia R.; et al. Identification of C5-NH2Modified Oseltamivir Derivatives as Novel Influenza Neuraminidase Inhibitors with Highly Improved Antiviral Activities and Favorable Druggability. J. Med. Chem. 2021, 64, 17992–18009. 10.1021/acs.jmedchem.1c01366. [DOI] [PubMed] [Google Scholar]
  275. Huang B.; Wang H.; Zheng Y.; Li M.; Kang G.; Barreto-de-Souza V.; Nassehi N.; Knapp P. E.; Selley D. E.; Hauser K. F.; et al. Structure-Based Design and Development of Chemical Probes Targeting Putative MOR-CCR5 Heterodimers to Inhibit Opioid Exacerbated HIV-1 Infectivity. J. Med. Chem. 2021, 64, 7702–7723. 10.1021/acs.jmedchem.1c00408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Hayashi H.; Barker J. E.; Cardenas Valdivia A.; Kishi R.; MacMillan S. N.; Gomez-Garcia C. J.; Miyauchi H.; Nakamura Y.; Nakano M.; Kato S. I.; et al. Monoradicals and Diradicals of Dibenzofluoreno[3,2-b]fluorene Isomers: Mechanisms of Electronic Delocalization. J. Am. Chem. Soc. 2020, 142, 20444–20455. 10.1021/jacs.0c09588. [DOI] [PubMed] [Google Scholar]
  277. Haghdoost M. M.; Golbaghi G.; Guard J.; Sielanczyk S.; Patten S. A.; Castonguay A. Synthesis, Characterization and Biological Evaluation of Cationic Organoruthenium(II) Fluorene Complexes: Influence of the Nature of the Counteranion. Dalton. Trans. 2019, 48, 13396–13405. 10.1039/C9DT00143C. [DOI] [PubMed] [Google Scholar]
  278. Duyar H.; Portakal H. S.; Yalcin E.; Kanat B.; Doluca O.; Seferoglu Z. Fluorene/Fluorenone Carboxamide Derivatives as Selective Light-up Fluorophores For C-Myc G-Quadruplex. Bioorg. Med. Chem. Lett. 2021, 36, 127824. 10.1016/j.bmcl.2021.127824. [DOI] [PubMed] [Google Scholar]
  279. Du H.; Zhang L.; Mao W.; Zhao Y.; Huang H.; Xiao Y.; Zhang Y.; He X.; Wang K. Ultrafine Fluorene-Pyridine Oligoelectrolyte Nanoparticles for Supersensitive Fluorescence Sensing of Heparin and Protamine. Chem. Commun. 2021, 57, 8304–8307. 10.1039/D1CC01969D. [DOI] [PubMed] [Google Scholar]
  280. Peng X.; Yang Y.; Luo B.; Wen S.; Huang P. Modular Tandem Mizoroki-Heck/Reductive Heck Reactions to Construct Fluorenes from Cyclic Diaryliodoniums. Adv. Synth. & Catal. 2021, 363, 222–226. 10.1002/adsc.202001080. [DOI] [Google Scholar]
  281. Peng X.; Yang J.; Qiu X.; Duan Y.; Bao Z.; Chen J. Synthesis of Dihydroanthracenes via Palladium-Catalyzed Tandem Mizoroki-Heck/Reductive Heck Reactions Using Cyclic Diaryliodoniums and Alkenes. Synlett. 2021, 32, 636–640. 10.1055/a-1343-5455. [DOI] [Google Scholar]
  282. Peng X.; Luo H.; Wu F.; Zhu D.; Ganesan A.; Wen S.; Huang P. Synthesis of Fluorenes with an All-Carbon Quaternary Center via Palladium-Catalyzed Dual Arylation using Cyclic Diaryliodonium Triflates. Adv. Synth. & Catal. 2017, 359, 1152–1156. 10.1002/adsc.201601260. [DOI] [Google Scholar]
  283. Swain S. S.; Pati S.; Hussain T. Quinoline Heterocyclic Containing Plant and Marine Candidates against Drug-Resistant Mycobacterium Tuberculosis: A Systematic Drug-Ability Investigation. Eur. J. Med. Chem. 2022, 232, 114173. 10.1016/j.ejmech.2022.114173. [DOI] [PubMed] [Google Scholar]
  284. Khwaza V.; Mlala S.; Oyedeji O. O.; Aderibigbe B. A. Pentacyclic Triterpenoids with Nitrogen-Containing Heterocyclic Moiety, Privileged Hybrids in Anticancer Drug Discovery. Molecules. 2021, 26, 2401. 10.3390/molecules26092401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  285. Kharb R. Updates on Receptors Targeted by Heterocyclic Scaffolds: New Horizon in Anticancer Drug Development. Anticancer. Agents. Med. Chem. 2021, 21, 1338–1349. 10.2174/1871520620666200619181102. [DOI] [PubMed] [Google Scholar]
  286. Kalaria P. N.; Karad S. C.; Raval D. K. A Review on Diverse Heterocyclic Compounds as the Privileged Scaffolds in Antimalarial Drug Discovery. Eur. J. Med. Chem. 2018, 158, 917–936. 10.1016/j.ejmech.2018.08.040. [DOI] [PubMed] [Google Scholar]
  287. Bootsma A. N.; Wheeler S. E. Stacking Interactions of Heterocyclic Drug Fragments with Protein Amide Backbones. ChemMedChem. 2018, 13, 835–841. 10.1002/cmdc.201700721. [DOI] [PubMed] [Google Scholar]
  288. Blakemore D. C.; Castro L.; Churcher I.; Rees D. C.; Thomas A. W.; Wilson D. M.; Wood A. Organic Synthesis Provides Opportunities to Transform Drug Discovery. Nat. Chem. 2018, 10, 383–394. 10.1038/s41557-018-0021-z. [DOI] [PubMed] [Google Scholar]
  289. Atukuri D.; Gunjal R.; Holagundi N.; Korlahalli B.; Gangannavar S.; Akkasali K. Contribution of N-Heterocycles towards Anti-Tubercular Drug Discovery (2014–2019); Predicted and Reengineered Molecular Frameworks. Drug. Dev. Res. 2021, 82, 767–783. 10.1002/ddr.21809. [DOI] [PubMed] [Google Scholar]
  290. Wu S. M.; Qiu X. Y.; Liu S. J.; Sun J. Single Heterocyclic Compounds as Monoamine Oxidase Inhibitors: From Past to Present. Mini. Rev. Med. Chem. 2020, 20, 908–920. 10.2174/1389557520666200302114620. [DOI] [PubMed] [Google Scholar]
  291. Marinozzi M.; Pertusati F.; Serpi M. lambda(5)-Phosphorus-Containing alpha-Diazo Compounds: A Valuable Tool for Accessing Phosphorus-Functionalized Molecules. Chem. Rev. 2016, 116, 13991–14055. 10.1021/acs.chemrev.6b00373. [DOI] [PubMed] [Google Scholar]
  292. Yamamoto K.; Kuriyama M.; Onomura O. Anodic Oxidation for the Stereoselective Synthesis of Heterocycles. Acc. Chem. Res. 2020, 53, 105–120. 10.1021/acs.accounts.9b00513. [DOI] [PubMed] [Google Scholar]
  293. Teja C.; Khan F. R. N. Radical Transformations towards the Synthesis of Quinoline: A Review. Chem. Asian. J. 2020, 15, 4153–4107. 10.1002/asia.202001156. [DOI] [PubMed] [Google Scholar]
  294. Saha P.; Saikia A. K. Ene Cyclization Reaction in Heterocycle Synthesis. Org. Biomol. Chem. 2018, 16, 2820–2840. 10.1039/C8OB00429C. [DOI] [PubMed] [Google Scholar]
  295. Munzeiwa W. A.; Omondi B.; Nyamori V. O. Architecture and Synthesis of P,N-Heterocyclic Phosphine Ligands. Beilstein. J. Org. Chem. 2020, 16, 362–383. 10.3762/bjoc.16.35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  296. Mohammadi Ziarani G.; Kheilkordi Z.; Gholamzadeh P. Ultrasound-Assisted Synthesis of Heterocyclic Compounds. Mol. Divers. 2020, 24, 771–820. 10.1007/s11030-019-09964-1. [DOI] [PubMed] [Google Scholar]
  297. Liu X.; He L. N. Synthesis of Lactones and Other Heterocycles. Top. Curr. Chem. 2017, 375, 21. 10.1007/s41061-017-0108-9. [DOI] [PubMed] [Google Scholar]
  298. Hosseini H.; Bayat M. Cyanoacetohydrazides in Synthesis of Heterocyclic Compounds. Top. Curr. Chem. 2018, 376, 40. 10.1007/s41061-018-0218-z. [DOI] [PubMed] [Google Scholar]
  299. Bandyopadhyay D.; Banik B. K. Synthesis of Medicinally Privileged Heterocycles through Dielectric Heating. Curr. Med. Chem. 2017, 24, 4596–4626. 10.2174/0929867324666170223152137. [DOI] [PubMed] [Google Scholar]
  300. Baccalini A.; Faita G.; Zanoni G.; Maiti D. Transition Metal Promoted Cascade Heterocycle Synthesis through C-H Functionalization. Chem.—Eur. J. 2020, 26, 9749–9783. 10.1002/chem.202001832. [DOI] [PubMed] [Google Scholar]
  301. Zhang X.; He H.; Xiang J.; Yin H.; Hou T. Selenium-Containing Proteins/Peptides from Plants: A Review on the Structures and Functions. J. Agric. Food. Chem. 2020, 68, 15061–15073. 10.1021/acs.jafc.0c05594. [DOI] [PubMed] [Google Scholar]
  302. Vinceti M.; Filippini T.; Cilloni S.; Crespi C. M. The Epidemiology of Selenium and Human Cancer. Adv. Cancer. Res. 2017, 136, 1–48. 10.1016/bs.acr.2017.07.001. [DOI] [PubMed] [Google Scholar]
  303. Rayman M. P. Selenium and Human Health. Lancet. 2012, 379, 1256–1268. 10.1016/S0140-6736(11)61452-9. [DOI] [PubMed] [Google Scholar]
  304. 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 Isocombretastatins and Phenstatins as Antitumor Agents. J. Med. Chem. 2017, 60, 7300–7314. 10.1021/acs.jmedchem.7b00480. [DOI] [PubMed] [Google Scholar]
  305. Liu H.; Xu H.; Huang K. Selenium in the Prevention of Atherosclerosis and Its Underlying Mechanisms. Metallomics. 2017, 9, 21–37. 10.1039/C6MT00195E. [DOI] [PubMed] [Google Scholar]
  306. Kielczykowska M.; Kocot J.; Pazdzior M.; Musik I. Selenium - A Fascinating Antioxidant of Protective Properties. Adv. Clin. Exp. Med. 2018, 27, 245–255. 10.17219/acem/67222. [DOI] [PubMed] [Google Scholar]
  307. Hariharan S.; Dharmaraj S. Selenium and Selenoproteins: It’s Role in Regulation of Inflammation. Inflammopharmacology. 2020, 28, 667–695. 10.1007/s10787-020-00690-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  308. Cheignon C.; Cordeau E.; Prache N.; Cantel S.; Martinez J.; Subra G.; Arnaudguilhem C.; Bouyssiere B.; Enjalbal C. Receptor-Ligand Interaction Measured by Inductively Coupled Plasma Mass Spectrometry and Selenium Labeling. J. Med. Chem. 2018, 61, 10173–10184. 10.1021/acs.jmedchem.8b01320. [DOI] [PubMed] [Google Scholar]
  309. An B.; Wang B.; Hu J.; Xu S.; Huang L.; Li X.; Chan A. S. C. Synthesis and Biological Evaluation of Selenium-Containing 4-Anilinoquinazoline Derivatives as Novel Antimitotic Agents. J. Med. Chem. 2018, 61, 2571–2588. 10.1021/acs.jmedchem.8b00128. [DOI] [PubMed] [Google Scholar]
  310. Wei R.; Xu R.; Zhang K.; Liang F.; Yao Y. Biological Enzyme Treatment of Starch-Based Lithium-Ion Battery Silicon-Carbon Composite. Nanotechnology. 2021, 32, 045605. 10.1088/1361-6528/abc114. [DOI] [PubMed] [Google Scholar]
  311. Susarla S.; Chilkoor G.; Kalimuthu J. R.; Saadi M.; Cui Y.; Arif T.; Tsafack T.; Puthirath A. B.; Sigdel P.; Jasthi B.; et al. Corrosion Resistance of Sulfur-Selenium Alloy Coatings. Adv. Mater. 2021, 33, 2104467 10.1002/adma.202104467. [DOI] [PubMed] [Google Scholar]
  312. Rehman W. U.; Wang H.; Manj R. Z. A.; Luo W.; Yang J. When Silicon Materials Meet Natural Sources: Opportunities and Challenges for Low-Cost Lithium Storage. Small. 2021, 17, 1904508 10.1002/smll.201904508. [DOI] [PubMed] [Google Scholar]
  313. Karaman D. S.; Sarparanta M. P.; Rosenholm J. M.; Airaksinen A. J. Multimodality Imaging of Silica and Silicon Materials In Vivo. Adv. Mater. 2018, 30, 1703651 10.1002/adma.201703651. [DOI] [PubMed] [Google Scholar]
  314. Hosnedlova B.; Kepinska M.; Skalickova S.; Fernandez C.; Ruttkay-Nedecky B.; Peng Q.; Baron M.; Melcova M.; Opatrilova R.; Zidkova J.; et al. Nano-Selenium and Its Nanomedicine Applications: A Critical Review. Int. J. Nanomedicine. 2018, 13, 2107–2128. 10.2147/IJN.S157541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Fan B.; Lin F.; Wu X.; Zhu Z.; Jen A. K. Selenium-Containing Organic Photovoltaic Materials. Acc. Chem. Res. 2021, 54, 3906–3916. 10.1021/acs.accounts.1c00443. [DOI] [PubMed] [Google Scholar]
  316. Parasram M.; Gevorgyan V. Silicon-Tethered Strategies for C-H Functionalization Reactions. Acc. Chem. Res. 2017, 50, 2038–2053. 10.1021/acs.accounts.7b00306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  317. Na M.; Han Y.; Chen Y.; Ma S.; Liu J.; Chen X. Synthesis of Silicon Nanoparticles Emitting Yellow-Green Fluorescence for Visualization of pH Change and Determination of Intracellular pH of Living Cells. Anal. Chem. 2021, 93, 5185–5193. 10.1021/acs.analchem.0c05107. [DOI] [PubMed] [Google Scholar]
  318. Jagannathan J. R.; Fettinger J. C.; Shaw J. T.; Franz A. K. Enantioselective Si-H Insertion Reactions of Diarylcarbenes for the Synthesis of Silicon-Stereogenic Silanes. J. Am. Chem. Soc. 2020, 142, 11674–11679. 10.1021/jacs.0c04533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  319. Zare B.; Nami M.; Shahverdi A. R. Tracing Tellurium and Its Nanostructures in Biology. Biol. Trace. Elem. Res. 2017, 180, 171–181. 10.1007/s12011-017-1006-2. [DOI] [PubMed] [Google Scholar]
  320. Tanini D.; Capperucci A.; Supuran C. T.; Angeli A. Sulfur, Selenium and Tellurium Containing Amines Act as Effective Carbonic Anhydrase Activators. Bioorg. Chem. 2019, 87, 516–522. 10.1016/j.bioorg.2019.03.062. [DOI] [PubMed] [Google Scholar]
  321. Ratti S. G.; Alvarez E. O. Tellurium Epigenetic Transgenerational Effects on Behavioral Expression of Coping Behavior in Rats. Prog. Brain. Res. 2019, 245, 247–261. 10.1016/bs.pbr.2019.03.003. [DOI] [PubMed] [Google Scholar]
  322. Medina-Cruz D.; Gonzalez M. U.; Tien-Street W.; Fernandez-Castro M.; Vernet-Crua A.; Fernandez-Martinez I.; Martinez L.; Huttel Y.; Webster T. J.; Garcia-Martin J. M. Synergic Antibacterial Coatings Combining Titanium Nanocolumns and Tellurium Nanorods. Nanomedicine. 2019, 17, 36–46. 10.1016/j.nano.2018.12.009. [DOI] [PubMed] [Google Scholar]
  323. Khanlarian M.; Rashchi F.; Saba M. A Modified Sulfation-Roasting-Leaching Process for Recovering Se, Cu, and Ag from Copper Anode Slimes at a Lower Temperature. J. Environ. Manage. 2019, 235, 303–309. 10.1016/j.jenvman.2019.01.079. [DOI] [PubMed] [Google Scholar]
  324. Dickson R. S.; Glowa G. A. Tellurium behaviour in the Fukushima Dai-ichi Nuclear Power Plant accident. J. Environ. Radioact. 2019, 204, 49–65. 10.1016/j.jenvrad.2019.03.024. [DOI] [PubMed] [Google Scholar]
  325. Choi K.; Medley J. K.; Konig M.; Stocking K.; Smith L.; Gu S.; Sauro H. M. Tellurium: An Extensible Python-Based Modeling Environment for Systems and Synthetic Biology. Biosystems. 2018, 171, 74–79. 10.1016/j.biosystems.2018.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  326. Chen S.; Xing C.; Huang D.; Zhou C.; Ding B.; Guo Z.; Peng Z.; Wang D.; Zhu X.; Liu S.; et al. Eradication of Tumor Growth by Delivering Novel Photothermal Selenium-Coated Tellurium Nanoheterojunctions. Sci. Adv. 2020, 6, eaay6825 10.1126/sciadv.aay6825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  327. Shi Z.; Cao R.; Khan K.; Tareen A. K.; Liu X.; Liang W.; Zhang Y.; Ma C.; Guo Z.; Luo X.; Zhang H.; et al. Two-Dimensional Tellurium: Progress, Challenges, and Prospects. Nano-Micro. Lett. 2020, 12, 99. 10.1007/s40820-020-00427-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  328. Ma C.; Yan J.; Huang Y.; Wang C.; Yang G. The Optical Duality of Tellurium Nanoparticles for Broadband Solar Energy Harvesting and Efficient Photothermal Conversion. Sci. Adv. 2018, 4, eaas9894 10.1126/sciadv.aas9894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  329. Kim C. K.; Ji J. M.; Zhou H.; Lu C.; Kim H. K. Tellurium-Doped, Mesoporous Carbon Nanomaterials as Transparent Metal-Free Counter Electrodes for High-Performance Bifacial Dye-Sensitized Solar Cells. Nanomaterials (Basel). 2020, 10, 29. 10.3390/nano10010029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  330. Ying P.; He R.; Mao J.; Zhang Q.; Reith H.; Sui J.; Ren Z.; Nielsch K.; Schierning G. Towards Tellurium-Free Thermoelectric Modules for Power Generation from Low-Grade Heat. Nat. Commun. 2021, 12, 1121. 10.1038/s41467-021-21391-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  331. Qiu G.; Huang S.; Segovia M.; Venuthurumilli P. K.; Wang Y.; Wu W.; Xu X.; Ye P. D. Thermoelectric Performance of 2D Tellurium with Accumulation Contacts. Nano Lett. 2019, 19, 1955–1962. 10.1021/acs.nanolett.8b05144. [DOI] [PubMed] [Google Scholar]
  332. Mao J.; Zhu H.; Ding Z.; Liu Z.; Gamage G. A.; Chen G.; Ren Z. High Thermoelectric Cooling Performance of N-Type Mg3Bi2-Based Materials. Science. 2019, 365, 495–498. 10.1126/science.aax7792. [DOI] [PubMed] [Google Scholar]
  333. Zhang N.; Zhao G.; Li L.; Wang P.; Xie L.; Cheng B.; Li H.; Lin Z.; Xi C.; Ke J.; et al. Magnetotransport Signatures of Weyl Physics and Discrete Scale Invariance in the Elemental Semiconductor Tellurium. Proc. Natl. Acad. Sci. U. S. A. 2020, 117, 11337–11343. 10.1073/pnas.2002913117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  334. Yang Y.; Xu M.; Jia S.; Wang B.; Xu L.; Wang X.; Liu H.; Liu Y.; Guo Y.; Wang L.; et al. A New Opportunity for the Emerging Tellurium Semiconductor: Making Resistive Switching Devices. Nat. Commun. 2021, 12, 6081. 10.1038/s41467-021-26399-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  335. He Z.; Yang Y.; Liu J. W.; Yu S. H. Emerging Tellurium Nanostructures: Controllable Synthesis and Their Applications. Chem. Soc. Rev. 2017, 46, 2732–2753. 10.1039/C7CS00013H. [DOI] [PubMed] [Google Scholar]
  336. Xu K.; Sharma A.; Kang J.; Hu X.; Hao Z.; Zhu W. Heterogeneous Electronic and Photonic Devices Based on Monolayer Ternary Telluride Core/Shell Structures. Adv. Mater. 2020, 32, 2002548 10.1002/adma.202002548. [DOI] [PubMed] [Google Scholar]
  337. Meier M. C.; Cheng W. H.; Atwater H. A.; Lewis N. S.; Carim A. I. Inorganic Phototropism in Electrodeposition of Se-Te. J. Am. Chem. Soc. 2019, 141, 18658–18661. 10.1021/jacs.9b10579. [DOI] [PubMed] [Google Scholar]
  338. Hirata J.; Itabashi D.; Aimoto M. Determination of Ultra-trace Tellurium in Steel by ID-ICP-MS/MS with Liquid-Liquid Extraction. Anal. Sci. 2021, 37, 1757–1763. 10.2116/analsci.21P143. [DOI] [PubMed] [Google Scholar]
  339. Bu Z.; Zhang X.; Hu Y.; Chen Z.; Lin S.; Li W.; Xiao C.; Pei Y. A Record Thermoelectric Efficiency in Tellurium-Free Modules for Low-Grade Waste Heat Recovery. Nat. Commun. 2022, 13, 237. 10.1038/s41467-021-27916-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  340. Zhang Z.; Khurram M.; Sun Z.; Yan Q. Uniform Tellurium Doping in Black Phosphorus Single Crystals by Chemical Vapor Transport. Inorg. Chem. 2018, 57, 4098–4103. 10.1021/acs.inorgchem.8b00278. [DOI] [PubMed] [Google Scholar]
  341. Yang W.; Lan X.; Wang Q.; Dong P.; Wang G. Selective Pre-leaching of Tellurium From Telluride-Type Gold Concentrate. Front. Chem. 2021, 9, 593888. 10.3389/fchem.2021.593888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  342. Jiang M.; Guo J.; Liu B.; Tan Q.; Xu B. Synthesis of Tellurium-Containing π-Extended Aromatics with Room-Temperature Phosphorescence. Org. Lett. 2019, 21, 8328–8333. 10.1021/acs.orglett.9b03106. [DOI] [PubMed] [Google Scholar]
  343. Kanazawa J.; Kitazawa Y.; Uchiyama M. Recent Progress in the Synthesis of the Monocarba-closo-dodecaborate(−) Anions. Chem.—Eur. J. 2019, 25, 9123–9132. 10.1002/chem.201900174. [DOI] [PubMed] [Google Scholar]
  344. Douvris C.; Ozerov O. V. Hydrodefluorination of Perfluoroalkyl Groups Using Silylium-Carborane Catalysts. Science. 2008, 321, 1188–1190. 10.1126/science.1159979. [DOI] [PubMed] [Google Scholar]
  345. Chen Z.; King R. B. Spherical Aromaticity: Recent Work on Fullerenes, Polyhedral Boranes, and Related Structures. Chem. Rev. 2005, 105, 3613–3642. 10.1021/cr0300892. [DOI] [PubMed] [Google Scholar]
  346. Akimoto G.; Otsuka M.; Miyamoto K.; Muranaka A.; Hashizume D.; Takita R.; Uchiyama M. One-pot Annulation for Biaryl-fused Monocarba-closo-dodecaborate through Aromatic B-H Bond Disconnection. Chem. Asian. J. 2018, 13, 913–917. 10.1002/asia.201800053. [DOI] [PubMed] [Google Scholar]
  347. Petersen T. B.; Khan R.; Olofsson B. Metal-Free Synthesis of Aryl Esters from Carboxylic Acids and Diaryliodonium Salts. Org. Lett. 2011, 13, 3462–3465. 10.1021/ol2012082. [DOI] [PubMed] [Google Scholar]
  348. Wang G.; Xiong B.; Zhou C.; Liu Y.; Xu W.; Yang C. A.; Tang K. W.; Wong W. Y. Copper-Catalyzed Diphenylation of P(O)-OH Bonds with Cyclic Diaryliodonium Salts. Chem. Asian. J. 2019, 14, 4365–4374. 10.1002/asia.201901284. [DOI] [PubMed] [Google Scholar]
  349. Ding M.; Hua W.; Liu M.; Zhang F. Pd-Catalyzed C(sp(3))-H Biarylation via Transient Directing Group Strategy. Org. Lett. 2020, 22, 7419–7423. 10.1021/acs.orglett.0c02353. [DOI] [PubMed] [Google Scholar]
  350. Fukuyama Y. Synthetic Studies on Small Molecule Natural Products with Neurotrophic Activity. Yakugaku Zasshi 2022, 142, 241–277. 10.1248/yakushi.21-00201. [DOI] [PubMed] [Google Scholar]
  351. Lombe B. K.; Feineis D.; Mudogo V.; Brun R.; Awale S.; Bringmann G. Michellamines A6 and A7, and Further Mono- and Dimeric Naphthylisoquinoline Alkaloids from A Congolese Ancistrocladus Liana and Their Antiausterity Activities against Pancreatic Cancer Cells. RSC Adv. 2018, 8, 5243–5254. 10.1039/C8RA00363G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  352. Wencel-Delord J.; Panossian A.; Leroux F. R.; Colobert F. Recent Advances and New Concepts for the Synthesis of Axially Stereoenriched Biaryls. Chem. Soc. Rev. 2015, 44, 3418–3430. 10.1039/C5CS00012B. [DOI] [PubMed] [Google Scholar]
  353. Xu M. M.; You X. Y.; Zhang Y. Z.; Lu Y.; Tan K.; Yang L.; Cai Q. Enantioselective Synthesis of Axially Chiral Biaryls by Diels-Alder/Retro-Diels-Alder Reaction of 2-Pyrones with Alkynes. J. Am. Chem. Soc. 2021, 143, 8993–9001. 10.1021/jacs.1c04759. [DOI] [PubMed] [Google Scholar]
  354. Wang G.; Shi Q.; Hu W.; Chen T.; Guo Y.; Hu Z.; Gong M.; Guo J.; Wei D.; Fu Z.; Huang W. Organocatalytic Asymmetric N-Sulfonyl Amide C-N Bond Activation to Access Axially Chiral Biaryl Amino Acids. Nat. Commun. 2020, 11, 946. 10.1038/s41467-020-14799-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  355. Wang C. S.; Wei L.; Fu C.; Wang X. H.; Wang C. J. Asymmetric Synthesis of Axially Chiral Naphthyl-C3-indoles via a Palladium-Catalyzed Cacchi Reaction. Org. Lett. 2021, 23, 7401–7406. 10.1021/acs.orglett.1c02574. [DOI] [PubMed] [Google Scholar]
  356. Jiang S.; Hu X.; Qiu J.; Guo H.; Yang F. A Fluorescent Sensor for Folic Acid Based on Crown Ether-Bridged Bis-Tetraphenylethylene. Analyst. 2019, 144, 2662–2669. 10.1039/C9AN00161A. [DOI] [PubMed] [Google Scholar]
  357. Kelber J. B.; Bensalah-Ledoux A.; Zahouani S.; Baguenard B.; Schaaf P.; Chaumont A.; Guy S.; Jierry L. Reversible Soft Mechanochemical Control of Biaryl Conformations through Crosslinking in a 3D Macromolecular Network. Angew. Chem., Int. Ed. Engl. 2020, 59, 23283–23290. 10.1002/anie.202010604. [DOI] [PubMed] [Google Scholar]
  358. Zuo Z.; Kim R. S.; Watson D. A. Synthesis of Axially Chiral 2,2’-Bisphosphobiarenes via a Nickel-Catalyzed Asymmetric Ullmann Coupling: General Access to Privileged Chiral Ligands without Optical Resolution. J. Am. Chem. Soc. 2021, 143, 1328–1333. 10.1021/jacs.0c12843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  359. Zhan B. B.; Wang L.; Luo J.; Lin X. F.; Shi B. F. Synthesis of Axially Chiral Biaryl-2-amines by Pd(II)-Catalyzed Free-Amine-Directed Atroposelective C-H Olefination. Angew. Chem., Int. Ed. Engl. 2020, 59, 3568–3572. 10.1002/anie.201915674. [DOI] [PubMed] [Google Scholar]
  360. Yu F.; Chen Y.; Jiang H.; Wang X. Recent Advances of BINOL-Based Sensors for Enantioselective Fluorescence Recognition. Analyst. 2020, 145 (21), 6769–6812. 10.1039/D0AN01225D. [DOI] [PubMed] [Google Scholar]
  361. Xu G.; Senanayake C. H.; Tang W. P-Chiral Phosphorus Ligands Based on a 2,3-Dihydrobenzo[d][1,3]oxaphosphole Motif for Asymmetric Catalysis. Acc. Chem. Res. 2019, 52, 1101–1112. 10.1021/acs.accounts.9b00029. [DOI] [PubMed] [Google Scholar]
  362. Xiang M.; Pfaffinger D. E.; Ortiz E.; Brito G. A.; Krische M. J. Enantioselective Ruthenium-BINAP-Catalyzed Carbonyl Reductive Coupling of Alkoxyallenes: Convergent Construction of syn-sec,tert-Diols via (Z)-sigma-Allylmetal Intermediates. J. Am. Chem. Soc. 2021, 143, 8849–8854. 10.1021/jacs.1c03480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  363. Wu X.; Han X.; Xu Q.; Liu Y.; Yuan C.; Yang S.; Liu Y.; Jiang J.; Cui Y. Chiral BINOL-Based Covalent Organic Frameworks for Enantioselective Sensing. J. Am. Chem. Soc. 2019, 141, 7081–7089. 10.1021/jacs.9b02153. [DOI] [PubMed] [Google Scholar]
  364. Wang D.; Liu W.; Tang M.; Yu N.; Yang X. Atroposelective Synthesis of Biaryl Diamines and Amino Alcohols via Chiral Phosphoric Acid Catalyzed para-Aminations of Anilines and Phenols. iScience. 2019, 22, 195–205. 10.1016/j.isci.2019.11.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  365. Sun G.; Deng Z.; Luo Z.; Wang Z.; Zhang J. Organocatalytic Asymmetric Arylation of p-Quinone Phosphonates: A Green Access to Biaryl Monophosphorus Ligands. Org. Lett. 2021, 23, 7630–7634. 10.1021/acs.orglett.1c02852. [DOI] [PubMed] [Google Scholar]
  366. Fatthalla M.; Grimblat N.; Brachet E.; Alami M.; Gandon V.; Le Bideau F.; Messaoudi S. Synthesis of Axially Chiral Biaryl Thioglycosides Through Thiosugar-Directed Pd-Catalyzed Asymmetric C-H Activation. Chem. Commun. 2021, 57, 10355–10358. 10.1039/D1CC03971G. [DOI] [PubMed] [Google Scholar]
  367. Shirakawa S.; Wu X.; Maruoka K. Kinetic Resolution of Axially Chiral 2-Amino-1,1’-biaryls by Phase-Transfer-Catalyzed N-Allylation. Angew. Chem., Int. Ed. Engl. 2013, 52, 14200–14203. 10.1002/anie.201308237. [DOI] [PubMed] [Google Scholar]
  368. Shen X.; Jones G. O.; Watson D. A.; Bhayana B.; Buchwald S. L. Enantioselective Synthesis of Axially Chiral Biaryls by the Pd-Catalyzed Suzuki-Miyaura Reaction: Substrate Scope and Quantum Mechanical Investigations. J. Am. Chem. Soc. 2010, 132, 11278–11287. 10.1021/ja104297g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  369. Kozlowski M. C.; Morgan B. J.; Linton E. C. Total Synthesis of Chiral Biaryl Natural Products by Asymmetric Biaryl Coupling. Chem. Soc. Rev. 2009, 38, 3193–3207. 10.1039/b821092f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  370. Tanaka K.; Nishida G.; Wada A.; Noguchi K. Enantioselective Synthesis of Axially Chiral Phthalides through Cationic [Rh(I)(H8-Binap)]-Catalyzed cross Alkyne Cyclotrimerization. Angew. Chem., Int. Ed. Engl. 2004, 43, 6510–6512. 10.1002/anie.200461533. [DOI] [PubMed] [Google Scholar]
  371. Shimada T.; Cho Y. H.; Hayashi T. Nickel-Catalyzed Asymmetric Grignard Cross-Coupling of Dinaphthothiophene Giving Axially Chiral 1,1’-binaphthyls. J. Am. Chem. Soc. 2002, 124, 13396–13397. 10.1021/ja0282588. [DOI] [PubMed] [Google Scholar]
  372. Kina A.; Miki H.; Cho Y. H.; Hayashi T. Palladium-Catalyzed Heck and Carbonylation Reactions of a Dinaphthaleneiodonium Salt Forming Functionalized 2-Iodo-1,1′-binaphthyls. Adv. Synth. & Catal. 2004, 346, 1728–1732. 10.1002/adsc.200404202. [DOI] [Google Scholar]
  373. Duan L.; Wang Z.; Zhao K.; Gu Z. Enantioselective Preparation of Atropisomeric Biaryl Trifluoromethylsulfanes via Ring-Opening of Cyclic Diaryliodoniums. Chem. Commun. 2021, 57, 3881–3884. 10.1039/D1CC00171J. [DOI] [PubMed] [Google Scholar]
  374. Hou M.; Deng R.; Gu Z. Cu-Catalyzed Enantioselective Atropisomer Synthesis via Thiolative Ring Opening of Five-Membered Cyclic Diaryliodoniums. Org. Lett. 2018, 20, 5779–5783. 10.1021/acs.orglett.8b02477. [DOI] [PubMed] [Google Scholar]
  375. Zhao K.; Yang S.; Gong Q.; Duan L.; Gu Z. Diols Activation by Cu/Borinic Acids Synergistic Catalysis in Atroposelective Ring-Opening of Cyclic Diaryliodoniums. Angew. Chem., Int. Ed. Engl. 2021, 60, 5788–5793. 10.1002/anie.202014127. [DOI] [PubMed] [Google Scholar]
  376. Xue X.; Gu Z. Synthesis of Bridged Biaryl Atropisomers via Sequential Cu- and Pd-Catalyzed Asymmetric Ring Opening and Cyclization. Org. Lett. 2019, 21, 3942–3945. 10.1021/acs.orglett.9b01062. [DOI] [PubMed] [Google Scholar]
  377. Xu S.; Zhao K.; Gu Z. Copper-Catalyzed Asymmetric Ring-Opening of Cyclic Diaryliodonium with Benzylic and Aliphatic Amines. Adv. Synth. & Catal. 2018, 360, 3877–3883. 10.1002/adsc.201800637. [DOI] [Google Scholar]
  378. Zhao K.; Duan L.; Xu S.; Jiang J.; Fu Y.; Gu Z. Enhanced Reactivity by Torsional Strain of Cyclic Diaryliodonium in Cu-Catalyzed Enantioselective Ring-Opening Reaction. Chem 2018, 4, 599–612. 10.1016/j.chempr.2018.01.017. [DOI] [Google Scholar]
  379. Wosinska-Hrydczuk M.; Skarzewski J. New Nitrogen, Sulfur-, and Selenium-Donating Ligands Derived from Chiral Pyridine Amino Alcohols. Synthesis and Catalytic Activity in Asymmetric Allylic Alkylation. Molecules. 2021, 26, 3493. 10.3390/molecules26123493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  380. Masuda R.; Yasukawa T.; Yamashita Y.; Kobayashi S. Nitrogen-Doped Carbon Enables Heterogeneous Asymmetric Insertion of Carbenoids into Amines Catalyzed by Rhodium Nanoparticles. Angew. Chem., Int. Ed. Engl. 2021, 60, 12786–12790. 10.1002/anie.202102506. [DOI] [PubMed] [Google Scholar]
  381. Mas-Rosello J.; Smejkal T.; Cramer N. Iridium-Catalyzed Acid-Assisted Asymmetric Hydrogenation of Oximes to Hydroxylamines. Science. 2020, 368, 1098–1102. 10.1126/science.abb2559. [DOI] [PubMed] [Google Scholar]
  382. Li C.; Pan Y.; Feng Y.; He Y. M.; Liu Y.; Fan Q. H. Asymmetric Ruthenium-Catalyzed Hydrogenation of Terpyridine-Type N-Heteroarenes: Direct Access to Chiral Tridentate Nitrogen Ligands. Org. Lett. 2020, 22, 6452–6457. 10.1021/acs.orglett.0c02268. [DOI] [PubMed] [Google Scholar]
  383. Chen C.; Peters J. C.; Fu G. C. Photoinduced Copper-Catalysed Asymmetric Amidation via Ligand Cooperativity. Nature. 2021, 596, 250–256. 10.1038/s41586-021-03730-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  384. Cheng F.; Duan D. S.; Jiang L. M.; Li B. S.; Wang J. X.; Zhou Y. J.; Jiao H. Y.; Wu T.; Zhu D. Y.; Wang S. H. Copper-Catalyzed Asymmetric Ring-Opening Reaction of Cyclic Diaryliodonium Salts with Imides. Org. Lett. 2022, 24, 1394–1399. 10.1021/acs.orglett.2c00247. [DOI] [PubMed] [Google Scholar]
  385. Guo W.; Gu J.; Gu Z. Catalytic Asymmetric Synthesis of Atropisomeric Nitrones: Versatile Intermediate for Axially Chiral Biaryls. Org. Lett. 2020, 22, 7622–7628. 10.1021/acs.orglett.0c02830. [DOI] [PubMed] [Google Scholar]
  386. Chao Z.; Ma M.; Gu Z. Cu-Catalyzed Site-Selective and Enantioselective Ring Opening of Cyclic Diaryliodoniums with 1,2,3-Triazoles. Org. Lett. 2020, 22, 6441–6446. 10.1021/acs.orglett.0c02256. [DOI] [PubMed] [Google Scholar]
  387. Xiao Y.; Sun Z.; Guo H.; Kwon O. Chiral Phosphines in Nucleophilic Organocatalysis. Beilstein. J. Org. Chem. 2014, 10, 2089–2121. 10.3762/bjoc.10.218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  388. Clayden J.; Moran W. J.; Edwards P. J.; LaPlante S. R. The Challenge of Atropisomerism in Drug Discovery. Angew. Chem., Int. Ed. Engl. 2009, 48, 6398–6401. 10.1002/anie.200901719. [DOI] [PubMed] [Google Scholar]
  389. Wang Y. B.; Tan B. Construction of Axially Chiral Compounds via Asymmetric Organocatalysis. Acc. Chem. Res. 2018, 51, 534–547. 10.1021/acs.accounts.7b00602. [DOI] [PubMed] [Google Scholar]
  390. Kumar R.; Wirth T. Asymmetric Synthesis with Hypervalent Iodine Reagents. Top. Curr. Chem. 2015, 373, 243–261. 10.1007/128_2015_639. [DOI] [PubMed] [Google Scholar]
  391. Ke J.; Zu B.; Guo Y.; Li Y.; He C. Hexafluoroisopropanol-Enabled Copper-Catalyzed Asymmetric Halogenation of Cyclic Diaryliodoniums for the Synthesis of Axially Chiral 2,2’-Dihalobiaryls. Org. Lett. 2021, 23, 329–333. 10.1021/acs.orglett.0c03833. [DOI] [PubMed] [Google Scholar]
  392. Dai Q.; Li W.; Li Z.; Zhang J. P-Chiral Phosphines Enabled by Palladium/Xiao-Phos-Catalyzed Asymmetric P-C Cross-Coupling of Secondary Phosphine Oxides and Aryl Bromides. J. Am. Chem. Soc. 2019, 141, 20556–20564. 10.1021/jacs.9b11938. [DOI] [PubMed] [Google Scholar]
  393. Cabre A.; Riera A.; Verdaguer X. P-Stereogenic Amino-Phosphines as Chiral Ligands: From Privileged Intermediates to Asymmetric Catalysis. Acc. Chem. Res. 2020, 53, 676–689. 10.1021/acs.accounts.9b00633. [DOI] [PubMed] [Google Scholar]
  394. Duan L.; Zhao K.; Wang Z.; Zhang F.-L.; Gu Z. Enantioselective Ring-Opening/Oxidative Phosphorylation and P-Transfer Reaction of Cyclic Diaryliodoniums. ACS Catal. 2019, 9, 9852–9858. 10.1021/acscatal.9b03454. [DOI] [Google Scholar]
  395. Zhu K.; Xu K.; Fang Q.; Wang Y.; Tang B.; Zhang F. Enantioselective Synthesis of Axially Chiral Biaryls via Cu-Catalyzed Acyloxylation of Cyclic Diaryliodonium Salts. ACS Catal. 2019, 9, 4951–4957. 10.1021/acscatal.9b00695. [DOI] [Google Scholar]
  396. Zhu C.; Yang B.; Mai B. K.; Palazzotto S.; Qiu Y.; Gudmundsson A.; Ricke A.; Himo F.; Backvall J. E. Highly Selective Palladium-Catalyzed Hydroborylative Carbocyclization of Bisallenes to Seven-Membered Rings. J. Am. Chem. Soc. 2018, 140, 14324–14333. 10.1021/jacs.8b08708. [DOI] [PubMed] [Google Scholar]
  397. Wang M.; Yin L.; Cheng L.; Yang Y.; Li Y. Straightforward Stereoselective Synthesis of Seven-Membered Oxa-Bridged Rings through In Situ Generated Cycloheptenol Derivatives. J. Org. Chem. 2021, 86, 12956–12963. 10.1021/acs.joc.1c01648. [DOI] [PubMed] [Google Scholar]
  398. Trost B. M.; Zuo Z.; Schultz J. E. Transition-Metal-Catalyzed Cycloaddition Reactions to Access Seven-Membered Rings. Chem.—Eur. J. 2020, 26, 15354–15377. 10.1002/chem.202002713. [DOI] [PubMed] [Google Scholar]
  399. Sasaki I.; Ohmura T.; Suginome M. Construction of Silicon-Containing Seven-Membered Rings by Catalytic [4 + 2 + 1] Cycloaddition through Rhodium Silylenoid. Org. Lett. 2020, 22, 2961–2966. 10.1021/acs.orglett.0c00690. [DOI] [PubMed] [Google Scholar]
  400. Kawai K.; Kato K.; Peng L.; Segawa Y.; Scott L. T.; Itami K. Synthesis and Structure of a Propeller-Shaped Polycyclic Aromatic Hydrocarbon Containing Seven-Membered Rings. Org. Lett. 2018, 20, 1932–1935. 10.1021/acs.orglett.8b00477. [DOI] [PubMed] [Google Scholar]
  401. Hoshino D.; Mori K. Divergent Access to Seven/Five-Membered Rings Based on [1,6]-Hydride Shift/Cyclization Process. Org. Lett. 2021, 23, 9403–9407. 10.1021/acs.orglett.1c03523. [DOI] [PubMed] [Google Scholar]
  402. Dordevic L.; Milano D.; Demitri N.; Bonifazi D. O-Annulation to Polycyclic Aromatic Hydrocarbons: A Tale of Optoelectronic Properties from Five- to Seven-Membered Rings. Org. Lett. 2020, 22, 4283–4288. 10.1021/acs.orglett.0c01331. [DOI] [PubMed] [Google Scholar]
  403. Doerksen R. S.; Hodik T.; Hu G.; Huynh N. O.; Shuler W. G.; Krische M. J. Ruthenium-Catalyzed Cycloadditions to Form Five-, Six-, and Seven-Membered Rings. Chem. Rev. 2021, 121, 4045–4083. 10.1021/acs.chemrev.0c01133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  404. Wang N.; Saidhareddy P.; Jiang X. Construction of Sulfur-Containing Moieties in the Total Synthesis of Natural Products. Nat. Prod. Rep. 2020, 37, 246–275. 10.1039/C8NP00093J. [DOI] [PubMed] [Google Scholar]
  405. Feng M.; Tang B.; Liang S. H.; Jiang X. Sulfur Containing Scaffolds in Drugs: Synthesis and Application in Medicinal Chemistry. Curr. Top. Med. Chem. 2016, 16, 1200–1216. 10.2174/1568026615666150915111741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  406. Blachier F.; Andriamihaja M.; Blais A. Sulfur-Containing Amino Acids and Lipid Metabolism. J. Nutr. 2020, 150, 2524S–2531S. 10.1093/jn/nxaa243. [DOI] [PubMed] [Google Scholar]
  407. Wu Y. J.; Meanwell N. A. Geminal Diheteroatomic Motifs: Some Applications of Acetals, Ketals, and Their Sulfur and Nitrogen Homologues in Medicinal Chemistry and Drug Design. J. Med. Chem. 2021, 64, 9786–9874. 10.1021/acs.jmedchem.1c00790. [DOI] [PubMed] [Google Scholar]
  408. Peng Y. H.; Liao F. Y.; Tseng C. T.; Kuppusamy R.; Li A. S.; Chen C. H.; Fan Y. S.; Wang S. Y.; Wu M. H.; Hsueh C. C.; et al. Unique Sulfur-Aromatic Interactions Contribute to the Binding of Potent Imidazothiazole Indoleamine 2,3-Dioxygenase Inhibitors. J. Med. Chem. 2020, 63, 1642–1659. 10.1021/acs.jmedchem.9b01549. [DOI] [PubMed] [Google Scholar]
  409. McConnell D. B. Biotin’s Lessons in Drug Design. J. Med. Chem. 2021, 64, 16319–16327. 10.1021/acs.jmedchem.1c00975. [DOI] [PubMed] [Google Scholar]
  410. Ma Z.; Rao L.; Bierbach U. Replacement of a Thiourea-S with an Amidine-NH Donor Group in a Platinum-Acridine Antitumor Compound Reduces the Metal’s Reactivity with Cysteine Sulfur. J. Med. Chem. 2009, 52, 3424–3427. 10.1021/jm900451y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  411. Ilardi E. A.; Vitaku E.; Njardarson J. T. Data-Mining for Sulfur and Fluorine: an Evaluation of Pharmaceuticals to Reveal Opportunities for Drug Design and Discovery. J. Med. Chem. 2014, 57, 2832–2842. 10.1021/jm401375q. [DOI] [PubMed] [Google Scholar]
  412. Beno B. R.; Yeung K. S.; Bartberger M. D.; Pennington L. D.; Meanwell N. A. A Survey of the Role of Noncovalent Sulfur Interactions in Drug Design. J. Med. Chem. 2015, 58, 4383–4438. 10.1021/jm501853m. [DOI] [PubMed] [Google Scholar]
  413. Yu Y.; Zhu S.; Hou Y.; Li J.; Guan S. Sulfur Contents in Sulfonated Hyaluronic Acid Direct the Cardiovascular Cells Fate. ACS Appl. Mater. Interfaces. 2020, 12, 46827–46836. 10.1021/acsami.0c15729. [DOI] [PubMed] [Google Scholar]
  414. Transue W. J.; Nava M.; Terban M. W.; Yang J.; Greenberg M. W.; Wu G.; Foreman E. S.; Mustoe C. L.; Kennepohl P.; Owen J. S.; et al. Anthracene as a Launchpad for a Phosphinidene Sulfide and for Generation of a Phosphorus-Sulfur Material Having the Composition P2S, a Vulcanized Red Phosphorus That Is Yellow. J. Am. Chem. Soc. 2019, 141, 431–440. 10.1021/jacs.8b10775. [DOI] [PubMed] [Google Scholar]
  415. Pal A.; Arshad F.; Sk M. P. Emergence of sulfur quantum dots: Unfolding their synthesis, properties, and applications. Adv. Colloid Interface Sci. 2020, 285, 102274. 10.1016/j.cis.2020.102274. [DOI] [PubMed] [Google Scholar]
  416. Liu S.; Cao Y.; Wu Z.; Chen H. Reactive Films Fabricated Using Click Sulfur(VI)-Fluoride Exchange Reactions via Layer-by-Layer Assembly. J. Mater. Chem. B 2020, 8, 5529–5534. 10.1039/D0TB00908C. [DOI] [PubMed] [Google Scholar]
  417. Lim J.; Pyun J.; Char K. Recent Approaches for the Direct Use of Elemental Sulfur in the Synthesis and Processing of Advanced Materials. Angew. Chem., Int. Ed. Engl. 2015, 54, 3249–3258. 10.1002/anie.201409468. [DOI] [PubMed] [Google Scholar]
  418. Hakkinen H. The Gold-Sulfur Interface at the Nanoscale. Nat. Chem. 2012, 4, 443–455. 10.1038/nchem.1352. [DOI] [PubMed] [Google Scholar]
  419. Crockett M. P.; Evans A. M.; Worthington M. J.; Albuquerque I. S.; Slattery A. D.; Gibson C. T.; Campbell J. A.; Lewis D. A.; Bernardes G. J.; Chalker J. M. Sulfur-Limonene Polysulfide: A Material Synthesized Entirely from Industrial By-Products and Its Use in Removing Toxic Metals from Water and Soil. Angew. Chem., Int. Ed. Engl. 2016, 55, 1714–1718. 10.1002/anie.201508708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  420. Buvignier A.; Peyre-Lavigne M.; Robin O.; Bounouba M.; Patapy C.; Bertron A.; Paul E. Influence of Dissolved-Aluminum Concentration on Sulfur-Oxidizing Bacterial Activity in the Biodeterioration of Concrete. Appl. Environ. Microbiol. 2019, 85, e00302-19 10.1128/AEM.00302-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  421. Wang M.; Fan Q.; Jiang X. Transition-Metal-Free Diarylannulated Sulfide and Selenide Construction via Radical/Anion-Mediated Sulfur-Iodine and Selenium-Iodine Exchange. Org. Lett. 2016, 18, 5756–5759. 10.1021/acs.orglett.6b03078. [DOI] [PubMed] [Google Scholar]
  422. Luo B.; Cui Q.; Luo H.; Hu Y.; Huang P.; Wen S. N-Benzyldithiocarbamate Salts as Sulfur Sources to Access Tricyclic Thioheterocycles Mediated by Copper Species. Adv. Synth. & Catal. 2016, 358, 2733–2738. 10.1002/adsc.201600405. [DOI] [Google Scholar]
  423. Li Y.-L.; Zhang C.; Jin X.; Yu X.-L.; Wang Z.; Zhang H.-Z.; Yang W.-Q. Copper-Catalyzed Asymmetric Intermolecular N-Monoarylation of Unprotected Sulfonamides via Desymmetrization of Diaryliodonium Salts at Room Temperature. Synlett. 2020, 31, 1077–1081. 10.1055/s-0040-1707121. [DOI] [Google Scholar]
  424. Zhu K.; Wang Y.; Fang Q.; Song Z.; Zhang F. Enantioselective Synthesis of Axially Chiral Biaryls via Copper-Catalyzed Thiolation of Cyclic Diaryliodonium Salts. Org. Lett. 2020, 22, 1709–1713. 10.1021/acs.orglett.9b04555. [DOI] [PubMed] [Google Scholar]
  425. Ye R.; Zhukhovitskiy A. V.; Deraedt C. V.; Toste F. D.; Somorjai G. A. Supported Dendrimer-Encapsulated Metal Clusters: Toward Heterogenizing Homogeneous Catalysts. Acc. Chem. Res. 2017, 50, 1894–1901. 10.1021/acs.accounts.7b00232. [DOI] [PubMed] [Google Scholar]
  426. Smith P. T.; Nichols E. M.; Cao Z.; Chang C. J. Hybrid Catalysts for Artificial Photosynthesis: Merging Approaches from Molecular, Materials, and Biological Catalysis. Acc. Chem. Res. 2020, 53, 575–587. 10.1021/acs.accounts.9b00619. [DOI] [PubMed] [Google Scholar]
  427. Singh B.; Gawande M. B.; Kute A. D.; Varma R. S.; Fornasiero P.; McNeice P.; Jagadeesh R. V.; Beller M.; Zboril R. Single-Atom (Iron-Based) Catalysts: Synthesis and Applications. Chem. Rev. 2021, 121, 13620–13697. 10.1021/acs.chemrev.1c00158. [DOI] [PubMed] [Google Scholar]
  428. Shi J.; Wang X.; Zhang S.; Xiao L.; Huan Y.; Gong Y.; Zhang Z.; Li Y.; Zhou X.; Hong M.; et al. Two-Dimensional Metallic Tantalum Disulfide as a Hydrogen Evolution Catalyst. Nat. Commun. 2017, 8, 958. 10.1038/s41467-017-01089-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  429. Mondelli C.; Gozaydin G.; Yan N.; Perez-Ramirez J. Biomass Valorisation over Metal-Based Solid Catalysts from Nanoparticles to Single Atoms. Chem. Soc. Rev. 2020, 49, 3764–3782. 10.1039/D0CS00130A. [DOI] [PubMed] [Google Scholar]
  430. Jiao L.; Xu W.; Wu Y.; Yan H.; Gu W.; Du D.; Lin Y.; Zhu C. Single-Atom Catalysts Boost Signal Amplification For Biosensing. Chem. Soc. Rev. 2021, 50, 750–765. 10.1039/D0CS00367K. [DOI] [PubMed] [Google Scholar]
  431. Chang B.; Zhang L.; Wu S.; Sun Z.; Cheng Z. Engineering Single-Atom Catalysts Toward Biomedical Applications. Chem. Soc. Rev. 2022, 51, 3688–3743. 10.1039/D1CS00421B. [DOI] [PubMed] [Google Scholar]
  432. Rajendran A.; Rajendiran M.; Yang Z. F.; Fan H. X.; Cui T. Y.; Zhang Y. G.; Li W. Y. Functionalized Silicas for Metal-Free and Metal-Based Catalytic Applications: A Review in Perspective of Green Chemistry. Chem. Rec. 2020, 20, 513–540. 10.1002/tcr.201900056. [DOI] [PubMed] [Google Scholar]
  433. Parvin T.; Yadav R.; Choudhury L. H. Recent Applications Of Thiourea-Based Organocatalysts In Asymmetric Multicomponent Reactions (Amcrs). Org. Biomol. Chem. 2020, 18, 5513–5532. 10.1039/D0OB00595A. [DOI] [PubMed] [Google Scholar]
  434. Yuan Y. C.; Mellah M.; Schulz E.; David O. R. P. Making Chiral Salen Complexes Work with Organocatalysts. Chem. Rev. 2022, 122, 8841–8883. 10.1021/acs.chemrev.1c00912. [DOI] [PubMed] [Google Scholar]
  435. Yoshida Y.; Ishikawa S.; Mino T.; Sakamoto M. Bromonium Salts: Diaryl-Lambda(3)-Bromanes as Halogen-Bonding Organocatalysts. Chem. Commun. 2021, 57, 2519–2522. 10.1039/D0CC07733J. [DOI] [PubMed] [Google Scholar]
  436. Wani A. A.; Chourasiya S. S.; Kathuria D.; Bharatam P. V. 1,1-Diaminoazines As Organocatalysts In Phospha-Michael Addition Reactions. Chem. Commun. 2021, 57, 11717–11720. 10.1039/D1CC04657H. [DOI] [PubMed] [Google Scholar]
  437. Polackova V.; Kristofikova D.; Nemethova B.; Gorova R.; Meciarova M.; Sebesta R. N-Sulfinylpyrrolidine-Containing Ureas And Thioureas As Bifunctional Organocatalysts. Beilstein. J. Org. Chem. 2021, 17, 2629–2641. 10.3762/bjoc.17.176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  438. Mahato C. K.; Mukherjee S.; Kundu M.; Pramanik A. Pyrrolidine-Oxadiazolone Conjugates as Organocatalysts in Asymmetric Michael Reaction. J. Org. Chem. 2019, 84, 1053–1063. 10.1021/acs.joc.8b02393. [DOI] [PubMed] [Google Scholar]
  439. Huang X.; Zhang W. Recyclable Fluorous Cinchona Organocatalysts For Asymmetric Synthesis Of Biologically Interesting Compounds. Chem. Commun. 2021, 57, 10116–10124. 10.1039/D1CC03722F. [DOI] [PubMed] [Google Scholar]
  440. Fallek A.; Weiss-Shtofman M.; Kramer M.; Dobrovetsky R.; Portnoy M. Phosphorylation Organocatalysts Highly Active by Design. Org. Lett. 2020, 22, 3722–3727. 10.1021/acs.orglett.0c01226. [DOI] [PubMed] [Google Scholar]
  441. An F.; Maji B.; Min E.; Ofial A. R.; Mayr H. Basicities and Nucleophilicities of Pyrrolidines and Imidazolidinones Used as Organocatalysts. J. Am. Chem. Soc. 2020, 142, 1526–1547. 10.1021/jacs.9b11877. [DOI] [PubMed] [Google Scholar]
  442. Wolf J.; Huber F.; Erochok N.; Heinen F.; Guerin V.; Legault C. Y.; Kirsch S. F.; Huber S. M. Activation of a Metal-Halogen Bond by Halogen Bonding. Angew. Chem., Int. Ed. Engl. 2020, 59, 16496–16500. 10.1002/anie.202005214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  443. Szell P. M. J.; Zablotny S.; Bryce D. L. Halogen Bonding as A Supramolecular Dynamics Catalyst. Nat. Commun. 2019, 10, 916. 10.1038/s41467-019-08878-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  444. Czarny R. S.; Ho A. N.; Shing Ho P. A Biological Take on Halogen Bonding and Other Non-Classical Non-Covalent Interactions. Chem. Rec. 2021, 21, 1240–1251. 10.1002/tcr.202100076. [DOI] [PubMed] [Google Scholar]
  445. Berger G.; Frangville P.; Meyer F. Halogen Bonding For Molecular Recognition: New Developments In Materials And Biological Sciences. Chem. Commun. 2020, 56, 4970–4981. 10.1039/D0CC00841A. [DOI] [PubMed] [Google Scholar]
  446. Il’in M. V.; Sysoeva A. A.; Novikov A. S.; Bolotin D. S. Diaryliodoniums as Hybrid Hydrogen- and Halogen-Bond-Donating Organocatalysts for the Groebke-Blackburn-Bienayme Reaction. J. Org. Chem. 2022, 87, 4569–4579. 10.1021/acs.joc.1c02885. [DOI] [PubMed] [Google Scholar]
  447. Heinen F.; Engelage E.; Cramer C. J.; Huber S. M. Hypervalent Iodine(III) Compounds as Biaxial Halogen Bond Donors. J. Am. Chem. Soc. 2020, 142, 8633–8640. 10.1021/jacs.9b13309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  448. Saito M.; Kobayashi Y.; Tsuzuki S.; Takemoto Y. Electrophilic Activation of Iodonium Ylides by Halogen-Bond-Donor Catalysis for Cross-Enolate Coupling. Angew. Chem., Int. Ed. Engl. 2017, 56, 7653–7657. 10.1002/anie.201703641. [DOI] [PubMed] [Google Scholar]
  449. Boelke A.; Kuczmera T. J.; Lork E.; Nachtsheim B. J. N-Heterocyclic Iod(az)olium Salts - Potent Halogen-Bond Donors in Organocatalysis. Chem.—Eur. J. 2021, 27, 13128–13134. 10.1002/chem.202101961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  450. Han J.; Zhang Y.; Liu Z. J. Diaryliodonium Salts as Efficient Lewis Acid Catalysts for Direct Three Component Mannich Reactions. RSC Adv. 2015, 5, 25485–25488. 10.1039/C5RA00209E. [DOI] [Google Scholar]
  451. Heinen F.; Engelage E.; Dreger A.; Weiss R.; Huber S. M. Iodine(III) Derivatives as Halogen Bonding Organocatalysts. Angew. Chem., Int. Ed. Engl. 2018, 57, 3830–3833. 10.1002/anie.201713012. [DOI] [PubMed] [Google Scholar]
  452. Labattut A.; Tremblay P. L.; Moutounet O.; Legault C. Y. Experimental and Theoretical Quantification of the Lewis Acidity of Iodine(III) Species. J. Org. Chem. 2017, 82, 11891–11896. 10.1021/acs.joc.7b01616. [DOI] [PubMed] [Google Scholar]
  453. Boelke A.; Kuczmera T. J.; Lork E.; Nachtsheim B. J. N-Heterocyclic Iod(az)olium Salts - Potent Halogen-Bond Donors in Organocatalysis. Chem.—Eur. J. 2021, 27, 13128–13134. 10.1002/chem.202101961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  454. Nishida Y.; Suzuki T.; Takagi Y.; Amma E.; Tajima R.; Kuwano S.; Arai T. A Hypervalent Cyclic Dibenzoiodolium Salt as a Halogen-Bond-Donor Catalyst for the [4 + 2] Cycloaddition of 2-Alkenylindoles. Chempluschem. 2021, 86, 741–744. 10.1002/cplu.202100089. [DOI] [PubMed] [Google Scholar]
  455. Zheng H.; Sang Y.; Houk K. N.; Xue X. S.; Cheng J. P. Mechanism and Origins of Enantioselectivities in Spirobiindane-Based Hypervalent Iodine(III)-Induced Asymmetric Dearomatizing Spirolactonizations. J. Am. Chem. Soc. 2019, 141, 16046–16056. 10.1021/jacs.9b08243. [DOI] [PubMed] [Google Scholar]
  456. Mondal P.; Pirovano P.; Das A.; Farquhar E. R.; McDonald A. R. Hydrogen Atom Transfer by a High-Valent Nickel-Chloride Complex. J. Am. Chem. Soc. 2018, 140, 1834–1841. 10.1021/jacs.7b11953. [DOI] [PubMed] [Google Scholar]
  457. Bower J. K.; Cypcar A. D.; Henriquez B.; Stieber S. C. E.; Zhang S. C(sp(3))-H Fluorination with a Copper(II)/(III) Redox Couple. J. Am. Chem. Soc. 2020, 142, 8514–8521. 10.1021/jacs.0c02583. [DOI] [PubMed] [Google Scholar]
  458. Voelkel M. H. H.; Wonner P.; Huber S. M. Preorganization: A Powerful Tool in Intermolecular Halogen Bonding in Solution. ChemistryOpen. 2020, 9, 214–224. 10.1002/open.201900355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  459. Yunusova S. N.; Novikov A. S.; Soldatova N. S.; Vovk M. A.; Bolotin D. S. Iodonium Salts as Efficient Iodine(III)-Based Noncovalent Organocatalysts for Knorr-Type Reactions. RSC Adv. 2021, 11, 4574–4583. 10.1039/D0RA09640G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  460. Yoshida Y.; Mino T.; Sakamoto M. Chiral Hypervalent Bromine(III) (Bromonium Salt): Hydrogen- and Halogen-Bonding Bifunctional Asymmetric Catalysis by Diaryl-λ3 -bromanes. ACS Catal. 2021, 11, 13028–13033. 10.1021/acscatal.1c04070. [DOI] [Google Scholar]

Articles from Chemical Reviews are provided here courtesy of American Chemical Society

RESOURCES