Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2025 Mar 7;125(6):3440–3550. doi: 10.1021/acs.chemrev.4c00808

Iodoarene Activation: Take a Leap Forward toward Green and Sustainable Transformations

Toshifumi Dohi 1,‡,*, Elghareeb E Elboray 1,§, Kotaro Kikushima 1, Koji Morimoto 1,, Yasuyuki Kita ‡,*
PMCID: PMC11951092  PMID: 40053418

Abstract

graphic file with name cr4c00808_0206.jpg

Constructing chemical bonds under green sustainable conditions has drawn attention from environmental and economic perspectives. The dissociation of (hetero)aryl–halide bonds is a crucial step of most arylations affording (hetero)arene derivatives. Herein, we summarize the (hetero)aryl halides activation enabling the direct (hetero)arylation of trapping reagents and construction of highly functionalized (hetero)arenes under benign conditions. The strategies for the activation of aryl iodides are classified into (a) hypervalent iodoarene activation followed by functionalization under thermal/photochemical conditions, (b) aryl–I bond dissociation in the presence of bases with/without organic catalysts and promoters, (c) photoinduced aryl–I bond dissociation in the presence/absence of organophotocatalysts, (d) electrochemical activation of aryl iodides by direct/indirect electrolysis mediated by organocatalysts and mediators acting as electron shuttles, and (e) electrophotochemical activation of aryl iodides mediated by redox-active organocatalysts. These activation modes result in aryl iodides exhibiting diverse reactivity as formal aryl cations/radicals/anions and aryne precursors. The coupling of these reactive intermediates with trapping reagents leads to the facile and selective formation of C–C and C–heteroatom bonds. These ecofriendly, inexpensive, and functional group-tolerant activation strategies offer green alternatives to transition metal-based catalysis.

1. Introduction

The activation of aryl–halide bonds through transition metal–catalyzed oxidative addition (Chart 1a) has enabled the development of various cross-coupling techniques for constructing C–C, C–O, and C–N bonds.18 The importance of these transformations, which are crucial for the synthesis of natural products, pharmaceuticals, compounds used in optical devices, and industrially useful starting materials,9,10 is reflected by the Nobel Prize awarded to Heck, Negishi, and Suzuki in 2010.46

Chart 1. Possible Nucleophile–Organoiodine Couplings.

Chart 1

Transition metal-catalyzed cross-couplings provide access to a wide range of products but exhibit several drawbacks, including the generation of byproducts and metal-containing waste and the high cost and air/moisture sensitivity of most catalysts. The cost factor is particularly critical when large quantities are required. The generation of byproducts and metal-containing waste also increases environmental and economic costs. Additionally, the isolated products often contain metal impurities, therefore requiring additional purification using specialized equipment and having a limited scope of pharmaceutical and industrial applications.11,12 Furthermore, the production of palladium, the primary transition metal used to promote cross-coupling reactions, is strongly influenced by social and environmental factors,13 which, together with the scarcity of this metal, considerably affects the cost of the corresponding catalysts. Therefore, novel ecofriendly and sustainable transformation strategies aligning with the advancement of modern synthetic chemistry are urgently required. The nucleophilic aromatic substitution reactions of aryl iodides enable straightforward arene functionalization (Chart 1b) but feature several limitations, such as narrow substrate scope, functional group sensitivity, and harsh conditions.14,15

In this review, we explore the activation methods of organoiodines and discuss their applications in the conversion of aryl–iodide bonds, revealing that these methods enable arylation reactions that may not be achievable under conventional conditions (Chart 1c). We highlight recent pioneering works and outline five key iodoarene activation strategies, as listed below (Chart 2). Although this work mainly focuses on the activation of iodoarenes, the presented concepts can be applied to other haloarenes.

  • a)

    Oxidative activation of Ar–I bond (Section 2): Activation to hypervalent organoiodines (e.g., diaryliodonium salts) enables transition metal-free bond formation via ligand coupling, enabling the arylation of diverse nucleophiles and formation of valuable heterocycles. The relatively low oxidation potential of the iodine atom presents a significant advantage for the oxidative activation of iodoarenes in comparison to other halogen atoms, which exhibit greater resistance to oxidation. Diaryliodonium species can act not only as aryl cation synthons but also as aryne and aryl radical equivalents, depending on the applied conditions.

  • b)

    Base-promoted Ar–I bond dissociation (Section 3): The activation of aryl iodides by bases with or without organocatalysts is included to shed light on the importance of transition metal-free approaches for C–H arylation, aliphatic carbon activation, addition to unsaturated bonds, hydrodehalogenation, dehalogenative deuteration, hydroxylation, carbonylation, formylation, etc.

  • c)

    Photoinduced Ar–I bond dissociation (Section 4): Photochemical iodoarene activation is discussed to address the significance of this green approach in related recent pioneering works. Several strategies have been developed to make this photoactivation chemistry more desirable and accessible even under visible-light irradiation, as exemplified by the direct excitation of haloarenes or facilitation of iodoarene reduction through single electron transfer (SET) from photoexcited catalysts, reaction components, bases, and electron donor–acceptor (EDA) complexes.

  • d)

    Electrochemical dissociation of Ar–I bond (Section 5): The electrochemical activation of aryl halides is a promising yet underexplored green route to functionalized arenes.

  • e)

    Electrophotoinduced activation of Ar–I bond (Section 6): Electrophotochemical activation enables the dissociation of the difficult-to-activate aryl–halide bonds.

Chart 2. Iodoarene and Haloarene Activation Strategiesa.

Chart 2

a ED = Electron Donor Species; PC = Photocatalyst.

Based on the above contents, the evaluations of iodoarene activations are discussed in Section 7 from the perspective of green sustainability. These evaluations include synthetic methods for iodoarenes, the design of diaryliodonium salts, and the comparison and assessment of reaction conditions in borylation and hydroxylation.

2. Oxidative Activation of Ar–I toward Arylation Reactions

The use of hypervalent iodine chemistry is a highly versatile and ecofriendly strategy enabling a wide range of transformations.1620 This approach exploits readily available nontoxic materials and does not rely on transition metals or ligands, thus presenting a sustainable alternative to traditional methods. The reactivity of hypervalent iodine reagents is based on the strong electrophilicity of the central iodine and high leaving ability of the aryl iodine group, which is ∼106 times greater than that of the triflate group.2123 These properties resemble those of heavy metal-based oxidants and transition metal species, inspiring the development of new synthetic transformations. Hypervalent iodine reagents are therefore used as green alternatives to heavy metals in modern organic synthesis,2426 and their ability to promote the formation of various bonds, such as C–C, C–N, C–X (X = halogen), and C–O, enables the construction of a wide range of organic skeletons.2737 This high oxidative coupling reactivity is also important for the total synthesis of biologically active natural products and their related scaffolds.3841 In addition to featuring excellent oxidant and electrophilic group-transfer abilities, hypervalent iodine reagents can promote different types of rearrangements,4244 such as the Hofmann, Beckmann, [1,2]-migration, [3,3]-sigmatropic/iodonio-Claisen-type, ring contraction, and ring expansion rearrangements, which further enhances their versatility as a synthetic tool.

The importance of hypervalent iodine chemistry as an efficient transition metal-free approach was highlighted after the independent discovery by Kita and Ochiai in the 1980s. Hypervalent iodine reagents are required in stoichiometric or excess amounts, which could have a negative environmental impact and conflict with the principles of green and sustainable chemistry. In 2005, these researchers demonstrated catalytic oxidative transformations facilitated by hypervalent iodine reagents, which were generated in situ through the oxidation of aryl iodides in the presence of sacrificial oxidants.45,46 This method was promptly applied to asymmetric synthesis using catalytic chiral hypervalent organoiodines, which led to the development and testing of various strategies and numerous chiral organoiodines.4753 This innovation has created new possibilities in the field of transition metal-free coupling.

Hypervalent organoiodines are categorized into two classes, λ3- and λ5-iodanes, based on the oxidation state of the central iodine atom (Figure 1). For instance, iodosoarenes, aryliodine carboxylates, and aryliodine organosulfonates are widely used as strong oxidizing reagents and are classified as λ3-iodanes. Among these compounds, phenyliodine diacetate (PIDA, PhI(OAc)2), phenyliodine bis(trifluoroacetate) (PIFA, PhI(OCOCF3)2), and [hydroxy(tosyloxy)iodo]benzene (HTIB, Koser’s reagent) are frequently employed in oxidative transformations. In contrast, aryliodine dihalides (X = F or Cl) are effective halogenating agents. Benziodoxoles are more stable than their acyclic counterparts and are valuable reagents for transferring functional (e.g., trifluoromethyl, ethynyl, cyano, and azido) groups. Diaryliodonium salts exhibit diverse reactivity because of the remarkable leaving ability of the iodoarene moiety and are, therefore, helpful aryl-group-transfer agents. Iodonium ylides and imides are excellent generators of carbene and nitrene species, respectively. The most essential λ5-iodanes in organic synthesis are 2-iodoxybenzoic acid (IBX) and the Dess–Martin periodinane (DMP), which are mild and highly selective reagents for the oxidation of alcohols.

Figure 1.

Figure 1

Hypervalent iodine reagents employed in organic synthesis.

Hypervalent iodine compounds, some of which are known for their explosive nature, have been discussed in review articles.1619 Pentavalent iodine reagents, such as IBX, DMP, and iodylarenes (ArIO2), can explode upon heating or impact. Various analogs of IBX have been developed and are summarized in the literature.18,19 In contrast, trivalent iodine reagents are relatively stable and less explosive. The frequently used reagents, such as PhIO, PIDA, PIFA, and HTIB, can be stored over long-term and are commercially available. However, drying PhIO at high temperatures causes disproportionation to PhI and PhIO2, which has the possibility of severe explosion. N-Tosyliminoiodanes are stable and can be stored for extended periods, whereas trifluoroacetate and mesyl derivatives are thermally sensitive and explosive. Benziodoxoles have higher thermal stability, enabling the development of various stable atom-transferring reagents including even those containing the azide group. In addition, diaryliodonium salts generally exhibit high stability, allowing them to be stored for extended periods. To the best of our knowledge, no explosive properties of typically used diaryliodonium salts have been reported so far.

Reactions involving hypervalent iodine reagents are initiated by the coordination of the nucleophile with the central iodine atom,54 which is highly Lewis-acidic and reactive owing to the presence of a region of positive electrostatic potential known as the σ-hole. This σ-hole not only affects the reactivity of hypervalent iodine but also facilitates molecular assembly through inter-/intramolecular secondary interactions, or halogen bonds, with nucleophilic regions. These anisotropic properties of iodine, particularly the electrophilic σ-holes, are essential for understanding the ability of hypervalent iodine species to engage in secondary interactions and halogen bonding. The number of σ-holes in λ3-iodanes can be one or two, whereas up to four are possible in λ5-iodanes (Figure 2).55

Figure 2.

Figure 2

(a) Classical σ-hole of λ3-iodane, (b) nonclassical σ-hole with two maxima, (c) interaction with bidentate nucleophiles, (d) activation of benziodoxoles by acids.

The classical σ-hole λ3-iodane is generated along with the extension of the primary σ-bond (Figure 2a). The σ-hole strength is determined by the nature of the primary two-center-two-electron σ-bond between the equatorial aryl group and the central I(III) atom, while the shape, direction, and splitting of the σ-hole are controlled by the three-center-four-electron hypervalent bond between the axial ligands and the I(III) atom. In contrast, the crystal structures of iodoarene dicarboxylates exhibit two additional secondary interactions (halogen bonds) between the carbonyl oxygens and the central iodine atom,5659 which are attributed to the presence of a nonclassical σ-hole with two maxima on the I(III) atom surface (Figure 2b).54,55 More than one nucleophile can bind to either the same σ-hole or the less electronegative area between two σ-holes (Figure 2c).60,61 When the bite angle falls within the range of 34.3–46.1°, a bidentate nucleophile can bind to one of the two σ-holes on the iodonium(III) cation. If the bite angle is 54.9–77.6°, a bidentate nucleophile can bind to two σ-holes or even one σ-hole and the region between the two σ-holes. Benziodoxoles, including Togni reagents (X = CF3),62 feature a weak interaction between the moderately strong σ-hole on the I(III) atom and the solvent, such as acetonitrile (Figure 2d).54 Upon activation by a Brønsted acid, the characteristics of the hypervalent bond change to induce the formation of a highly electrophilic disubstituted iodonium cation with two localized σ-bonds that generate two strong σ-holes. One of these σ-holes forms an intramolecular halogen bond with the generated OH group, while the other can more tightly coordinate the incoming solvent with excellent proximity to the coupling partner X+.

The groups of Mayer and Legault developed a Lewis acidity scale for diaryliodonium salts based on the equilibrium constants for the complexation of a broad range of iodonium salts by Lewis bases, such as halides, carboxylates, phenolates, amines, and tris(p-anisyl)phosphine.63,64 Notably, cyclic diaryliodonium salts exhibited Lewis acidities approximately 2 orders of magnitude greater than those of their acyclic counterparts. The Lewis acidities of diaryliodonium salts were found to be comparable with that of Schreiner’s thiourea, while the theoretically predicted Lewis acidity of common cationic iodine(III) species (PhI+X, X = OH, Cl, F, OAc, OTs, OTf) was found to be similar or even stronger than those of widely used Lewis-acidic catalysts, such as BF3·OEt2 and TiCl4. The characteristics of λ3-iodanes explain the high electrophilicity/Lewis acidity of the central iodine, conservation of the configuration around the I(III) atom, and strong complexation in the crystal lattice and solution during the reaction, which considerably influences the susceptibility to reductive elimination.6569 Consequently, diaryliodonium salts have been employed as arylation reagents or halogen bond–donor and/or Lewis-acidic organocatalysts7073 to achieve diverse transformations.

This section focuses on the transition metal-free activation of aryl–iodide bonds through the oxidative activations of aryl iodides into λ3-iodanes. Among all possible λ3-iodanes, diaryliodonium salts can undergo transition metal-free dissociation of the Ar–I(III) bond and transfer aryl groups through the reductive elimination of hypernucleofuge aryl-λ3-iodane groups (hyperleaving groups).2123 Aryliodonium ylides can also transfer aryl groups, as demonstrated by several examples in this section.

2.1. Breakthroughs in the Utilization of Diaryliodonium Salts

Until about 35 years ago, hypervalent iodine reagents were reported to exhibit reactivity similar to that of many metal oxidants, and numerous organic synthetic reactions were reported using these reagents.16,7476 In the mid-1980s, Kita et al. used hypervalent iodine reagents, particularly PIDA and PIFA, as environmentally friendly oxidants to replace toxic heavy metal oxidants, such as mercuric diacetate, thallium triacetate, and lead tetraacetate. They also revealed the unique reactivity of hypervalent iodine and its potential as an alternative to transition metal catalysts. Using these reactions, they synthesized numerous natural products, leading to a paradigm shift in the use of hypervalent iodine.77,78 Additionally, Kita et al. found that PIFA in 1,1,1,3,3,3-hexafluoroisopropanol (HFIP)79 enables the oxidation of aromatic compounds via SET-induced aryl C–H functionalization with various nucleophiles (Scheme 1a).8085 The involvement of aromatic cation radicals was confirmed by ultraviolet (UV) and electron spin resonance (ESR) spectroscopies.81,82 The same group extended the metal-free oxidative coupling to intra- and intermolecular biaryl synthesis via C–H arylation using a combination of PIFA and a Lewis acid, such as boron trifluoride or trimethylsilyl bromide (Me3SiBr), in dichloromethane (Scheme 1b).8690 Intermolecular biaryl cross-coupling was also achieved by choosing substrates that avoided homobiaryl formation (Scheme 1c).9193

Scheme 1. Evolution of Transition Metal-Free Aryl C–H Functionalization Involving Hypervalent Iodine by Kita et al.

Scheme 1

The combination of PIFA with a Lewis acid induced homocoupling of thiophenes and pyrroles. The resulting products were a mixture of regioisomers containing head-to-tail (H-T) and head-to-head (H-H) biaryls (Scheme 1d).9497 After careful optimization of the reaction conditions, they disclosed that the use of HTIB and Me3SiBr in HFIP induced the regioselective homocoupling of 3-substituted thiophenes to generate H-T biaryls as a single isomer (Scheme 1e).98,99 Furthermore, the regioselective cross-coupling of unfunctionalized aromatic compounds was achieved to afford heterobiaryls (Scheme 1f).100 During the investigation of this reaction system, the authors noticed the generation of phenylthienyliodonium salts as the reaction intermediates, which react with the other arenes via SN2′-type substitution to furnish the coupling product. The isolated phenylthienyliodonium salts underwent biaryl cross-coupling in the presence of Me3SiBr to afford the same product. At nearly the same time, the authors discovered ipso-substitution of diaryliodonium bromides with electron-rich aromatic compounds in the presence of Me3SiOTf (Scheme 1g).101,102 In this reaction, the diaryliodonium salt serves as the SET oxidant to generate an aromatic cation radical, which was observed by ESR and UV spectroscopic measurements.

Beringer initially reported research on the synthesis of diaryliodonium salts and their use in arylation in the 1950s.103,104 However, it received little attention for a long time. Around 2010, when Kita’s discoveries were reported in the literature, diaryliodonium salts saw renewed interest, and numerous studies on arylation reactions have since been conducted. Transition metal-catalyzed arylations of aromatic compounds were reported by Sanford and Gaunt, wherein diaryliodonium salts serve as both aryl source and oxidant for the catalysts to generate high-valent aryl–metal species via oxidative addition.105113 By contrast, transition metal-free arylation using diaryliodonium salts involves various activated aryl species, such as aryl electrophiles, aryl radicals, and aryne equivalents, which can be captured by a range of nucleophiles to form C–C and C–heteroatom bonds, as demonstrated by several chemists (Scheme 2). Numerous reviews published in the past decade have covered topics related to transition metal-free and metal-catalyzed reactions, in which diaryliodonium salts act as efficient substrates,114117 including the double functionalization of C–I(III) and ortho C–H bonds.118,119

Scheme 2. Diaryliodonium Salts Act as Aryl Cation-like Species, Aryl Radicals, or Aryne Precursors under Transition Metal-Free Conditions.

Scheme 2

The current section summarizes the latest progress in transition metal-free arylations with diaryliodonium salts as aryl cation-like species (Section 2.3), aryl radicals (Section 2.4), and aryne equivalents (Section 2.5). Arylation using diaryliodonium salts generates a stoichiometric amount of iodoarenes, thereby reducing the atom economy. We also introduce two strategies to minimize waste: the incorporation of the iodoarene into final products via intramolecular or sequential reactions (Section 2.6) and the recyclable use of iodoarene auxiliaries (Section 2.7).

2.2. Oxidative Activation of Aryl Iodide for Diaryliodonium Salt Synthesis

Diaryliodonium salt synthesis typically begins with the oxidative activation of aryl iodides followed by the coupling of the hypervalent iodine(III) species with arene (Ar-H) or organometallic arene (Ar-M) reagents. Various approaches have been developed to synthesize diaryliodonium salts under mild and accessible conditions.115117,120,121 Hence, the following part describes the commonly used and updated strategies for producing diaryliodonium salts.

The generation of hypervalent iodine from aryl iodide has been summarized by Zhdankin and Stang in previously reported reviews.17,18 Oxidation of iodoarenes has been classically performed via oxidative chlorination to generate aryliodine dichloride (Scheme 3a-i),122 which is an unstable yellow solid unsuitable for long-term storage at low temperatures. Relatively stable aryliodine diacetates are generally synthesized via oxidative diacetoxylation using peracetic acid in acetic acid (Scheme 3a-ii).123 Alternatively, the combination of acetic acid with various oxidants, such as sodium periodate,124 sodium percarbonate,125 potassium peroxodisulfate,126 sodium perborate,127129 and m-chloroperbenzoic acid (mCPBA), is used.130,131 Hydrolysis of aryliodine dichloride and diacetate under basic conditions affords the corresponding iodosoarenes (Scheme 3a-iii).132,133 [Hydroxy(tosyloxy)iodo]arene, another common and essential hypervalent iodine(III) compound, has been classically synthesized via the generation of aryliodine diacetate followed by treatment with p-toluenesulfonic acid monohydrate (pTsOH·H2O).134 One-pot synthesis via oxidation of iodoarene by mCPBA in the presence of pTsOH·H2O provides a more convenient procedure (Scheme 3a-iv).135,136

Scheme 3. Oxidative Activation of Aryl Iodide and Preparation of Diaryliodonium Salts.

Scheme 3

Dehydrative condensation of aromatic compounds with these hypervalent iodine(III) compounds, such as iodosoarenes,101,137,138 (diacetoxyiodo)arenes,139141 and [hydroxy(tosyloxy)iodo]arenes,142146 affords diaryliodonium(III) salts, which is the simplest and cleanest synthesis (Scheme 3b-i). Given that these reactions involve electrophilic substitution, the starting materials are limited to electron-rich aromatic compounds. A metal–I(III) exchange strategy using organometallic aryl nucleophiles, such as aryllithiums, -silanes, -stannanes, -boronic acids, and -boronates, has been developed to prepare various diaryliodonium salts (Scheme 3b-ii).147155

One-pot syntheses of diaryliodonium salts including the oxidation of the corresponding iodoarenes followed by dehydrative condensation have also been developed. The combination of an organic oxidant, such as mCPBA, and a strong acid (TfOH, TsOH, or TFA) was used to achieve the one-pot syntheses of a wide range of commonly used diaryliodonium salts in high yields (Scheme 3c-i).141,156162 The choice of mCPBA as the organic oxidant enabled the purification of iodonium salts by simple trituration with diethyl ether. The K2S2O8/TFA combination was employed for the in situ generation of diaryliodonium trifluoroacetate followed by anion exchange with TfONa to afford the corresponding triflate.163165 In addition, urea–H2O2/Tf2O and Oxone/TfOH (oxidant/acid) combinations were successfully used for the one-pot preparation of diaryliodonium triflates.166,167 NaBO3·H2O/Ac2O/H2SO4 and Oxone/H2SO4 combinations were employed for the one-pot synthesis of diaryliodonium bromides after anion exchange of the generated diaryliodonium hydrogen sulfates in the presence of KBr.168,169 Dohi et al. developed a practical synthesis of diaryliodonium acetates bearing the trimethoxyphenyl group as an aryl group, which involves sequential oxidation of iodoarene with peracetic acid followed by condensation with trimethoxybenzene in 2,2,2-trifluoroethanol (TFE).170 These one-pot strategies for the preparation of diaryliodonium salts, including the oxidation of iodoarenes and condensation with aromatic compounds, have been applied to flow and electrochemical syntheses.171178 Olofsson et al. designed a one-pot reaction of iodoarenes with arylboronic acids in the presence of m-CPBA/BF3-OEt2 for the regiospecific synthesis of diaryliodonium salts (Scheme 3c-ii).179 They also updated the one- and two-pot protocols to be more tolerant of electron-rich arenes, electron-deficient iodoarenes/unactivated arenes, and electron-deficient iodoarenes/electron-deficient arylboronic acids in the syntheses of iodonium triflates, tosylates, and tetrafluoroborates, respectively.180 Additional anion exchange may be crucial for preparing difficult-to-access iodonium salts or for changing the counteranions of the iodonium salts to modify their physical and chemical properties.181 Cyclic diaryliodonium salts are typically prepared through oxidation of the ortho-iodoarene scaffold in the presence of an acid, generating a hypervalent iodine intermediate followed by an intramolecular electrophilic aromatic substitution reaction with the neighboring aryl moiety. This topic has been comprehensively discussed in several recent reviews.118,182185

2.3. Arylation of Carbon and Heteroatom Nucleophiles via Ligand Coupling

The investigation of transition metal-free arylation can be traced back to the seminal report of Beringer in 1953, who used diaryliodonium salts to arylate various nucleophiles, such as hydroxide, alkoxides, phenoxides, benzoate, nitrite, sulfonamides, amines, diethyl oxalacetate, sulfite, sulfinate, and cyanide.103,104 This method enabled the formation of a wide range of C–C and C–heteroatom bonds from a limited range of substrates, and the reactions were conducted in refluxing polar or mixed solvents, affording arylation products in poor to good yields. The findings of this work provided new opportunities for transition metal-free cross-arylations based on the development of sustainable and ecofriendly methods for constructing diverse chemical bonds under benign conditions.

The related recent studies have primarily aimed to develop chemoselective iodonium salts using a cost-effective and widely available dummy group to decrease costs as well as recycle aryl iodide waste, conducting the process under mild conditions (base, solvent, and temperature) with stoichiometric amounts of reactants, and identify conditions compatible with a broad range of nucleophiles and iodonium salts. Furthermore, the significance of this chemistry for delivering organic compounds suitable for further functionalization, addressing challenges associated with previously reported strategies, and transferring the process from the synthetic community to the manufacturing society for multigram-scale reactions, synthesis of pharmaceutical drugs/ingredients, and late-stage functionalization of valuable compounds is also targeted. The following sections commence with a succinct introduction followed by a discussion of work published largely after 2017.

2.3.1. Arylation via Ligand Exchange and Coupling Mechanism

The reactions between diaryliodonium salts and nucleophiles afford aryl–nucleophile bonds under transition metal-free conditions (Scheme 4).15,117,186 These reactions involve the exchange of the counteranion of the iodonium salt with the nucleophile, which affords two T-shaped Ar1(Ar2)I-Nu intermediates in equilibrium via pseudorotation. The nucleophile is transferred to the more approachable equatorial aryl group through a concerted ipso-substitution mechanism involving the formation of a three-center-four-electron transition state followed by the reductive elimination of the aryl iodide and coupling product formation.187189 The aryl group selectivity is thought to be determined during the ligand coupling step, specifically through the migration of the nucleophile from the I(III) atom to the adjacent ipso-carbon of the equatorial aryl group.

Scheme 4. Possible Mechanism of the Reaction of Diaryliodonium Salts with Nucleophiles.

Scheme 4

The reactions of diaryliodonium salts with nucleophiles are sustainable alternatives to metal catalyst-promoted C–C and C–heteroatom bond formation. Considerable progress has been made in transition metal-free arylations using diaryliodonium salts under benign conditions, although this approach suffers from the production of stoichiometric amounts of aryl iodides as waste, which can be costly if expensive starting materials are used. The chemoselectivity of transition metal-free cross-coupling reactions involving diaryliodonium salts with two different aryl ligands can be challenging to predict and is primarily influenced by the nature of the nucleophile and diaryliodonium salt.190 Generally, electron-deficient aryl ligands are preferred for electrophilic transfer because of their ability to better stabilize the negative charge formed in the transition state of reductive elimination.191 In addition, other factors, such as steric effects, including the ortho- and anti-ortho-effects, should be considered when discussing the chemoselectivity of this process.186,190,192

Several auxiliary or dummy aryl groups have been used to increase chemoselectivity (Scheme 5a). The selection of dummy ligands depends on the other aryl or heteroaryl group of the diaryliodonium salt. To date, auxiliary diaryliodonium salts with one dummy aryl ligand, such as thienyl, p-methoxyphenyl (PMP), mesityl (Mes), and trimethoxyphenyl (TMP) groups, have been employed.156,189,192194 Among then, TMP was found to be the most electron-rich and, hence, most suitable for enhancing chemoselectivity in the direct reactions of nucleophiles. This dummy ligand is derived from a commercially available and inexpensive compound, making diaryliodonium salt synthesis more facile and productive. Additionally, the dummy ligand allows for the chemoselective transfer of the aryl group during arylation to form the desired product in a controlled manner and produce an inexpensive dummy aryl iodide coproduct instead of a wasteful aryl iodide.

Scheme 5. Dummy Ligand Strategy Used for the Unified Selective Arylation of Nucleophiles.

Scheme 5

In this review, aryliodonium salts incorporating these dummy ligands are expressed using abbreviations to facilitate visual understanding (Scheme 5b). Specifically, aryl(4-phenyl)iodonium salt is represented as aryl(PMP)iodonium salt or Ar(PMP)I-L, aryl(2,4,6-trimethylphenyl)iodonium salt as aryl(Mes)iodonium salt or Ar(Mes)I-L, and aryl(2,4,6-trimethoxyphenyl)iodonium salt as aryl(TMP)iodonium salt or Ar(TMP)I-L, where L represents the counteranion.

2.3.2. C–O Bond Formation

Considerable progress has been achieved in the development of diaryliodonium salts as effective agents for the arylation of diverse OH groups under transition metal-free mild conditions. Olofsson et al. expanded the arylation scope to include phenols and iodonium salts under both aqueous and nonaqueous conditions (Schemes 6a-i and -ii).195197 Gaunt et al. emphasized the essential role of the fluoride counteranion in the iodonium salt in activating the phenolic OH group through hydrogen bonding, which allows O-arylation reactions to be carried out using a weak base, NaHCO3 (Scheme 6a-iii).198 Olofsson et al. aimed to provide optimal aqueous and nonaqueous conditions for the OH-arylation of various types of alcohols, including primary, secondary, tertiary, benzylic, and allylic ones, as well as carbohydrates (Schemes 6b-i and -ii).196,199,200 Additionally, sodium hexamethyldisilazide (NaHMDS) was found to be a more effective base than KOtBu for coupling highly sterically congested systems comprising both tertiary alcohols and iodonium salts with different aryl groups (Scheme 6b-iii).201 Stuart et al. used aryl(Mes)iodonium bromides or aryl(TMP)iodonium tosylates for the highly chemoselective arylation of diverse aliphatic and aromatic alcohols in the presence of NaH as a base (Scheme 6b-iv).202,203 The arylation of heteroaromatic carboxylic acids, such as indolecarboxylic acids, can be effectively achieved using simple iodonium salts at elevated temperatures to produce aryl esters (Scheme 6c-i).204 Additionally, KOtBu was found to be an effective base for mediating the O-arylation of aromatic and aliphatic carboxylic acids, as well as N-hydroxysuccinimide and phthalimide (Schemes 6c-ii and -iii).197,205,206 This approach was later extended to include N-hydroxybenzotriazoles and 3-hydroxybenzotriazin-4-ones, enabling the [3,3]-sigmatropic rearrangements of the resulting O-arylated products (Scheme 6d).207209 The transfer of the sulfonate counterion was successful in moving to the adjacent aryl moiety within the diaryliodonium sulfonate salt, resulting in the production of aryl sulfonate esters in high yields (Scheme 6e-i). In contrast, the two-component reaction of arylsulfonic acids with diaryliodonium triflates afforded coupling products in moderate yields (Scheme 6e-ii).197 Furthermore, diaryliodonium triflates were used to arylate various -P(O)-OH substrates, including phosphinic acid, hydrogen phosphate, and hydrogen phosphonate, which afforded the corresponding aryl esters in excellent yields (Scheme 6f).210

Scheme 6. General Conditions for O-Arylations with Diaryliodonium Salts.

Scheme 6

The Dohi and Kita group has developed a convenient base-free and step-economical strategy for the O-arylation of carboxylic acids involving the in situ generation of aryl(TMP)iodonium carboxylates (Ar(TMP)I-OCOR) via the one-pot reaction of iodosobenzene with 1,3,5-trimethoxybenzene (TMP-H) and a carboxylic acid (Scheme 7).138 The subsequent heating of the thus generated aryl(TMP)iodonium carboxylates led to ligand coupling and the formation of the desired arylcarboxylate ester. This process was applicable to a wide range of carboxylic acids, including electron-rich, electron-deficient, and sterically congested (hetero)aryl ones, as well as aliphatic carboxylic acids and natural products, such as cholic acid, even in the presence of multiple hydroxyl groups.

Scheme 7. O-Arylation of Carboxylic Acids via the In Situ Generation of Iodonium Carboxylates.

Scheme 7

The direct arylation of hydroxide ions with diaryliodonium salts is challenged by the frequent formation of regioisomeric aryl ethers instead of the intended phenols.187 The highly basic conditions of the reaction facilitate competing processes, such as SNAr reactions and the generation of aryne intermediates. This process is also accompanied by the formation of undesired aryl ethers due to the propensity of phenols to undergo further arylation under the employed conditions. To address this problem, Zhou et al. used oximes, which are nucleophiles commonly employed in one-pot, two-step cross-coupling reactions with diaryliodonium salts as a hydroxide surrogate (Scheme 8a).211216 Initially, diaryliodonium salts (Ar = Ar′) were used in C–O couplings with oximes to generate O-arylated oximes in situ, with subsequent treatment with Cs2CO3 upon heating producing the desired phenols. The Olofsson group used hydrogen peroxide and silanol as hydroxide surrogates in coupling reactions with diaryliodonium salts to generate phenols (Schemes 8b and 8c).217 The reaction conditions were optimized to achieve exclusive regioselectivity for the desired phenols. When hydrogen peroxide was used, the generated phenoxide intermediate and resulting phenol could not undergo further arylation under the applied conditions. Only iodonium salts with moderately electron-donating substituents and electron-deficient iodonium salts with only one phenyl group were tolerated, whereas aryl transfer groups or electron-rich dummy ligands were not suitable.

Scheme 8. Hydroxide Equivalents Used for Coupling with Iodonium Salts and Phenol Formation.

Scheme 8

Onomura and Kuriyama developed an efficient catalytic method for synthesizing phenols by coupling diaryliodonium salts with water in the presence of an organocatalyst, N-benzylpyridin-2-one, under mild conditions (Scheme 8d).218 This process commences with the O-arylation of the catalyst to form a 2-aryloxypyridinium intermediate, which is subsequently hydrolyzed by water to produce the desired phenol and regenerate the catalyst. The addition of fused silica is essential for rendering the reaction catalytic, and nBu4NBF4 is necessary for improving the product yield.

Aryl(TMP)iodonium tosylates (Ar(TMP)I-OTs) were utilized as effective reagents for the arylation of phenols, including a broad range of natural products (Scheme 9a).219 The TMP group in these salts facilitated their preparation and chemoselective aryl transfer during coupling, enabling the highly selective transfer of various electron-rich, electron-poor, sterically hindered, and heterocyclic aryl groups to produce the corresponding diaryl ethers. The one-pot sequential reaction, involving in situ aryl iodide oxidation, aryl(TMP)iodonium salt formation, and coupling with phenols, further expanded the versatility of this process.

Scheme 9. (a) Arylation of Phenols Using Ar(TMP)I-OTs, (b) O-Arylation of Phenols with Ar(TMP)I-OAc, (c) Ambigols Synthesized Using Diaryliodonium Salts.

Scheme 9

Dohi and Kita reported the O-arylation of phenols with aryl(TMP)iodonium acetates (Ar(TMP)I-OAc), which acted as highly reactive aryl precursors.220,221 The TMP group, with its two ortho methoxy groups coordinating the central I(III) atom, facilitated the dissociation of the acetate counterion, enhancing acetate basicity and phenol nucleophilicity via deprotonation (Scheme 9b). The reaction demonstrated scalability, and substrates with various functional groups, including aliphatic alcohols and boronic esters, were well tolerated under the employed conditions. Gulder et al. used the O-arylation of phenols using TMP iodonium salts to prepare ambigols, which are natural products derived from cyanobacteria and exhibit intriguing biological activities (Scheme 9c).222

Olofsson et al. discovered a method for efficiently O-functionalizing carbohydrate derivatives under simple conditions using protected furanose and pyranose structures and successfully arylated them by aryl groups derived from diaryliodonium salts (Scheme 10a).200,223 The versatility of this process was showcased through the 2- and 7-fold arylations of glucose diol and cyclodextrin derivatives, respectively.

Scheme 10. O-Arylation of Unactivated Carbohydrates with Diaryliodonium Salts.

Scheme 10

Gilmour et al. extensively researched the O-arylation of the anomeric hydroxyls of unactivated sugars (Scheme 10b).224 Specifically, diaryliodonium salts were used to arylate the lactol positions of mono-, di-, and trisaccharides, with preferences for α- and β-anomers observed for benzyl- and p-methoxybenzyl-protected substrates, respectively. The stereochemistry of the starting materials was transferred to the product with complete stereoretention.

Motivated by the high reactivity of diaryliodonium salts as transition metal-free arylating reagent, Xu et al. designed an effective process for arylating DNA-conjugated libraries of (hetero)aryl phenols and naphthols (Scheme 11).225 In this DNA-encoded library synthesis, the O-arylation of phenolic groups was accomplished in high yields and with high DNA fidelity in the presence of additional amidic groups. The scope of this method was expanded to include the late-stage O-arylation of tyrosine on DNA-conjugated peptides and synthesis of DNA-conjugated analogues of sorafenib.

Scheme 11. On-DNA O-Arylation of Phenols with Diaryliodonium Salts.

Scheme 11

Gao and Tan designed an O-arylation of N-arylhydroxylamines using diaryliodonium salts for the synthesis of 2-amino-2′-hydroxy-1,1′-binaphthyl (NOBIN)-type biaryls (Scheme 12).226,227 The O-arylation of N-arylhydroxylamines produced N,O-diarylhydroxylamines, which subsequently underwent a [3,3]-sigmatropic rearrangement and rearomatization to produce the corresponding biaryls.228

Scheme 12. O-Arylation of N-Arylhydroxylamine and NOBIN-Type Biaryl Formation.

Scheme 12

2.3.3. C–N Bond Formation

Since the discovery of diaryliodonium salts by Beringer, their versatility as transition metal-free N-arylating agents has been significantly expanded to include a broad range of substrates and more convenient conditions.104 The arylation of aqueous ammonia under basic conditions was reported to produce only the related aniline and not di- or triarylation products (Scheme 13a). The nitrite anion was successfully N-arylated with a variety of diaryliodonium salts under base-free conditions (Scheme 13b-i).229,230 Additionally, useful aryl azides were prepared by the arylation of the azide anion with iodonium salts, such as electronically diverse aryl(TMP)iodonium tosylates, which acted as highly chemoselective arylating agents (Schemes 13b-ii–iv).230233 The arylation of anilines with diaryliodonium salts can be achieved at high temperatures in the absence of bases (Scheme 13c).234,235 In contrast, the arylation of aliphatic cyclic secondary amines with aryl(TMP)iodonium trifluoroacetate can be accomplished under basic conditions, affording N-arylation products (Scheme 13d).236,237 Under basic conditions, aryl/alkyl cyanamides can be effectively arylated with diaryliodonium salts in aqueous or nonaqueous environments to generate diverse N-arylated cyanamides (Scheme 13e).238,239 Muñiz et al. used aryl(phenyl)iodonium salts and Mes2IOTf to arylate cyclic imides and amides and synthesize sterically hindered aniline derivatives (Scheme 13f).240 In this process, the bulky aryl group was transferred with excellent chemoselectivity. Additionally, Stuart et al. arylated phthalimide potassium salts with aryl(TMP)iodonium tosylates to yield N-arylation products with good chemoselectivity (Scheme 13F).241 Acyclic secondary amides were arylated by iodonium salts in the presence of bases to produce N-arylated amides (Scheme 13g).242,243 Specific N-heteroarenes, such as pyrazoles and 1,2,3-triazoles, were successfully reacted with diaryliodonium triflates under milder conditions to form the corresponding N-aryl heterocycles (Scheme 13h).244,245 According to density functional theory (DFT) calculations, the presence of a basic nitrogen atom adjacent to the arylated NH center is crucial for enhancing the reactivity required to carry out the reaction under mild basic conditions.246

Scheme 13. General Conditions for N-Arylations with Diaryliodonium Salts.

Scheme 13

Karchava et al. reported the successful arylation of a tertiary amine (DABCO) with electronically and sterically diverse aryl(Mes)iodonium triflates (Ar(Mes)I-OTf) (Scheme 14a).247 The obtained N-aryl-DABCO triflates afforded 1,4-disubstituted piperazines and flibaserin drugs when reacted with nucleophiles to induce ring opening. Olofsson et al. reported a method for the N-arylation of primary and secondary amines compatible with a wide range of diaryliodonium salts and amine nucleophiles (Scheme 14b).248,249

Scheme 14. N-Arylation of DABCO and Formation of Quaternary Salts.

Scheme 14

Motivated by the reactivity of diaryliodonium salts as transition metal-free arylating reagents, Prakash et al. developed a method for the regioselective arylation of 1,2,3-triazoles at the challenging N(2) position (Scheme 15a).250 Under optimal conditions (aryl(TMP)iodonium salt, Na2CO3, toluene, and heating at 100 °C), diverse substituted and unsubstituted-1,2,3-triazoles were regioselectively arylated to produce N2-arylated products. The results of DFT calculations were in good agreement with the experimentally observed N2-regioselectivity. Similarly, N-tetrazoles were selectively arylated at the N2-position using diaryliodonium salts (Scheme 15b).251 The optimal conditions were compatible with a wide range of 5-substituted tetrazoles. Han and Wang reported a chemoselective N-arylation of pyridazin-3-ones under base-free conditions (Scheme 15c).252 The authors examined the reactivity of diaryliodonium hexafluorophosphate by demonstrating the remarkable selectivity of the N-arylation of a wide range of pyridazinones. Contrary to previous ionic mechanisms, these mechanistic investigations uncovered a free radical reaction pathway.

Scheme 15. Regioselective N2-Arylation of Triazoles/Tetrazoles with Diaryliodonium Salts.

Scheme 15

The potential of aryl(TMP)iodonium acetates (Ar(TMP)I-OAc) for the N-arylation of N-methoxysulfonamides and N,O-protected hydroxylamines was investigated by Dohi and Kita (Scheme 16).253 The use of iodonium salts with TMP as the dummy ligand and acetate as the counteranion was critical for achieving high yields under mild conditions, while reactions with other counteranions, such as OTs, OTf, and OCOCF3, were unsuccessful or led to poor yields.

Scheme 16. N-Arylation of Sulfonamides and Protected Hydroxylamines with Ar(TMP)I-OAc.

Scheme 16

Postnikov et al. examined the influence of N-heterocyclic substituents on the reactivity and chemoselectivity of ortho-azole-tethered diaryliodonium salts (Scheme 17).254 The gram-scale arylation of NaNO2 with a range of electronically diverse iodonium salts afforded ortho-nitroarene derivatives and could be carried out as a one-pot sequential reaction starting from the related iodoarenes. The azole iodonium salts demonstrated high reactivity and chemoselectivity when reacted with a variety of halogen, nitrogen, oxygen, and sulfur nucleophiles, which was attributed to the stabilization of these salts via coordination with the adjacent azole substituent. However, the reactivity and chemoselectivity of these salts were reduced when the ortho-azole group was subjected to N-methylation or N-protonation.

Scheme 17. Nitration of Diaryliodonium Salts Bearing N-Heterocyclic Aryl Groups.

Scheme 17

Olofsson et al. used diaryliodonium salts to achieve the N-arylation of secondary acyclic amides under mild conditions (Scheme 18a),242 suggesting that this reaction proceeds via ligand exchange and the formation of T-shaped intermediates, which undergo ligand coupling through [1,2]- and [2,3]-rearrangements to yield the desired N-arylated tertiary amide. Wang et al. prepared N-arylated secondary acyclic amides through the in situ generation of aryne intermediates from iodonium salts followed by nucleophilic attack by the deprotonated amide and protonation, which afforded the desired N-arylated products (Scheme 18b).243 This mechanism was supported by the formation of two regioisomeric N-arylated amides.

Scheme 18. N-Arylation of Acyclic Amides with Diaryliodonium Salts.

Scheme 18

2.3.4. N- vs O-Arylation of Ambident Nucleophiles

The arylation of aldoximes and ketoximes with diphenyliodonium triflates under basic conditions (Cs2CO3, MeCN, rt, 6 h) resulted in the exclusive formation of O-arylated oximes in high yields.213 This process was the subject of extensive studies conducted by Mo and colleagues to investigate the influence of the oxime structure and applied conditions for N- vs O-arylation (Scheme 19).255,256 The reaction of diaryliodonium salt with ketoxime in the presence of KOtBu (conditions a) generated an N-arylation product (nitrone) and O-arylation product. The reaction of dibenzylideneacetone oximes with KOH as a base (conditions b) led to the formation of nitrone as the sole product. Similarly, oxindole oxime yielded the N-aryl oxindole nitrone as the only product under the optimum conditions.

Scheme 19. N- vs O-Arylations of Oximes with Diaryliodonium Salts.

Scheme 19

The impact of substituents on the 2-pyridone skeleton and the preference for N- or O-arylation with diaryliodonium salts was thoroughly examined (Scheme 20a).257 Under optimal conditions, unsubstituted and C3 to C5 electron-donor/acceptor substituted 2-pyridones displayed a preference for N-arylation over O-arylation. In contrast, C-6 substituted 2-pyridones exclusively produced O-arylated products. It is believed that the O-selectivity was due to the steric interaction that occurred during and after the N-attack of the 6-substituted-2-pyridone on the I(III) center of the iodonium salt, which was not the case with O-attack. Kumar investigated the N- vsO-arylation of quinolones and related substrates using conventional and microwave heating in aqueous NaOH solvent to produce the same selectivity but with improved productivity and shorter reaction times under microwave heating (Scheme 20b).258,259 The arylation of 4-quinolone and 4-methylquinolin-2(1H)-one with microwave heating resulted in the exclusive formation of N-arylated products. This N-arylation protocol is beneficial for large-scale synthesis and can be applied to structurally relevant substrates like acridin-9-one, quinoxaline-2-one, and benzoimidazol-2-one. Notably, 2-substituted quinoline-4-ones reacted under the same microwave conditions with diaryliodonium triflates produced only O-arylation products due to the sterically hindered nitrogen center.

Scheme 20. a) Substitutions-Controlled Chemoselective N- vs O-Arylation of Pyridin-2-ones. b) N- vs O-Arylation of Quinolones under MW Conditions.

Scheme 20

The Onomura and Kuriyama group emphasized the importance of selecting the optimal base for achieving chemoselective N- and O-arylation of pyridin-2-one using diaryliodonium triflate (Scheme 21).260,261 They discovered that N,N-diethylaniline and quinoline were the most suitable bases for N- and O-arylation of pyridin-2-one, respectively, affording exceptional chemoselectivity and high productivity. Quinoxalin-2-one also exhibited chemospecific O-arylation under various conditions, particularly in the presence of Cs2CO3 as a base. The diverse substitutions on both pyridin-2-one/quinoxalin-2-one and diaryliodonium salts were found to be well-tolerated. Although the X-ray crystallography structure of the diphenyliodonium salt with the amidate counteranion of 5-trifluoromethylpyridin-2-one was confirmed, the reaction mechanism of this alternative selectivity is still unclear.

Scheme 21. Chemoselective N-/O-Arylation of Pyridine-2-one and Quinoxaline-2-one.

Scheme 21

Recently, the investigation of the O- versus N-arylation of N-alkoxyamides with diaryliodonium salts has been conducted (Scheme 22).262 Among the various N-substituted benzamides and iodonium salts, N-tert-butoxyamide and aryl(TMP)iodonium acetate (Ar(TMP)I-OAc) demonstrated the highest reactivity and O-arylation selectivity. The reaction conditions were suitable for gram-scale synthesis and could accommodate iodonium acetates and N-tert-butoxyamides with electron-donating/withdrawing and sterically congested groups. By adjusting the substituents on the starting materials, high-to-excellent O-selectivity could be achieved. Remarkably, N-methoxy-4-nitrobenzamide yielded only O-arylation products when reacted with sterically congested iodonium acetates.

Scheme 22. Chemoselective O-Arylation of Amides with Iodonium Acetates.

Scheme 22

2.3.5. C–C Bond Formation via C–H Arylation

Diaryliodonium salts have shown remarkable versatility as transition metal-free arylating reagents, capable of reacting with a range of functional groups, including Csp3 and Csp2 nucleophiles. Csp2–Csp3 bond formation has been achieved using silyl enol ethers of aliphatic ketones, which were coupled with diphenyliodonium fluoride to produce mono- or diphenyl ketones (Scheme 23a).263,264 Active methylene compounds, such as 2-substituted cyanoacetates,265 2-substituted malonates,190,266 2-substituted malononitriles,267 ethyl acetoacetates,268 and nitroalkanes,269 have been found to be highly effective in arylation reactions with diaryliodonium salts in the presence of KOtBu or NaH as bases (Scheme 23b–23e). Shibata and colleagues utilized cyclic β-keto esters/amides as benchmark substrates for pentafluorophenylation, triflyl-(hetero)arylation, and pentafluorosulfanylarylation by related iodonium salts with high chemoselectivity.270272

Scheme 23. General Conditions for the Csp3-Arylation of Enolizable Substrates.

Scheme 23

The study of asymmetric versions of arylation reactions was initially conducted by Ochiai and colleagues, who utilized a chiral iodonium salt ((S)-(−)-1,1-binaththyl-2-yl(phenyl)iodonium tetrafluoroborates) for arylation of 2-(alkoxycarbonyl)-1-indanone with moderate enantioselectivity (up to 53% ee) (Scheme 23f).152 Aggarwal and Olofsson employed Simpkins’ (R,R)-base for the desymmetrization of cyclohexanones, followed by coupling with dipyridyliodonium chloride to produce pyridyl derivatives with high enantioselectivity (up to 94% ee), which were subsequently used in the total synthesis of (−)-epibatidine alkaloid (Scheme 23g).273 The scope of Csp3-arylation with iodonium salts was extended to heterocycles; N-acetyl-3-acyloxy-2-oxindoles underwent C3-arylation with aryl(Mes)iodonium hexafluorophosphates,274 whereas N-acetyl-3-indolinones afforded a C2-arylated product in moderate yield (Scheme 23h and i).275 However, aryl(Mes)iodonium triflates showed that the Mes-group transferred more preferentially than the aryl group, resulting in a mixture of C2-aryl and C2-mesity products in moderate yields. C5-Arylation of 1-methyl-5-alkylbarbiturtaes was also achieved (Scheme 23j) and then applied to the synthesis of the general anesthetic mephobarbital drug.276 Other heterocycles, such as azlactones and 4-substituted pyrazolin-5-ones, were potently arylated by iodonium salts,277,278 wherein aryl(Mes)iodonium salts exhibited poor and excellent aryl-transfer selectivity with azlactones and pyrazolin-5-ones, respectively.

The Olofsson group conducted extensive density functional theory (DFT) studies to propose a plausible mechanism for Csp3-arylation using diaryliodonium salts (Scheme 23k).190,269,279 This process involves deprotonation of the activated methylene group of the substrate under basic conditions to generate an enolate intermediate, which undergoes ligand exchange with the anion of the iodonium salt to form C- and O-enolate iodonium intermediates, respectively. The reductive eliminations via [1,2]- and [2,3]-rearrangements through three- and five-membered transition states, respectively, result in the formation of the desired α-arylated product and the extrusion of aryl iodide. DFT calculations demonstrated that both pathways were viable for the formation of the product, and the [2,3]-rearrangement process was found to be more energetically favorable than the [1,2]-rearrangement.

The research conducted by Zhang and colleagues extended the range of C-arylation by employing 2-nitoketones for the synthesis of tertiary 2-aryl-2-nitoketones (Scheme 24a).280 This reaction was compatible with a range of 5-, 6-, 7-, and 12-membered cyclic 2-nitroketones, benzocyclic nitroketones, and acyclic aryl/alkyl nitroketones. Wang and Li proposed a mild approach for the C-3 arylation of 3-acetyloxy-2-oxindoles using diaryliodonium salts, resulting in the formation of 3-aryl-3-acetyloxy-2-oxindoles (Scheme 24b).274

Scheme 24. a) α-Arylation of α-Nitroketones with Iodonium Salts. b) C-3 Arylation of Oxindoles with Diaryliodonium Salts.

Scheme 24

The application of activated C–H substrates as C-nucleophiles in coupling reactions with iodonium salts under ambient conditions was investigated with the aim of forming Csp3–Csp2 bonds.281283 The compounds, including tertiary amides of α-fluoro-α-nitroacetamides and α-fluoro-α-cyanoacetamides, were found to be effective in forming α-arylated products through α-arylation with electronically and sterically diverse aryl/heteroaryl groups derived from diaryliodonium tosylates with a mesityl dummy ligand (Scheme 25a).281 Although secondary amides of these fluorinated acetamides showed moderate reactivity, the reaction scope of α-substituted-α-fluoroacetamides was expanded to include secondary amides of α-fluoroacetoacetamides, which effectively coupled with electron-rich aryl groups to produce fully substituted benzylic products (Scheme 25b).283 However, a reaction with electron-deficient iodonium aryl groups resulted in the formation of α-arylated products containing an electrophilic α-acetyl group, which is prone to spontaneous deacetylation under basic conditions, leading to the formation of the related α-aryl-α-fluoroacetamide. Regrettably, the tertiary amide of α-fluoroacetoacetamide was not reactive under these conditions. Prakash et al. utilized fluorobis(phenylsulfonyl)methane, which comprised Csp3-nucleophiles with fluoro groups, for coupling with diaryliodonium salts (Scheme 25c).284 This led to the synthesis of fluorobis(phenylsulfonyl)methylarenes, which served as useful scaffolds for further reduction steps and the synthesis of biologically attractive Ar-CH2F and Ar-CD2F derivatives.

Scheme 25. C-Arylation of Fluoroacetamides with Diaryliodonium Salts.

Scheme 25

The difluoro enol silyl ethers, which were prepared by reacting trifluoromethylketones with TMSCl/Mg, have proven to be useful nucleophilic synthons for merging α,α-difluorocarbonyl moieties with a variety of electrophiles (Scheme 26a).285 Diaryliodonium salts effectively served as precursors for arylating a small library of difluoro enol ethers, resulting in the formation of corresponding α-arylated products.

Scheme 26. α-Arylation of α,α-Difluoromethyl Ketone.

Scheme 26

The Dohi and Kita group first developed a transition metal-free method for the decarboxylative coupling of aryl(TMP)iodonium tosylate (Ar(TMP)I-OTs) with α,α-difluoro-β-ketoacid sodium salt, which afforded useful fluorine-containing scaffolds without requiring hazardous reagents (Scheme 26b).286 Mechanistic studies concluded that decarboxylation would occur to generate iodonium difluoroenolate salt. Further ligand coupling produced the aryldifluoromethyl ketone. Interestingly, the reaction of uracil iodonium salt resulted in C-6 difluoromethylation by cine substitution rather than the anticipated C-5 product.

Yorimitsu and his colleagues have disclosed a method for generating 2-hydroxy-2′-iodobiaryls through dehydrogenative coupling of phenols with electron-rich diacetoxyiodoarenes.287 The Kalek group reported a complementary process for coupling 2-naphthols with diaryliodonium salts, resulting in the formation of related biaryls (Scheme 27).288 By using cyclohexane and an inorganic base, the deprotonation of 2-naphthol was avoided, leading to the selective formation of C1-arylated 2-naphthols rather than O-arylation, and O-/C1-double arylation products as previously obtained by Quideau.289 Efficient coupling with substituted 2-naphthols required diaryliodonium salts with electron-deficient aryl groups. A plausible mechanism was proposed based on experimental and DFT studies, suggesting that the association of 2-naphthol with diaryliodonium salts formed diaryliodonium naphthoxide. C–C bond formation through rearrangement produced the dearomatized ketone, which spontaneously rearomatized via tautomerization to give C1-arylated naphthol. The other possible O-arylation product was found to be energetically unfavorable in the presence of protonated 2-naphthol.

Scheme 27. C2-Arylation of 2-Naphthols with Diaryliodonium Salts.

Scheme 27

Treating pyridine with Tf2O and a secondary amine afforded azahexatrienes (Zincke imines), which were used by Greaney’s group as C-nucleophiles for the regiodivergent arylation of pyridines under transition metal-free and -catalyzed conditions.290 When symmetrical diaryliodonium salts were used as arylating reagents for coupling with a Zincke imine under transition metal-free conditions followed by cyclization, the related meta-arylated pyridines were obtained exclusively (Scheme 28).

Scheme 28. Regioselective Arylation of Pyridine with Diaryliodonium Salts.

Scheme 28

2.3.6. C–S Bond Formation

Diaryliodonium salts have proven to be valuable reagents for the arylation of various sulfur-containing substrates, affording C–S bonds. Sandin and Beringer conducted the S-arylation of thioglycolic acid, thiophenol, cysteine, and sulfinate salt under aqueous conditions (Scheme 29a-i and g-i).104,291 The combination of Cs2CO3 and toluene promoted the completion of the reaction within 10 min (Scheme 29a-ii).292 Alkylthiol was converted in situ to the sodium salt, then reacted with diphenyliodonium triflates to give hexyl phenyl sulfide (Scheme 29a-iii).293 Polymer-bound diaryliodonium salts were used to S-arylate sodium benzenethiolate, and the resulting polymer-bound aryl iodide could be recycled (Scheme 29a-iv).294 Sanford reported a different approach for synthesizing aryl sulfides by reacting alkyl/aryl thiols and thioethers with diaryliodonium trifluoroacetates under acidic conditions (Scheme 29a-v).295 KSCN combined with diaryliodonium triflates led to the formation of aryl thiocyanates (Scheme 29b).296,297 Potassium and sodium derivatives of thiocarboxylic, thiosulfonic, and dithiocarbamic acid underwent S-arylation with diaryliodonium salts in moderate to satisfactory yields (Scheme 29c–e).298300 Ciufolini and colleagues developed a copper-free method for synthesizing triarylsulfonium triflates by reacting diaryl sulfides with iodonium triflates (Scheme 29f).301 The reaction was effective, and the iodonium salts with two different aryl groups generated a mixture of arylation products with a preference for transferring electron-rich aryl groups to electron-poor and thienyl groups. The Dohi group investigated the use of diaryliodonium triflates containing a mesityl group as a dummy ligand for fully chemoselective S-arylation of diaryl sulfides and the formation of triarylsulfonium triflates under both copper-free and copper-catalyzed conditions.302 A study conducted by Kumar and Manolikakes successfully synthesized diaryl and aryl-heteroaryl sulfones by reacting (hetero)arylsulfinic acid sodium salts with diaryliodonium salts under thermal and microwave conditions (Scheme 29g-ii and -iii).303306 Xu and co-workers employed readily available sodium glycosyl sulfinates for the glycosyl sulfonation of diaryliodonium salts and the synthesis of glycosyl aryl sulfones in the presence of DMSO or H2O as solvent.307

Scheme 29. General Conditions for S-Arylations with Diaryliodonium Salts.

Scheme 29

The Kalek group utilized diaryliodonium salts for the S-arylation of phosphorothioate diesters, preserving the stereochemistry of the phosphorus atom (Scheme 30).308 A variety of O,O-diaryl and O,O-dialkyl phosphorothioate diesters reacted efficiently under the specified reaction conditions. Phosphorus substrates such as phosphorodithioates and phosphonothioates were also found to be effective to yield S-arylation products.

Scheme 30. S-Arylation of Phosphorothioate Diesters with Iodonium Salts.

Scheme 30

A process for the S-arylation of thioamides under basic conditions was established by Olofsson and colleagues (Scheme 31).309 Aromatic thioamides and pyridine-2-thiol displayed excellent S-site selectivity during the reaction. While aliphatic thioamides and pyrrolidine-2-thione produced mixtures of S- and N-arylations, 3,4-dihydroisoquinoline-1(2H)-thione only yielded an N-arylation product. This selectivity was attributed to the conjugation within the thioamide group, where efficient conjugation in aromatic thioamides led to increased S-nucleophilicity. In contrast, less efficient conjugation, along with the constraints in cyclic thioamides, resulted in more N-nucleophilicity. Mechanistic studies ruled out the radical mechanism and the generation of aryne intermediates.

Scheme 31. Arylation of Thioamides with Diaryliodonium Salts.

Scheme 31

Apart from the methods previously reported for arylation of thiols using diaryliodonium salts,292,295,310 Kalek et al. have developed the arylation of heterocyclic thiols under mild conditions (Scheme 32a).311 This S-arylation with a specific base demonstrated success with a wide range of 5-/6-membered heterocyclic thiols, aliphatic thiols, and diverse mercaptobenzazoles. A more effective protocol for the S-arylation of heterocyclic thiols using diaryliodonium salts was reported by the Thakur group (Scheme 32b).312 Notably, no interference from N-arylation was observed in the screened thiol substrates, even with acidic N–H groups, and the related S-arylation products were exclusively formed.

Scheme 32. S-Arylation of Thiols Using Diaryliodonium Salts.

Scheme 32

Wang, Han, and collaborators demonstrated that the reaction of ortho-OTf-substituted diaryliodonium salts with thiophenols afforded the corresponding diaryl sulfides (Scheme 32c), whereas this type of iodonium salts underwent intramolecular aryl migration in the absence of thiols.313 The obtained ortho-OTf-substituted diaryl sulfides were employed for further coupling reactions, wherein the OTf group served as the reaction point.

The early work of Sandin et al.291 on the S-arylation of Cys-amino acids with diphenyliodonium chloride prompted the Payne group to explore the chemoselective arylation of peptides and proteins at the Cys-position using more functionalized diaryliodonium salts (Scheme 32d).314 To accomplish the late-stage functionalization of MUC1 VNTR peptide, affibody zEGFR, and histone H2A proteins, alkyne-, keto-, and mPEG-derived diaryliodonium salts were synthesized.

In addition to aryl sulfinate sodium salt,304N,N′-disulfonylhydrazine has been used as an effective aryl sulfonyl anion precursor due to its stability and solubility in common solvents.315 In sequential one-pot reactions, disulfonylhydrazine with triethylamine generated ammonium sulfinate, which then reacted with added diaryliodonium salt to produce the related diaryl sulfone (Scheme 33a). This reaction was compatible with various N,N′-disulfonylhydrazines. Alternatively, treating sulfonyl hydrazides with base produced sulfonyl anions that reacted with diaryliodonium salts to similarly yield the corresponding diaryl sulfones (Scheme 33b).316

Scheme 33. Aryl Sulfonyl Anion Equivalents for Coupling with Diaryliodonium Salts.

Scheme 33

Ye, Wu, Chen, and co-workers developed a direct approach for the synthesis of γ-ketosulfones via a 3-component reaction of cyclopropanol, SO2-surrogate (DABSO), and diaryliodonium salt in H2O as the only solvent under additive-, catalyst-, and oxidant-free conditions (Scheme 34).317 Mechanistic investigations indicated the in situ generation of γ-ketosulfinate intermediate, which underwent ligand coupling with diaryliodonium salt to give substituted γ-keto sulfones. Preliminary antitumor activity of these aryl-substituted γ-ketosulfones showed potent inhibition of SBC-2 and 16HBE cell lines.

Scheme 34. 3-Component Reaction of Diaryliodonium Salt with DABSO and Cyclopropanol in Water.

Scheme 34

The Bolm group has introduced a new method for synthesizing sulfoxides, which involves the in situ generation of sulfinate anions from β-sulfinyl esters under basic conditions and the subsequent reaction with diaryliodonium triflates (Scheme 35a).318 A broad range of S-(hetero)aryl-β-sulfinyl esters were successfully reacted to produce diaryl sulfoxides. In a related study, Zhang and colleagues reported a method for synthesizing S,S-disubstituted sulfoxides using diaryliodonium tetrafluoroborate, which produced the desired product in high yield (Scheme 35b).319,320 The proposed mechanism for the reaction commenced with a retro-Michael reaction of the β-sulfinyl ester to generate a sulfinate anion, which then underwent ligand exchange followed by ligand coupling to form the sulfoxide product (Scheme 35c).

Scheme 35. Synthesis of Sulfoxides through Generation of Sulfinate Anion and Coupling with Iodonium Salts.

Scheme 35

The reaction of diaryliodonium salts with dithiocarbamic acid sodium salt provided the corresponding S-aryl dithiocarbamates.300 A multicomponent protocol, developed by Murarka et al., was a convenient process for the synthesis of S-aryl dithiocarbamate (Scheme 36a).321 This new method involved a cascade reaction of iodonium salts with cyclic/acyclic aliphatic amines and carbon disulfide under additive-free conditions, leading to the production of a diverse range of S-aryl dithiocarbamates. The synthesis of aryl thiols using diaryliodonium salts is a challenging process, which Karchava and colleagues addressed by developing a new approach for synthesizing aryl thiol surrogates through the reaction of iodonium salts with potassium O-alkylxanthates (Scheme 36b).322 This approach utilized mild conditions that prevented further functionalization and over-reaction of the resulting S-aryl O-alkylxanthate products. Based on these results, the same group developed a new thiol-free process for the preparation of alkyl aryl thioethers from the corresponding S-aryl O-alkylxanthates.323

Scheme 36. (a) One-Pot 3-Component Synthesis of S-Aryl Dithiocarbamates. (b) Coupling of Iodonium Salt with Xanthates.

Scheme 36

Lu, Yang, and Wu reported independently base-promoted S-arylations of sulfenamides with diaryliodonium salts and the synthesis of the related sulfilimines with excellent chemoselectivity (Scheme 37).324326 The reaction conditions tolerated sulfenamides with N-aryl/alkyl acyl groups and S-alkyl/aryl substituents.

Scheme 37. S-Arylations of Sulfenamides with Diaryliodonium Salts.

Scheme 37

2.3.7. Miscellaneous Bond Formations

The versatility of diaryliodonium salts extends beyond their ability to form C–C, C–N, and C–S bonds as arylating reagents. They also facilitate the formation of other chemical bonds, such as C–P bonds. The reaction of diaryliodonium salts with phosphite anions resulted in the synthesis of arylphosphonates in high yields (Scheme 38a).327 Electron-rich and electron-deficient diaryliodonium salts were used for covalent functionalization of few-layer black phosphorus, offering excellent ambient stability and the potential to tune electronic properties.328

Scheme 38. Ar–P and Ar–B Bonds Formations with Diaryliodonium Salts.

Scheme 38

The formation of C–B bonds was the focus of research conducted by Muñiz and colleagues, who developed a useful methodology for the reaction of diaryliodonium acetates with diboron reagents in methanol as a solvent under base-free conditions (Scheme 38b).329 The acetate counteranion and methanol solvent played a crucial role, interacting as Lewis bases with the diboron reagent and umpolunging the electrophilic character. This resulted in the generation of a nucleophilic boron center capable of reacting with diaryliodonium salts and forming aryl boronic esters. Arylation of various diboron substrates was also demonstrated.

Chalcogens with properties similar to sulfur, such as selenium and tellurium, have been observed to perform the same type of reactions with similar reactivity.301,330336 Townsend and colleagues demonstrated the reactivity of potassium selenocyanate (KSeCN) with diaryliodonium salts to produce corresponding aryl selenocyanates (Scheme 39a).337 Upon treatment of the resulting products with sodium borohydride, selenium anions were generated, which are reactive toward various iodonium salts and glycosyl halide electrophiles. The one-pot and sequential reaction of KSeCN with two different diaryliodonium salts resulted in the formation of unsymmetrical diaryselenides in high overall yields (Scheme 39b). The Zhang group utilized (Me4N)SeCF3, a readily available nucleophile, to generate aryl trifluoromethylselenoethers (Scheme 39c).338 Soldatova, Postnikov, and collaborators designed a one-pot strategy for double arylation of KSeCN with the two aryl ligands of the aryl(TMP)iodonium triflate (Ar(TMP)I-OTf) (Scheme 39d).339 The arylation of KSeCN nucleophile generates ArSeCN and TMP-I as an electrophile and electron-rich arene intermediate, which reacted together in the presence of HFIP as a crucial solvent340 to yield the desired diarylselenide.

Scheme 39. Arylation of KSeCN and (Me4N)SeCF3 by Diaryliodonium Salts.

Scheme 39

The formation of C–halogen bonds can be achieved through the thermolysis of an iodonium salt with a halide counteranion, either preformed or prepared in situ. This process involves a pseudo SN2-type mechanism, where the halide anion is transferred to the equatorial aryl group of the iodonium salt, resulting in the production of the corresponding aryl halide.341345 This same process can be used to synthesize radiolabeled haloarenes,346349 including radiofluorination of iodonium salts. This process has been extensively studied and practical application in positron emission tomography (PET) was found after the pioneering work of Pike and colleagues.350361 The use of diaryliodonium salts and aryliodonium ylides as arylating reagents provides a useful expansion for the radio-functionalization of a broad library of electron-rich and electron-deficient (hetero)aryl moieties, which are not accessible through traditional methods. Indeed, the nucleophilic radiofluorination process is a rapid and effective method for synthesizing various 18F-labeled compounds and radiotracer molecules for biomolecular imaging with PET. As this topic has been summarized in many reviews before,356361 we will focus on recent examples of radio-functionalization of iodonium salts/ylides.

The Matsunaga group achieved the successful synthesis of astatine-211 radiolabeled multifunctionalized molecules by employing a formal SNAr reaction between an aryliodonium ylide and astatide nucleophile (211At) under reducing conditions (Scheme 40a).362 This radiolabeling process was applicable to the functionalization of various substrates, including estrone natural product, fibrate drug, phenylalanine amino acid, and other (hetero)arenes with 211At. The nucleophilic radioastatination process was expanded by Guérard et al., who screened various aryliodonium ylides under different conditions (Scheme 40b).363 Aryliodonium ylides with electron-withdrawing substituents performed the reaction smoothly at room temperature (conditions a), while electron-rich aryliodonium ylides required modified conditions (conditions b). Under conditions b, the presence of the radical scavenger TEMPO was crucial to avoid degradation of the iodonium ylide and achieve a high radiochemical yield and molar reactivity. Aryliodonium ylides serve as excellent reagents for arylating 125I and 211At, which resolved the previously identified aryl selectivity issue with aryl(PMP)iodonium salts.364

Scheme 40. Synthesis of 211At-Labeled Compounds Using Aryliodonium Ylides.

Scheme 40

DDT = dithiothreitol, K222 = kryptofix, and TEMPO = 2,2,6,6-tetramethylpiperidin-1-yl)oxyl.

Aryl(TMP)iodonium salts (Ar(TMP)I-X) played a pivotal role in achieving superior aryl transfer selectivity and radiochemical yields (RCY) during the radiosynthesis of [18F]fluoroarenes (Scheme 41a).365 The utilization of these salts facilitated the synthesis of a broad range of electron-rich/deficient [18F]fluoro(hetero)arenes, as well as [18F]fluoroarenes with additional functional groups, which are valuable for constructing potential PET radiopharmaceuticals. Pike et al. evaluated 18F-fluorination of aryliodonium ylides under the normal radiofluorination conditions (Scheme 41b).366 4-substituted aryl iodonium ylides with alkoxy and halogen groups afforded mixtures of regioisomeric 18F-labeled products in moderate ylides when MeCN was used as a solvent, whereas the reactions in DMF yielded lower ylides with significant decrease or absence of the undesired 3-18F products. These results indicated the generation of aryne intermediate as a competing pathway with the direct ipso-nucleophilic mechanism367,368 and declared the role of DMF solvent as an aryne quencher.

Scheme 41. Radiofluorination of Aryliodonium Precursors.

Scheme 41

2.3.8. Synthesis of Heterocycles via Intramolecular Cyclization

Diaryliodonium salts are noteworthy for their ability to form diverse carbocyclic and heterocyclic structures through a range of coupling reactions with nucleophiles. These salts can facilitate arylation via intramolecular cyclization, by directly generating the desired cyclic product or providing an intermediate that can undergo further functionalization and cyclic skeleton formation in a single or multistep process. For instance, Chi and colleagues designed diaryliodonium tosylates that linked the ortho-position of one aryl group to a nucleophile, while the other aryl ligand was substituted with electron-withdrawing and donating groups (Scheme 42a).369

Scheme 42. Diaryliodonium Salts for the Synthesis of Indolines, Benzofurans, and Benzoxazoles.

Scheme 42

Intramolecular ligand coupling resulted in the production of N-protected indolines in high yield. A critical factor in this process was the inclusion of TEMPO as a radical scavenger for achieving optimal results. Kürti and Olofsson independently synthesized benzo[b]furans by two methods: O-arylation of ketoxime using diaryliodonium salts followed by [3,3]-rearrangement and cyclization under acidic conditions, and chemospecific O-arylation of ethyl acetohydroxamate using diaryliodonium salts, followed by the addition of ketones and subsequent treatment with aqueous hydrogen chloride. Kürti’s approach involved O-arylation of ketoximes with diaryliodonium salts to give O-arylketoximes, which were then treated with aqueous hydrogen chloride to produce a diverse range of benzofurans (Scheme 42b).211 Togo et al. also reported a similar sequential reaction using one-pot procedure.212 Olofsson’s method involved a sequence of hydrolysis/transoximation, [3,3]-rearrangement, and cyclization, ultimately yielding the substituted benzofurans (Scheme 42c).213 Mo et al. extended the synthesis of substituted benzoxazoles by using a two-step protocol involving selective O-arylation of amidoxime with diaryliodonium salts and treatment with trifluoroacetic acid to achieve [3,3]-rearrangement and cyclization, resulting in the final benzoxazole products (Scheme 42d).370

Mo and colleagues have developed a one-pot cascade reaction of alkyne-tethered-oximes, which proved effective in synthesizing 2,3-quaternary fused indolines with high diastereoselectivity. This reaction sequence involved N-arylation of oxime, 1,3-dipolar cycloaddition, and [3,3]-rearrangement (Scheme 43a).371

Scheme 43. a) Synthesis of Fused Indolines via N-Arylation, [3+2] Cycloaddition, and [3,3]-Rearrangement. b) Synthesis of 2-Hydroxytetrahydroquinolines from O-Cyclopropyl-hydroxylamine via N-Arylation, Rearrangement, Cyclization, and Aromatization.

Scheme 43

The use of an ether-linker (X = O) in the reaction led to the production of only fused indoline products through [3,3]-rearrangement. The Kürti team proposed that N-arylation of O-cyclopropyl hydroxylamine could be a suitable precursor for [3,3]-sigmatropic rearrangement and the formation of diverse heterocycles (Scheme 43b).372 To test this hypothesis, a library of O-cyclopropyl hydroxylamines was reacted with diaryliodonium triflates, resulting in the formation of N-aryl-O-cyclopropyl-hydroxamates. Treating the N-arylation products with triethylamine successfully resulted in the formation of 2-hydroxytetrahydroquinolines via a cascade reaction involving [3,3]-rearrangement, cyclization, and rearomatization.

Acridine derivatives were synthesized through a one-pot reaction of substituted ortho-acylanilines with diaryliodonium triflate, which involved N-arylation of the amino group followed by intramolecular Friedel–Crafts cyclization (Scheme 44a).235 The same hypothesis was applied to the synthesis of 9-arylxanthenes using a cascade reaction of 2-(aryl(hydroxy)methyl)phenol with diaryliodonium triflates to generate an O-arylation product, which subsequently cyclized under acidic conditions to form 9-arylxanthenes (Scheme 44b).373

Scheme 44. Synthesis of Acridine/Xanthene via Arylation and Friedel-Crafts Cyclization.

Scheme 44

Ogura and co-workers synthesized a range of substituted indolines and indoles by using ortho-N-alkyl-O-TBS-hydroxylamine tethered diaryliodonium tosylates and treating with TBAF (Scheme 45a).374 When an electron-withdrawing N-Boc group was introduced to the substrate, the related functionalized indole was afforded instead (Scheme 45b). The reaction involved the initial C–N bond formation followed by Boc-migration to generate Boc-protected N-hydroxy indoline, which underwent extrusion of tert-butyl hydrogen carbonate and isomerization to afford the indole product.

Scheme 45. Intramolecular Cyclization of Iodonium Salts into N-Alkoxyindolines and Indoles.

Scheme 45

2.4. Arylation via Generation of Aryl Radical

Aryldiazonium salts are utilized as precursors for aryl radicals; however, they are limited in accessibility due to their instability, explosiveness, and the difficulty associated with their handling. Additionally, aryl halides require a high reduction potential for activation.375 In contrast, diaryliodonium salts offer a more suitable alternative since they are easy to prepare, less toxic, and have a lower reduction potential compared to aryl halides, displaying reactivity comparable to aryldiazonium salts. They are highly reactive and stable, making them ideal candidates for the construction of C–C and C–heteroatom bonds through formal aryl cations and arynes. Additionally, diaryliodonium salts can exhibit radical reactivity by homolysis of the weak hypervalent bond under thermal or light-mediated conditions.375,376 Transition metal-free aromatic C–H arylations with hypervalent iodines, including diaryliodonium salts, have gained significant attention due to their environmentally friendly and sustainable nature, making them attractive alternatives to metal-catalyzed processes.39,377383

Kita and his group also discovered a mechanistically different approach for the construction of biaryl that depended on the reaction conditions, and particularly the Lewis acid activator used (Scheme 46).101 Cross-coupling of aryl(heteroaryl)iodonium bromide with electron-rich arene in the presence of TMSOTf led to the synthesis of heteroaromatic biaryl with a unique ipso-substitution of the heteroaryl moiety on the iodonium salt. The reaction scope was expanded to include diverse diaryliodonium salts with electron-rich aryl moieties in addition to aryl(heteroaryl)iodonium bromide and a broad range of electron-rich (polycyclic)aromatic substrates.101,102 UV–vis and ESP spectroscopic studies indicated the generation of radical cation species during the reaction and consequently supported the SET reaction mechanism via EDA complex. Distinctive ipso-substitution of the iodonium salt through the highly chemoselective transfer of the more electron-rich thienyl moiety to the radical cation resulted in the formation of the final biaryl product after a rearomatization step.

Scheme 46. Oxidative Couplings and Biaryl Formation via a Unique SET Mechanism.

Scheme 46

Arylating aromatic hydrocarbons using diaryliodonium salts under base and solvent-free microwave conditions was developed by Rodríguze et al. (Scheme 47a).384 Ackermann reported on the base-free site-selective C3-arylation of various substituted indoles (Scheme 47b).385 Oligopeptides containing indole-3-acetamide moieties were effectively arylated (Scheme 47c).386 Zhang, Yu, and their colleagues selectively arylated N-heteroarenes, such as pyrroles and pyridines, with electron-rich/poor diaryliodonium salts under basic conditions (Scheme 47d and e).387 Quinones such as 1,4-benzoquenone and 1,4-naphthoquenone were effectively arylated at the C2-carbon with a broad range of electron-rich diaryliodonium salts (Scheme 47f).388 The challenge of aryl transfer selectivity in the radical arylation of heteroarenes was tackled through the design of diaryliodonium salts (Scheme 47g).389 Aryl(TMP)iodonium triflates (Ar(TMPI-OTf) were identified to be the optimal design for base-induced radical couplings with N-heteroaromatic substrates, leading to the formation of aryl-heteroaryl products with selective transfer of the aryl group rather than the TMP group. The scope of heteroarenes included pyrrole, imidazole, and N-heteroaromatics of pyridine, pyrimidine, and pyridazine, resulting in a mixture of regioisomeric products. Notably, a one-pot sequential arylation reaction via in situ preparation of aryl(TMP)iodonium salt from iodosobenzene was also found to be applicable.

Scheme 47. General Conditions for Csp2-Arylation with Diaryliodonium Salts.

Scheme 47

The C3 arylation of imidazopyridines and imidazothiazoles using diaryliodonium chloride was achieved in the presence of KOtBu (Scheme 48a).390 When the C3-position in the starting imidazopyridine was substituted, no reaction was observed at any other C–H positions. In contrast, the unsubstituted imidazopyridine yielded a relatively low (38%) amount of the desired C3-arylation product. Experimental studies and related reports suggested two possible free radical pathways and ionic mechanisms involving the generation of an aryne intermediate to explain the observed behavior. Successful arylation at the C3-position of quinoxaline-2-ones was achieved using diaryliodonium salts under mild conditions (Scheme 48b).391 Productivity of the process can be influenced by the electronic and steric effects of substituents, with electron-rich quinoxalinones giving better results than electron-deficient ones. High yields were only attained when the N4-imine was present, as it coordinated with the I(III) center and facilitated the generation of aryl radical species. In addition, a free radical mechanism was proposed, based on radical scavenger experiments and the ability of iodonium salts to generate aryl radical intermediates.

Scheme 48. C–H Arylation of Imidazo-pyridine/Thiazole with Diaryliodonium Salts.

Scheme 48

According to Jiao’s research, diaryliodonium salts were utilized for the direct and regioselective α-arylation of boron dipyrromethenes (BODIPYs) under mild reaction conditions (Scheme 49).392 As the equivalents of diaryliodonium salts increased, 3,5-diarylated BODIPYs were obtained with high regioselectivity. The high selectivity of α-arylation, as opposed to the expected β- or meso-arylation of BODIPYs, was supported by DFT calculations. A free radical mechanism, initiated by the decomposition of diaryliodonium salt to generate aryl radical, was proposed.

Scheme 49. Regioselective α-Arylation of BODIPY Dyes with Diaryliodonium Salts.

Scheme 49

Indole-based iodonium salts could potentially be employed in cross-couplings; however, their limited availability and low stability prompted Waser and colleagues to explore more stable cyclic indole- and pyrrole-benziodoxolones for couplings with electron-rich (hetero)arenes under the conditions established by the Kita group, including TMS-Cl/Br and HFIP (Scheme 50).393 Notably, the coupling of 2- or 3-indolyl-benziodoxol(on)e with electron-rich (hetero)arene nucleophilic molecules resulted in the formation of related 2-(hetero)arylated indoles with high regioselectivity. Additionally, 2- or 3-pyrrolyl-benziodoxolones reacted with 1,3,5-trimethoxybenzene to generate the related 2-aryl pyrroles. The authors proposed three possible mechanisms initiated by the Lewis acid: SET from a charge-transfer complex as defined by Kita, an SNAr process through either the general ligand exchange/reductive elimination, or participation of the N-atom of the indolyl moiety in the generation of the iodonium ylide.

Scheme 50. (Hetero)arylation through Oxidative Cross-coupling of (Hetero)arenes with Heteroaryl-benziodoxol(on)es.

Scheme 50

Wang, Han, and co-workers have investigated ortho-functionalized diaryliodonium salts for intramolecular aryl migration and the construction of valuable skeletons (vide infra).394398 They discovered that treating N-alkyl/aryl sulfonamide-substituted diaryliodonium salt with triethylamine resulted in desulfonylative aryl migration and the formation of sterically congested biarylamines (Scheme 51).398 Cyclic sulfonamides could also be produced using this method, depending on the Ar3 substituents of the aryl sulfonyl group. The reaction was initiated by electron-donor–acceptor complexation of triethylamine with iodonium salt, followed by internal SET to generate the aryl radical. This radical subsequently underwent [1,5]-substitution followed by desulfonylation and H-abstraction to ultimately yield the biaryl product. The cyclic amides were generated via [1,6]-addition and subsequent aromatization.

Scheme 51. ortho-Substituted Diaryliodonium Salts and Construction of Hindered Biarylamines.

Scheme 51

Cyclic diaryliodonium triflates were designed to undergo intramolecular annulations catalyzed by alkylamine and deliver polycyclic aromatic hydrocarbon frameworks (Scheme 52).399 Cyclic iodonium salts underwent intramolecular ring contraction in the presence of tBuNH2 to give the corresponding polycyclic fused systems. The catalytic cycle was initiated by SET from tBuNH2 to the iodonium salt and the generation of a ring-closed radical, which reacted with the amine cation radical to generate a polycyclic cation intermediate. In the final step, aromatization proceeded in the presence of a base to give the fluorene derivative. Various cyclic iodonium salts successfully afforded the corresponding polycyclic arenes.

Scheme 52. Base-Promoted Intramolecular Annulation of Cyclic Diaryliodonium Salts.

Scheme 52

Rohde, Hong, and co-workers proposed BHAS (base-promoted homolytic aromatic substitution) as an alternative strategy to the failed annulative π-extension (APEX) approach for the construction of polycyclic N-heteroarenes from the reaction of azine with cyclic diaryliodonium salt (Scheme 53).400 In addition to the reaction as substrate, azines participated as a promoter with KOtBu to in situ generate the organic electron donors (OEDs) that are required for the dissociation of the iodonium salt. Pyrazines, pyridines, and quinoxalines were used as azine substrates, and unactivated benzene was also used in the presence of pyrazine (2 equiv) as a promoter to give the corresponding polycyclic aromatic products. The reaction was initiated by a reaction of pyrazine 2 with KOtBu to in situ generate dianion and radical anions species as OEDs. Reduction of the iodonium salt via SET from OEDs produced aryl radical via C–I bond homolysis. Minisci-type addition of an aryl radical to azine generated an aryl-heteroaryl radical. Subsequent deprotonation followed by intramolecular SET yielded a poly cyclic radical, which underwent deprotonation followed by SET to afford iodonium salt and then the desired product.

Scheme 53. Base-Promoted Annulation of Azines with Iodonium Salt and Formation of N-Doped Polycyclic Aromatic Compounds.

Scheme 53

Various aryl- and heteroaryl-flanked cyclic iodonium salts have been synthesized and transformed via S, Se, Te, and SO2 exchange with I to give the corresponding diary-annulated frameworks under transition metal-free conditions.332,401,402 The pioneering work of Jiang and co-workers included the synthesis of diaryl-annulated sulfides and selenides by reaction of cyclic diaryliodonium salt with elemental sulfur and selenium under basic conditions, respectively (Scheme 54a).331 This process allowed the synthesis of π-conjugated sulfide and selenide cyclic compounds as promising motifs in organic field-effect transistors (OFET). The reaction mechanism was supported by mechanistic and electron paramagnetic resonance (EPR) studies (Scheme 54b). The deep blue solution of the reactive trisulfur radical anion species was obtained by reaction of elemental sulfur with a base. During the anion exchange process, an iodonium-trisulfur intermediate was formed, followed by radical transfer from the trisulfur moiety to the aryl group and dissociation of Ar–I bond. The thus-generated aryl radical intermediate coupled with another trisulfur radical anion, and the corresponding thiophenol anion underwent intramolecular cyclization via nucleophilic substitution to afford the final product. Elemental selenium can be also employed to synthesize the corresponding diarylannulated selenides in the presence of KOtBu (Scheme 54c).

Scheme 54. Synthesis of Annulated Sulfide and Selenide π-Conjugates via S/I and Se/I Exchange.

Scheme 54

The addition of TEMPONa promotes the generation of aryl radical species from aryl-λ3-iodanylidene malonates, and was utilized for a three-component cascade reaction forming 1,2-oxyarylated products, reported by the Studer group (Scheme 55a).376,403405 The reaction of iodanylidene malonates gave the desired oxyarylation products with excellent aryl selectivity through fragmentation of the aryl radical. The reaction was well-tolerated by various styrene substrates, but poor yields were obtained with internal alkenes and aliphatic alkenes. The reaction mechanism involved the generation of an aryl radical through either the association of TEMPONa with iodine(III), followed by inner-sphere SET or direct reduction through outer-sphere SET to generate the aryl radical. The addition of the aryl radical to the alkene resulted in the formation of an alkyl radical, which coupled with the persistent TEMPO radical to produce the observed product. When unsymmetrical diaryliodonium salts (Ar ≠ Ar’) were employed as the aryl source, mixtures of oxyarylation products were observed (Scheme 55b).

Scheme 55. Single Electron Reduction of Iodine(III) Compounds with TEMPONa and Cascade Radical Alkene Oxyarylations.

Scheme 55

The arylation reaction between vinyl pinacolboronates and diaryliodonium triflates resulted in the formation of trans-arylvinylboronates through the unique involvement of wet carbonates (Scheme 56).406 The reaction mechanism commenced with the ionization of K2CO3 by water to generate the carbonate anion (CO32–), which led to the generation of an aryl radical and [Ph-I+][CO32–]. Meanwhile, the aryl radical reacted with CO32–-activated vinyl boronate to give α-boronate radical, which underwent SET with [Ph-I+][CO32–] followed by deprotonation and the subsequent intramolecular release of bicarbonate to afford the final product.

Scheme 56. Wet Carbonate Promoted the Arylation of Vinylboronates with Iodonium Salt.

Scheme 56

In addition to the former protocols for functionalization of 2-naphthol with iodonium salts,220,288,289 Solorio-Alvarado and co-workers discovered a new activation mode for diaryliodonium salts during C- and O-double arylation of 2-naphthols in the presence of the radical precursor TMP2O [1,1′-oxybis(2,2,6,6-tetramethylpiperidine)] under base-free conditions (Scheme 57).407 Experimental and DFT studies indicated a radical mechanism, initiated by TMP2O via spontaneous N–O bond homolysis to give TMP and TEMPO. HAT from 2-naphthol to TMP provided a C-radical, which reacted with diaryliodonium triflate to generate a C-arylated intermediate. The second equivalent of diaryliodonium salt delivered the diarylated product.

Scheme 57. TMP2O Mediated Double Arylation of 2-Naphthol with Diaryliodonium Triflates.

Scheme 57

Radical O-arylation of N-hydroxyindazoles with diaryliodonium salts and subsequent [3,3] rearrangement was observed to be condition-dependent (Scheme 58).408 3 equiv of diaryliodonium salt and KOtBu produced the corresponding and highly preferred 3-(2-hydroxyaryl)indazoles, whereas 6 equiv of diaryl iodonium salts and KOtBu gave the corresponding dehydrogenative cross-coupling products N-(tetrahydrofyran-2-yl)-3-(2-hydroxyaryl)indazoles. Based on mechanistic studies, the postulated mechanism was initiated by the generation of aryl and tBuO radicals, which then reacted with N-hydroxyindazole to generate phenoxy and indazole radicals. Subsequently, radical–radical coupling followed by rearomatization afforded the desired product.

Scheme 58. Radical O-Arylation and [3,3] Rearrangement to Afford C3-Hydroxyarylindazoles.

Scheme 58

Li and co-workers replaced the traditional reaction conditions with liquid-assisted ball-milling conditions for N-arylation of aniline derivatives with diaryliodonium triflates (Scheme 59).409 In contrast to the popular nucleophilic substitution mechanism of this transformation, this reaction proceeded via a radical mechanism. Oxidation of aniline by diaryliodonium triflate afforded azobenzene, which attacked via the phenyl radical generated from the homolytic cleavage of Ar–I(III) to give an amidyl radical. In the presence of H2O, the amidyl radical was protonated to give a hydrazine derivative, which collapsed to afford the desired product along with nitosobenzene.

Scheme 59. N-Arylation of Anilines with Iodonium Salt Using the Ball-Milling Technique.

Scheme 59

2.5. Arylation via Generation of Arynes

In addition to acting as aryl cation and aryl radical precursors, diaryliodonium salts are superior candidates for aryne precursors due to the high electrophilicity of the central IIII atom and the hypernucleofuge capability of aryl-λ3-iodane (hyperleaving group) which is a million times more than the triflate group.410412 The generated aryne intermediates provide double functionalization of arenes in a single step, which is a highly desirable synthetic approach.116

Originally, Beringer et al. noticed the decomposition of diphenyliodonium-2-carboxylate (also known as phenylbenziodoxole) at high temperatures (>180 °C) and generation of benzyne that was trapped by broad range of arynophiles (Scheme 60a).413 The ability of diaryliodonium salt to generate an aryne intermediate through deprotonation of ortho-hydrogens with strong base was discovered by Akiyama.414 The reaction of di(4-tolyl)iodonium bromide in the presence of NaOtBu under reflux conditions afforded a 27% combined yield of 3-tolyl- and 4-tolyl-ethers in a 1:1 ratio through the in situ generation of the aryne (Scheme 60b). Kitamura and co-workers designed (phenyl)[o-(trimethylsilyl)aryl]iodonium triflate as an aryne precursor to solve the previously associated challenges of the generation of more than one aryne intermediate, the use of a strong base and the requirement of harsh conditions (Scheme 60c).415417 Controlling the reaction temperature led to a chemoselective generation of the aryne functionality. Thus, double cycloadditions with two different arynophiles in a sequential one-pot reaction produced complex polycyclic aromatic compounds. This study represented the unique reactivity and hypernucleofugality of the aryliodonium group in the presence of a triflate-leaving group.2123 Yoshimura, Zhdankin, and co-workers developed a substrate analogous to Kitamura’s substrates for the generation of aryne under benign conditions (Scheme 60d).418,419 The presence of ortho-B(OH)2 to the hypernucleofugic aryliodonium group made the boronic acid group more oxophilic and allowed it to be triggered by water as the only activator.

Scheme 60. Generation of the Aryne from ortho-(Un)functionalized Diaryliodonium Salts.

Scheme 60

Aryne generation from ortho-functionalized diaryliodonium salts was applied to C–P and C–S bond formations (Scheme 61). Chlorodiphenylphosphine was used to trap the benzyne intermediate generated from the collapse of diphenyliodonium-2-carboxylate, then oxidation with hydrogen peroxide yielded the ortho-chlorophenyldiphenylphosphine oxide (Scheme 61a).420 Kitamura and Stang’s benzyne precursor reacted with methyl phosphorodiamidite in the presence of TBAF to produce phenylphosphonic diamide in a quantitative yield (Scheme 61b).421 Yoshimura, Zhdankin, and co-workers used pseudocyclic arylbenziodoxaborole triflate for a reaction with organic sulfides and the synthesis of the related sulfonium salts with high regioselectivity (Scheme 61c).422 A broad library of cyclic/acyclic dialkyl, aryl alkyl, and diaryl-sulfides were tolerated under the reaction conditions to give the corresponding meta-fluorophenyl-substituted sulfonium salts. Additionally, dimethyl and methyl phenyl sulfoxides were incorporated to give ortho-hydroxy-substituted sulfonium salts via stepwise formation of four-membered cyclic intermediate (Scheme 61d).

Scheme 61. Ar–P and Ar–S Bonds Formation with Diaryliodonium Salts.

Scheme 61

Using unsymmetrical diaryliodonium salts with ortho-hydrogens on the two aryl ligands for the generation of arynes through a C–H deprotonation approach is an uncontrollable strategy because of the possibility for generation of up to four chemo- and regioselective arynes. Therefore, Stuart and co-workers employed, under basic LiHMDS conditions, the synthetically accessible aryl(Mes)iodonium tosylate (Ar(Mes)I-OTs) for chemospecific generation of arynes and extrusion of mesityl iodide, which can be recycled (Scheme 62).237,423 The generated arynes were trapped through [4 + 2] cycloaddition with furan to evaluate the regioselectivity of the deprotonation process. When the aryl groups have electron-rich/poor substituents at C-4, the desirable oxabicyclic products were produced in high yields. Notably, iodonium salts substituted with inductively withdrawing groups at the C-3 position led to deprotonation of the hydrogen at the C-2 position with high regioselectivity (>20:1). However, substitution of the aryl ligand of the iodonium salt with an electron-donating and sterically hindered methyl group at the C-3 position afforded mixture of regioisomers.

Scheme 62. Selective Generation of Arynes from Diaryliodonium Salts.

Scheme 62

In the same vein, Wang and collaborators used Ar(TMP)I-OTs for the generation of arynes in the presence of LHMDS (1.5 equiv), and cycloaddition with N-arylpyrroles to yield the corresponding bridged amine cycloadducts (Scheme 63a).424 The observed trend of chemo- and regio-selectivity was in accordance with Stuart’s trend. The same group developed a direct N-arylation of tertiary amines by using aryl(mesityl)iodonium tosylates under LiHMDS/toluene/110 °C conditions (Scheme 63b);425 the reaction worked well with aliphatic tertiary amines and electron-rich aromatic tertiary amines. A possible mechanism was suggested, beginning with the base-promoted generation of aryne from the iodonium salt, followed by nucleophilic attack by the tertiary amine to generate a zwitterion intermediate, and finally a proton transfer to afford the final product (Scheme 63c).

Scheme 63. Reaction of Aryne with Pyrrole or Tertiary Amines.

Scheme 63

Stuart et al. treated 1,2-disubstituted-aryl(Mes)iodonium tosylates with NaOtBu to generate arynes that subsequently reacted in situ with arynophiles to finally afford 1,2,3,4-tetrasubstituted arenes (Scheme 64).426 The strong electron-withdrawing effect and superior leaving ability of the iodonium group, facilitated by the substituents at the 1- and 2-positions, allowed a concerted regioselective deprotonation at the 3-position and the chemo- and regioselective generation of the aryne. The limitations of aryl(Mes)iodonium tosylates motivated the same group to develop aryl(TMP)iodonium tosylates as a relatively more reactive species and achieve control of the reaction pathway for the selective generation of arynes.427 Reactions of a broad range of aryl(TMP)iodonium salts and arynophiles were performed in the presence of an appropriate base. The reaction of 3-chlorophenyl(TMP)iodonium tosylate and furan was used as an ideal system for one-pot competition reactions between various arynophiles and furan to measure the relative reactivity of these arynophiles and parametrize them under a single arynophilicity (A) value scale.428

Scheme 64. Reaction of Arynes with Various Arynophiles.

Scheme 64

Furthermore, Stuart and co-workers reported aryne generation under mild basic conditions using K3PO4.429 Similarly, Han and collaborators substituted aryl(Mes)iodonium salts with meta-OTf groups to generate an aryne intermediate under mild K2CO3 basic conditions.430 Various meta-OTf iodonium salts reacted smoothly with (substituted)furans, azide, amine, and ethylene arynophiles. Meta-substituted iodonium salts showed a higher reactivity than their para-substituted counterparts with excellent regioselectivity. These conditions tolerated a broad array of iodonium salts and arynophiles with sensitive functional groups such as benzyl halide/alcohol, boronate esters, and ketones. Quantitative analysis of functional group compatibility demonstrated that this approach is more functional group compatible than other known methods.

Takenaga and Dohi studied the potency of uracil(aryl)iodonium salts for the highly chemoselective generation of aryne/heteroaryne. Treating the uracil(aryl)iodonium salt with LiHMDS in the presence of arynophile exclusively gave the uracil-adduct via in situ generation of the heteroaryne analogue of uracil (uracilyne) (Scheme 65).121,431,432 Optimization of the iodonium salt revealed that 2-trifluoromethylphenyl group and tosylate counteranion were important for improving the chemical properties of the iodonium salt and increasing the yields. The reactivity of the generated uracilyne tolerated a broad range of cycloaddition and σ-bond insertion reactions. Unsymmetrical arynophiles were converted to the corresponding cycloadducts with excellent chemo- and regio-selectivity.

Scheme 65. Generation of Uracil-heteroaryne from Uracil-iodonium Salt and Participation in Cycloaddition and Insertion Reactions.

Scheme 65

The Li group designed densely substituted 3-sulfonyloxyaryl(mesityl)iodonium triflates as efficient and high-functional-group-economy 1,2-benzdiyene precursors (Scheme 66a).433 The presence of the mesityliodonium moiety at the meta-position of the sulfonyloxy group allowed the ortho-deprotonative elimination strategy to generate 1,2-aryne under weakly basic conditions, in addition to predicting the site-elective generation of aryne and the subsequent aryne transformation. Consequently, deprotonation of the ortho-proton led to regioselective generation of an aryne intermediate. The presence of the 3-sulfonyloxy group directed the nucleophile to attack the meta-position and released 2,3-aryne, which subsequently participated in the cascade process to finally afford the multisubstituted arene product. The mild conditions applied tolerated 1,2-benzdiyne precursors and were applicable to diverse tandem reactions besides the two-step [2 + 2] cycloaddition and Grob-fragmentation process, accessing chemically and biologically useful skeletons (Scheme 66b).

Scheme 66. Sequential Transformation of Aryne Generated from Diaryliodonium Salts.

Scheme 66

2.6. Transformations to Iodine-Containing Products

Diaryliodonium salts are superior reagents in terms of reactivity and stability for arylating diverse substrates and forming C–C and C–heteroatom bonds. The traditional reaction of diaryliodonium salt involves transferring one of its aryl ligands to the substrate to afford the product, while the other aryl ligand is reductively eliminated in a stoichiometric amount as aryl iodide waste along with the counteranion. Therefore, the atom economy of this process was approximately 10–20%, as estimated by Nachtsheim and co-workers.434 This disadvantage could be overcome by using various recyclable hypervalent iodine techniques, such as polymer-supported, nonpolymeric, ionic-liquid/ion-supported, and metal–organic framework (MOF)-hybridized reagents.19,434

Several protocols were developed to improve the overall atom economy of the process and afford highly functionalized products in a one-pot reaction.435438 These elegant strategies relied on the incorporation of the two aryl ligands of the acyclic iodonium salt in the final product through a one-pot cascade reaction, in which the aryl iodide waste from the first step successively participated as a reagent in the second step. Additionally, connecting the two aryl ligands of the iodonium salt in one system (cyclic diaryliodonium salt) was a sophisticated strategy for designing high atom-economical processes. These earlier atom-efficient transformations of acyclic and cyclic diaryliodonium salts were catalyzed by transition metals.183,402,434,439,440 Herein, we will cover the recent transition metal-free strategies which were developed to improve the sustainability, practicability, and applicability of these atom-economical processes.

ortho-Functionalized diaryliodonium salts were designed by Wang, Han, and co-workers to undergo intramolecular aryl migration and deliver useful products with available ortho-iodo-substituents for further derivatization.394398 The strategy started with diaryliodonium salts bearing ortho-trifluoromethylsulfonyloxy groups designed to undergo intramolecular aryl migration and deliver the corresponding othro-iodo diaryl ethers (Scheme 67a).394 The reaction proceeded through an intramolecular SNAr mechanism directed by the sulfonyl group at the ortho-position. The base was important for trapping the liberated anion and accelerating the dissociation of the sulfonyl group. The same hypothesis was applied to a two-step protocol involving C–H activation of complex aromatic hydrocarbons and synthesis of diverse ortho-OTf substituted diaryliodonium salts followed by site-selective O-arylation via an intramolecular aryl migration mechanism.395 Further investigations delivered a N-acetyl sulfonamide substituted diaryliodonium salt as an ultimate scaffold for intramolecular aryl migration in the presence of DMAP to afford ortho-iodo N-aryl sulfonamides (Scheme 67b).397 The reaction mechanism started with deacetylation of the N-Ac group using DMAP followed by SNAr via a spiro-Meisenheimer complex to afford the final product.

Scheme 67. Intramolecular Aryl Migration of Iodonium Salt and the Formation of Iodo-products.

Scheme 67

The same group designed ortho-sulfonylmethylene-substituted diaryliodonium triflates for an efficient intramolecular SNAr (Truce-Smiles rearrangement) and the synthesis of ortho-iodo-diarylmethylene sulfones (Scheme 67c).441 Ionic liquid (IL) was important for stabilization of the transition states and achieving the reaction under mild conditions with high yields.

Olofsson et al. combined the benefits of the SNAr methodology and arylation with diaryliodonium salt in one cascade process to difunctionalize primary amines, ammonia, and water (Scheme 68).442,443Ortho-fluorinated diaryliodonium salts with an additional electron-withdrawing group (EWG) substituted para to the fluoro-group were designed and synthesized to react with amine/ammonia/water nucleophiles under mild conditions. The presence of two EWGs in the ortho- and para-positions to the fluorine atom caused the SNAr reaction with N- and O-nucleophiles to generate the new iodonium salt intermediate, which underwent intramolecular aryl transfer via I(III)–Ar bond dissociation to afford the diarylated products. Furthermore, the retained iodine atom in the products allowed further derivatization as demonstrated by the conversion to NMP-7.

Scheme 68. Atom-Economic Diarylation of Primary Amines, Ammonia, and Water.

Scheme 68

In addition, Olofsson and colleagues used the previous diaryliodonium salts for developing a novel strategy for diarylation of a sulfur nucleophile (potassium ethyl xanthogenate) and to access an iodo-substituted unsymmetrical diaryl sulfides under base- and thiol-free mild conditions (Scheme 69).444 The reaction conditions tolerated a broad range of functional groups and were employed for the synthesis of complex products using diaryliodonium salts derived from heterocycles, bioactive compounds, and drug molecules. The proposed mechanism was supported by experimental and theoretical DFT studies. The reaction of diaryliodonium salts with S-nucleophiles underwent SNAr followed by nucleophilic cleavage of the xanthogenate moiety. Intramolecular aryl transfer produced the desired diaryl sulfides.

Scheme 69. Atom-Economic Diarylation of a Sulfur Nucleophile with Diaryliodonium Salts.

Scheme 69

After the atom-efficient diarylation of N-, O-, and S-nucleophiles with the two aryl groups of diaryliodonium salts, Olofsson et al. moved forward and designed a novel von Braun-type reaction for the diarylation/ring opening of cyclic amines to give highly functionalized diaryl amines (Scheme 70).445 An iodonium salt with a highly activated ortho-fluoroaryl group was required for the first step of the SNAr reaction with an unstrained cyclic amine to give ortho-amino-substituted diaryliodonium salts in high yields. A one-pot sequential reaction of intramolecular arylation generated the diarylammonium intermediate, then nucleophilic substitution with a nucleophile afforded C- and N-functionalized diaryl amines. The reaction tolerated various iodonium salts and 5/6-membered cyclic amines in addition to a variety of amine, phenol, carboxylic acid, thiol, and halogen nucleophiles.

Scheme 70. Atom-Economic N,N-Diarylation and C-Functionalization through Ring Opening of Cyclic Amines.

Scheme 70

Cyclic diaryliodonium salts were used by Zhang, Wu, and co-workers as aryne precursors for meta-selective halogenation and O-arylation (Scheme 71).446,447 The scope of the phenol included diverse (hetero)aromatic phenols to give high selectivity of the related 2-iodo-3′-phenoxy-1,1′-biaryl products (meta:ortho, 90:1–99:1). The substituents on the aryl ring of the cyclic iodonium salt controlled the site of O-arylation process. 2-Substituted and 2,4-disubstituted iodonium salts afforded meta-functionalization of the less hindered aryl ring (meta:ortho, 75:25–99:1). In contrast, 3′-inductively electron withdrawing substituents (e.g., OMe, CF3) and 6,6′-disubstituted iodonium salts with high torsional strain gave ortho-derivatizations as the major products (meta:ortho, 1:99), involving the ligand coupling with the phenol not via aryne generation.

Scheme 71. Selective O-Arylation of a Cyclic Iodonium Salt with Phenol.

Scheme 71

2.7. C–H Functionalization of (Hetero)arenes via In Situ Formation of Iodonium Salt

The generation of recyclable diaryliodonium salts by recovering the generated aryl iodide waste and reconverting it to the required diaryliodonium salts is not only a useful technique to overcome the disadvantages of the diaryliodonium salt but also could provide an attractive approach for C–H derivatization of (hetero)arenes without prefunctionalization. As there are different techniques used for recycling hypervalent iodines, in this section we discuss only the simplest approach for recycling diaryliodonium salt reagents.448

Transition metal-free C–H derivatization of inactivated arenes/heteroarenes is a highly challenging but desirable green and sustainable strategy.449,450 Diaryliodonium salt could provide a cascade approach to achieve this challenging transformation via a sequence of Ar1–I oxidative activation, coupling with Ar2–H to give Ar1(Ar2)IX, and finally cross-coupling with a nucleophile (Scheme 72). However, this iodonium salt-mediated strategy is also difficult due to the possibility of regio- and chemo-selective products during the second and third cross-coupling steps, respectively. Thus, finding suitable conditions to achieve this process in a regio- and chemo-selective manner is highly desirable as it will reproduce Ar1–I as a coproduct, which could be recycled, minimizing the cost and environmental impact of the process. Also, this hypothesis could help the hypervalent iodine community to develop the catalytic transformation and consequently overcome the main drawback of diaryliodonium salt transformations.230

Scheme 72. Sustainable C–H Arylation Strategy Mediated by Diaryliodonium Salt Formation.

Scheme 72

Kita’s group discovered the first oxidative cross-coupling strategy based on in situ generation of diaryliodonium tosylate from the condensation of [hydroxy(tosyloxy)iodo]benzene (Koser’s reagent) with a heteroaromatic compound followed by ligand exchange of the tosylate with the bromide by the addition of TMSBr; this would give aryl(heteroaryl)iodonium bromide as the key reaction intermediate as there is no reaction with the tosylate salt (Scheme 73a).100 Cross-coupling of the generated iodonium bromide with an aromatic nucleophile, via the formal hydroarylation, produced the related biaryl with an excellent degree of chemo- and regio-selectivity. The presence of hexafluoroisopropanol (HFIP)340 as a solvent and trimethylsilyl bromide (TMSBr) as an activator appeared crucial. This unique oxidative cross-coupling protocol can take place in a stepwise or one-pot sequential reaction and showed applicability for the coupling of thiophene and pyrrole with electron-rich aromatic nucleophiles, such as methoxy-substituted arenes, pyrroles, and thiophenes.100,451,452 Furthermore, this process successfully cross-coupled (hetero)arene C–H with azole N–H, and achieved the construction of C–N biaryls.453 Further extension of this unique oxidative coupling achieved the synthesis of unsymmetrical head-to-tail (H-T) bithiophenes as the sole regioisomeric product (Scheme 73b).98 Under the previously optimized conditions, 3-alkyl- and 3-alkoxy-thiophenes participated as both the iodonium salt component and aromatic nucleophile. The exclusive H-T connection of the bithiophenes was attributed to the high electron-richness of the 2-position of the thiophene substrates that facilitated, in part, the highly regioselective synthesis of the phenyl(thienyl)iodonium salt. The other unreacted part of the thiophene substrate attacked, through its reactive 2-position, the free 5-position on the thienyl moiety of the iodonium salt to give the unsymmetrical thiophene dimer. This methodology was further expanded99 and successfully applied to the concise synthesis of the efficient photovoltaic donor–acceptor oligothiophene MK-2, a high-performance organic dye for application in solar cells.454

Scheme 73. Oxidative Coupling of C–H Thiophene with C–H Arene Nucleophile Mediated by the Formation of Diaryliodonium Bromide.

Scheme 73

Olofsson and co-workers designed a new strategy for C–H nitration of arenes via a one-pot sequential oxidation of Ar2–I, coupling with Ar1–H and formation of diaryliodonium salt, and finally nitration with NaNO2 to produce the functionalized nitroarene (Scheme 74).230 However, the scope of this C–H derivatization was limited to symmetrical diaryliodonium salts and arene substrates with halogen- and alkyl-substituents in addition to the unsubstituted benzene.

Scheme 74. C–H Functionalization of Arene Mediated by In Situ-Generated Diaryliodonium Salt.

Scheme 74

Iodoarylation of heterocycles through the in situ generation of iodonium salt/ylide was developed by Cheng et al. as an efficient one-pot two-step strategy for highly economical and useful transformations.455,456 For example, the reaction of NH-pyrazole with aryliodine diacetate gave N-aryl-4-iodopyrazole via a sequence of in situ generation of pyrazolyl(aryl)iodonium salt under acidic conditions followed by intermolecular regioselective N-arylation (Scheme 75).457 In this reaction, the iodine atoms were incorporated in the final product.

Scheme 75. 1,4-Difunctionalization of Pyrazole via In Situ Generation of Iodonium Salt.

Scheme 75

Ishikawa et al. synthesized an electron-rich indole using a one-pot reaction of electron-rich aromatic-tethered boron-masking N-hydroxylamides with PIFA, followed by treatment with triethylamine (Scheme 76).458 The reaction mechanism involved the in situ generation of diaryliodonium salt, which was then treated with triethylamine to produce the indole product after a sequence of deborylation, acyl migration, intramolecular cyclization, and desorption of carboxylic acid before tautomerization. The use of a boron-masked scaffold, together with the careful design of electron-rich aryl substrates, was essential for the unique regioselective diaryliodonium salt formation.

Scheme 76. Synthesis of Indoles via In Situ Generation of Iodonium Salt and Intramolecular C–N Bond Formation.

Scheme 76

Dohi and co-workers used iodomesitylene as a recyclable iodoarene for C–H functionalization of arenes through the synthesis of aryl(Mes)iodonium salts (Ar(Mes)I-X) via a stepwise or one-pot strategy (Scheme 77a).141 The reaction tolerated a broad range of functional groups and worked well with arene and heteroarene nucleophiles to produce the related iodonium salts with excellent regioselectivity. These aryl(mesityl)iodonium salts were screened in transition metal-free transformations such as SET-induced coupling of 2,4,6-trimethoxyphenyl(mesityl)iodonium tosylate with 1,4-dimethoxynaphthalene and production of the corresponding biaryl with exclusive chemoselective transfer of the trimethoxyphenyl moiety from the iodonium salt (Scheme 77b). Aryne generation from the 4-methoxyphenyl(mesityl)iodonium tosylate in the presence of LiHMDS and [3 + 2] cycloaddition with furan gave a bridged cycle product (Scheme 77c). Recently, the Stuart group designed a one-pot formal dehydroboration of aryl boron reagents via in situ generation of aryl(Mes)iodonium salt and aryne intermediates.459 This sequence was applied to the formal synthesis of the investigational Aurora Kinase inhibitor PF-03814735.

Scheme 77. Recyclable Diaryliodonium Salt-Mediated C–H Functionalization of Arenes.

Scheme 77

Zhao and co-workers developed a novel two-step strategy for C–H functionalization of arenes mediated by in situ generated isoxazole diaryliodonium salt (Scheme 78a).460 The incorporation of 4-iodo-3,5-dimethylisoxazole (DMIX-I) as a recyclable iodoarene for the synthesis of aryl(DMIX)iodonium salts provided superior reactivity and selectivity over the well-established aryl(TMP)iodonium salts. This discovery provided an alternative and more practical approach for the preparation of diaryliodonium salts by introducing the target C–H arene with a high site selectivity at a late-stage after the synthesis of DMIX-I(OH)OTs and DMIX-I(OAc)2 intermediates. The scope of aryl(DMIX)iodonium salt was extended to various arenes including a bioactive compound with regioselective C–H derivatization. Ligand coupling of the aryl(DMIX)iodonium salt by a broad range of nucleophiles including C-, N-, O-, and S-nucleophiles afforded the related functionalized arenes with high selectivity (Scheme 78b). In addition to the two-step strategy, a one-pot procedure for the regioselective C–H λ3-iodanation of arenes followed by chemospecific arylation of a nucleophile was also possible.

Scheme 78. C–H Derivatization of (Hetero)arenes through the Formation of Diaryliodonium Salt Intermediates.

Scheme 78

The usefulness of the C–H functionalization of arenes mediated by diaryliodonium salt was assessed in drug design and discovery through C–H functionalization of naproxen and gemfibrozil drugs.461,462 Coupling of a C–H drug substrate, such as naproxen or a gemfibrozil derivative, with ArI(OH)OTs or ArI(OAc)2 in the presence of fluorinated solvent afforded the corresponding diaryliodonium salt (Scheme 79). These iodonium salts are suitable for diverse functionalization with excellent chemoselectivity under transition metal- free conditions to provide a library of modified drugs.

Scheme 79. C–H Functionalization of Drugs Mediated by Iodonium Salt.

Scheme 79

3. Base-Promoted Ar–I Bond Dissociation

Transition metal-free C- and heteroatom-arylations via direct cross-coupling with aryl halides provide atom- and step-economical processes for the syntheses of functionalized arene skeletons.463466 In addition to diaryliodonium salts and their unique hypervalent bond (Ered ≈ −0.65 V vs SCE),467,468 which are highly efficient arylating reagents under transition metal-free benign conditions, aryl halides are attractive arylating reagents because they can be used directly without further modification and are commercially available, easily synthesized, inexpensive, and bench-stable starting materials. The BHAS or organocatalytic C–H arylation reaction has recently garnered significant interest owing to the possibility of the construction of diverse skeletons via inter- and intramolecular BHAS reactions under mild reaction conditions without requiring sophisticated transition metals and ligands.377,469,470 In these reactions, aryl halides are effective aryl radical precursors in the arylation of a broad range of arenes under base-catalyzed conditions. Cleaving the aryl–halogen bond without external thermal, photo- or electrochemical stimuli is challenging mainly because of a higher energy barrier, that results from a high reduction potential (Scheme 80a).471,472 Among all aryl halides, iodoarenes are characterized by their reduction potential, bond dissociation energy, polarizability, and different Ar–X bond cleavage mechanism,473 which enable the transition metal-free arylation process to be performed under benign conditions with a broad substrate scope and functional group tolerance. For base-promoted Ar–I bond dissociation, the transformation requires either a strong base or combination of base and ligand (also named promoter, activator, or catalyst). The general mechanism of this strategy is based on the generation (in situ) of an OEDs with sufficient redox potential to initiate a single-electron transfer (SET) process to the aryl halide and affords a reactive aryl radical species for arylation reactions (Scheme 80b). Notably, single electron donors with potentials as low as −1.3 V versus Fc/Fc+ (−0.8 V vs SCE) still have potency for activating aryl iodide substrates. The high energy barrier of the SET process (uphill by 0.6–0.8 V) can be overcome by the irreversible conversion of [Ar–I]•– to Ar and I. The SET process is a technique that can be used to access previously inaccessible aryl halides or accomplish chemical transformations under relatively mild conditions; furthermore, inexpensive and commercially available starting substrates can be used.474491 This transformation may be recognized as an “electron-catalyzed reaction”,466,492,493 which played a crucial role in modern radical chemistry and delivered a promising route to obtain radical intermediates.

Scheme 80. (a) Reduction Potentials of Arylating Reagents. (b) Aryl Radical Generation via Single-Electron-Transfer (SET) Process.

Scheme 80

The pioneering work of Itami et al. emphasized the importance of using KOtBu for the arylation of azines (i.e., pyrazine) with aryl iodides (Scheme 81a).494,495 The efficiency of this type of transformation was maximized by the addition of a catalytic amount of organic compound as an additive,469 which coordinated MOtBu (M = Na, K) and facilitated the initiation step of the chain process (Scheme 81b). A plausible mechanism for the radical reaction was proposed by Studer and Curran and involves initiation and propagation processes (Scheme 81c).492 The initiation process was clarified by Murphy et al.469 to include the reaction of MOtBu with an organic additive or in situ generation of an OED as a promoter of the initiation step via the single electron transfer (SET) reduction of the aryl halide into the aryl halide radical anion, which subsequently dissociates to the aryl radical and halide anion. The generated aryl radical reacts with aromatic compounds to produce cyclohexadienyl radical species, which are subsequently deprotonated by a base to generate the keystone biaryl radical anion, which serves as a powerful reducing agent (reductant upconversion470). Propagation of the process via a SET to the aryl halide produces the biaryl product and regenerates another radical anion to continue the radical chain process.

Scheme 81. Transition Metal-Free Cross-Coupling of Aryl Halide with Arenes.

Scheme 81

Murphy et al. performed extensive experimental and theoretical studies to propose a unified mechanism in which various organocatalysts served as precursors of the electron donor species. Thus, the organocatalysts reacted first with tert-butoxide salt to generate the electron donors in situ; the electron donors transfer an electron to halobenzene and form the aryl radicals.469,496503

3.1. Organocatalytic Activation of Ar–I Bond

Activation of the reaction between (hetero)aryl halides and unactivated (hetero)arenes using a combination of tert-butoxide salt and organocatalyst is the most significant and practical alternative to using transition metals from an environmental, economic, and safety perspective.382,493,504 After the first transition metal-free biaryl coupling was discovered by Itami et al.,494,495 various additive were found to be crucial for coupling of aryl halides to nonactivated arenes (Scheme 82). In 2010, three research groups described organocatalysts as being efficient activators for the C–H arylation of unactivated arenes with haloarenes in the presence of tert-butoxide salts as no reactions occur in the absence of organocatalysts. Kwong and Lei505507 reported dimethyl ethylenediamine (DMEDA) and the groups led by Shirakawa and Hayashi508 and Shi509 extolled the importance of 1,10-phenanthroline (Phen) derivatives as promoters of the C–H arylation of arenes. 1,10-Phenanthrolines became popular organocatalysts for the activation of aryl halides to generate aryl radicals which cross couple with (hetero)arene via inter- and intramolecular approaches.510513 Interestingly, commercially available and inexpensive compounds, such as phenyl hydrazine,514,515 amino acid (l-proline),516 alcohol (nBuOH),517,518 and urea (1,3-diethylurea)519 were found to be efficient catalysts for activating the cross-coupling between C–H (hetero)arene and (hetero)aryl iodide. Aniline derivatives (N-methylaniline520 and indoline521) and pyridine derivatives (8-hydroxyquinoline522 and 2-pyridyl carbinol496,523) were also used as catalysts. Similar to the previously known behavior of the organocatalyst/KOtBu combination, aminoalcohols,524,525 sulfonylhydrazide,526 quinoline-1-amino-2-carboxylic acid,527 diketopiperazine,528531 9-methylamino-phenalen-1-one (phenalenyl ligand),532 2,6-bis(imino)pyridine,533 and bis(arylimino)acenaphthene534 were used catalytically in the presence of KOtBu to promote the C(sp2)-H arylation of unactivated arene with aryl halide to afford the corresponding biaryls. Carbene precursors are also efficient organocatalysts for the C–H arylation of arenes in the presence of a strong base.497,535,536 Porphyrin,537 macrocyclic aromatic pyridone pentamer,538 and foldamer-based pyridone dimer,539 in addition to other promoters,540546 are potent organocatalysts for activating the cross-coupling between unfunctionalized C–H arenes and aryl iodide. In addition to the previously mentioned homogeneous organocatalysts, heterogeneous promoters such as graphene oxide,547N-doped porous carbon nanotubes,548 and glucose-derived carbonaceous549 and porous phenanthroline-based polymers550 displayed a high catalytic activity and recyclability for the direct C–H arylation of unactivated arenes with aryl iodide.

Scheme 82. Transition Metal-Free Organocatalytic Arylation of Unactivated C–H Arenes with Aryl Iodide.

Scheme 82

Zhou and Zeng et al. designed crown ether-containing monopeptides as organocatalysts for activating the direct C–H arylation of (hetero)arene by aryl iodides in the presence of KOtBu (Scheme 83, conditions a).551 Association of the organocatalyst through noncovalent interactions enables multiple crown ether moieties to efficiently bind K+ ions and subsequently allows the tBuO anion to facilitate the activation of aryl iodides via the SET protocol. The size of the crown ether and hydrophobicity of the side chain are crucial for the high performance and lower catalytic loading of the organocatalyst. The catalyst activates only aryl iodides as haloarenes. From among the N,N-/O,O-/N,O-bidentate organocatalysts, Chaudhary et al. introduced 2,3-di(pyridine-2-yl)pyrazine as a new organocatalyst for the direct cross-coupling of haloarene with unactivated benzene (Scheme 83, conditions b).552 Diverse aryl halides and heteroaryl halides were used to afford the corresponding biaryls.

Scheme 83. Crown Ether-Derived Supramolecular Organocatalyst to Promote the C–H Arylation of Arenes with Aryl Halides: 2,3-Di(pyridine-2-yl)pyrazine Promoted Haloarene for the Arylation of Benzene.

Scheme 83

The cascade reaction of benzhydrol and 2-iodoaniline in the presence of KOtBu/1,10-phenathroline, via a controlled sequence of the oxidation of an alcohol to a ketone followed by the direct condensation with the aniline-NH2 group and radical annulation, provided a direct process for the synthesis of phenanthridine derivatives (Scheme 84a).553 Under the reaction conditions, KOtBu played an important role in the aerobic oxidation of the alcohol and the homolytic aromatic substitution (HAS). The scope of 2-iodoaniline extended to include electron-rich/deficient and biaryl substrates. Substituted benzhydrols afforded regioisomeric mixtures based on the mode of action together with the annulation with both aromatic rings. The proposed mechanism involves the reduction of the in situ formed imine by the generated super electron donor species (Scheme 84b). Thus-generated aryl radical cyclized to afford thus-generated aryl radical which cyclized to afford the tricyclic radical, which converted to the final product via the SET process. Wu et al. developed a cascade reaction of 2-halobenzaldehydes and indolin-2-ones in the presence of Cs2CO3 for the synthesis of naphtho[3,2,1-cd]indol-5(4H)-ones (Scheme 84c).554 Indolin-2-ones served as the substrates and electron donor precursors. The proposed mechanism initiated with indolin-2-ones by Cs2CO3 to generate the electron donor enolate, which reacted with the aldehyde to form the condensed intermediate. The SET from the electron donor enolate to the condensed intermediate provided the aryl radical, which underwent 6-endo cyclization to produce the final product.

Scheme 84. Cascade Synthesis of Phenanthridine Derivatives: Indoin-2-one as Substrate and Promoter for the Synthesis of Tetra-Fused Skeletons.

Scheme 84

Murphy et al. have long been involved in the development of neutral organic super electron donors for the activation of various substrates including aryl halides.555564 However, these super electron donors are used in stoichiometric amounts or in excess to promote the reactions and this factor impedes the use of these reagents in radical chain reactions. Recently, the same group solved this problem by altering the nature of the super electron donors.565 Thus, using benzimidazolium salt led to the catalytic radical cyclization of N-allyl- and propargyl-2-iodoaniline through 5-exo-trig and 5-exo-dig cyclization and the cyclized indoline and indolenine products (Scheme 85a) were obtained. Computational and experimental studies revealed that the generated benzimidazole radical is among the most potent electron donor reducing agents (Eox = −1.86 V vs SCE). The proposed mechanism for the reaction involves the reduction of the aryl iodide substrate by the benzimidazole radical via the SET process to generate an aryl radical, which undergoes 5-exo-trig cyclization to afford a radical intermediate, followed by the hydrogen atom transfer (HAT) process to furnish the final product (Scheme 85b). Similarly, Kondo et al. developed an intermolecular cross-coupling of aryl halide with styrene (1,1-diarylethelene) in the presence of NaH/1,10-phenanthroline, which generated reduced anion species as super electron donors and H-sources.566 The hydroarylation of styrene proceeds in an anti-Markovnikov fashion to afford 1,1,3-triarylethane.

Scheme 85. Catalytic Reductive Radical Cyclization Reactions with Super Electron Donors.

Scheme 85

Cheong and Lee reported the combination of 2-iodoaniline derivatives and ketones in the presence of KOtBu and 4,7-diphenylphenanthroline for the construction of diverse indoles under mild conditions (Scheme 86a).567 DFT calculations supported the mechanism shown in Scheme 86b. Deprotonation of both substrates followed by the SET process from the electron donor generated the aryl radical complex. Radical-enolate coupling via 7-endo-trig, followed by the SET process, protonation, and condensation afforded the final indole product.

Scheme 86. Synthesis of Indoles via Radical-Enolate Coupling.

Scheme 86

The Breslow intermediate derived from N-heterocyclic carbene organocatalysis enabled various incredible transformations and successfully radical NHCs were applied to the HAT strategy for the direct activation of aliphatic C–H bonds.568570 Bertrand et al. demonstrated that the reaction of aryl iodide, aryl aldehyde, and styrene under mesoionic carbene-catalyzed conditions afforded the corresponding acylarylation products (Scheme 87a, conditions a).571 Ohmiya et al. used thiazolium salt as an NHC precursor for the three-component arylacylation of styrenes (Scheme 87a, conditions b).572 The catalytic cycle began with the in situ formation of the Breslow intermediate (−1.93 V vs SCE), which donates a single electron to aryl iodide to generate an aryl radical and the Breslow intermediate-derived ketyl radical (Scheme 87b). After the addition of the aryl radical to styrene to generate the reactive alkyl radical, intermolecular acyl transfer from the ketyl radical intermediate proceeds to afford the final product and regenerates the NHC catalyst.

Scheme 87. Mesoionic Carbene Catalyzed Acylarylation of Alkenes.

Scheme 87

The unique characteristics of phenalenyl-based molecules and in particular its more reactive doubly reduced anionic redox state prompted Mandal et al. to design efficient catalytic processes for the activation of aryl iodide during Mizoroki-Heck-type cross-coupling with styrenes, C–H arylation of arenes, and borylation under solvent-free ball-milling conditions.502,503,573,574 Metallic potassium was used to tune the redox state of the catalyst and generate the reactive anion species (Ered = −1.82 in acetonitrile), and was key to the activation of the Mizoroki-Heck-type cross-coupling between aryl iodide and styrene (Scheme 88a).574 The proposed catalytic cycle involves the consecutive injection of two electrons to the cationic phenalenyl to generate the anion, which reduced aryl iodide via the SET process to afford the aryl radical (Scheme 88b). The aryl radical was trapped by styrene and after deprotonation the radical anion intermediate was generated; the radical anion intermediate was reduced via the SET process to form the stilbene product regenerate the catalyst.

Scheme 88. Phenalenyl-Anion Catalyzed Heck-Type Cross-Coupling of Aryl Iodide with Styrene.

Scheme 88

The same group developed a unique Buchwald–Hartwig-type cross-coupling of aryl halides with amine using a phenalenyl-based organocatalyst (Scheme 89a).575 The bispyridinium salt and phenalenyl ligand generate the phenalenyl anion, which is responsible for performing this radical-mediated C–N cross-coupling at room temperature. Interestingly, (hetero)aryl dihalides and primary aliphatic/aromatic amines reacted with high selectivity to afford the related monofunctionalized products. The generated aryl radical binds with the amine via a concerted deprotonation and C–N bond formation to generate the radical anion, which is oxidized by the phenalenyl radical to afford the final aryl amine and the phenalenyl anion.

Scheme 89. Reduced Phenalenyl-Catalyzed Buchwald-Hartwig-Type Cross-Coupling of Aryl Halide with Amine.

Scheme 89

Furthermore, phenalenyl C was used for the catalytic hydrodehalogenation and dehalogenative deuteration of a broad array of aryl/heteroaryl halides under ambient conditions (Scheme 90a).502,503,572,576 The combination of the precatalyst, catalytic metallic potassium, and KOtBu is essential to enable the desired transformation in high yields. The doubly reduced phenalenyl anion (E21/2 = −1.88 V) undergoes the thermodynamically favorable SET with aryl iodide to generate the aryl radical, which accepts a hydrogen atom from DMSO to afford the dehalogenated products and the DMSO-radical (Scheme 90a). The thus-generated radical reacts with KOtBu to generate the radical anion, which serves as the reductant to regenerate phenalenyl anion.

Scheme 90. Reduced Phenylenyl Catalyzed Transformation of Ar-X to Ar-H/D.

Scheme 90

Borylation of aryl iodides by using B2(OR)2 has been achieved in the presence of Cs2CO3 under heating conditions.577 The combination of KOtBu and 1,10-phenanthroline KOtBu also enables this reaction (Scheme 91a, conditions a).578 The organic promoter N,N′-diboryl-4,4′-bipyridinylidene was also utilized with KOMe to activate the aryl halide (Scheme 91a, conditions b).579 The reaction mechanism is initiated by the SET process using 1,10-phenanthroline and KOtBu or diboryl-bipyridinylidene to generate the aryl radical, which reacts with the in situ generated boronate anion to afford the aryl boronic acid pinacol ester (Scheme 91b).

Scheme 91. Dehalogenative Borylation Aryl Halides Using 1,10-Phenanthroline or Diboryl-bipyridinylidene Promoter.

Scheme 91

Jiao et al. developed efficient catalytic conditions using 4-phenylpyridine to activate the cross-coupling of haloarenes with diboron reagents and synthesis of arylboronates (Scheme 92a).580584 The combination of diboron with the pyridine derivative in the presence of an appropriate base generates the boryl-pyridine complexes, which are well-known potent super electron donors (SEDs), that can be used effectively to reduce aryl halides under thermal and visible light conditions (Scheme 92b).

Scheme 92. Dehalogenative Borylation Aryl Halides Utilizing Boryl-pyridine Complex.

Scheme 92

In addition to its use in borylation, the versatility of boryl-pyridine complexes was expanded to induce cross-coupling of aryl halides with thiols (Scheme 93a).585 A broad array of (hetero)aryl iodides and aliphatic/aromatic thiols reacted well, and a wide range of functional groups was tolerated. The aryl radical generated by the boryl-pyridine complex reacts with the thio alkoxide to afford the thio ether radical anion, which undergoes the SET process to afford the corresponding thioether (Scheme 93b).

Scheme 93. Pyridine-Boryl SEDs Promoting Cross-coupling of Aryl Halides with Thiols.

Scheme 93

3.2. Organocatalyst-Free Dissociation of Ar–I Bond

As mentioned earlier, Itami et al. reported the first transition metal-free arylation of electron-deficient N-heterocyclic compounds (used as solvent) with aryl halides and the synthesis of heterobiaryl products using KOtBu as the sole promoter under microwave conditions (Scheme 94a).494,495 Charette et al. reported an efficient intramolecular arylation in the presence of KOtBu in pyridine under microwave conditions that did not require additional organocatalysts (Scheme 94b).586 Furthermore, Wilden et al. observed that organic promoters are not essential for cross-coupling of unactivated benzene with aryl iodides as the reaction can be activated with KOtBu alone (Scheme 94c).587,588 The reaction proceeded via the SET to initiate the generation of the aryl radical; however, the exact mechanism of this step was not clear.

Scheme 94. Organocatalyst-Free C–H Arylation of Arenes with Aryl Iodides.

Scheme 94

Bisai et al. noted the possibility of performing the intramolecular BHAS reaction under organocatalyst and organocatalyst-free conditions in the presence of KOtBu using mesitylene as a solvent.589,590 Murphy et al. studied in-depth the above-mentioned reactions, in which pyridine was used as a substrate and/or solvent with KOtBu as the only promoter.497,591,592 As butoxide is not capable of directly initiating the radical reaction, the formed radicals must be generated in another way. Thus, after extensive computational and experimental studies, Murphy et al. observed the generation of super electron donor species (Scheme 94d) which are assumed to be responsible for the reduction of the aryl halides in Schemes 94a and 94b.

Additionally, for the coupling reaction between iodobenzenes and benzene using KOtBu as the sole additive, Murphy et al. proposed a different mechanism from that suggested by Wilden. The researchers assumed a more sluggish activation mode can exist in the absence of the organic additive (Scheme 94e).497,591,592 Intensive experimental and theoretical studies unambiguously indicated the formation of benzyne which can be used for activating the process. In the presence of KOtBu, aryl iodide can in situ generate benzyne in very low concentrations. Benzyne can work as a diradical, reacting with benzene to form distal diradicals, which serve as super electron donor species.

A combination of hexamethyldisilazane (HMDS) and tetramethylammonium fluoride (TMAF) was used by Kondo et al. to in situ generate an amide base to promote the cross-coupling of aryl iodide with C–H heteroarenes (Scheme 95a).593 Electronically and sterically diverse (hetero)aryl iodides reacted effectively with pyrazine to afford the corresponding heterobiaryls. Other C–H heteroarenes such as quinoxaline, pyridazine and thiophene afforded mixtures of regioisomers in moderate yields. The reaction mechanism involves the HMDS reacting with TMAF to generate the [(TMS)HN Me4N+] base, which transfers a single electron to aryl iodide, thereby promoting the following reaction (Scheme 95b).

Scheme 95. Amide Base Generated In Situ to Promote Csp2-H Arylation with Aryl Iodide.

Scheme 95

Electron-deficient aryl iodides were reacted with arenes in DMSO in the presence of tetramethylammonium fluoride or Cs2CO3 to afford the corresponding cross-coupling products (Scheme 96).594,595 Under these conditions, these additives participated as promoters to give a single electron to the deactivated aryl iodides and generate aryl radicals for cross-coupling with arenes.

Scheme 96. Base-Promoted Synthesis of Biaryls from Aryl Iodides.

Scheme 96

Fukuoka et al. developed a radical initiator-free protocol for the carbonylation of aryl iodides with CO in the presence of aryloxide, base/H2O or aryl carboxylic acid salt for the synthesis of aryl acid derivatives via a probable SET initiated radical mechanism (Scheme 97a).596 Similarly, Ryu et al. performed the intermolecular Heck aminocarbonylation of aryl iodides597 and intramolecular Heck carbonylation of 2-iodobenzyl alcohols and 2-iodobenzyl amines with CO for the synthesis of benzolactone and benzolactam derivatives, respectively (Scheme 97b).598,599 The scope of the alcohol substrate encompasses primary, secondary, and tertiary 2-iodobenzylalcohols in addition to 2-(2-iodophenyl)ethanol and 3-(2-iodophenyl)-1-propanol for the synthesis of 5–7 membered benzolactones. In reactions with 2-iodobenzylamine, DABCO was replaced Et3N to avoid the formation of N-ethyl side products. The postulated reaction mechanism starts with the SET from Et3N to aryl iodide to generate the aryl radical, which reacts with CO to afford the acyl radical. The following intramolecular nucleophilic addition by the OH group generates the cyclized intermediate and the SET process affords the final product.

Scheme 97. Organocatalyst-Free Carbonylation of Aryl Iodides.

Scheme 97

Han et al. demonstrated a transition metal-free carbonylative coupling of aryl iodide with aryl trifluoroborate salt under CO using simple basic conditions (Scheme 98a).600 In addition, the same group used CHCl3 or N-formylsaccharin as the CO surrogates (Scheme 98b).601,602 The high reaction temperature and base are responsible for the dissociation of the aryl-to-iodide bond and generation of the aryl radical which is trapped by the generated CO from the reaction of CHCl3 or N-formylsaccharin with base (Scheme 98c). The obtained acyl radical reacts with aryl boronic acid in the presence of base to generate the biaryl ketone radical anion, which undergoes the SET process to aryl iodide to afford the biaryl ketone.

Scheme 98. Carbonylative Coupling of Aryl Iodide and Aryl Boronic Acid with CO.

Scheme 98

Taillefer et al. discovered the capability of additive-free KOtBu/DMF to initiate the SRN1 coupling reaction between aryl halides and enolizable aryl ketones and the synthesis of C(sp3)-arylation products (Scheme 99a).603 The combination of KOtBu/DMF appears to be crucial as no reaction occur if one of these reagents is replaced. A radical chain mechanism was postulated based on the performed theoretical and experimental studies (Scheme 99b). Deprotonation of DMF by KOtBu generates the corresponding carbamoyl anion, which reacts with DMF to afford the dianion. These anion species can serve as single electron donors to aryl iodide.591 The thus-generated aryl radical reacts with enolate to afford the aryl ketone radical anion, which exchanges an electron with aryl iodides to afford the final product.

Scheme 99. KOtBu/DMF Initiated α-Arylation of the Enolizable Aryl Ketone by Aryl Iodide.

Scheme 99

The halogen bond assisted dissociation of the aryl-iodide bond was found to be an efficient strategy to construct heterocycles under mild conditions. Thus, ortho-iodothioanilide in the presence of KOtBu and 1,10-phenanthroline (additive) at room temperature underwent intramolecular C–S cross-coupling to afford a broad variety of 2-substituted benzothiazoles (Scheme 100a).604 The control experiment revealed that the free-NH of the thioanilide is key to this type of cyclization. XRD, DFT, NMR and UV studies indicated the presence of intermolecular (I···S) halogen bonding, that was responsible for activating the aryl-iodide bond via the SET process (Scheme 100b). A rational mechanism was initiated by deprotonating the NH-group to afford the corresponding anion, which interacted with the iodine atom of another molecule via halogen bonding to generate the XB complex. Aryl-iodide bond activation through the SET from the thiolate anion generated the aryl radical. Deprotonation by added base followed by intramolecular cyclization produced the benzothiazole radical anion, which underwent the SET process with another substrate to afford the final product.

Scheme 100. Halogen-Bond Assisted Ar–I Bond Dissociation and Intramolecular Cyclization.

Scheme 100

Walsh et al. reported that the 2-azaallyl anion generated from ketimine and strong base serves as a super electron donor for aryl iodides. The SET process proceeds to generate aryl and 2-azaallyl radicals, which undergo radical coupling to form C–C bonds (Scheme 101a).605607 The same group combined this reaction with an intramolecular cyclization process to develop the constructions of various heteroaromatics bearing ethylamine moieties (Scheme 101b).608611 Iodoarenes bearing an allenyl moiety undergo the SET process induced by the 2-azaallyl anion to generate the corresponding aryl radical. Intramolecular cyclization followed by radical coupling with the 2-azaallyl radical affords the final products. This strategy was applied to the construction of benzofuran, isochromen, isoquinoline, and indole derivatives bearing an ethyl amine moiety (Scheme 101c).

Scheme 101. 2-Azallyl Anion Initiated Radical Cyclization and Tandem Formation of Heterocyclic Skeletons.

Scheme 101

James et al. discovered that oxime anions serve as dual SRN1 initiators and hydroxide surrogates in the dehalogenative hydroxylation of aryl halides to afford phenols under mild conditions (Scheme 102a).612 The SRN1 chain mechanism was supported by further experimental and computational studies. Either the oxime anion or charge-transfer (CT) complex could transfer a single electron to aryl halide under heat or light stimulation (Scheme 102b). The generated aryl radical interacts with the oxime anion via the formation of a weak 2-center-3-electron σ-bonded intermediate, which undergoes intramolecular SET into the adjacent π* orbitals to afford the delocalized radical anion (E1/2 = −2.14 vs SCE). Electron exchange with another aryl halide generates the arylate oxime, which transforms to the corresponding phenol under basic conditions.

Scheme 102. Oxime as Super Electron Donor and Hydroxide Equivalent in Dehalogenative Hydroxylation of Aryl Halides.

Scheme 102

Wang et al. discovered rongalite (sodium hydroxymethylsulfinate) as a novel precursor of the super electron donor sulfoxylate anion radical (SO2•–) species which initiates homolysis of aryl-halide bonds and generates the aryl radical for the arylation of (hetero)arenes, thiolates, sulfides, sulfinates, phosphines, phosphites, etc., through the BHAS or SRN1 mechanism (Scheme 103a).613 The radical chain cycle begins with the generation of SO2•– species from rongalite or Na2S2O4 under thermal conditions (Scheme 103b). SO2•– transfers a single electron to the aryl halide to generate the aryl radical, which reacts with other arenes or nucleophiles to afford the corresponding radical anion intermediates. The SET process proceeds to afford the coupling products. Laha et al. demonstrated the intramolecular arylation of 2-halobenzenesulfonamide and generated biarylsultams induced by the sulfoxylate anion radical (Scheme 103c).614

Scheme 103. Sulfoxylate Anion Radical Effectively Induces the Generation of the Aryl Radical from Aryl Halides.

Scheme 103

3.3. Dissociation of Ar–I to Aryl Anion

In addition to the dissociation of the Ar–I bond to generate the aryl radical; the bond can be cleaved to generate the aryl carbanion. KOMe was applied to the transition metal-free borylation of aryl halides in the presence of silylborane via the in situ generation of aryl carbanions, under ambient reaction conditions (Scheme 104a).615 Combining with hexamethyldisilane (TMS-TMS) enables dehalogenative deuteration in CDCN solvent (Scheme 104b).616 Additionally, KOMe/TMS-TMS reagents were involved in the dehalogenative formylation of aryl iodides with DMF to afford the corresponding aryl aldehydes (Scheme 104c).617 A plausible mechanism was suggested based on control experiments and extensive NMR studies. The methoxy anion coordinates the two silyl groups of TMS-TMS to generate hypercoordinate silane species as a strong base, which reacts with Ar–I to generate the aryl carbanion. The addition of Mg(OTf)2 improved the yield slightly via coordination/activation of the carbonyl group of DMF for the reaction with the aryl carbanion nucleophile.

Scheme 104. KOMe-Induced Dehalogenative Functionalization of Aryl Iodides.

Scheme 104

Baik and Cho et al. reported the borylation of aryl iodides using commercially available 1,1-bis[(pinacolato)boryl]methane as a potent precursor to activate the Ar–I bond and as a boron source in the presence of NaOtBu (Scheme 105a).618 Extensive theoretical and experimental studies supported the postulated mechanism in Scheme 105b. The reaction of pinB-CH2–Bpin with tert-butoxide anion forms the tert-butoxyboronate ester and borylmethyl anion. The latter and aryl iodide form a Lewis acid/base complex, that produces the sodium aryl anion. The generated aryl anion attacks the Lewis acidic boron center to form the borate species, which collapsed to afford the final product.

Scheme 105. Borylation of Aryl Iodide with 1,1-Bis[(pinacolato)boryl]methane.

Scheme 105

Sodium hydride exhibited potency alone or as an iodide composite for the hydrodehalogenation of aryl halides.619,620 Recently, Zhang et al. established the NaH-mediated activation of ortho-diiodoarenes and generated arynes in a controlled manner, that enabled their insertion into the C–C σ-bonds of unactivated ketones and the synthesis of vicinal difunctionalized arenes (Scheme 106).621 DFT calculations revealed the crucial role of the two adjacent iodines in ortho-diiodoarene for the interaction with NaH, which facilitates the generation of the aryl anion followed by that of the aryne. The enolate generated from the ketone reacts with the aryne with high regioselectivity for substrates comprising bulk substituents. The formed anion intermediate undergoes intramolecular cyclization and a Fries-type rearrangement to afford the final product. This strategy is compatible with diverse aryl alkyl and alkyl–alkyl ketones and unsymmetrical ortho-diiodoarenes to afford regioisomeric products with selectivity depending on the type and position of the substituents.

Scheme 106. NaH-Activated Generation of Aryne for Double C–C Bond Formation.

Scheme 106

Analogously, a NaH/o-diiodoarene combination was used for the N-arylation of a broad array of secondary amides and amines in addition to N-heterocycles, hydrazine, urea, lactam, and sulfonamide derivatives (Scheme 107).622 In this strategy, o-diiodoarene acted as an aryne precursor and electrophilic iodine donor.

Scheme 107. NaH-Activated Generation of Aryne for C–N and C–I Bond Formations.

Scheme 107

4. Photoinduced Ar–I Bond Dissociation

Photochemistry affords powerful opportunities to expand the potential of organic chemistry as harnessing the readily accessible light energy instead of the traditional thermal force is an abundant, renewable, and clean energy approach for performing novel molecular transformations and producing unique functional molecules. Therefore, significant progress of light-induced Ar–I bond dissociation has occurred and various transformations can be performed under benign conditions.623 Light-induced aryl-halogen bond dissociation, in the presence/absence of photocatalyst, allows for new green and sustainable avenues in aromatic substitution reactions.375,624643 Upon absorbing light, the exciting components provide intriguing transformations that are relatively complicated or even not possible using other ground state approaches. The direct transformation of aryl halide into the related aryl radical represents a crucial approach to access a reactive synthetic intermediate to use in subsequent chemical transformations. The development of aryl radical-mediated transformations has played an essential role in shaping the development of synthetic radical chemistry.623 Generating aryl radicals from the related aryl halides under photoirradiation conditions can be accomplished via Ar–X bond homolysis, SET, and halogen-atom transfer (XAT) pathways (Scheme 108). The classic photoactivation via homolytic cleavage of the Ar–X bond usually requires irradiation with high-energy UV-light (Scheme 108a). This direct activation approach requires specialized setups and displays low functional group compatibility, which consequently limit its application in organic synthesis. The addition of electron donor species (ED) assists the activation of aryl halides via the formation of an EDA complex (Scheme 108b). Photoirradiation leads to the injection of an electron to the unoccupied orbital of aryl halide, followed by an intramolecular electron transfer from the π-orbital to the antibonding σ* orbital and cleavage of Ar–X bond. The SET to aryl iodide and Ar–I bond cleavage is much more thermodynamically and kinetically, respectively, favorable than that to other aryl halides.644 The XAT-based activation method involves the direct homolysis of an Ar–X bond by abstraction of the halogen atom with a suitable radical species (R), which is generated from alkylamine under photoirradiation conditions. The abstraction step proceeds via a colinear arrangement controlled by several factors (Scheme 108c).645 In this section, we highlight the recent progress in this flourishing area by discussing the recent light-induced arylation with aryl iodide with or without organophotocatalysts, and their mechanisms.

Scheme 108. General Approaches for the Generation of Aryl Radical from Aryl Iodide.

Scheme 108

4.1. Photocatalyst-Free Dissociation of Ar–I Bond

The photocatalyst-free light-induced Ar–I bond cleavage and construction of the C–C and C–heteroatom bonds via an initiation step of either direct excitation of an aryl halide or the SET process, is a highly desired transformation as it provides an operationally simple, inexpensive, and environmentally benign transformation under mild conditions with high functional group tolerance. The SET process could be possible if one of the reaction components can absorb light or when the reactants form photoactive EDA complexes, comprising an electron acceptor aryl iodide and electron donor compound (Scheme 109).646657 The UV–vis absorption spectra of the EDA complex have a new broad absorption peak called the CT band, that has shifted to longer wavelengths. Photoirradiation at the wavelength of the CT band generates the excited EDA complex, which undergoes intramolecular SET from the donor to the Ar–I to generate a pair of radical ions trapped in the solvent cage. Diffusion radical ions from the solvent cage affords the aryl iodide radical anion which collapses to generate the reactive aryl radical. The presence of the iodide leaving group causes the irreversible C–I bond cleavage to occur faster than the competing back electron transfer reverse process within the radical ion pair. Iodoarene-based EDA complex photochemistry has been noted as a robust tactic for broadening the potential of light-induced aryl radical synthetic chemistry. Thus, light could serve as an essential and unique energy source to cleave the Ar–I bonds, and chemical transformations that are impossible or difficult to conduct under alternative conditions become feasible.

Scheme 109. EDA Complexation and SET for the Generation of Aryl Radical from Aryl Iodide.

Scheme 109

4.1.1. C–Chalcogen Bond Formation

The feasibility of Ar–halide bond homolysis under UV irradiation persuaded many research groups to hypothesize photoinduced cross-coupling processes for the construction of C–chalcogen bonds.658660 The reaction of an aryl halide with diselenide, disulfide, and ditelluride under the optimized UV-light conditions formed the corresponding selenoether, thioether, and diaryltelluride, respectively (Scheme 110a).658 Aliphatic and aromatic dichalcogenide substrates and various aryl halides bearing OH, NH2, and allyl groups were well-tolerated. Radical trapping experiments and electron spin resonance spectroscopy revealed the generation of a carbon centered free radical during the reaction, thus supporting the reaction mechanism in Scheme 110b. UV irradiation of the reaction mixture led to homolysis of the diselenide to the selenide radical and excitation of the aryl halide to generate an aryl radical, which reacted via radical–radical coupling to afford the aryl selenoether.

Scheme 110. Direct Homolysis of Aryl Halide Bond under UV-Light and the Construction of C–Se, C–S, and C–Te Bonds.

Scheme 110

Laulhé et al. recognized the importance of electron-donor–acceptor (EDA) complex generation during visible-light-induced cross-coupling of an aryl iodide with diaryl dichalcogenides in the presence of KOtBu/DMSO (Scheme 111a).661 The experimental and theoretical studies suggested the responsibility of an EDA complex between the in situ generated dimsyl anion and aryl iodide for the observed chemical reactivity. The EDA complex leads to the formation of the aryl radical via charge transfer; the aryl radical was trapped by the disulfide to afford the final product.

Scheme 111. Photoinduced Charge Transfer from Dimsyl Anion for Ar–I Bond Dissociation.

Scheme 111

Photoinduced aryl-halide bond dissociation along with the in situ liberation of SO2 using commercially available SO2-equivalents provide an attractive combination for the construction of diverse aryl sulfonyl-containing skeletons.662664 Accordingly, Wu et al. envisioned a protocol for the synthesis of 2-(arylsulfonyl)acetonitrile via a one-pot cascade reaction of aryl iodide, SO2-equivalent (DABCO.(SO2)2), and 3-azido-2-methylbut-3-en-2-ol under UV-irradiation conditions (Scheme 112a).664 Mechanistic investigations indicated that UV irradiation of the aryl iodide provides an aryl radical, which undergoes SO2-insertion to afford an arylsulfonyl radical. Reaction of the generated radical with the alkene bearing azide group followed by release of N2 produces a nitrogen radical. Further C–C bond cleavage in the presence of an iodo-radical and DABCO delivers the final product and acetone.

Scheme 112. Photoinduced Synthesis of 2-Sulfonylacetonitriles via Insertion of SO2.

Scheme 112

4.1.2. C–Chalcogen Bond Formation via EDA Complex with Chalcogen Anions

Phenolate and thiolate anions are among various organic anion species whose physicochemical properties can be markedly altered upon photoexcitation.626,665,666 These organic anions (which absorb light in the ground state) behave as strong reducing agents in the excited states and can generate reactive radicals from their stable precursors via the direct photoexcitation and SET processes. An alternative mechanism involved the affinity of electron-rich organic anions (which cannot absorb light) to associate with electron-deficient radical precursor substrates and form ground state EDA complexes (active toward light) which have the tendency to excite with light, undergo SET, and finally generate the reactive radical species.

The photochemistry of phenolate and thiolate anions as dual photocatalysts and reagents or only as photocatalysts is discussed in the section headed “Photoinduced Ar–I bond dissociation”.667,668 We discuss herein the versatility of thiolate anions as photocatalysts and reagents for cross-couplings and the construction of C–S bonds.

Sulfur anions exhibit versatile reactivity for promoting molecular transformations under transition metal-free, operationally simple, and green visible light irradiation conditions.669671 Miyake et al. discovered C–S cross-coupling between aryl halides and aryl thiols under white LED irradiation even in the absence of reducing organic photoredox catalysts (Scheme 113a).672 Spectroscopic and theoretical studies indicated that the electron-poor aryl halide and electron-rich thiolate anion associated via π–π interactions and the formation of the EDA complex. Visible light irradiation of the EDA complex resulted in the generation of the thiyl and aryl radicals via the SET process, which coupled together to afford the final C–S cross-coupling product. This transition metal and photoredox-free pioneering work tolerated diverse functional groups and worked well with aliphatic/aromatic thiols and diverse aryl halides including heterocycles and pharmaceutical ingredients. Analogously, Yue et al. developed a convenient approach to construct the Ar–Se bond via in situ generation of the EDA complex between an aryl halide and aryl selenolate anion under blue light conditions (Scheme 113b).673 The reaction is compatible with a broad range of electron-deficient (hetero)aryl halides. Sekar et al. attempted C–S cross-couplings via halogen-bond assisted generation of EDA or CT complexes,674 followed by SET under visible light irradiation. This strategy was applied to the one-pot synthesis of heteroaryl thioethers via in situ formation of heteroaryl iodide, followed by cross-coupling with aryl/alkyl thiols in the presence of KOtBu/DMSO under visible light conditions.675

Scheme 113. Photoinduced Coupling of Aryl Halides with Thio-/Seleno-phenols via an EDA Complex as the Key Intermediate.

Scheme 113

Xanthate was effectively incorporated as a dual sulfur surrogate and photocatalyst in the photoinduced cross-coupling reaction with 2-iodochalcones to produce thiochroman-4-one derivatives (Scheme 114a).676 Various control reactions, spectroscopic experiments, and DFT calculations were performed to sustain the interaction between the xanthate anion and aryl iodide via the formation of a halogen bond followed by the formation of the EDA complex. Visible light induced the SET from the xanthate anion to Ar–I to generate aryl and thiyl radicals, which underwent radical–radical cross-coupling to generate the xanthate ester. Thioester cleavage in the presence of xanthate anion furnishes thiolate, which undergoes intramolecular Michael addition to afford the final product. The reaction scope was expanded to a three-component reaction of xanthate, 2-iodobenzaldehyde, and chalcone or crotonaldehyde to synthesize the corresponding thiochroman-4-ol or thiochromene, respectively (Scheme 114b).

Scheme 114. Visible Light Induced C–S Bond Formation via a Sequence of Halogen-Bond Interaction, EDA Complexation and Photoinduced SET Process.

Scheme 114

The importance of the EDA complexation of an aryl halide with a thiolate anion for the generation of aryl radical prompted Li et al. to design a modified strategy to synthesize benzothiazoles with high regioselectivity under oxidant-free conditions.677 Direct visible light photolysis of ortho-halothiobenzanilide under basic conditions afforded the related benzothiazole product via intramolecular dehalogenative C–S cyclization (Scheme 115a).678 In addition to CFL light, blue LED and natural sunlight were effective and afforded comparable yields even with gram scale experiments. Photoirradiation of the deprotonated starting material generates the excited state species, which undergoes intramolecular SET from the thiolate anion to the N-aryl ring followed by dehalogenative cyclization to finally afford the cyclization product. The cascade cyclization reaction of 1-iodo-2-isothiocyanatobenzene with 2-isocyanoacetate under visible light photolysis produces the corresponding benzo[d]imidazo[5,1-b]thiazole (Scheme 115b).679 The proposed mechanism begins with [3 + 2] cycloaddition in the presence of base to generate the imidazole intermediate, which isomerizes to the thiolate anion. An internal EDA complex forms between the thiolate anion and aryl moiety, followed by SET under light irradiation to generate the diradical intermediate. The cyclization that follows affords the final product.

Scheme 115. Photoinduced Ar–I Bond Dissociation and Intramolecular Cyclization.

Scheme 115

The UV-light promoted coupling of sodium sulfinate with an aryl halide in the presence of Cs2CO3 was reported by Yan and Zhang as a complementary strategy for the synthesis of sulfones (Scheme 116a).680 The UV–vis study indicated the formation of an EDA complex between the sulfinate anion and Ar–I. Thus, the proposed mechanism began with the formation of the EDA complex, followed by UV irradiation-promoted SET from the sulfinate anion to Ar–X; the complex then collapsed into an aryl and a sulfonyl radical. Subsequent radical–radical coupling afforded the cross-coupling sulfone product. This type of transformation is compatible with electron-rich/poor aryl sulfinate, alkyl sulfinate, and electron-deficient aryl halides; however, electron-rich aryl halides afforded lower yields even after a prolonged reaction time. The success of this protocol encouraged the same group to develop on-DNA radical cross-couplings using visible-light promoted SET within EDA complexes and reverse adsorption to support solid support (RASS) strategies.681 This type of transformation was applied, for example, to the cross-coupling of on-DNA-heteroaryl halides with alkyl/aryl sulfinates and synthesis of DNA-tagged heteroaryl sulfones (Scheme 116b).

Scheme 116. Light-Induced Cross-coupling of Sodium Sulfinate with Aryl Halides and Production of Sulfone.

Scheme 116

Zhang et al. established an efficient method for the visible-light induced trifluoromethylselenolation of aryl halides with [Me4N][SeCF3] under benign conditions (Scheme 117).682 Mechanistic studies suggested the formation of an EDA complex between the electron acceptor aryl halide and electron donor selenium anion. Photoexcitation of the EDA complex led to intracomplex SET to finally afford an aryl and a trifluoromethyl selenium radical, which are rapidly cross-coupled to afford an aryl selenoether product. Alternatively, DABCO can also form the EDA complex and convert to an N-centered radical cation via the SET process, which oxidizes the trifluoromethyl selenium anion to generate the trifluoromethyl selenium radical. The coupling that follows, affords the final product.

Scheme 117. Photoinduced Cross-coupling of Aryl Halide with [Me4N][SeCF3].

Scheme 117

4.1.3. C–C Bond Formation

Chiba et al. developed a process for the cross-coupling of aryl iodides with various alkenes under irradiation with violet light (390 nm) (Scheme 118a).683 The process is relevant for the synthesis of allyl, prenyl, acetomethyl, lactone, and spirocyclic epoxide derivatives of the corresponding arenes. The radical cascade reaction is initiated by the irradiation of Ar–I with violet light to transfer it to the photoexcited state, where it undergoes Ar–I bond homolysis to produce an aryl radical and atomic iodine (I). For the reaction with allyl silane, the irreversible addition of an aryl radical to an alkene affords a β-silyl radical (Scheme 118b). Oxidation generates a β-silyl carbocation, which furnishes the allylation product after removal of the silyl group. A possible alternate pathway is that which occurs via the formation of an alkyl iodide followed by the elimination of the silyl group.

Scheme 118. Photoinduced Cross-coupling of Aryl Iodide with Various Alkenes.

Scheme 118

Parasram et al. used the results obtained by Chiba to design a cascade approach for the synthesis of phenanthrene derivatives via the photoexcitation of aryl iodides in the presence of styrenes (Scheme 119a).684 Unsymmetrical phenanthridines bearing useful functional groups were prepared under the reaction conditions which are applicable to gram scale synthesis. The reaction mechanism was postulated based on UV–vis studies and control experiments. Light-induced homolysis of the aryl iodide generated the aryl radical, which was captured by the styrene double bond to form the corresponding benzylic radical (Scheme 119b). Thereafter, iodination of the radical intermediate occurred followed by elimination of HI to generate a stilbene derivative. Subsequently, the photoinduced Mallory-type cyclization furnished dihydrophenanthrene, which subsequently oxidized in the presence of I2/air to afford the final product.

Scheme 119. Photoinduced Arylation/Cyclization Cascade Generating Phenanthrene Derivatives.

Scheme 119

Xia et al. used the phenolate anion as a strong photoreductant catalyst in the oxyarylation of olefins with aryl halides and TEMPOH under visible light irradiation.685 In addition, the designed phenolate anion photocatalyst was used for the intramolecular dearomative spirocyclization of 4-hydroxyaryl-tethered-2-haloarenes and synthesis of spirocyclohexadienones.686 A photocatalyst-free process was designed using vinylphenolate anions as both the reaction component and redox mediator for the reductive activation of aryl halides under mild conditions (Scheme 120a).687 The Heck-type arylation of vinylphenols with aryl halides under visible light irradiation conditions afforded the corresponding multisubstituted alkenes. A plausible mechanism begins with the excitation of the phenolate anion under light irradiation (Scheme 120b). The redox potential of the excited species (−2.48 V vs SCE) implies it can undergo SET with Ar–X to afford phenoxy and aryl radicals. The generated aryl radical reacts with the vinylphenol to generate a radical anion intermediate, which transfers a hydrogen atom to the phenolate to afford the desired product. Alternatively, the radical anion intermediate undergoes a SET to the other Ar–X to afford the product along with the aryl radical, which propagates the chain process.

Scheme 120. Vinylphenolate as Photoreductant and Heck-Type Reagent in Reaction with Aryl Halides to Afford Multisubstituted Alkenes.

Scheme 120

Wu et al. synthesized sulfonated cyclic compounds by reacting aryl iodides tethering N-methylmethacrylamide with silyl enolate and an SO2 source (DABCO(SO2)2) under ultraviolet irradiation (Scheme 121).688 The reaction mechanism is initiated by UV-induced Ar–I bond cleavage to generate an aryl radical, which undergoes intramolecular 5-exo-cyclization to afford the cyclized alkyl radical. Further sulfonation and subsequent trapping by the silyl enolate, followed by radical–radical coupling with the iodine radical generate an alkyl iodide intermediate, which undergoes desilylation to afford the final product.

Scheme 121. Photoinduced Synthesis of Sulfonated Cyclic Compounds via SO2-Insertion.

Scheme 121

Paixão et al. developed an intramolecular reductive cyclization protocol to convert N-propargyl-2-halobenzenesulfonamides and N-protected 2-halophenylacrylamides into the related N-heterocycle indoles and oxindoles, respectively, in the presence of TMS3Si-H and visible light as an efficient promoter system.689 Further extension was performed for the synthesis of highly substituted indolines and 2,3-dihydrobenzofurans from the related N-allyl-N-(2-iodophenyl)acetamide and 1-allyloxy-2-iodobenzene, respectively (Scheme 122a).690 The reaction mechanism begins with irradiation of the reaction mixture with visible light and generation of a reactive excited complex as an EDA complex or exciplex (Scheme 122b). Either energy or electron transfer from the generated complex leads to homolytic cleavage of the Si–H bond and generation of a silicon radical, which abstracts the iodine from the starting aryl iodide to generate an aryl radical. Intramolecular radical cyclization followed by H-abstraction from TMS3Si-H affords the final product.

Scheme 122. (TMS)3SiH/Visible Light Promoted Intramolecular Reductive Cyclization.

Scheme 122

The carbon dioxide radical anion CO2•– has emerged as a strong single electron donor to reduce recalcitrant aryl halides under photocatalytic conditions.691693 Recently, Hou et al. used their oxygen-format salt-DMSO system for photopromoted oxyarylation and hydroarylation of alkenes with aryl halides (Scheme 123a).694 The hydroarylation pathway required electron-deficient alkenes for reaction with electronically and sterically diverse (hetero)aryl halides under the same conditions. Mechanistic investigations indicated the importance of CO2•– and the EDA complex that forms between the dimsyl anion and aryl halide to generate an aryl radical under blue light irradiation (Scheme 123b). For the reaction with electron-rich styrene, the generated aryl radical is trapped by styrene, which reacts with O2 and subsequently abstracts a hydrogen atom from the formate salt to afford the final product after extrusion of H2O. By contrast, the reaction with an electron-deficient alkene does not involve the reaction with O2 before the HAT process.

Scheme 123. Photoreduction of Aryl Halides with O2/Format Salt/DMSO System.

Scheme 123

Preparation of ketones under transition metal-free conditions by the direct addition of the generated radical species to carbonyl compounds is challenging owing to the formation of the thermodynamically unfavorable alkoxy radical and possibility of competing H-abstraction processes. Li et al. successfully designed a photochemical protocol for the synthesis of aryl ketones by the direct coupling of aryl halides with dicarbonyl compounds under UV-light/base conditions (Scheme 124a).695 Various aryl halides and symmetrical diketones were used to deliver the related aryl ketone; however, the incorporation of 2,3-pentanedione as an unsymmetrical diketone afforded a mixture of aryl ketones. Experimental studies indicated the importance of UV-light and using N-methylpiperidine as a base to generate the aryl radical via the SET process (Scheme 124b). Addition of the aryl radical to the diketone produced the corresponding alkoxy radical, which eliminated the radical to afford the final aryl ketone.

Scheme 124. Photopromoted Cross-coupling of Aryl Halides with Diketones and the Formation of Aryl Ketones.

Scheme 124

The same group established the photoinduced cross-coupling of hydrazones with aryl iodides to form diarylmethanes under benign conditions (Scheme 125a).696 Mechanistic and DFT studies supported the formation of an EDA complex via π–π interactions between the deprotonated hydrazone and aryl iodide (Scheme 125b). The photoinduced SET process generated aryl and benzylic radicals, which undergo radical–radical cross-coupling to generate the benzhydryldiazene intermediate. Subsequent deprotonation liberates N2, followed by protonation by solvent/H2O to afford the final product.

Scheme 125. Light-Promoted Coupling of Aryl Iodide with Hydrazone and the Formation of Diarylmethane.

Scheme 125

Xia et al. designed a photochemical approach for the C-3 arylation of oxindole with aryl halides via the in situ generation of oxindole enolate as a strong electron donor (Scheme 126a).697 Mechanistic studies revealed the contribution of oxindole enolate as an electron donor in EDA complex formation; furthermore, the quantum yield of the process (Φ = 11.1) indicated the possibility of radical chain propagation (Scheme 126b). Thus, the reaction begins with the formation of an EDA complex, which, under visible light irradiation promotes the SET process to generate an oxindole and aryl radical. Trapping the generated aryl radical by the enolate affords a ketyl radical, which is a strong reductant (Epred = −2.31 V vs Ag/Ag+ in MeCN) and transfers a single electron to the aryl halide to generate the final product and propagates the radical chain process.

Scheme 126. Oxindole Enolate Induced C-3 Arylation of Oxindole with Aryl Halide via EDA Complex Formation.

Scheme 126

Jiang and Karchava independently used the EDA complex photoactivation strategy for Csp3-arylation of activated methylene compounds with aryl halides under photoredox conditions.698,699 KOH/DMSO/blue LED conditions were designed by Jiang et al. to activate cross-coupling of various β-keto esters with electron-deficient aryl halides and the in situ deacetylation reaction for the synthesis of α-aryl ester derivatives (Scheme 127a). Milder reaction conditions were devised by Karchava to tolerate a wider substrate scope of activated methylene compounds, electron-deficient aryl halides, and heteroaryl halides (Scheme 127b). Mechanistic investigations supported the aggregation of the deprotonated active methylene and aryl halide via EDA complex formation and generation of aryl and alkyl radicals under visible light irradiation.

Scheme 127. Photoinduced Arylation of Activated Methylene Compounds with Aryl Halide.

Scheme 127

The cross-coupling reaction of electron-deficient haloarenes with electron-rich (hetero)arenes under solely violet light irradiation was studied by Yeow and Chiba to understand the crucial role played by photoinduced SET from electron-rich (hetero)arenes to photoexcited aryl halides for the formation of the desired biaryl products.700 Additionally, Fang et al. studied the interaction between aryl halides and a Lewis base (Et3N) during the activation of aryl halide bond dissociation under visible light irradiation (Scheme 128a).701 Base and light are essential for this type of activation and the generated aryl radical was confirmed by trapping with TEMPO radical trapping reagent. This protocol was applied to the C–H arylation of pyrroles with various deactivated aryl halides and radical-initiated polymerization. Bhalla et al. developed stepwise and one-pot protocols to synthesize biaryls and diarylalkynes via a sequence of UV-light mediated aromatic Finkelstein reactions to convert aryl bromides to aryl iodides followed by UV-light/KOtBu promoted cross-coupling with unactivated arenes and terminal alkynes, respectively (Scheme 128b).702 The combination of UV-light and KOtBu is essential to activate aryl iodide. These reactions tolerate electronically different aryl halides and C–H arenes in addition to phenyl/pyridyl-acetylenes to afford the target products with complete chemo- and regio-selectivity. The cross-coupling mechanism initiated by the photoirradiation of aryl iodides in the presence of KOtBu, which induced the SET to the aryl iodide and generation of the reactive aryl radical for coupling with arenes or alkynes.

Scheme 128. Photoinduced C–H Arylation of Heteroarenes and Terminal Alkynes.

Scheme 128

Sekar et al. extended their interest in halogen-bond activated aryl halides under visible light irradiation and designed a regioselective C–H arylation of 2-arylimidazo-[1,2-a]pyridines with aryl iodides (Scheme 129a).703 The proposed mechanism begins with the generation of aryl radicals induced by KOtBu under light irradiation; the aryl radical couples with imidazopyridine to generate the radical intermediate (Scheme 129b). Deprotonation by KOtBu and the subsequent SET process with aryl iodide affords the final product and the aryl radical.

Scheme 129. Halogen Bond Promoted Aryl Iodide for C-3 Arylation of Imidazopyridines.

Scheme 129

A combination of KOtBu, DMSO, and light suffices to promote the intramolecular C–H arylation of (2-halobenzyl) phenyl ether and (2-halophenyl) benzyl ether and formation of the related benzo[c]chromenes and dibenzo[c,f]chromenes, respectively (Scheme 130a).704 UV–vis studies and DFT calculations proposed the in situ formation of the EDA complex (Scheme 130b). The reaction of DMSO with KOtBu generates the dimsyl anion, which forms an EDA complex with the substrate. Photoirradiation leads to an excited state with a charge transfer character to generate the aryl radical, which undergoes 6-endo-cyclization followed by deprotonation and the subsequent SET to afford the final product.

Scheme 130. Intramolecular Homolytic Aromatic Substitution under Blue Light Irradiation.

Scheme 130

Chen et al. photoactivated ortho-(5-Csp3-H)-tethered aryl halides via the in situ formation of EDA complexes for cascade reactions and the formation of various heterocyclic skeletons (Scheme 131).705,706 Various ortho-anilide aryl halides and ortho-halobenzamides were reacted under two different photochemical conditions for the formation of oxindole, isoindolinone, and phenanthridinone derivatives. Experimental and theoretical studies supported three weak interactions that occurred via the participation of DBU or KOtBu, which serves as the electron donor for the complexation of aryl halides and formation of photoactive EDA complexes. Blue light irradiation of these complexes promoted the SET and generation of aryl radicals which underwent 1,5-HAT followed by radical cyclization to finally afford the desired product.

Scheme 131. Base and Solvent Anion Activated Ar–X Bond Photodissociation and Formation of Heterocycles.

Scheme 131

4.1.4. Hydration and Hydroxylation of Aryl Halides

Halogenated aromatic compounds are persistent organic pollutants as their scaffolds are hydrophobic and nonbiodegradable and consequently accumulate in the environment and living organisms.644,707 Thus, these highly carcinogenic compounds cause severe environmental and health problems. Replacing C–halogen with C–H or any other benign group is an important organic transformation in organic synthesis as it provides a pathway for detoxification/degradation of environmentally hazardous organic halides. Hydrodehalogenation and dehalogenative deuteration of aryl halides are desired processes to modify the properties of the candidate pharmaceutical and agrochemical compounds; the activity and metabolic stability can be improved, and the toxicity may be reduced.708712 Interestingly, Rossi et al. used a combination of visible light, KOtBu, and DMSO solvent for the hydrodehalogenation of aryl halides under benign conditions (Scheme 132a).713 Alternatively, Lanterna and Scaiano et al. reported the importance of halogen bond interaction between electron-deficient haloarenes and the methoxide base for the photodissociation of the Ar–X bond and generation of the aryl radical; subsequently, the C–H arene was produced under mild UVA irradiation conditions (Scheme 132b).714 Ryu and Wu et al. used Et3N as the electron- and hydrogen-donor to reduce various aryl halides to their related arenes under UV-light or thermal conditions (Scheme 132).715 By using a flow photomicroreactor with DABCO (1.5 equiv) as a base, the hydrodehalogenation reaction was complete after 20 min and afforded the related C–H product in quantitative yield.

Scheme 132. Light/Additive-Induced Hydrodehalogenation of Aryl Halides.

Scheme 132

The Meerwein-Ponndorf-Verley (MPV) reduction of ketones to the related alcohols via hydrogen atom transfer from isopropanol motivated Zeng et al. to design a hydrodehalogenation process for aryl halides via the formation of a 6-membered ring transition state between the excited aryl halide and isopropanol (Scheme 133).716 The reaction of the aryl halide with isopropanol (as reducing reagent and solvent) under UV-light irradiation afforded the corresponding arene. Mechanistic studies and DFT calculations indicated (i) an inappropriate free radical mechanism, (ii) HAT from the C2 of isopropanol and (iii) an intramolecular pathway involving a simultaneous HAT and Ar–X bond cleavage via a 6-membered transition state.

Scheme 133. MPV-Type Reduction of Aryl Halides with Isopropanol under UV-Light.

Scheme 133

Yao et al. developed a direct reduction process for the hydrodefunctionalization and/or defunctionalization deuteration of functionalized arenes (Scheme 134a).717 Particularly, aryl iodides were subjected to deiodination deuteration under UV-light in the presence of CD3OD/Na2CO3 and formed deuterated arenes. The proposed mechanism involves the HAT process from the deuterated solvent to the aryl radical, which generated the product. Similarly, Qu and Kang used visible light/KOtBu/DMF as a successful combination for hydrodehalogenation (Scheme 134b).718 DMF served as a hydrogen source as indicated by performing a control experiment using DMF-d7 as a solvent to afford the related deuterated product. The rational mechanism begins with deprotonated DMF transferring a single electron to Ar–X under visible light irradiation to generate the aryl radical, which abstracts a hydrogen atom from DMF to afford the product and the DMF radical. This radical species undergoes deprotonation by KOtBu followed by the single electron reduction of the aryl halide to furnish the carbamate. Hou and Li et al. reported a hydrodehalogenation that uses the oxygen-format salt-DMSO system (Scheme 134c).694 The reaction involves the reduction of an aryl halide by CO2•–, which is generated via the HAT process between formate and the dimsyl radical. The thus-generated aryl radical abstracts a hydrogen atom from formate to afford the final product. Alternatively, the aryl radical can also be generated via the EDA complex between the dimsyl anion and aryl halide.

Scheme 134. Hydrogenation/Deuteration of Aryl Halides via Photoinduced Ar–X Bond Cleavage.

Scheme 134

Gevorgyan et al. used the photoinduced hydrodehalogenation reaction for an α-C–H borylation involving the 1,5-HAT process (Scheme 135).719,720 The reaction of an amine derivative bearing a 2-iodobenzoly group with bis(catecholato)diboron afforded the corresponding α-borylated amine derivative. The photoinduced generation of the aryl radical followed by 1,5-HAT generates an α-aminoalkyl radical, which after being trapped by the boronate ester afforded the catecholboronate. Transesterification with pinacol delivered the desired pinacol boronate product.

Scheme 135. α-Csp3-H Borylation of N-(2-Iodobenzoyl)amine via In Situ Generation of Radical Translocating Group.

Scheme 135

The viability of the homolytic cleavage of the Ar–X bond under ultraviolet irradiation enabled Liu et al. to design a new strategy for the hydroxylation of aryl halides via in situ trapping of the generated aryl radicals by molecular oxygen (Scheme 136).721 The addition of a catalytic amount of NaI enabled the recalcitrant aryl bromide/chloride substrates to react effectively under the UV conditions.722,723 Aryl halides with diverse functional groups and bioactive moieties were amenable to reacting under the irradiation conditions. The postulated mechanism begins with the homolysis of the Ar–X bond under UV irradiation to generate the aryl radical.721 For aryl bromide/chloride, the liberated Br/Cl are reduced with the I of NaI. Trapping the generated aryl radical by O2 and subsequent reaction with Et3N affords the phenol derivative. The thus-generated oxygenated triethyl amine reduces the iodine radical to afford acetamide and the iodide anion.

Scheme 136. Hydroxylation of Aryl Halides with O2 under UV Irradiation.

Scheme 136

4.1.5. Borylation and Phosphinylation of Aryl Halides

Larionov et al. achieved the additive-free UV light–induced borylation of aryl halides under batch and continuous-flow conditions (Scheme 137a).724 Electronically diverse aryl halides were coupled with tetrahydroxydiboron and other diboron regents on a gram scale to afford the corresponding arylboronic acid, esters, and trifluoroborates (formed by treating arylboronic acids with KHF2). The haloarene reactivity increased with the decreasing dissociation energy of the Ar–X bond, which could be used to realize chemoselectivity. When the solvent was changed to isopropanol or HFIP, C–X/C–H diborylation was observed, and 1,2- and 1,3-diborylation could be regioselectively achieved by adjusting the solvent and haloarene substituents (Scheme 137b).725 1,2-Diborylation was favored by isopropanol and para electron-withdrawing and meta alkyl groups or halogen atoms, whereas 1,3-diborylation was favored by HFIP and ortho alkyl groups/halogen atoms or meta electron-withdrawing groups. Scheme 137c illustrates the 1,2- and 1,3-diborylation mechanisms. The photoinduced homolytic cleavage of an Ar–X bond generates an aryl radical, which reacts with the diboron reagent to form an arylboronate and a boryl radical. The cage effect of isopropanol facilitates the subsequent addition reaction, which regioselectively yields a diborylated radical because of the stabilizing effect of the conjugated boryl group. Alternatively, the photoinduced heterolytic cleavage of an Ar–X bond in the presence of HFIP340 produces a triplet aryl cation, which reacts with an activated diboron anion to afford a borylated aryl radical cation. The recombination of these species regioselectively affords a 1,3-diborylated intermediate rather than other cationic intermediates, as the latter are destabilized by the π-accepting boryl group. Notably, para electron-deficient substituents favored 1,2-diborylation even in HFIP.

Scheme 137. Additive-Free Photoinduced Borylation.

Scheme 137

Li et al. achieved the photoinduced borylation of aryl halides in the presence of N,N,N′,N′-tetramethyldiaminomethane (TMDAM) under batch and continuous-flow conditions (Scheme 138a).726,727 The combination of UV light and TMDAM was important for achieving high yields, with lower yields observed in the absence of either factor. According to the proposed mechanism, the UV irradiation–induced homolysis of the [Ar–I]* bond or reduction of the excited aryl iodide with TMDAM via SET generates an aryl radical. Concomitantly, TMDAM promotes the addition of water to B2pin2 to form a hydroxyborate intermediate, which reacts with the aryl radical to afford the desired product. The generated boryl radical anion is oxidized via SET to give a hydroxyborane. Studer et al. developed a mild protocol for photoinduced borylation with bis(catecholato)diboron (B2cat2) in the presence of DMF (Scheme 138b).728,729 In this scenario, the generated aryl radical reacts with B2cat2 to produce an aryl–boron bond, which is activated by DMF to afford an intermediate with a weak B–B one-electron σ-bond. The subsequent collapse affords the desired boronate, which is unstable and undergoes in situ transesterification with pinacol in the presence of Et3N to give the required pinacol boronate.

Scheme 138. Additive- or Solvent-Assisted Photoinduced Borylation of Aryl Halides.

Scheme 138

Itoh et al. achieved photoinduced borylation by activating Ar–X bonds through halogen bonding.730 The reaction of haloarenes with B2pin2 in the presence of 2-naphthol (2-NpOH) as a halogen bond acceptor and K2CO3 under visible-light irradiation produced arylboronic esters (Scheme 139). According to the results of control experiments and spectroscopic studies, halogen bonding between Ar–X and naphthoxide affords an EDA complex, which is excited upon photoirradiation to generate an aryl radical via SET. The subsequent iodide elimination furnishes an aryl radical that reacts with the activated borate to form an arylboronic ester.

Scheme 139. Photoinduced Borylation of Aryl Halides Involving EDA Complex Formation.

Scheme 139

Yu et al. phosphinylated heteroaryl halides under visible-light irradiation in the presence of diarylphosphine oxides (Scheme 140a).731 Zeng et al. developed a more practical protocol with a broad scope of substrates, including electronically and sterically diverse (hetero)aryl halides and H-phosphonates (dialkylphosphonates and diarylphosphine oxides) (Scheme 140b).732 According to the proposed mechanism, Ar–X bond photolysis generates aryl and halogen radicals (path a). The phosphonate anion is oxidized by the halogen radical via SET to generate a phosphonate radical, which couple with the aryl radical to afford the desired product. Alternatively, the aryl and phosphonate radicals can be produced by the photoexcitation of the Ar–X bond via a five-membered-ring transition state followed by intramolecular SET (path b). Hassan et al. synthesized (hetero)arylphosphonates via a UV light–induced photo-Arbuzov reaction between (hetero)aryl halides and trimethylphosphite under additive-free conditions (Scheme 140c).733 This reaction proceeds via the photoinduced generation of an aryl radical and its subsequent trapping by trimethylphosphite, with the subsequent methyl radical elimination affording the desired product.

Scheme 140. Photoinduced Phosphinylation/Phosphonylation Aryl Halides.

Scheme 140

After the development of the dimsyl anion as an efficient electron donor for the formation of EDA complexes with electron-accepting aryl iodides and activation of Ar–I bond dissociation under visible light,661,704,705 Laulhé et al. revealed that the solvent anions of DMF and MeCN can photoactivate the borylation and phosphonation of aryl iodides through the formation of EDA complexes (Scheme 141).734 A broad range of aryl iodides could be reacted with either B2pin2 or P(OEt)3 to afford arylboronates or arylphosphonates, respectively. Mechanistic studies proposed the formation of solvent anion–aryl iodide EDA complexes, which generate aryl radicals through photoexcitation followed by SET. The subsequent radical trapping by DMF–B2pin2 or P(OEt)3 affords the desired products.

Scheme 141. Photoinduced Borylation and Phosphonation of Aryl Halides via EDA Complex Formation.

Scheme 141

Roy et al. reported the carbonate anion–assisted photoactivation of aryl halides relying on anion−π interactions (Scheme 142).735 Several salts, particularly AcOK and K3PO4, effectively promoted the photoinduced dissociation of Ar–X bonds, with the highest activity observed for K2CO3. This simple protocol was successfully tested on a gram scale and applied to borylation and phosphonylation. The results of experimental and spectroscopic studies suggested that the key aryl radical intermediate is generated via an anion−π interaction between the carbonate anion and aryl halide and ruled out the involvement of interactions between the electron-rich phosphite and aryl halide in Ar–X bond dissociation.

Scheme 142. Carbonate Anion-Assisted Photodissociation of Aryl Halides and Formation of Ar–P and Ar–B Bonds.

Scheme 142

4.2. Organophotocatalytic Activation of Ar–I Bonds

Photocatalytic reactions have been used to complement/access failed reactions and/or overcome the drawbacks associated with alternative transformations.736740 Ru- and Ir-based organophotocatalysts are well suited for promoting arylation by aryl iodides but suffer from the scarcity of their metal components, high cost, low sustainability, and high toxicity. Therefore, transition metal-free organic photoredox catalysts hold promise as alternatives to their transition metal-based counterparts for aryl iodide activation, enabling robust, green, and sustainable arylation for modern organic synthesis.626,630,636,741749 Organic photocatalysts are readily available, less toxic, and inexpensive and can utilize low-energy (e.g., visible) light-absorbing molecules, the structures of which can be tuned to improve their photophysical properties and photocatalytic performance. These organic compounds are privileged photocatalysts, avoiding the disadvantages of their precious metal-based counterparts and offering cheap and sustainable access to unique transformations and a broad array of substrates that are unreactive in most synthetic contexts. Additionally, the diversity of organophotocatalysts holds promise for the discovery and optimization of novel reactions.

The interaction of an organic photocatalyst (PC) with light affords an excited state (PC*) suitable for reducing an aryl iodide via SET in an oxidative quenching cycle (Scheme 143a). Alternatively, PC* is reduced by an electron donor to generate PC•– via SET, which is followed by SET-based aryl iodide reduction in a reductive quenching cycle (Scheme 143b). The organophotocatalyst promotes SET, providing access to previously inaccessible substrates (e.g., aryl halides) and transformations under mild conditions, thereby fostering the use of cheap and commercially available starting substrates.474491 XAT is a promising alternative to SET for generating aryl radicals from aryl iodides via homolytic Ar–I bond cleavage, which involves iodine abstraction by an in situ generated alkyl radical under photoredox conditions (Scheme 143c).645,750,751 The advantages of these techniques have enabled the development of elegant visible light-induced activations of the challenging Csp2–X bonds in aryl halides without the participation of expensive and toxic transition metals under mild conditions.

Scheme 143. Generation of Aryl Radicals from Aryl Iodides Promoted by Organophotocatalysts.

Scheme 143

Chiba et al. demonstrated the catalytic reduction of aryl halides based on the formation of aryl radicals promoted by polysulfide anions as photocatalysts and their subsequent participation in the anti-Markovnikov hydroarylation of alkenes (in the presence of Hantzsch ester (HEH) as a reductant), borylation, hydrogenation, and biaryl formation (Scheme 144a).752,753 K2Sx was used to generate the photoactive S3•–, S42–, and S32– species, with the ground-state redox potentials of S4•–/S42– and S3•–/S32– couples estimated at around −0.85 and −1.35 V vs SCE, respectively. For example, biaryl cross-coupling starts with the photoexcitation of S42–, and the excited form reduces the aryl halide via SET to generate S4•– and an aryl radical via the transient formation of an aryl halide radical anion (Scheme 144b). Charge transfer between S4•– and S32– generates ground-state S42– and S3•–. The aryl radical adds to N-methylpyrrole to give a dearomatized biaryl radical, which is then oxidized by the photoexcited [S3•–]* to afford the desired product.

Scheme 144. Polysulfide Anion-Photocatalyzed Functionalization of Aryl Halides.

Scheme 144

Among a small library of N,N- and N,O-coordinated organocatalysts, Das et al. explored pyridone amide as a potent organophotocatalyst for Ar–I bond activation inducing inter/intramolecular biaryl formation in the presence of tBuOK under UV irradiation (Scheme 145a).754 The reaction tolerated a broad range of (hetero)aryl iodides and provided access to tricyclic lactam and sultam derivatives via intramolecular C–C bond formation. According to the proposed mechanism (Scheme 145b), the catalytic cycle is initiated by the deprotonation of the organocatalyst, with subsequent photoirradiation generating an excited complex. Subsequently, SET to Ar–I generates an aryl radical and the pyridone amide radical cation. The aryl radical is trapped by the arene substrate to give a biaryl radical, which is oxidized by the pyridone amide radical cation via SET to produce a biaryl cation and cationic amide complex. In the presence of a base, the biaryl cation is deprotonated to afford the coupling product.

Scheme 145. Photoinduced C(sp2)–H Arylation and Formation of Biaryls and Tricyclic Fused Lactams and Sultams.

Scheme 145

10-Phenylphenothiazine (Ph-PTH) was used as an organophotocatalyst for the chemoselective hydrodehalogenation and C–C cross-coupling of polyhalogenated arenes.755 Larionov et al. developed phenothiazine (H-PTH) as an organic photocatalyst for the visible light–induced borylation of aryl iodides (Scheme 146a).756 In the presence of Cs2CO3, the reduction potential of phenothiazine reached approximately −3 V vs SCE, which enabled the photoborylation of Ar-O/N/Cl/Br/I substrates. The process was suitable for gram-scale synthesis and the borylation of structurally complex substrates, including bioactive ingredients and natural products, showing a high functional group tolerance. The photoactivation mechanism was based on photoinduced proton-coupled electron transfer, as verified by experimental and theoretical studies (Scheme 146b). According to the proposed mechanism, a complex consisting of H-PTH and Cs2CO3 held together by hydrogen bonding was converted to a singlet excited state under photoirradiation, with subsequent SET generating an aryl radical and a phenothiazinyl radical–hydrogen carbonate complex. The aryl radical reacts with B2pin2 to afford the desired borylation product along with a boryl radical, which reacts with the phenothiazinyl radical–hydrogen carbonate complex to regenerate H-PTH.

Scheme 146. Visible Light/Phenoxide-Induced Borylation of Ar–X Bonds via Photoinduced Proton-Coupled Electron Transfer.

Scheme 146

Phenothiazine (H-PTH) photocatalysts were used for the phosphonation of aryl halides to form aromatic phosphonates under ambient conditions (Scheme 147a).757 According to the mechanism proposed based on the results of radical trapping and fluorescence quenching experiments (Scheme 147b), the photoirradiation of H-PTH generates an excited species, which reduces the aryl halide via SET to form an aryl radical and the H-PTH radical cation. The aryl radical reacts with phosphite to generate a phosphoranyl radical, which is oxidized by the H-PTH radical cation via SET to generate a phosphonium cation. The subsequent reaction with DBU affords the desired product.

Scheme 147. Phenothiazine-Photocatalyzed Phosphonation of Aryl Halides.

Scheme 147

Maestro and Alemán constructed valuable heterocyclic skeletons via the photoinduced generation of aryl radicals in the presence of Ph-PTH followed by trapping by S, P, and Si atoms (Scheme 148a).758 The suggested mechanism starts with the reduction of the aryl halide by the photoexcited Ph-PTH (Ph-PTH*) through SET, and the thus generated aryl radical is attacked by the sulfur lone pair. The subsequent homolytic cleavage of the tBu bond affords a cyclic product (Scheme 148b). The catalytic cycle is completed by the reduction of Ph-PTH•+ via SET in the presence of DIPEA, which regenerates the active catalyst (Ph-PTH). This strategy was applied to the synthesis of cyclic sulfinamides, sultines, sulfoxides, phosphonates, and silyl ethers.

Scheme 148. Photocatalytic Intramolecular Homolytic Substitution-Based Cyclization or Aryl Iodides.

Scheme 148

Budén et al. hydrodehalogenated aryl halides using Hantzsch ester (HEH) as a visible light-absorbing electron-donating photoredox reagent (Scheme 149a).759 Li et al. used HEH to couple haloarenes and arylsulfinates under visible light (Scheme 149b).760 HEH was also used in the photocatalytic formation of C–S bonds in diarylsulfones, and this process was successfully performed on a preparative scale under sunlight. According to the proposed mechanism, deprotonated Hantzsch ester (HE) interacts with Ar–X to form an EDA complex, the excited state of which has a reduction potential (Ered = −2.49 V vs SCE) sufficient for converting Ar–X into an aryl radical with the concomitant formation of the HE radical. The coupling of the aryl radical with an arylsulfinate gives a sulfone radical anion, which forms an EDA complex with the HE radical. Finally, SET to another aryl halide molecule affords the desired product.

Scheme 149. Hantzsch Ester-Photocatalyzed Hydrodehalogenation and Sulfonation of Haloarenes.

Scheme 149

Leonori et al. generated aryl radicals under photoredox conditions via the in situ generation of α-aminoalkyl radicals as Ar–X bond activators via XAT.645,744746 Besides the Giese alkylation and allylation of aryl iodides using tertiary amines as α-aminoalkyl radical precursors and 4CzIPN as a photocatalyst under blue-light irradiation (Scheme 150a),761 this strategy was successfully applied to the arylation of pyrroles and phosphites with aryl halides under photoredox conditions.762 Experimental and computational studies pointed to the importance of 4CzIPN and the generated α-aminoalkyl radicals as initiators for the radical chain propagation mechanism. The excited form of the catalyst produced upon irradiation cannot directly activate Ar–X bonds via SET but rather oxidizes the tertiary amine to generate an α-aminoalkyl radical as the key intermediate for Ar–X bond activation via XAT. The thus generated aryl radical reacts with pyrrole to form a biaryl radical, which engages in XAT with the other aryl halides to afford the desired product and regenerate the aryl radical for the propagation cycle (Scheme 150b).

Scheme 150. Aminoalkyl Radical-Mediated Cross-Coupling of Aryl Halides Based on Halogen-Atom Transfer (XAT).

Scheme 150

Baidya et al. used α-aminoalkyl radicals as powerful XAT agents to generate aryl radicals, with subsequent intramolecular cyclization affording biologically relevant alkaloids (Scheme 151).763 For example, the irradiation of ortho-iodo-substituted N-arylbenzamides in the presence of 4CzIPN/nBu3N afforded phenanthridinone derivatives, including natural products from the Amaryllidaceae family.

Scheme 151. Aminoalkyl Radical-Mediated Intramolecular Cyclization of Aryl Halides Based on XAT.

Scheme 151

Nicewicz et al. developed various dehalogenative transformations, such as hydrodehalogenation and radiofluorination via halide/18F exchange, facilitated by acridinium photocatalysts (Mes-Acr+) as efficient photochemical reductants (Scheme 152).764767 In these systems, the photoexcited catalyst (Mes-Acr+*) engages in SET with a tertiary amine to generate a Mes-Acr radical, which is photoexcited to afford a strongly reducing twisted intramolecular charge-transfer state that reacts with an aryl halide to give an aryl radical and regenerate Mes-Acr+.

Scheme 152. Hydrodehalogenation and Radiofluorination Promoted by Mes-Acr Photocatalysts.

Scheme 152

Gianetti et al. used a green light-absorbing acridinium photocatalyst for the efficient α-arylation of cyclic ketones with aryl halides in the presence of pyrrolidine to form α-aryl ketones (Scheme 153a).768 This method was applied to the multigram-scale syntheses of a selective β3-adrenergic agonist precursor and other biologically valuable intermediates. Mechanistic investigations revealed the possibility of the concurrent contributions of SET and XAT (Scheme 153b). Upon green-light irradiation, the catalyst is converted to an excited state (Acr+*) capable of engaging in oxidative/reductive quenching cycles (E1/2(C2+•/C+*) = −1.85 V and (E1/2(C+*/C) = +1.15 V vs SCE). In the oxidative cycle, Acr+* acts as a strong reductant and engages in SET to the desired aryl halide to generate an aryl radical and the oxidized catalyst. The aryl radical reacts with an enamine intermediate via radical chain propagation or radical–radical coupling to afford the desired product after hydrolysis, while the oxidized catalyst engages in SET with the enamine to regenerate the acridinium salt. In the case of the reductive quenching cycle, the excited catalyst accepts an electron from the employed amine to provide an amine radical cation, which generates an aryl radical via XAT. The aryl radical initiates the radical propagation cycle to afford the desired product.

Scheme 153. α-Arylation of Cyclic Ketones Promoted by Acridinium Photocatalysts.

Scheme 153

Wolf et al. replaced precious Ir photocatalysts with a cheap organic alternative (3DPAFIPN) and used it to arylate HPPh2 with aryl iodides (Scheme 154a).769,770 The metal-free process showed superior productivity and selectivity toward quaternary phosphonium salts at base and catalyst loadings lower than those required by Ir catalysts. This more versatile metal-free synthetic method featured a broader scope, including the arylation of arylphosphines (e.g., H2PPh) and white phosphorus (P4) and affording the related asymmetrical/symmetrical products with excellent selectivity. Regarding the aryl iodide substituents, ortho substituents resulted in a preference for tertiary phosphines, whereas meta and para substituents resulted in the exclusive formation of quaternary phosphonium salts. According to the proposed mechanism (Scheme 154b), the photoexcited 3DAPAFIPN* engages in SET with Et3N to generate the 3DAPAFIPN radical anion [E1/2 (PC/PC•–) = −1.59 V vs SCE], which reduces the aryl iodide to give an aryl radical and regenerate the catalyst. The aryl radical abstracts a hydrogen atom from the phosphine to give a phosphine radical, which dimerizes to generate a diphosphine that is converted into a tertiary phosphine which is attacked by the aryl radical.

Scheme 154. 3DPAFIPN-Catalyzed Arylation of Phosphines and Phosphorus under Visible-Light Irradiation.

Scheme 154

Shang et al. designed diarylamides and (thio)phenolates with nitrogen, sulfur, oxygen, and ortho-diphenylphosphino groups as strongly reducing photocatalysts for activating challenging Ar–X and other bonds by visible light.771773 The scope of this process included the coupling of aryl halides with diboron reagents, N-methylpyrrole, and triethylphosphite, which afforded the corresponding arylboronates, 2-arylpyrroles, and arylphosphonates, respectively (Scheme 155a).773 According to the suggested mechanism, the irradiation of the deprotonated phenolate catalyst generates an excited state (phenolate*) that activates the aryl halide via SET to afford aryl and phenoxy radicals (Scheme 155b). The reaction of the phenoxy radical with the activated diboronate affords the persistent borate radical anion, which reacts with the aryl radical to afford the desired product and regenerate the phenolate catalyst (mechanism a). In the case of arylation, the aryl radical is trapped by pyrrole to generate a biaryl radical, which is oxidized by the phenoxy radical to generate the desired biaryl after deprotonation by a base (mechanism b).

Scheme 155. ortho-Phosphinophenolates as Potent Photoredox Catalysts for the Borylation, Arylation, and Phosphonation of Aryl Halides.

Scheme 155

Molander et al. used thiophenolate as a photoredox catalyst to mediate the photoactivation of (hetero)aryl halides followed by Giese addition, thus obtaining 3,3′-disubstituted oxindoles (Scheme 156).774 The postulated mechanism starts with the photolysis of the disulfide catalyst to generate a thiyl radical, which abstracts hydrogen from HCO2Na to produce thiophenol. The deprotonation of thiophenol or the single-electron reduction of the disulfide catalyst delivers the thiolate, which is photoexcited to a strongly reducing state. SET between the excited thiolate and aryl halide generates an aryl radical, which undergoes Giese addition to acrylamide followed by SET with the thiyl radical to generate the desired oxindole after deprotonation.

Scheme 156. Thiophenolate-Catalyzed Photoactivation of Aryl Halides for Addition to Aryl Acrylamides.

Scheme 156

Chen et al. employed an N-heterocyclic nitrenium salt (NHN) as a potent photoredox catalyst for the reduction of aryl halides that generated aryl radicals and enabled a variety of radical transformations.775 Aryl halides bearing amide moieties were transformed into the corresponding cyclization products, such as isoindoline-1-ones, oxindoles, and phenanthridine-6(5H)-ones, via an aryl halide reduction/1,5-HAT/cyclization sequence in the presence of NHN under blue light (Scheme 157a). Other radical reactions, such as hydrodehalogenation, cascade cyclization/hydrodehalogenation, and biaryl cross-coupling, were also successfully carried out using NHN photocatalysis. Two pathways were proposed based on the results of experimental studies (Scheme 157b). In the first pathway, NHN photoexcitation results in intramolecular charge transfer, which generates an aminyl radical that reduces the aryl halide via a CT complex to generate an aryl radical. The following 1,5-HAT and intramolecular cyclization and aromatization via deprotonation afford the desired product. Alternatively, the aryl radical can be generated via SET from the photoexcited NHN/TMEDA CT complex.

Scheme 157. N-Heterocyclic Nitrenium Salts as Effective Photoredox Catalysts for Haloarene Activation.

Scheme 157

Diazaphospholene and diazaphosphinane organocatalysts enable the radical functionalization of aryl halides.776779 Cramer et al. used 1,3,2-diazaphospholene hydride (DAP-H) generated from the corresponding phosphine oxide as a pivotal catalyst in the presence of DBU and HBpin to achieve the photoinduced radical cyclization of olefinic group–bearing aryl halides and thus access diverse cyclic skeletons (Scheme 158a).780 Various aryl halides were used to construct dihydrobenzofurane, chromane, and indoline heterocycles. According to the proposed mechanism, the reduction of the phosphine oxide precatalyst by HBpin affords DAP-H, which dimerizes to (DAP)2 under irradiation (Scheme 158b). The dissociation equilibrium between (DAP)2 and 2DAP enables halogen abstraction from the aryl halide to initiate the radical chain process. The obtained DAP-X regenerates DAP-H in the presence of DBU/HBpin, whereas the aryl radical undergoes a 5-exo-trig cyclization to generate an alkyl radical. HAT between the generated radical and DAP-H affords the desired product and DAP, which propagates the radical chain process.

Scheme 158. Diazaphospholene-Photocatalyzed Cyclization of Organohalides.

Scheme 158

Abe et al. designed a library of pyrrolo[2,1-a]isoquinolines as multifunctional photoredox organocatalysts (Scheme 159).781 A photocatalyst with advantageous S1 (E*oxS = −2.06 V vs SCE) and T1 ((E0,0T) = 2.58 eV (59.5 kcal/mol) states was used to achieve photocatalyzed SET and energy-transfer reactions. For example, this organophotocatalyst promoted the couplings of aryl halides with (hetero)arenes, B2pin2, and triethylphosphite under visible light. The photoexcitation of this catalyst affords a singlet state with a highly negative oxidation potential, which results in the generation of an aryl radical, with the following coupling reaction affording the desired product.

Scheme 159. Pyrroloisoquinoline as an Efficient Photoredox Catalyst for Ar–X Bond Activation.

Scheme 159

The merging of synthetic photoredox catalysis with natural proteins through genetic code expansion inspired Wang et al. to devise an artificial reductive photodehalogenase (RPDase) by encoding 4-benzoylphenylalanine at the 66th position of the superfolder yellow fluorescent protein followed by posttranslational intramolecular cyclization and oxidation to form a benzophenone-imidazolinone chromophore (Scheme 160a).782 This RPDase, in the presence of light and H(D)COONa as a sacrificial reductant, efficiently promoted the hydro(deutero)dehalogenation of sterically and electronically diverse (hetero)aryl halides and bioactive derivatives. Intriguingly, whole-cell biocatalysis using RPDase-expressing recombinant Escherichia coli enabled the efficient hydrodehalogenation of bioactive aryl halides. Control and spectroscopic experiments supported the mechanism outlined in Scheme 160b. According to this mechanism, the benzophenone moiety of the RPDase undergoes photoexcitation from S0 to S1 and subsequent intersystem crossing to T1 followed by reductive quenching with formate to afford the benzophenone radical anion (RPDase•–, E0 = −1.49 V in DMF) and CO2•– (Ered = −2.2 V vs SCE). The reduction of the aryl halide by RPDase•– or CO2•– generates an aryl radical, which abstracts hydrogen from formate to yield the arene product and regenerate CO2•– to propagate the chain mechanism.

Scheme 160. Artificial Photodehalogenase for the Hydrogenation and Deuteration of Aryl Halides.

Scheme 160

Adhikari et al. showed that in the presence of KOtBu and light, fluorene acts as an efficient radical initiator the for C–C coupling of aryl iodides with arenes producing (hetero)biaryls (Scheme 161a).783 The radical chain mechanism is initiated by the in situ conversion of fluorene to carbanion/radical anion species capable of SET in their excited states, which results in the reductive cleavage of the Ar–I bond (Scheme 161b). The remaining part of the mechanism is chain propagation, as in the case of BHAS.

Scheme 161. Fluorene-Mediated Reductive Cleavage of Ar–I Bonds in Biaryl Synthesis.

Scheme 161

Besides the documented ground-state catalytic efficiency of phenalenyl-based scaffolds, Roy et al. explored the behavior of this system in the excited state, revealing that the excited phenalenyl radical anion is a potent reductant for the reductive cleavage of Ar–X bonds (Scheme 162a).784 This phenalenyl-based photocatalytic system was successfully used to functionalize aryl halides through hydrodehalogenation, coupling with unactivated arenes, and C–P and C–B bond formation. According to the results of electrochemical, spectroscopic, and DFT studies (Scheme 162b), the phenalenyl photocatalyst forms a complex with DBU, and subsequent photoexcitation and SET generate a solvated radical pair. The second photoexcitation of phenalenyl•– generates an excited state (phenalenyl•–*) with a higher potential for ArX reduction. The thus generated aryl radical reacts with the coupling partner to afford the desired product, whereas the starting photocatalyst is obtained from phenalenyl•–* via SET.

Scheme 162. Phenalenyl-Photocatalyzed Reductive Functionalization of Aryl Halides.

Scheme 162

Blakely et al. developed a photoredox multicomponent reaction of aryl iodides with olefins and O2 mediated by silyl radicals as halogen acceptors via XAT (Scheme 163a).785 The differential nucleophilicity of the in situ generated aryl radicals enabled the preferential reaction with olefins and O2 for sequential C–C and C–O bond formation, respectively. This protocol was applied to a wide range of alkenes to obtain the corresponding hydroxyarylated products. The hypothesized mechanism starts with the oxidation of silanol by the photoexcited Cl-4CzIPN to generate a silyl radical via the Brook rearrangement (Scheme 163b). XAT between the aryl halide and silyl radical affords an aryl radical, which adds to the olefin, and the subsequent reaction with triplet O2 affords a peroxy radical. The following single-electron reduction by the Cl-4CzIPN radical anion and reduction via a photocatalytic cycle induced by silanol afford the desired product.

Scheme 163. XAT-Based Three-Component Reaction of Aryl Iodides with Olefins and O2.

Scheme 163

Guo et al. used Li2S and K2S as sulfur sources and electron donors for the formation of EDA complexes, inducing the selective transformation of aryl halides into diaryl sulfides and diaryl disulfides under light irradiation (Scheme 164).786 In the presence of a diaryl sulfide photocatalyst and base, an EDA complex consisting of the photocatalyst and S2– engages in SET upon irradiation to generate S•– and a photocatalyst radical anion, which reduces the aryl halide via SET to form an aryl radical (Scheme 164a). The thus generated aryl radical reacts with S•– to generate a thiolate anion, which forms an EDA complex with the aryl halide. Subsequent SET gives aryl and thiyl radicals, which undergo radical–radical coupling to afford the desired diaryl sulfide. When the reaction was conducted with excess K2S in the absence of a photocatalyst, diaryl disulfides were obtained instead of diaryl sulfides (Scheme 164b). The proposed mechanism starts with the formation of an EDA complex comprising S2– and an aryl halide, with subsequent photoinduced SET affording an aryl radical and S•–. The following radical–radical coupling produces a thiolate anion, which interacts with the aryl halide to form an EDA complex that is converted into aryl and thiyl radicals under irradiation. The generated thiyl radical dimerizes to afford the diaryl disulfide, whereas the aryl radical is trapped by S•– to form a thiolate anion.

Scheme 164. Sulfide Anion-Induced Photodissociation of Aryl Halides and C–S Bond Formation.

Scheme 164

4.3. Photoinduced Dissociation of Aryl–Iodonium Bonds

Light-induced chemical transformations have drawn considerable attention as elegant and ecofriendly approaches in organic chemistry.787792 UV–vis light is a clean, green, convenient, and renewable energy source used to promote chemical reactions through the absorption-induced generation of reactive excited states that engage in reactions or are converted into products inaccessible under traditional thermal conditions.

The photolysis of diaryliodonium salts produces aryl radicals, which can participate in a series of transformations.793796 We discuss the recently reported metal- and organometallic photocatalyst-free visible light-induced arylation reactions involving diaryliodonium salts and affording C–C and C–heteroatom bonds. The versatility of diaryliodonium salts as photoinitiators for the synthesis of polymers has been extensively reviewed and is outside the scope of the present work.797809

4.3.1. Photocatalyst-Free Dissociation of Aryl–Iodonium Bonds

The photochemical reactions of diaryliodonium salts are initiated by an interaction between the iodonium center and an electron donor, which affords an EDA complex with photochemical properties superior to those of the parent iodonium substrate. The photoexcitation of the EDA complex enables internal SET, which generates an aryl radical that subsequently undergoes radical–radical coupling or is trapped by a radical acceptor to give the desired product.

Tobisu and Chatani reported that the arylation of N-methylpyrrole with diaryliodonium salts affords 2-arylpyrroles under visible light in the absence of photocatalysts (Scheme 165).810 The mechanism proposed based on the results of UV–vis spectroscopic analysis indicates an interaction between pyrrole and the iodonium salt through the formation of a CT complex. The photoexcitation of this complex generates a pyrrole radical cation and an aryl radical along with an aryl iodide via SET. The subsequent coupling reaction affords a 2-arylpyrrole cation, which is deprotonated to furnish an arylated pyrrole.

Scheme 165. Photoinduced Arylation of Pyrroles with Diaryliodonium Salts.

Scheme 165

Yadav et al. synthesized quaternary CF3-substituted oxindoles via an arylation–cyclization cascade under visible-light irradiation, using diaryliodonium salts as aryl radical sources for the reaction with N-aryl-2-(trifluoromethyl)acrylamide (Scheme 166).811 According to the proposed mechanism, the photoexcitation of the iodonium salt followed by SET with the acrylamide generates an aryl radical and acrylamide radical cation. Coupling followed by cyclization and aromatization affords the desired product.

Scheme 166. Visible Light-Induced Synthesis of Oxindoles Involving an Arylation–Cyclization Cascade.

Scheme 166

Manolikakes et al. developed a visible light-induced three-component reaction between a diaryliodonium salt, an arylpropynoate, and DABSO affording the corresponding sulfonylated coumarin (Scheme 167a).812 Sunlight irradiation provided a comparable yield. In the proposed mechanism, DABSO serves as an SO2 source and diaryliodonium salt activator. The interaction between DABSO and the iodonium salt results in the formation of a CT complex, which engages in SET under irradiation to generate an aryl radical. This radical reacts with SO2 to generate an arylsulfonyl radical, which is intercepted by an alkynyl ester to generate an alkenyl radical that undergoes spirocyclization followed by oxidation to generate a cationic spirocyclized intermediate. The following 1,2-ester migration and sequential rearomatization and deprotonation afford the coumarin product. Analogously, the one-pot reaction of a diaryliodonium salt, N-arylacrylamide, and DABSO under visible-light irradiation delivered the corresponding sulfonylated oxindole (Scheme 167b).813

Scheme 167. Photoinduced Three-Component Synthesis of Sulfonylated Coumarins and Oxindoles.

Scheme 167

Volla et al. performed the spirocyclization of alkenyl and alkynyl amides using diaryliodonium salts and DABSO under white-light irradiation to generate the corresponding spirocycles (Scheme 168a).814 A four-component cascade reaction was reported by the same group. The reaction of diaryliodonium triflates, DABSO, an alkynyl cyclohexadienone, and diphenyldiselenide under white-light irradiation afforded substituted dihydrochromenones with high regio- and diastereoselectivities (Scheme 168b).815

Scheme 168. Photoinduced Multicomponent Synthesis of Sulfonylated Bicyclic Compounds.

Scheme 168

Karchava et al. reported the arylation of phosphine derivatives with aryl(Mes)iodonium triflate (Ar(Mes)I-OTf) affording arylphosphonium salts under visible-light irradiation (Scheme 169a).816 The proposed mechanism starts with the formation of an EDA complex between the Mes-iodonium salt and tertiary phosphine, which engages in SET under photoirradiation to generate aryl and phosphorus-centered radicals. The combination of these radicals affords the desired arylphosphonium salt. Further expansion was performed by carrying out the one-pot coupling of aminophosphines with Ar(Mes)I-OTf to generate phosphonium salts under blue-light irradiation, with subsequent hydrolysis affording the corresponding arylphosphine oxides (Scheme 169b).817 The same group developed an indirect route to unsymmetrical tertiary arylphosphines (Scheme 169c).818 The one-pot reaction of (2-cyanoethyl)diphenylphosphine with Ar(Mes)I-OTf under visible-light irradiation followed by the treatment of the thus generated quaternary phosphonium salts with DBU furnished the corresponding tertiary arylphosphines via a retro-Michael reaction.

Scheme 169. Photoinduced Arylation of Phosphines with Iodonium Salts.

Scheme 169

Lakhdar et al. synthesized arylphosphonates from diaryliodonium salts and trialkylphosphites under blue-light irradiation (Scheme 170).819 According to the mechanism proposed based on the results of experimental and theoretical studies, the low electron-donating ability of phosphites leads to the formation of weak EDA complexes with iodonium salts. SET under blue-light irradiation affords an aryl radical and a phosphorus-centered radical cation, which combine to generate arylphosphonium salts. Nucleophilic displacement in the presence of a base affords the desired product.

Scheme 170. Light-Promoted Reaction of Iodonium Salts with Trialkylphosphites Affording Arylphosphonates.

Scheme 170

4.3.2. Organophotocatalyzed Aryl–Iodonium Bond Dissociation

Manolikakes et al. synthesized N-aminosulfonamides via the photoinduced three-component reactions of diaryliodonium salts, substituted hydrazines, and SO2 surrogates (DABSO or K2S2O5/TFA) in the presence of perylenediimide (PDI) as a photoredox organocatalyst (Scheme 171).820 The reaction mechanism starts with the photoexcitation of PDI to generate PDI*. Concomitantly, the liberated SO2 interacts with hydrazine to afford a zwitterion, which is oxidized by PDI* via SET and deprotonated to generate a sulfonyl radical. SET from the reduced catalyst (PDI•–) to the aryl halide affords an aryl radical, which undergoes radical–radical coupling with the sulfonyl radical to afford the desired product.

Scheme 171. Perylenediimide-Promoted Reactions of Iodonium Salts with Hydrazine and SO2 under Visible-Light Irradiation.

Scheme 171

Zhang et al. developed a three-component cascade reaction of diaryliodonium salts with N-propargyl aromatic amines and DABSO catalyzed by Eosin Y under visible-light irradiation for the synthesis of 3-arylsulfonylquinolines (Scheme 172).821 According to the proposed mechanism, Eosin Y is photoexcited upon irradiation with green LED light to produce Eosin Y*, which engages in SET with an iodonium salt to generate an aryl radical and Eosin Y•+. The aryl radical reacts with DABSO to give an arylsulfonyl radical, which is regioselectively trapped by the triple bond of propargylamine to generate the corresponding alkenyl radical. Intramolecular cyclization followed by deprotonation furnishes a radical anion, which is oxidized by Eosin Y•+ or the iodonium salt to regenerate Eosin Y or the aryl radical and yield the desired product.

Scheme 172. Photocatalytic Synthesis of 3-Arylsulfonylquinolines under Green-Light Irradiation.

Scheme 172

Piguel et al. reported a three-component reaction of iodonium salts with DABSO and an imidazoheterocycle in the presence of EosinY under green-light irradiation, wherein the C-3 sulfonylation of the imidazoheterocycles proceeded via C–H arylsulfonylation (Scheme 173).822 The reaction mechanism is analogous to that shown in Scheme 171. An arylsulfonyl radical generated by Eosin Y* reacts with the imidazoheterocycle at C-3 to form a radical intermediate. SET with Eosin Y•+ and subsequent deprotonation afford the desired product.

Scheme 173. Photoinduced Selective C–H Arylsulfonylation of Imidazoheterocycles.

Scheme 173

Ma et al. applied arylsulfonyl radical generation upon irradiation to the fluorosulfonylation of diaryliodonium salts using DABSO and KHF2 in the presence of camphorquinone (CQ) as an organophotocatalyst (Scheme 174).823 According to the proposed mechanism, blue-LED irradiation converts CQ to an excited state, which reduces the employed iodonium salt to an aryl radical and generates CQ•+. The coupling of the aryl radical with SO2 from DABSO affords an arylsulfonyl radical, whereas CQ•+ reduces DABCO to the DABCO radical cation. These species combine in the presence of KHF2 to form a sulfonate intermediate, which loses DABCO to generate an arylsulfonyl fluoride.

Scheme 174. Photocatalytic Synthesis of Arylsulfonyl Fluorides from Iodonium Salts, DABSO, and KHF2.

Scheme 174

Murarka et al. used methylene blue trihydrate as an efficient photoredox organocatalyst for the arylsulfonylation of Morita–Baylis–Hillman acetates by diaryliodonium triflates in the presence of DABSO (Scheme 175a).824 The proposed mechanism starts with the photoexcitation of the catalyst, which is followed by reductive quenching with DIPEA via SET to generate the highly reducing MB•– radical anion, which reduces the diaryliodonium salt to regenerate MB and form an aryl radical. The sequential addition of this radical to SO2 and the double bond of allyl acetate affords the corresponding alkyl radical, which expels an acetate radical to afford the desired product. When the reaction was carried out in the absence of DABSO, the aryl radical directly reacted with allyl acetate to generate the corresponding allylarene (Scheme 175b).

Scheme 175. Photoinduced Arylsulfonation and Arylation of Allyl Acetate with Diaryliodonium Triflates.

Scheme 175

Quinoline and pyridine N-oxides were selectively arylated at C-2 with diaryliodonium tetrafluoroborates in the presence of Eosin Y as a photocatalyst under visible light to afford the corresponding N-heterobiaryls (Scheme 176).825 The proposed mechanism starts with the excitation of Eosin Y by blue light, and the thus generated Eosin Y* is converted into an aryl radical and EosinY•+. The addition of the aryl radical to quinoline N-oxide followed by oxidation with EosinY•+ generates an N-oxide cation that is deprotonated to generate the desired 2-arylated product. 1,4-Benzoquinone and K2S2O8 were found to be important additives promoting deprotonation and oxidizing EosinY, respectively.

Scheme 176. Photoinduced Arylation of Quinoline and Pyridine N-Oxides with Diaryliodonium Salts.

Scheme 176

Aryliodonium ylides can also generate aryl radicals18,826828 under irradiation in the presence of EosinY, which was applied to the direct C(sp2)-H arylation of diverse heterocycles (Scheme 177a).829 Aryliodonium ylide screening indicated the broad applicability of this reaction and its excellent compatibility with diverse functional groups and bioactive compounds. The ability of 4CzIPN to induce the generation of aryl radicals from aryliodonium ylides under irradiation was used for the sequential arylation/cyclization of 2-isocyanobiaryls to generate 6-arylated phenanthridines (Scheme 177b).830 The proposed mechanism starts with the photoexcitation of 4CzIPN, with the subsequent SET generating an aryl radical and 4CzIPN•+. The addition of the aryl radical to the isocyano group affords an imidoyl radical, which is converted to a cyclized radical. The following oxidation by 4CzIPN•+ via SET affords the desired product after deprotonation.

Scheme 177. Generation of Aryl Radicals from Iodonium Ylides under Photoredox Conditions.

Scheme 177

Roy et al. designed a general photoredox system comprising NaI, Ph3P, and N,N,N′,N′-tetramethylethylenediamine (TMEDA) for the activation of diaryliodonium salts and generation of aryl radicals, which engaged in the C–H arylation of heterocycles (Scheme 178).831,832 This protocol was compatible with a broad array of aromatic and nonaromatic heterocycles, i.e., azauracils, quinoxaline-2-ones, cinnolinones, imidazopyridines, indazoles, pyrazinones, pyrazines, quinoline N-oxides, isoquinolines, and indoles. Detailed mechanistic investigations indicated the formation of a tetrameric EDA complex between the diaryliodonium salt, NaI, Ph3P, and TMEDA, with the irradiation-based excitation of this complex followed by SET generating an aryl radical, NaI, Ph3P, and TMEDA•+. The solvent (HFIP/H2O)-assisted nucleophilic addition of the aryl radical to azauracil affords a radical intermediate that engages in SET with NaI, Ph3P and TMEDA•+ to regenerate the photoredox system and produce a cationic intermediate. The deprotonation of the latter in the presence of the triflate anion gives the desired product.

Scheme 178. Photoactivation of Iodonium Salts by an Electron-Donor Triad (Na/Ph3P/N,N,N′,N′-Tetramethylethylenediamine (TMEDA)) for C–H Arylation.

Scheme 178

5. Electrochemical Dissociation of Ar–I Bonds

The conventional chemical transformations used to realize arylation with aryl halides suffer from a narrow substrate scope and the use of expensive metals and/or ligands or strong bases beside other environmental and practical concerns. Hence, efficient, green, and sustainable alternative synthetic transformations under benign conditions are urgently required. In this context, electrochemistry attracts growing interest and has emerged as a versatile and green strategy for broadening the scope of and complementing the current arylation approaches to realize challenging arylation reactions by applying electric current as a driving force.624,625,736,833844 The high tunability of electrochemical processes via applied current or potential adjustment makes them well suited for numerous redox reactions. These processes are ecofriendly, as they employ electrons as green redox reagents instead of the environmentally deleterious chemical reagents, which lowers the associated risks, costs, and waste production while increasing the atom/step economy.845853 Classical electrochemical reactions are initiated by SET from the electrode surface to the substrate in a heterogeneous process (direct electrolysis) (Scheme 179). Alternatively, a chemical substance (mediator, M) can be used in catalytic amounts as a redox reagent to shuttle SET (through a homogeneous process) to the substrate and initiate the electrochemical transformation (indirect electrolysis). Indirect electrolysis is superior to its direct counterpart, allowing one to (i) avoid substrate overreduction and mitigate electrode passivation, (ii) eliminate kinetic inhibition and overpotentials and thus accelerate the electrochemical process, (iii) achieve a high and/or different selectivity, and (iv) conduct electrolysis at a potential lower than the redox potential of the substrate, which leads to the reaction proceeding under milder conditions and exhibiting a broad functional group tolerance.

Scheme 179. Electrochemical Processes for Aryl Halide Activation.

Scheme 179

The reduction potential, bond dissociation energy, polarizability, and Ar–X bond cleavage mechanism484 of iodoarenes favor the dissociation of Ar–I bonds under mild conditions, which results in a broad substrate scope and sustainability. In this part of the review, we demonstrate the crucial role of electric current as a green energy source for aryl iodide reduction/activation at the cathode and the generation of aryl radicals suitable for further functionalization.844,854

5.1. Direct Dissociation of Ar–X Bonds

The electrochemical reductive transformations of aryl halides can be achieved by direct electrolysis or indirect electrolysis mediated by redox catalysts.624,625,845,849,850 Direct electrolysis is facilitated by the cathode-to-ArX SET and generation of highly reactive radical ions and radical intermediates, requiring the subsequent oxidation of a sacrificial reductant (i.e., anode erosion or use of cheap reductants with lower potentials, such as Et3N) to maintain the charge balance.

Ke et al. synthesized phenols via electrochemically driven Ar–X bond dissociation in the presence of H2O/Et3N/air (Scheme 180).855 The electrochemical process is hypothesized to start with the cathodic reduction of an aryl iodide to generate a radical anion, which undergoes Ar–I bond mesolysis to generate an aryl radical and liberate an iodide anion. The aryl radical reacts with molecular oxygen to form a peroxy radical, which is converted into the corresponding phenol in the presence of Et3N as a sacrificial reductant. The anodic oxidation of iodide affords molecular iodine.

Scheme 180. Electrochemical Hydroxylation of Aryl Halides.

Scheme 180

The Pan and Chi group reported the electrochemical hydrodehalogenation of aryl halides, showing that unlike that described previously, it did not require a divided cell and metallic electrode but featured a narrow substrate scope (Scheme 181, conditions a).856 Guo et al. designed a more applicable and controllable electrolysis protocol for haloarene hydrodehalogenation (conditions b).857 Ethanol, added as a cosolvent, competed with electrosensitive substrate moieties and consequently prevented overreduction and decomposition. In both reaction systems, the main feature of electrolysis on the cathode is the reduction of the aryl iodide to a radical anion, which undergoes Ar–I bond mesolysis to generate an aryl radical. The aryl radical is preferentially reduced to the corresponding aryl anion, which abstracts a proton from the solvent, Et3N, or moisture to afford the hydrogenated product. At the anode, the tertiary amine acts as a sacrificial reductant and provides electrons for the cathodic reduction.

Scheme 181. Electrochemical Hydrodehalogenation of Aryl Halides.

Scheme 181

This electrochemical hydrodehalogenation of aryl halides can be applied to reductive deuteration (Scheme 182). Lei et al. used an undivided cell with Pt and Pb electrodes (conditions a).858 In this case, the proposed mechanism involves the concomitant reduction of an aryl halide and D2O at the cathode to generate aryl and deuterium radicals, respectively, which undergo radical–radical coupling to afford a deuterated arene. Zhang et al. fabricated a copper nanowire array (NWA) cathode in situ through the electrochemical reduction of CuO NWAs on Cu foil and used it for the electrochemical deuteration of aryl halides with D2O as the sole deuterium source (conditions b).859 In view of its large surface area, the Cu NWA cathode outperformed Cu foil, Pt, carbon paper, and Ni foam cathodes. Interestingly, the one-pot conversion of Ar–H bonds to Ar–D bonds through a sequence of halogenation and deuterodehalogenation was efficiently applied to the site-specific deuteration of C–H (hetero)arenes and pharmaceuticals. Furthermore, paired reactions at the cathode and anode without additional oxidants are also possible.

Scheme 182. Electrochemical Dehalogenative Deuteration of Aryl Halides with D2O.

Scheme 182

Mo et al. borylated aryl iodides through the electrochemical in situ generation of aryl radicals as key intermediates (Scheme 183a).860 A plausible radical mechanism was suggested based on electron paramagnetic resonance and cyclic voltammetry indicating the generation of aryl radicals (Scheme 183b). According to this mechanism, the aryl iodide is reduced at the cathode to form an aryl radical, which reacts with B2pin2 to form a C–B bond in the presence of a base, with subsequent B–B bond dissociation generating Ar–Bpin and a borate radical anion.

Scheme 183. Electrochemical Borylation of Aryl Iodides.

Scheme 183

Wang et al. electrochemically phosphorylated aryl halides by trialkylphosphites in an undivided cell with a graphite cathode and Ni sacrificial anode (Scheme 184a).861 Trialkylphosphites worked well, although no reaction was observed for trimethyl- and triphenylphosphites. Mechanistic studies ruled out the possibility of Ni(II)/Ni(0) and Ni(III)/Ni(I) cycles, which are the key processes in Ni-catalyzed electrochemical reactions.862,863 According to a plausible mechanism, the reaction starts with the reduction of an aryl halide at the cathode to generate an aryl radical, which couples with a phosphite to form a phosphoranyl radical (Scheme 184b). The oxidation of the phosphoryl radical by Ni(II) generated through the erosion of the Ni anode followed by the reaction with nucleophiles in the reaction medium affords the desired phosphorylation product.

Scheme 184. Electrochemical Phosphorylation of Aryl Halides.

Scheme 184

5.2. Indirect Dissociation of Ar–X Bonds

The direct electrolysis of aryl halides, particularly complex ones, can be hindered by reactivity and selectivity problems. When direct electrolysis is not efficient, an alternative electrochemical strategy that involves the use of soluble redox mediators to shuttle electrons from the cathode to the aryl halide in the solution, becomes more appropriate. Therefore, the use of redox-active organic mediators (catalysts) can make electrosynthetic transformations proceed under more benign conditions via unique mechanistic pathways and thus result in a high reactivity, selectivity, and functional group compatibility, broad substrate scope, etc.624,625,845,849,850

Early examples of organic mediators include the use of phenanthrene and 9,9-disubstituted fluorenes for the reductive radical cyclization of ortho alkenylhaloarenes to the related five- and six-membered fused skeletons in an undivided cell equipped with a Pt cathode and Mg sacrificial anode (Scheme 185a).864866 In these reactions, stoichiometric amounts of redox mediators were used. Mitsudo et al. reported the electroreductive deuteration of haloarenes in the presence of 9-fluorenone as a mediator using an undivided cell equipped with a Pt cathode and Zn sacrificial anode (Scheme 185b).867 In contrast, Zhu et al. demonstrated that a catalytic amount of perylene bisimide serves as indirect mediator for the direct coupling of aryl halides with pyrroles to generate heterobiaryls (Scheme 185c).868 The reaction worked well with 1-ethyl-3-methylimidazolium bis((trifluoromethyl)sulfonyl)imide ([EMIM]NTf2) as an important electrolyte in an undivided cell equipped with a glassy carbon cathode and Zn sacrificial anode.

Scheme 185. Electroreductive Functionalization of Aryl Halides in the Presence of Redox Mediators.

Scheme 185

Brown et al. realized an electroreductive radical cyclization using an undivided electrochemical flow cell equipped with a stainless steel (SS) anode and glassy carbon (GC) cathode (Scheme 186).869 This reaction system had the advantages of not requiring a sacrificial anode, operating with catalytic mediator amounts, and suitability for gram-scale synthesis. The proposed mechanism starts with the one-electron reduction of phenanthrene on the cathode to generate the phenanthrene radical anion, which reduces the aryl halide via SET to form an aryl radical. After cyclization with an alkenyl moiety, the reduction of the resulting radical by the phenanthrene radical anion followed by protonation affords the desired product. The presence of phenanthrene (mediator) and its radical anion results in the formation of a detached reaction layer between the cathode and Ar–X. In this layer, the mediator acts as a charge shuttle by fluxing the phenanthrene radical anion outward from the cathode to induce SET to Ar–X and thus obtain Ar–X•–. Thus, the direct two-electron reduction of Ar–X to Ar–H is prevented.

Scheme 186. Phenanthrene-Catalyzed Electroreductive Radical Cyclization.

Scheme 186

Wang et al. employed cyanoarenes as aryl radical precursors and redox mediators for the electroreductive 1,2-diarylation of alkenes using a Zn sacrificial anode and reticulated vitreous carbon (RVC) cathode (Scheme 187).870 This protocol was compatible with various electron-withdrawing group–substituted cyanoarenes and the gram-scale synthesis and late-stage functionalization of natural and bioactive compounds. According to the mechanism proposed based on the results of DFT calculations, dicyanoarene reduction on the cathode generates a dicyanoarene radical anion, which reduces the employed aryl halide to form an aryl radical via SET. The anti-Markovnikov addition of the aryl radical to an alkene generates the corresponding alkyl radical, which combines with the dicyanoarene radical anion to afford the desired product after cyanide expulsion.

Scheme 187. Regioselective 1,2-Diarylation of Alkenes with Aryl Halides and Aryl Nitriles.

Scheme 187

The trapping of aryl or alkyl radicals by CO2 followed by protonation affords the corresponding carboxylic acids. Senboku et al. developed an electroreductive radical cyclization induced by substoichiometric amounts of methyl tert-butylbenzoate as a redox mediator, with the subsequent CO2 trapping affording bicyclic carboxylic acids (Scheme 188a).871,872 The reaction with ortho-alkenylbromoarenes generated monocarboxylic acids, whereas dicarboxylic acids were obtained when ortho-alkynylbromoarenes were used. In these cascade reactions, two or three electrons were sequentially transferred to bromoarenes for multiple bond formation. Xue et al. demonstrated the electroreductive carboxylation of aryl halides with CO2 generating carboxylic acid derivatives in the presence of 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD), wherein a catalytic amount of naphthalene was used as a mediator (Scheme 188b).873 Electronically and sterically diverse aryl halides were tolerated, and the late-stage carboxylation of aryl halides derived from natural products and drugs could be realized. The proposed mechanism is initiated by the one-electron reduction of naphthalene on the cathode, and the thus generated naphthalene radical anion engages in SET to the aryl halide to afford an aryl radical. The reduction of the aryl radical by the naphthalene radical anion or cathode generates a radical anion, which reacts with CO2 to furnish the corresponding benzoic acid derivative after protonation.

Scheme 188. Electroreductive Radical Generation Followed by CO2 Trapping.

Scheme 188

6. Electrophotochemical Activation of Aryl Halides

The benefit and power of electricity and light can be merged in elegant electrophotocatalytic (or photoelectrocatalytic) strategies to solve the critical problems of chemical reactions and enable unprecedented chemical transformations.624,625,874885 During consecutive photoinduced electron transfer (conPET), the photocatalyst is photoexcited and then reduced by an additional electron donor to generate a photocatalyst radical anion (PC•–), which is further photoexcited to afford a strongly reducing excited radical anion (PC•–*) (Scheme 189a).490,746,764,886896 The sequential generation of PC•–* from the ground state of the photocatalyst can be replaced with the combination of an electrochemical process and photoirradiation; the one-electron reduction of the photocatalyst proceeds at the electrode surface to generate PC•–, which is photoexcited to afford PC•–* (Scheme 189b). This synergistic strategy, i.e., the electrophotochemical activation of aryl halides, avoids the use of additional sacrificial reagents or strong bases and, hence, side reactions. In addition, the scope of potential catalysts is considerably broadened, as a limited number of compounds can absorb light and act as photocatalysts in their neutral states, whereas their radical ions are more suitable for this process. Therefore, the electrophotochemical strategy offers a general platform for transcending the limitations of electro- and photochemical protocols. This elegant hypothesis was successfully applied to the Ar–X bond activation of aryl halides with very negative reduction potentials.

Scheme 189. (a) Consecutive Photoinduced Electron Transfer and (b) Electrophotochemical Activation of Aryl Halides.

Scheme 189

Lambert et al. developed a powerful reductive electrophotocatalytic strategy mediated by dicyanoanthracene (DCA) as an organocatalyst (Scheme 190).897 The electrochemical reduction of DCA at the cathode generates a photoactive radical anion, DCA•–, which is excited by visible light to form a strongly reducing radical anion, DCA•–* (Ered = −3.2 V vs SCE). This protocol was applied to the reductive borylation, stannylation, and (hetero)arylation of aryl halides with very negative reduction potentials under blue-light irradiation using an H-type divided cell equipped with a carbon foam cathode and Zn-plate sacrificial anode. Challenging (hetero)aryl halides with reduction potentials of −1.9 to −2.9 V and potentially sensitive functional groups were tolerated. The reduction of the employed aryl halide by DCA•–* via SET leads to the regeneration of DCA and formation of an aryl radical, which is trapped by a radical acceptor (i.e., coupling partner) to afford the desired product. Mechanistic studies ruled out the formation of EDA complexes from DCA•– and aryl halides, and a dual photocatalytic mechanism was also found to be unlikely, as DCA is not excited by blue-LED light.

Scheme 190. Electrophotocatalytic Functionalization of Aryl Halides Induced by Dicyanoanthracene (DCA).

Scheme 190

Wickens et al. developed an electrophotocatalytic process promoted by N-aryl-1,8-naphthalimide (NpMI) for the reductive hydrodehalogenation, phosphorylation, and (hetero)arylation of challenging aryl chlorides (Scheme 191).898 According to the proposed mechanism, the priming of NpMI with electrons before visible-light excitation leads to the generation of the strongly reducing NpMI•–*, which has a reduction potential beyond that of Na0 and comparable with that of Li0. This approach enabled the reduction of highly electron-rich aryl chlorides (Ered = −3.4 V vs SCE) and was compatible with various aryl chlorides bearing potentially sensitive functional groups. Moreover, the adopted strategy showed robust reactivity and selectivity compared with the traditional photochemical and electrochemical approaches to the activation of aryl chlorides and radical cross-coupling with pyrrole over hydrodehalogenation.

Scheme 191. Electrophotocatalytic Reduction of Aryl Chlorides through Electron-Primed N-Aryl-1,8-naphthalimide (NpMI) as a Photoredox Catalyst.

Scheme 191

Wu et al. designed a coupled photoelectrochemical/photoredox strategy for the efficient and selective reductive functionalization of aryl halides under visible–near-infrared (NIR) light in the presence of a perylene-based (PDI) photocatalyst in an undivided cell equipped with an Sb2(S,Se)3 photocathode, Pt anode, Ag/AgCl reference electrode, and nBu4NOAc electrolyte (Scheme 192a).899 PDI, the photocathode, vis–NIR light, and the acetate anion of the electrolyte were required for the transformation to take place and maximize the yield of the coupling product. A broad range of electron-deficient (hetero)aryl halides could be coupled with diverse pyrrole derivatives, P(OEt)3, and B2pin2 for the chemoselective construction of C–C, C–P, and C–B bonds, respectively, in preference to the competitive hydrodehalogenation. Cyclic voltammetry and spectroscopic experiments supported the mechanism shown in Scheme 192b. According to this mechanism, the vis–NIR photoexcitation of the Sb2(S,Se)3 photocathode generates electron–hole pairs, and the electrons reach the cathode surface and reduce PDI to PDI•–. The acetate anion of the electrolyte interacts with the π-acceptor PDI via anion−π interactions and accelerates light capture and SET to generate PDI•–. The second photoexcitation of PDI•– affords the strongly reducing PDI•–* (Ered = −1.86 V vs SCE), which reduces the aryl halide to an aryl radical and is converted into PDI. The thus generated aryl radical is trapped by the coupling partner to afford the cross-coupling product.

Scheme 192. Tandem Photoelectrochemical/Photoredox Reductive Functionalization of Aryl Halides.

Scheme 192

Li et al. envisaged a novel multicomponent protocol for the electrophotocatalytic 1,2-diarylation of alkenes with aryl halides and cyanoaromatics (Scheme 193a).900 The synergistic nature of the electrophotocatalytic process enabled the reductive generation of ArX•– and ArCN•– for alkene 1,2-diarylation and process termination. According to the mechanism proposed based on the results of mechanistic studies and DFT calculations (Scheme 193b), the cathodic reduction of phenanthrene generates the corresponding radical anion, which engages in SET to form an aryl radical. Analogously, the DABCO radical cation generated by the photoexcited 4-DPAIPN via SET abstracts hydrogen from HCO2Na to afford the reactive CO2•–, which reacts further to generate an aryl radical. The trapping of this aryl radical by an alkene followed by coupling with ArCN•–, which is generated through the cathodic and photocatalytic reduction of ArCN, affords three-component coupling products after decyanation.

Scheme 193. Electrophotocatalytic Three-Component Cascade Reaction Used for the 1,2-Diarylation of Alkenes.

Scheme 193

Zhang et al. used thioxanthone as a precatalyst for the electrophotocatalytic hydrogenation of imines and reductive derivatization of aryl halides (Scheme 194).901 The addition of TfOH to the catalytic system enabled the efficient hydrogenation of imines at low potentials and thus precluded competitive reactions. In the absence of TfOH, the catalytic system generated a strongly reducing active species with a potency comparable with that of Na0 and Li0 and therefore enabled the photocatalytic reductive functionalization of aryl halides with very negative reduction potentials. The catalytic system enabled the hydrogenation, borylation, stannylation, and arylation of exceptionally challenging aryl halides and construction of Csp2–H, Csp2–B, Csp2–Sn, and Csp2–Csp2 bonds. These transformations were performed in an undivided cell under reductant-free conditions (the generated radical intermediates were used as sacrificial reagents) and exhibited high faradaic efficiencies.

Scheme 194. Electrophotocatalytic Reductive Functionalization of Aryl Halides.

Scheme 194

7. Assessing the Features of Iodoarene Transformations

Green chemistry is defined as the “design of chemical products and processes to reduce or eliminate the use and generation of hazardous substances.”902905 To contribute to this goal, synthetic chemists have developed chemical processes that utilize low-toxicity abundant resources, employ sustainable and renewable feedstocks, minimize waste and hazardous product formation, and use smaller amounts of energy. Anastas and Warner defined the 12 principles of green chemistry in 1998 as (1) waste prevention, (2) atom economy, (3) less hazardous chemical syntheses, (4) designing safer chemicals, (5) safer solvents and auxiliaries, (6) design for energy efficiency, (7) use of renewable feedstocks, (8) reduce derivatives, (9) catalysis, (10) design for degradation, (11) real-time analysis for pollution prevention, (12) inherently safer chemistry for accident prevention.

To evaluate the chemical transformations from a green chemistry perspective, Atom Economy (AE) and Environmental Impact Factor (E-factor) are the simplest and most widely adopted green metrics in both industrial and academic chemistry (Figure 3).906,907 AE, proposed by Trost in 1991, is calculated as the molecular weight of the desired product divided by the total molecular weight of all reactants appearing in the stoichiometric equation, expressed as a percentage (Figure 3a).908,909 This metric assumes the stoichiometric use of starting materials and a theoretical chemical yield, making it a practical tool for assessing alternative synthetic routes to a target molecule without requiring experimental data. In contrast, the E-factor, proposed by Sheldon in 1992, quantifies the actual waste generated in a process and is determined by dividing the total mass of waste by the mass of the final product (Figure 3b).910,911 Thus, an ideal E-factor is zero, with higher values indicating greater waste generation and environmental impact. The E-factor values differ depending on the industry segment, as described in the literature by Sheldon (Table 1). Waste is broadly defined as anything that is not the desired product, including all auxiliary materials such as solvents and chemicals used in workup. Originally, water was excluded from waste calculations because the use of water was considered unlikely to cause an environmental impact. However, as water disposal or reuse often necessitates pretreatment, the current trend includes water as part of the waste. A major source of waste in chemical manufacture is solvent and water losses. Therefore, the use of new related E-factors, complete E-factors (cEF) and simple E-factors (sEF), has been suggested.906 The cEF accounts for all process materials including solvents and water, whereas the sEF metric excludes water and solvents from the calculation (Figure 3a-i). The sEF is particularly suited for evaluating processes during the early development stage, where solvent and water use may not yet be optimized (Figure 3a-ii).

Figure 3.

Figure 3

Green metrics. (a) Atom Economy (AE) and (b) Environmental Impact Factor (E-factor).

Table 1. E-Factors in the Chemical Industry Described by Sheldon.906,907,910,911.

industry segment product tonnage E-factor (kg waste/kg product)
oil refining 106–108 <0.1
bulk chemicals 104–106 <1–5
fine chemicals 102–102 5–>50
pharmaceuticals 10–103 25–>100

The loss of iodine in reactions using aryl iodides as the starting materials results in waste generation and reduced atom economy because of the large atomic weight of iodine (∼160). Nevertheless, aryl iodides are more reactive than other aryl halides, such as aryl bromides and chlorides, and enable various transformations even in the absence of transition metal catalysts, thus helping avoid exposure to rare and toxic metal species. Aryl iodides are oxidized to generate hypervalent iodine reagents, which can be used as nonmetal oxidants. Diaryliodonium salts are stable and easily accessible compounds that can be converted into various products via ligand coupling or aryne generation. In addition to their oxidative transformations, aryl iodides can be activated by relatively low-energy techniques, such as irradiation or electronic treatment, to promote the generation of aryl radical intermediates.

In this section, we select representative bond formations using iodoarenes as starting materials and introduce the developments concerning green sustainable chemistry. Specifically, we compare the related preparation methods and efficiency of phenol O-arylation, the benefits of using diaryliodonium salts as aryne precursors, and the sequential double functionalization affording benzofuran skeletons. For the direct activation of iodoarenes inducing bond-forming reactions, we selected borylation and hydroxylation initiated by irradiation or electric stimuli and compared their conditions from a green chemistry perspective. The reactions using diaryliodonium salts or iodoarenes described here contribute to safer processes, waste reduction, high yields, selective transformations, short-step synthesis, energy efficiency, and green solvent use.

7.1. Iodoarene Syntheses Based on Green Sustainability

Numerous methods for aryl iodination have been developed to date.912914 Although various combinations of iodine sources with activating additives have been employed, this review highlights selected methods emphasizing green sustainability without the use of transition metals (Scheme 195).

Scheme 195. Transition Metal-Free Synthesis of Iodoarenes.

Scheme 195

For the iodination of the aryl C–H bond, electrophilic aromatic substitution of arenes has been recognized as a fundamental and widely accessible approach (Scheme 195a).914,915 Elemental iodine, as a cost-effective and atom-economical iodine source, is advantageous (Scheme 195a-i). However, its low reactivity renders it incapable of iodinating aromatic rings on its own, and it generates hydrogen iodide as a byproduct during the reaction. To enhance its reactivity, various acids and/or oxidizing reagents are typically employed. Acids promote the polarization of elemental iodine to increase its electrophilicity, whereas oxidants generate highly reactive iodonium cation-like species. Under oxidative conditions, iodide anions released during the reaction are further oxidized into active species, improving iodine atom economy. Heavy metal-based oxidants, including compounds of silver,916,917 mercury,918,919 lead,920 chromium,921 cerium,922 iron,923,924 and copper,925 have been commonly used. However, from the perspective of green and sustainable chemistry, these reagents should be avoided. Instead, nonmetallic or earth-abundant alkali metal oxidants, such as nitric acid,926 sodium nitrates (NaNO3),927 sodium peroxosulfates (Na2S2O8),928,929 hydrogen peroxide,930932 Selectfluor,933,934 and iodine(III, V, or VII) compounds,935937 have been employed as alternatives. Unsubstituted aromatic hydrocarbons and electron-rich aromatic compounds undergo iodination in the presence of elemental iodine and these oxidants. In contrast, electron-deficient arenes bearing nitro and carbonyl groups require severe reaction conditions using concentrated sulfuric acid due to their low nucleophilicity.938,939 Similarly, the iodination of heteroarenes depends on their electronic density, and the reaction conditions are adjusted accordingly.

Iodide salts, such as alkali metal or ammonium iodides, are also effective iodine sources generating iodonium cationic-species under oxidative conditions (Scheme 195a-ii). These salts are advantageous because of their ease of handling and lower volatility compared to sublimable iodine.914,915 In addition to common oxidants,940942 nitrosonium tetrafluoroborate (NOBF4),943 potassium bromate (KBrO3),944 potassium chlorate (KClO3),945 potassium iodate (KIO3),946 sodium hypochlorite (NaOCl),947 Oxone,948 and dimethyl sulfoxide949 have been employed to oxidize iodide salts.

N-Iodo compounds, such as N-iodosuccinimide (NIS)950954 and 1,3-diiodo-5,5-dimethylhydantoin,955957 serve as efficient iodine sources, though they generate stoichiometric amounts of waste during the iodination process. The combination of NIS with suitable acids, such as triflic acid (CF3SO3H), boron trifluoride etherate (BF3·OEt2), or trifluoroacetic acid, enables the iodination of electron-deficient arenes under milder conditions compared to methods involving elemental iodine or iodide salts (Scheme 195a-iii).951953 In addition to these iodinating agents, other electrophilic iodinating reagents, such as iodochloride,958 bis(pyridine)iodonium tetrafluoroborate,959,960 and N-iodosaccharin,961 have also been developed for aryl C–H bond iodinations. However, these reagents suffer from drawbacks, including high energy consumption during their preparation and the generation of substantial amounts of waste during the reaction.

Aryl iodides can be also synthesized through the substitution of other functional groups with iodine atoms, starting from prefunctionalized aromatic compounds (Scheme 195b). This approach allows for the introduction of iodine atoms based on the substituent positions, thereby enabling the synthesis of a diverse range of iodoarenes. For instance, treatment of arylamines with nitrate leads to Griess diazotization, producing aryl diazonium salts that react with alkali metal iodides to furnish the corresponding aryl iodides in a one-pot procedure (Scheme 195b-i).962965 Aromatic carboxylic acids undergo decarboxylative iodination to form aryl iodides. Although heavy-metal (mercury or lead) reagents have been employed in the past,966,967 this transformation can be induced by nonmetallic reagents, such as Burton ester968 and hypervalent iodines (Scheme 195b-ii),969971 albeit with significant amounts of waste. Moreover, decarboxylative iodination under simple conditions using elemental iodine and inorganic bases has been reported (Scheme 195b-iii).972 Halogen exchange reactions, such as the substitution of aryl bromides or chlorides, offer another efficient route to iodoarenes. Nucleophilic aromatic substitution973 and the transition metal-catalyzed Finkelstein reaction974976 provide the iodoarene synthesis, although these methods suffer from substrate limitation or reliance on transition metals. Recent advances demonstrated a photoinduced Finkelstein reaction using sodium iodide and catalytic amount of elemental iodine, providing a more sustainable alternative (Scheme 195b-iv).977 In addition, aryl boronic acids,978,979 trifluoroborate,980 and triflate981 are effective precursors for aryl iodide synthesis, although their synthesis may consume energy and exhibit low atom economy during the iodination process.

Direct iodination of aromatic compounds using elemental iodine presents the most straightforward and cost-effective approach, offering ideal highly atom-economic transformation. However, elemental iodine is a known irritant to the eyes, skin, and respiratory tract,982 and its activation requires strong acids and/or oxidants. In contrast, iodide salts are easy-to-handle iodine sources that generate iodonium cation-like species in the presence of oxidants. Despite these challenges, the industrial production of various iodoarenes has been established, making them commercially available. NIS serves as a complementary iodinating agent for less reactive arenes and is available commercially, even though it generates waste and thus exhibits a low atom economy. Indirect iodination of prefunctionalized arenes also provides an efficient approach for synthesizing diverse iodoarenes, particularly when the starting materials are easily accessible, though this method requires additional reaction steps.

7.2. Reagent Design for Efficient Diaryliodonium Salts

7.2.1. Ligand Control for Highly Reactive and Selective Arylation

Diaryliodonium salts are efficient arylation reagents for various nucleophiles to form aryl–carbon or aryl–heteroatom bonds in the absence of transition metal catalysts. These reactions involve ligand exchange between the nucleophile and diaryliodonium salt followed by ligand coupling between the nucleophile and one aryl group. These types of reactions were first reported in the 1950s, as exemplified by the reaction of diphenyliodonium bromide with sodium phenoxide to afford diphenyl ether.103,104 In this section, we compare the reported O-arylations of phenols with various types of diaryliodonium salts and their preparation methods (Scheme 196).

Scheme 196. Comparison of Various Diaryliodonium Salts: Preparation and Phenol O-Arylation.

Scheme 196

Diaryliodonium bromides, which were employed in early reports, are unstable and can be prepared using iodosobenzene or iodoxybenzene, which are explosive oxidants.103 Numerous works have aimed to develop easily accessible and stable diaryliodonium salts bearing various counteranions and protocols for the efficient arylation of various nucleophiles under mild conditions. Olofsson et al. reported the O-arylation of phenol derivatives with diaryliodonium triflates or tetrafluoroborates in the presence of bases, such as KOtBu, NaH, or NaOH.195197 These iodonium salts are relatively stable and can be prepared using a mild oxidant, mCPBA. Gaunt et al. used fluoride as the counteranion of diaryliodonium salts, revealing that the reaction of diaryliodonium fluoride with phenol derivatives proceeds effectively even in the presence of a weak base, NaHCO3 and suggesting that the fluoride anion assists the activation of hydroxy groups.198 When the diaryliodonium salts have two different aryl groups, both of them can participate in bond formation during arylation. To achieve the unified selectivity of aryl transfer, TMP group-bearing iodonium salts have been used.190,219 In this case, the other aryl group is transferred during arylation, and the TMP group is converted to TMP–I. The synthesis of TMP-iodonium salts has been well studied and can be achieved using one-pot procedures.156,161,162 Kita and Dohi achieved a highly reactive arylation using TMP-iodonium acetates, wherein the TMP ligand and acetate anion assisted the activation of a hydroxy group and accelerated the following ligand coupling.220 Diaryliodonium acetates can be prepared using 9% peracetic acid in acetic acid as a green oxidant rather than mCPBA; peracetic acid generates acetic acid upon oxidation, whereas mCPBA forms chlorobenzoic acid as waste.170,221

7.2.2. Easily Accessible Aryne Precursors

Arynes are useful synthetic intermediates for constructing functionalized aromatic rings via carbon–carbon or carbon–heteroatom bond formation (Scheme 197).983 Among the available aryne generation methods, those relying on o-silylaryl triflates are most frequently employed, generating aryne intermediates under almost neutral conditions.984 Nevertheless, the preparation of these precursors generally requires multistep reactions involving the halogen–metal exchange of o-halophenols or ortho-metalation of phenol derivatives. In addition, both silyl and triflate groups serve as auxiliaries for aryne generation, which results in the production of considerable amounts of waste. Various aryne precursors have been developed to improve preparation accessibility and reduce waste. Iodoarenes and diaryliodonium salts also serve as aryne precursors. o-Iodoaryl sulfonates are converted into arynes via halogen–metal exchange with n-BuLi or i-PrMgCl followed by the elimination of sulfonate anion985,986 and can be prepared in one step from the corresponding 2-iodophenols, although the use of strong bases imposes limitations on substrate diversification. In contrast, when diaryliodonium salts are employed as aryne precursors, the iodonium moieties serve as leaving groups after ortho activation. Diaryliodonium salts bearing ortho-silyl or boryl groups generate arynes in the presence of the fluoride anion or water as activators, respectively.418,987 However, the syntheses of these precursors require multiple steps. Diaryliodonium salts without ortho functional groups also generate arynes via ortho deprotonation by appropriate bases, such as LDA or KOtBu.414 In this case, aryne precursors can be prepared from relatively simple arenes or iodoarenes without the introduction of activating groups. Stuart et al. used Mes-iodonium salts as stable aryne precursors easily accessible by well-established methods.237,423 In addition, the iodomesitylene generated during aryne formation can be reused in the preparation of aryne precursors. This approach provides the step economy and waste reduction process for the synthesis of highly functionalized arenes.

Scheme 197. Comparison of Various Aryne Precursors: Generation and Versatility.

Scheme 197

7.2.3. Double Functionalization of Diaryliodonium Salts

The above-mentioned aryne generation from Mes-iodonium salts represents the double functionalization of aryliodines(III) and ortho aryl–H bonds attached to one aryl group of diaryliodonium salts,120 which provides a step-economical synthesis of multifunctionalized arene derivatives. In this section, we take the synthesis of the benzofuran skeleton as an example and compare it with other approaches. Benzofurans are important motifs found in various bioactive compounds and can be synthesized via several approaches.988993 When constructing benzofurans from benzene rings, one needs to incorporate carbon and oxygen atoms in mutual ortho positions. 2-Halophenols are useful starting materials, featuring a preinstalled oxygen atom and aryl halide moieties enabling aryl–carbon bond formation. For example, the most adopted approach involves transition metal–catalyzed coupling with terminal alkynes generating 2-alkynylphenols followed by intramolecular cyclization,9941001 which has also been applied to a one-pot procedure (Scheme 198a).10021009 Relatively simple phenols not bearing halogen atoms at ortho positions can be also converted into benzofuran derivatives via transition metal–catalyzed aryl C–H bond functionalization (Scheme 198b).10101012 When arylboronic acids or haloarenes are employed as the starting materials, sequential transformations involving the generation of O-aryl oximes via transition metal–catalyzed coupling, [3,3]-sigmatropic rearrangements, and cyclization result in efficient benzofuran construction.10131015 In these approaches, the starting materials can be prepared using established and step-economical methods, and numerous benzofuran derivatives are therefore accessible. The use of diaryliodonium salts as starting materials allows the initial O-aryl oxime formation to be carried out under transition metal–free conditions (Scheme 198c). In addition, these sequential transformations can be realized in one pot, as independently reported by three groups at almost the same time.211213

Scheme 198. Comparison of Benzofuran Synthesis Methods.

Scheme 198

7.3. Development of Aryl Iodide Borylation Methods

Arylboronates are robust and stable building blocks that can be converted into functionalized arenes via transition metal-catalyzed coupling and other reactions.6,10161020 These are typically prepared via the generation of aryl metal species followed by their reaction with trialkyl borates.10211023 Transition metal-catalyzed borylations, developed as an alternative method, require a stoichiometric amount of highly reactive organometallics or expensive and toxic transition metals.10241031 Considerable efforts have been devoted to developing cost-effective green sustainable methods for synthesizing arylboronates. Based on the breakthrough discoveries of the boron–boron bond cleavage by Hoveyda1032 and Fernández,1033 several transition metal-free borylation reactions using diboron compounds have been developed. The reaction of aryl iodides with pinacol diboron (4.0 equiv) in the presence of inorganic bases, such as cesium carbonate, in MeOH as a green solvent under reflux conditions affords arylboronic acid pinacol esters, with the sEF value was calculated to be 13.3 (Scheme 199a).577 The addition of a 4-phenylpyridine catalyst improves the corresponding yields by stabilizing the generated boryl intermediates, with the sEF value decreasing to 4.0 owing to the reduced amount of B2pin2 (2.0 equiv) (Scheme 199b).580,581

Scheme 199. Comparison of Synthetic Approaches to Arylboronic Esters.

Scheme 199

The sEF values were calculated based on the information provided in the literature. Reagents required for work-up were excluded.

Estimated values for Ph-Bcat intermediate.

To construct an energy-effective system without heating conditions, light energy can activate aryl halides and/or diboron reagents at lower temperatures.1034 The combination of compact fluorescent light (CFL) with a metal-based photocatalyst1035 or high-pressure mercury lamp induces borylation in the presence of organic bases with low sEF values (Scheme 199c).726,727 In the latter case, the reaction was applied to a continuous processing system. Furthermore, additive-free photoinduced borylations were demonstrated under UV (254 nm) or blue LED irradiation, which reduced final waste (Scheme 199d).724,728,729 The former (using only 2.0 equiv of B2pin2 and resulting in a high yield) exhibits a lower sEF value, whereas the latter (using 4 equiv of B2cat2 and leading to a moderate yield) shows a higher sEF value. The electrochemical borylation of aryl iodides was also reported, although the addition of an inorganic base and heating conditions were still required (Scheme 199e).860 In this case, 4.0 equiv of B2pin2 and cesium carbonate were employed, resulting in an increased sEF value of 11.1.

7.4. Development of Aryl Iodide Hydroxylation Methods

Phenol derivatives are one of the fundamental motifs of natural products, bioactive compounds, and organic functional materials;10361038 thus, various synthetic approaches to these derivatives have been developed.1039,1040 The transition metal-catalyzed coupling reactions of aryl halides with hydroxide anions from inorganic bases or hydroxide surrogates are robust approach,10411048 although the use of strong bases, such as the hydroxide anion, results in undesirable side reactions, whereas the reactions of hydroxide surrogates generate chemical waste. Considering the principles of green and sustainable chemistry, effective hydroxylation using alternative hydroxide sources under milder conditions is desired. Molecular oxygen was used as an alternative hydroxide source in the photoinduced hydroxylation of aryl iodides based on the photoinduced homolytic cleavage of aryl–iodine bonds used in borylation and phosphinylation reactions (Scheme 200a).721 In this reaction, aryl radicals were generated and trapped by molecular oxygen to generate aryl peroxyl radicals, which were converted into hydroxyarenes in the presence of triethylamine. Triethylamine is the only additive, resulting in low sEF values.

Scheme 200. Comparison of Synthetic Approaches to Hydroxyarenes.

Scheme 200

The sEF values were calculated based on the information provided in the literature. Reagents required for the work-up were excluded.

Furthermore, hydroxylation methods using water as a hydroxide source were developed, providing a safer protocol for accident prevention by avoiding the use of combustion-supporting oxygen gas. Aryl halides reacted with water to generate hydroxyarenes in a hybrid catalytic system comprising graphitic carbon nitride and a nickel complex under xenon lamp irradiation via the activation of both aryl halides and water (Scheme 200b).1049 Under these reaction conditions, although the number of additives increases, high reaction yields result in low sEF values. This semiheterogeneous catalytic system could potentially be applied to continuous processing. The electrochemical hydroxylation of a broad variety of aryl iodides with water was also reported (Scheme 200c).855 The sEF value is slightly higher due to the addition of ammonium hexafluorophosphate. Nevertheless, this reaction proceeded in water, the most recommended green solvent.

8. Summary and Outlook

Transition metal–catalyzed arylations suffer from several drawbacks, including their environmental and economic impacts. Therefore, considerable attention has been drawn to the development of green and sustainable strategies for constructing diverse aromatic frameworks under mild conditions. This review discusses the rapidly growing fields of (hetero)aryl halide activation for the direct (hetero)arylation of trapping reagents and the construction of chemical bonds under benign conditions, focusing on the popular strategies applicable to the activation of aryl halides. The hypervalent activation of iodoarenes via the formation of diaryliodonium salts provides a unique strategy for dissociating Ar–I bonds and generating highly reactive intermediates. The reactivity of diaryliodonium salts is attributed to the presence of σ-hole(s) on the I(III) surface, which allows these salts to act as Lewis acids in addition to being reactive arylating reagents. The feasibility of Ar–I(III) bond cleavage and the generation of formal aryl cation, aryl radical, and aryne intermediates enables the construction of diverse C–C and C–heteroatom bonds under transition metal-free benign (thermal and photochemical) conditions. Symmetrical diaryliodonium salts were first used as aryl transfer reagents to increase chemoselectivity. Interestingly, diaryliodonium salts bearing sterically congested and/or electron-rich aryl groups as dummy (auxiliary) ligands were found to be suitable for achieving the highly chemoselective transfer of the aryl electrophile. In view of their exceptionally high reactivity in SNAr reactions of diaryliodonium salts can become common arylating reagents for the synthesis of radiolabeled molecules for biomolecular imaging (PET).

Despite the usefulness of this transition metal-free direct arylation protocol, it suffers from the production of aryl iodide waste (and therefore exhibits a low atom economy of 10–20%) and potential purification problems. Therefore, we discussed the recent successful strategies for tackling this issue and maximizing the arylation atom efficiency. One strategy makes use of the two aryl groups of the iodonium salt and their incorporation into the final product through one-pot cascade reactions. Alternatively, one can recycle the aryl iodide byproduct. The recycling of iodoarene waste can not only solve the main problem of this approach and provide an attractive transition metal-free strategy for the C–H functionalization of unactivated (hetero)arenes but also enable the use of aryl iodides in catalytic amounts. Thus, we hope that this review will be of interest to researchers aiming to develop new methods of solving the problems associated with this piece of chemistry.

Base-induced Ar–I bond dissociation through SET from a ground-state electron-donating intermediate under organic promoter and organic promoter–free conditions was overviewed. The addition of an organic promoter significantly improves the rate of the transformation through the cascade reaction of a base (normally KOtBu) and this promoter to in situ generate strong electron donors capable of reducing Ar–X through SET. Under organic promoter–free conditions, the electrons could originate from the combinations of base/high reaction temperature, reaction of a base with the solvent, and generation of strongly electron- donating intermediates or base-activated halogen bond–assisted Ar–X bond cleavage. Notably, the generation of strong electron donors even in very low yields is sufficient to initiate the chain mechanism. These protocols employ unique organic electron donors with tunable reduction ability for the conversion of aryl halides into the highly reactive aryl radical and aryl anion intermediates, which participate in the formation of diverse bonds and heterocyclic skeletons under benign transition metal-free conditions. The advantages of this sustainable chemistry will promote further studies to (a) fully understand the reaction mechanism and provide a full structural picture of the in situ generated organic electron-donating intermediates to guide the design of more powerful promoters; (b) overcome the drawbacks of regioselectivity and the use of substrates in large molar excess (as solvents); (c) investigate other ways of accessing the SET process; (d) establish general methods with broad substrates scope for the highly chemo- and regioselective syntheses of highly functionalized products; (e) develop novel organic electron donors from easily accessible starting materials capable of reducing not only the electronically and sterically diverse aryl iodides but also the more challenging aryl bromides, chlorides, and fluorides.

In addition to the ground-sat thermal ways of activations, the situation is more attractive for the excited-state reactions, with photoinduced aryl–halogen bond cleavage playing a vital role as a complementary/alternative sustainable strategy of transition metal–catalyzed processes for the derivatization of haloarenes under mild conditions. In the context of this review, we highlighted the possible transition metal-free sustainable strategies for the photoactivation of haloarenes. The most convenient approach to light-induced aryl–iodine bond cleavage under metal- and photocatalyst-free conditions is the direct excitation of aryl iodides with UV light or SET. For SET to occur under photocatalyst-free conditions, the substrate should act as a photoreductant and transfer a single electron to the desired aryl halide under irradiation; alternatively, the aryl iodide can interact with an electron donor and form an EDA complex capable of absorbing visible light and afford an excited state where SET takes place. Another transition metal-free photochemical strategy for activating aryl halides relies on organic photocatalysts. In such reactions, the excited photocatalysts can reduce aryl halides via SET. Moreover, XAT activation is a useful methodology employed for Ar–I bond cleavage with no need for strong reductants. These activation reactions result in Ar–X bond cleavage and generate aryl radicals for C–C and C–heteroatom bond formation with high functional group tolerance under environmentally benign conditions.

Despite the progress in and advantages of the transition metal-free photoactivation of aryl halides, further studies in the following directions should develop valuable alternatives to transition metal–catalyzed processes: (a) Expand the present strategies and discover more tactics for the metal- and photocatalyst-free photoactivation of aryl halides; (b) design new transition metal-free organic photocatalysts with suitable redox potentials for improving the catalytic performance of photochemical transformations, overcome the challenges of aryl halide substrates, and enhance the scope and diversity of the photochemical process; (c) develop novel photochemical transformations applicable to the synthesis of challenging products, including natural products and pharmaceuticals; (d) understand the mechanistic details of aryl halide photoactivation to design efficient photochemical cross-couplings; (e) expand the light absorption range to enable visible light–driven reactions; (f) improve the quantum efficiency of the time-consuming photochemical processes; (g) avoid the use of sacrificial electron donors, (h) reduce the molar excess, improve the chemo- and regioselectivity, and broaden the scope of arenes in the C–H arylation of unactivated arenes with aryl halide.

Electrochemical synthesis has emerged as an efficient and green method of achieving challenging transformations using electric current as a driving force. In this case, electrons are used as a potent, controllable, traceless, and green redox reagent instead of the environmentally hazardous chemical substances, which allows one to lower the associated risks and costs, reduce waste generation, hinder side reactions, and elevates the atom and step economy of the process. Electrochemical reductive functionalization at the cathode is relatively rare compared with oxidative reactions at the anode. Considering the aim of this review, we highlighted two possible ways of inducing the electrochemical dissociation of aryl halides at the cathode and generating aryl radicals as very reactive intermediates suitable for various cross-couplings. Direct electrolysis involves direct SET from the surface of the cathode to the aryl halide and suffers from reactivity and selectivity problems. Therefore, a more convenient indirect electrolysis strategy was developed for the electroreductive functionalization of aryl halides under more benign conditions with greater reactivity, selectivity, substrate scope, functional group tolerance, etc. In this case, an organic mediator acts as a redox-active catalyst to shuttle the electrons from the cathode to aryl halide.

As reductive functionalization at the cathode is underdeveloped, we expect the synthetic community to make efforts at broadening the reaction scope, overcoming the challenges associated with the reductive dissociation of aryl–halogen bonds, and developing more selective, efficient, sustainable, and practical electrochemical approaches. These targets could be fulfilled by, i.e., (a) designing benign electroreductive strategies through, e.g., the fabrication of more efficient electrode materials and/or use of new organic mediators with suitable redox potentials; (b) designing novel organocatalysts with unique catalytic pathways and discovering new catalytic concepts; (c) avoiding the use of sacrificial reagents and electrodes through the development of paired electrolysis systems for the activation of aryl halides; (d) developing mild electrochemical processes for the late-stage functionalization of complex substrates.

Similar to conPET methodologies for the reductive functionalization of haloarenes but with no need for sacrificial electron-donating additives, electrophotochemical activation strategies were developed for haloarenes with extremely negative reduction potentials. This new strategy enables the further expansion of the electro- and photochemical approaches by combining the two chemistries in one step while maximizing their advantages and minimizing their limitations. As the electrophotochemical strategy is still in its infancy, we expect more discoveries in the near future, particularly in relation to the development of new commercially available, cheap, and potent organic mediators.

In general, we highlighted the pioneering works and discussed the recent progress in the activation of haloarenes via hypervalency, bases, and organocatalysts under the action of heat, radiation, and electricity. Further studies should deal with weaknesses of these processes to make them competitive sustainable alternatives to traditional and metal-catalyzed approaches.

Acknowledgments

We acknowledge financial support from the Japan Society for the Promotion of Science (JSPS), the Ministry of Education, Culture, Sports, Science and Technology (MEXT), and the New Energy and Industrial Technology Development Organization (NEDO). T.D., K.K., and K.M. acknowledge support from JSPS KAKENHI grant number 19K05466 (T.D.) 23K04827 (K.K.), and 23K06060 (K.M.), JST CREST grant number JPMJCR20R1, and the Ritsumeikan Global Innovation Research Organization (R-GIRO) project.

Biographies

Toshifumi Dohi received his MS degree in 2002 (Prof. S. Murai) from the Graduate School of Engineering of Osaka University, Japan, and his PhD degree in 2005 (Prof. Y. Kita) from the Graduate School of Pharmaceutical Sciences of Osaka University, where he studied new reactivities of transition metal catalysts and synthetic chemistry using hypervalent iodine reagents. After completing his PhD research, he subsequently became an Assistant Professor at Osaka University. In 2008, he relocated to Ritsumeikan University, where he was promoted to an Associate Professor in 2014 and to a full Professor in 2019. His current research interests focus on reagent/catalyst design and the development of new reactions using hypervalent iodine reagents. T.D. received the IUPAC-ICOS 15 Poster Award for excellence in presentation, the Pharmaceutical Society of Japan (PSJ) Award for Young Scientists (2009), the Banyu Chemist Award (2013), the Thieme Chemistry Journal Award (2014), the GSC Encouragement Award (2015), and the International Congress on Pure & Applied Chemistry (ICPAC) Lecture Award (2019).

Elghareeb E. Elboray received his MS degree in 2006 from South Valley University, Egypt, under the supervision of Prof. M. F. Aly, and his PhD in 2013 from the School of Chemistry, University of Leeds, UK, under the supervision of Prof. R. Grigg. In 2017, he joined Prof. T. Suzuki’s lab at Kyoto Prefectural University of Medicine, Japan, as a JSPS postdoctoral fellow. In 2019 he moved to Prof. T. Furuta’s lab at Kyoto Pharmaceutical University. Since April 2021, he has been a postdoctoral researcher in Prof. T. Dohi’s lab at Ritsumeikan University. He has broad interests, starting with the generation of dipoles and 1,3-dipolar cycloadditions, Pd-catalyzed allene chemistry and multicomponent cascades, design and synthesis of bioactive compounds, design of axially chiral catalysts and incorporation into asymmetric synthesis. He is currently focusing on hypervalent iodine chemistry and, in particular, the versatility of diaryliodonium salts.

Kotaro Kikushima received his PhD in 2010 from Osaka University, under the supervision of Prof. Toshikazu Hirao. In 2010, he joined the group of Prof. Brian M. Stoltz at the California Institute of Technology as a postdoctoral fellow, supported by a JSPS Research Fellowship for Young Scientists. In 2011, he became an Assistant Professor at Nagoya City University (Prof. Seiichi Nakamura’s group) and then moved to Okayama University as a Research Assistant, working with Prof. Yuta Nishina, in 2012. From 2014 to 2018, he was a Specially Appointed Assistant Professor at Osaka University, working with Prof. Sensuke Ogoshi. Since April 2018, he has served as an Assistant Professor at Ritsumeikan University, working with Prof. Toshifumi Dohi. In 2022, he joined the group of Prof. Richmond Sarpong at University of California Berkeley as a visiting researcher. His current research interests include developing cost-effective methods for the synthesis of organofluorine and heterocyclic compounds. He has received the Tosoh Corporation Award in Synthetic Organic Chemistry, Japan (2012), the CSJ Presentation Award (2015), the seventh International Conference on Green and Sustainable Chemistry and the fourth JACI/GSC Symposium Poster Award (2015), and the Award for Young Scientists from the Pharmaceutical Society of Japan, Kansai Branch (2023).

Koji Morimoto studied organic chemistry at the Graduate School of Pharmaceutical Sciences, Osaka University (Prof. Y. Kita, 2000–2009), where he obtained his MS and PhD degrees, while studying synthetic chemistry using hypervalent iodine reagents. After obtaining his PhD, he worked as a researcher at Ritsumeikan University with Professor Yasuyuki Kita. In 2016, he became an Assistant Professor at Ritsumeikan University. He received the Award for Young Scientists at the Pharmaceutical Kinki Branch (2010) and the Award for the Division of Organic Chemistry, the Pharmaceutical Society of Japan (2017). His current research interests are in the field of coupling reactions of aromatic compounds having heteroatoms using hypervalent iodine reagents.

Yasuyuki Kita received his PhD (1972) from Osaka University and subsequently became a member of the Faculty of Pharmaceutical Sciences of the University. After two years (1975–1977) of postdoctoral work with Professor George Büchi at MIT, he moved back to Osaka University. He was promoted to Associate Professor in 1983 and to Full Professor of Osaka University in 1992. In 2008, he retired from Osaka University and joined Ritsumeikan University as the Dean of the Faculty of Pharmaceutical Sciences. From 2011 to 2015, he was Vice President of the Research Organization of Science and Technology, Ritsumeikan University. Since April 2015, he has served as an Invited Research Professor and Director of the Research Center for Drug Discovery and Pharmaceutical Development Science at the same university. His research interests cover a wide range of topics in synthetic organic chemistry, including the development of new asymmetric synthesis and new reagents and the total synthesis of biologically active natural products. His current research interest is in hypervalent iodine chemistry. He has published more than 510 original papers. His awards include the Pharmaceutical Society of Japan (PSJ) Award for Young Scientists (1986), the PSJ Award for Divisional Scientific Contribution (1997), the PSJ Award (2002), the Japanese Society for Process Chemistry (JSPC) Award for Excellence (2005), the Society of Iodine Science (SIS) Award (2007), the E.C. Taylor Senior Award (2017), and the Hoansha Award for the 70th Anniversary (2024).

Author Contributions

Elghareeb E. Elboray and Kotaro Kikushima contributed equally to the work, and their order is listed alphabetically. Toshifumi Dohi conceived the project and provided supervision and prepared The Proposal of this review with Dr. Elboray. Elghareeb E. Elboray drafted the original manuscript for Sections 1–6 and 8. Kotaro Kikushima drafted the original manuscript for Section 7 and a part of Section 2, modified graphics in all sections, and revised and edited all sections. Koji Morimoto reviewed the manuscript. Yasuyuki Kita conceived the project and provided supervision. CRediT: Toshifumi Dohi conceptualization, project administration, supervision, writing - original draft, writing - review & editing; Elghareeb E. Elboray investigation, writing - original draft; Kotaro Kikushima investigation, writing - original draft, writing - review & editing; Koji Morimoto investigation, writing - review & editing; Yasuyuki Kita project administration, supervision.

The authors declare no competing financial interest.

References

  1. de Meijere A.; Bräse S.; Oestreich M.. Metal-Catalyzed Cross-Coupling Reactions and More; Wiley: 2013. [Google Scholar]
  2. Colacot T. J.New Trends in Cross-Coupling: Theory and Applications; Royal Society of Chemistry: 2014. [Google Scholar]
  3. Palani V.; Perea M. A.; Sarpong R. Site-Selective Cross-Coupling of Polyhalogenated Arenes and Heteroarenes with Identical Halogen Groups. Chem. Rev. 2022, 122, 10126–10169. 10.1021/acs.chemrev.1c00513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Heck R. Palladium-catalyzed reaction of organic halides with olefins. Acc. Chem. Res. 1979, 12, 146–151. 10.1021/ar50136a006. [DOI] [Google Scholar]
  5. Negishi E. Palladium- or nickel-catalyzed cross coupling. A new selective method for carbon-carbon bond formation. Acc. Chem. Res. 1982, 15, 340–348. 10.1021/ar00083a001. [DOI] [Google Scholar]
  6. Suzuki A. Organoborates in new synthetic reactions. Acc. Chem. Res. 1982, 15, 178–184. 10.1021/ar00078a003. [DOI] [Google Scholar]
  7. Wolfe J. P.; Wagaw S.; Marcoux J. F.; Buchwald S. L. Rational Development of Practical Catalysts for Aromatic Carbon-Nitrogen Bond Formation. Acc. Chem. Res. 1998, 31, 805–818. 10.1021/ar9600650. [DOI] [Google Scholar]
  8. Hartwig J. F. Carbon-Heteroatom Bond-Forming Reductive Eliminations of Amines, Ethers, and Sulfides. Acc. Chem. Res. 1998, 31, 852–860. 10.1021/ar970282g. [DOI] [Google Scholar]
  9. Nicolaou K. C.; Bulger P. G.; Sarlah D. Palladium-Catalyzed Cross-Coupling Reactions in Total Synthesis. Angew. Chem., Int. Ed. 2005, 44, 4442–4489. 10.1002/anie.200500368. [DOI] [PubMed] [Google Scholar]
  10. Corbet J.-P.; Mignani G. Selected Patented Cross-Coupling Reaction Technologies. Chem. Rev. 2006, 106, 2651–2710. 10.1021/cr0505268. [DOI] [PubMed] [Google Scholar]
  11. Gallagher W. P.; Vo A. Dithiocarbamates: Reagents for the Removal of Transition Metals from Organic Reaction Media. Org. Process Res. Dev. 2015, 19, 1369–1373. 10.1021/op500336h. [DOI] [Google Scholar]
  12. Forfar L.C.; Murray P.M., Meeting Metal Limits in Pharmaceutical Processes. In Organometallics in Process Chemistry. Colacot T.; Sivakumar V., Eds.; Topics in Organometallic Chemistry; Springer, Cham, 2018, Vol. 65, pp 217–252. [Google Scholar]
  13. McCarthy S.; Braddock D. C.; Wilton-Ely J. D. E. T. Strategies for sustainable palladium catalysis. Coord. Chem. Rev. 2021, 442, 213925. 10.1016/j.ccr.2021.213925. [DOI] [Google Scholar]
  14. Kwan E. E.; Zeng Y.; Besser H. A.; Jacobsen E. N. Concerted Nucleophilic Aromatic Substitutions. Nat. Chem. 2018, 10, 917–923. 10.1038/s41557-018-0079-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Rohrbach S.; Smith A. J.; Pang J. H.; Poole D. L.; Tuttle T.; Chiba S.; Murphy J. A. Concerted Nucleophilic Aromatic Substitution Reactions. Angew. Chem., Int. Ed. 2019, 58, 16368–16388. 10.1002/anie.201902216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Stang P. J.; Zhdankin V. V. Organic. Polyvalent Iodine Compounds. Chem. Rev. 1996, 96, 1123–1178. 10.1021/cr940424+. [DOI] [PubMed] [Google Scholar]
  17. 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]
  18. 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]
  19. 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]
  20. Writh T., Ed. Hypervalent iodine chemistry. Top. Curr. Chem., vol 373; Springer, Switzerland, 2016. [Google Scholar]
  21. Ochiai M.Reactivities, Properties and Structures. In Hypervalent Iodine Chemistry Top. Curr. Chem., vol 224; Wirth T., Eds.; Springer, Berlin, Heidelberg, 2003; pp 5–68. [Google Scholar]
  22. Okuyama T.; Takino T.; Sueda T.; Ochiai M. Solvolysis of Cyclohexenyliodonium Salt, a New Precursor for the Vinyl Cation: Remarkable Nucleofugality of the Phenyliodonio Group and Evidence for Internal Return from an Intimate Ion–Molecule Pair. J. Am. Chem. Soc. 1995, 117, 3360–3367. 10.1021/ja00117a006. [DOI] [Google Scholar]
  23. Ochiai M.Organic Synthesis Using Hypervalent Organiiodines. In Chemistry of Hypervalent Compounds; Akiba K.-y., Ed.; Wiley-VCH, New York, 1999; pp 359. [Google Scholar]
  24. Dohi T.; Kita Y.. Oxidizing Agents. In Iodine Chemistry and applications; Kaiho T., Ed.; John Wiley & Sons, Hoboken, 2015; Ch. 16, pp 277–301. [Google Scholar]
  25. Liu J.; Xiong X.; Chen J.; Wang Y.; Zhu R.; Huang J. Double C-H Activation for the C-C Bond Formation Reactions. Curr. Org. Synth. 2018, 15, 882–903. 10.2174/1570179415666180720111422. [DOI] [Google Scholar]
  26. Yusubov M. S.; Zhdankin V. V. Iodine catalysis: A green alternative to transition metals in organic chemistry and technology. Resource-Efficient Technologies 2015, 1, 49–67. 10.1016/j.reffit.2015.06.001. [DOI] [Google Scholar]
  27. Hyatt I. F. D.; Dave L.; David N.; Kaur K.; Medard M.; Mowdawalla C. Hypervalent iodine reactions utilized in carbon-carbon bond formations. Org. Biomol. Chem. 2019, 17, 7822–7848. 10.1039/C9OB01267B. [DOI] [PubMed] [Google Scholar]
  28. Kandimalla S. R.; Parvathaneni S. P.; Sabitha G.; Reddy B. V. S. Recent Advances in Intramolecular Metal-Free Oxidative C-H Bond Aminations Using Hypervalent Iodine(III) Reagents. Eur. J. Org. Chem. 2019, 2019, 1687–1714. 10.1002/ejoc.201801469. [DOI] [Google Scholar]
  29. Maiti S.; Alam M. T.; Bal A.; Mal P. Nitrenium Ions from Amine-Iodine(III) Combinations. Adv. Synth. Catal. 2019, 361, 4401–4425. 10.1002/adsc.201900441. [DOI] [Google Scholar]
  30. Muñiz K. Promoting Intermolecular C-N Bond Formation under the Auspices of Iodine(III). Acc. Chem. Res. 2018, 51, 1507–1519. 10.1021/acs.accounts.8b00137. [DOI] [PubMed] [Google Scholar]
  31. Arnold A. M.; Ulmer A.; Gulder T. Advances in Iodine(III)-Mediated Halogenations: A Versatile Tool to Explore New Reactivities and Selectivities. Chem.—Eur. J. 2016, 22, 8728–8739. 10.1002/chem.201600449. [DOI] [PubMed] [Google Scholar]
  32. Kohlhepp S. V.; Gulder T. Hypervalent iodine(III) fluorinations of alkenes and diazo compounds: new opportunities in fluorination chemistry. Chem. Soc. Rev. 2016, 45, 6270–6288. 10.1039/C6CS00361C. [DOI] [PubMed] [Google Scholar]
  33. Han Z.-Z.; Zhang C.-P. Fluorination and Fluoroalkylation Reactions Mediated by Hypervalent Iodine Reagents. Adv. Synth. Catal. 2020, 362, 4256–4292. 10.1002/adsc.202000750. [DOI] [Google Scholar]
  34. Murarka S.; Antonchick A. P. Oxidative Heterocycle Formation Using Hypervalent Iodine(III) Reagents. Top. Curr. Chem. 2015, 373, 75–104. 10.1007/128_2015_647. [DOI] [PubMed] [Google Scholar]
  35. Budhwan R.; Yadav S.; Murarka S. Late stage functionalization of heterocycles using hypervalent iodine(III) reagents. Org. Biomol. Chem. 2019, 17, 6326–6341. 10.1039/C9OB00694J. [DOI] [PubMed] [Google Scholar]
  36. Singh F. V.; Shetgaonkar S. E.; Krishnan M.; Wirth T. Progress in Organocatalysis with Hypervalent Iodine Catalysts. Chem. Soc. Rev. 2022, 51, 8102–8139. 10.1039/D2CS00206J. [DOI] [PubMed] [Google Scholar]
  37. Yoshimura A.; Saito A.; Zhdankin V. V. Recent Progress in Synthetic Applications of Cyclic Hypervalent Iodine(III) Reagents. Adv. Synth. Catal. 2023, 365, 2653–2675. 10.1002/adsc.202300275. [DOI] [Google Scholar]
  38. Dohi T.; Kita Y.. Hypervalent Iodine-Induced Oxidative Couplings (New Metal-Free Coupling Advances and Their Applications in Natural Product Syntheses). In Hypervalent Iodine Chemistry. Top. Curr. Chem., vol 373; Wirth T., Eds.; Springer, Cham, 2016. [DOI] [PubMed] [Google Scholar]
  39. Dohi T.; Kita Y.. Hypervalent Iodine. In Iodine Chemistry and applications; Kaiho T., Ed.; John Wiley & Sons, Hoboken, 2015; Ch. 7, pp 103–157. [Google Scholar]
  40. Kumar R.; Sharma N.; Prakash O. Hypervalent Iodine Reagents in the Synthesis of Flavonoids and Related Compounds. Curr. Org. Chem. 2020, 24, 2031–2047. 10.2174/1385272824999200420074551. [DOI] [Google Scholar]
  41. Wang Z. Innovation of hypervalent(III) iodine in the synthesis of natural products. New J. Chem. 2021, 45, 509–516. 10.1039/D0NJ05078D. [DOI] [Google Scholar]
  42. Chen W. W.; Cuenca A. B.; Shafir A. The Power of Iodane-Guided C-H Coupling: A Group Transfer Strategy in Which a Halogen Works for Its Money. Angew. Chem., Int. Ed. 2020, 59, 16294–16309. 10.1002/anie.201908418. [DOI] [PubMed] [Google Scholar]
  43. Zhang B.; Li X.; Guo B.; Du Y. Hypervalent iodine reagent-mediated reactions involving rearrangement processes. Chem. Commun. 2020, 56, 14119–14136. 10.1039/D0CC05354F. [DOI] [PubMed] [Google Scholar]
  44. Maertens G.; Canesi S. Rearrangements Induced by Hypervalent Iodine. Top. Curr. Chem. 2015, 373, 223–242. 10.1007/128_2015_657. [DOI] [PubMed] [Google Scholar]
  45. Dohi T.; Maruyama A.; Yoshimura M.; Morimoto K.; Tohma H.; Kita Y. Versatile Hypervalent-Iodine(III)-Catalyzed Oxidations with m-Chloroperbenzoic Acid as a Cooxidant. Angew. Chem., Int. Ed. 2005, 44, 6193–6196. 10.1002/anie.200501688. [DOI] [PubMed] [Google Scholar]
  46. Ochiai M.; Takeuchi Y.; Katayama T.; Sueda T.; Miyamoto K. Iodobenzene-Catalyzed α-Acetoxylation of Ketones. In Situ Generation of Hypervalent (Diacyloxyiodo)benzenes Using m-Chloroperbenzoic Acid. J. Am. Chem. Soc. 2005, 127, 12244–12245. 10.1021/ja0542800. [DOI] [PubMed] [Google Scholar]
  47. 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]
  48. Flores A.; Cots E.; Bergès J.; Muñiz K. Enantioselective Iodine(I/III) Catalysis in Organic Synthesis. Adv. Synth. Catal. 2019, 361, 2–25. 10.1002/adsc.201800521. [DOI] [Google Scholar]
  49. Fujita M. Enantioselective Heterocycle Formation Using Chiral Hypervalent Iodine (III). Heterocycles 2018, 96, 563–594. 10.3987/REV-17-877. [DOI] [Google Scholar]
  50. Claraz A.; Masson G. Asymmetric iodine catalysis-mediated enantioselective oxidative transformations. Org. Biomol. Chem. 2018, 16, 5386–5402. 10.1039/C8OB01378K. [DOI] [PubMed] [Google Scholar]
  51. Ghosh S.; Pradhan S.; Chatterjee I. A survey of chiral hypervalent iodine reagents in asymmetric synthesis. Beilstein J. Org. Chem. 2018, 14, 1244–1262. 10.3762/bjoc.14.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Kumar R.; Wirth T. Asymmetric Synthesis with Hypervalent Iodine Reagents. Top. Curr. Chem. 2015, 373, 243–262. 10.1007/128_2015_639. [DOI] [PubMed] [Google Scholar]
  53. Shetgaonkar S. E.; Singh F. V. Catalytic stereoselective synthesis involving hypervalent iodine-based chiral auxiliaries. Org. Biomol. Chem. 2023, 21, 4163–4180. 10.1039/D3OB00057E. [DOI] [PubMed] [Google Scholar]
  54. de Magalhães H. P.; Togni A.; Lüthi H. P. Importance of Nonclassical σ-Hole Interactions for the Reactivity of λ3-Iodane Complexes. J. Org. Chem. 2017, 82, 11799–11805. 10.1021/acs.joc.7b01716. [DOI] [PubMed] [Google Scholar]
  55. Kirshenboim O.; Kozuch S. How to Twist, Split and Warp a σ-Hole with Hypervalent Halogens. J. Phys. Chem. A 2016, 120, 9431–9445. 10.1021/acs.jpca.6b07894. [DOI] [PubMed] [Google Scholar]
  56. Lee C.-K.; Mak T. C. W.; Li W.-K.; Kirner J. F. Iodobenzene Diacetate. Acta Crystallogr. 1977, B33, 1620–1622. 10.1107/S0567740877006694. [DOI] [Google Scholar]
  57. Alcock N. W.; Countryman R. M.; Esperås S.; Sawyer J. F. Secondary Bonding. Part 5. The Crystal and Molecular Structures of Phenyliodine(III) Diacetate and Bis(dich1oroacetate). J. Chem. Soc., Dalton Trans. 1979, 854–860. 10.1039/DT9790000854. [DOI] [Google Scholar]
  58. Stergioudis G. A.; Kokkou S. C.; Bozopoulos A. P.; Rentzeperis P. J. Structure of bis(trifluoroacetato)(phenyl)iodine(III), C10H5F6IO4. Acta Crystallogr. 1984, C40, 877–879. 10.1107/S0108270184006089. [DOI] [Google Scholar]
  59. Alcock N. W.; Harrison W. D.; Howes C. Secondary Bonding. Part 13. Aryl-tellurium(IV) and -iodine(III) Acetates and Trifluoroacetates. The Crystal and Molecular Structures of Bis-(p- methoxyphenyl) tellurium Diacetatae, μ-oxo-bis[diphenyltrif1uoro-acetoxytellurium] Hydrate, and [Bis(trifluoroacetoxy)iodo] benzene. J. Chem. Soc., Dalton Trans. 1984, 1709–1716. 10.1039/DT9840001709. [DOI] [Google Scholar]
  60. Suslonov V. V.; Soldatova N. S.; Ivanov D. M.; Galmés B.; Frontera A.; Resnati G.; Postnikov P. S.; Kukushkin V. Yu.; Bokach N. A. Diaryliodonium Tetrachloroplatinates(II): Recognition of a Trifurcated Metal-Involving μ3-I···(Cl,Cl,Pt) Halogen Bond. Cryst. Growth Des. 2021, 21, 5360–5372. 10.1021/acs.cgd.1c00654. [DOI] [Google Scholar]
  61. Aliyarova I. S.; Ivanov D. M.; Soldatova N. S.; Novikov A. S.; Postnikov P. S.; Yusubov M. S.; Kukushkin V. Yu. Bifurcated Halogen Bonding Involving Diaryliodonium Cations as Iodine(III)-Based Double-σ-Hole Donors. Cryst. Growth Des. 2021, 21, 1136–1147. 10.1021/acs.cgd.0c01463. [DOI] [Google Scholar]
  62. Charpentier J.; Früh N.; Togni A. Electrophilic Trifluoromethylation by Use of Hypervalent Iodine Reagents. Chem. Rev. 2015, 115, 650–682. 10.1021/cr500223h. [DOI] [PubMed] [Google Scholar]
  63. 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]
  64. 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]
  65. Bauer A.; Maulide N. Recent discoveries on the structure of iodine(III) reagents and their use in cross-nucleophile coupling. Chem. Sci. 2021, 12, 853–864. 10.1039/D0SC03266B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. 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]
  67. Sreenithya A.; Sunoj R. B. On the activation of hypercoordinate iodine(III) compounds for reactions of current interest. Dalton Trans. 2019, 48, 4086–4093. 10.1039/C9DT00472F. [DOI] [PubMed] [Google Scholar]
  68. Soldatova N. S.; Postnikov P. S.; Suslonov V. V.; Kissler T. Yu.; Ivanov D. M.; Yusubov M. S.; Galmés B.; Frontera A.; Kukushkin V. Yu. Diaryliodonium as a double σ-hole donor: the dichotomy of thiocyanate halogen bonding provides divergent solid state arylation by diaryliodonium cations. Org. Chem. Front. 2020, 7, 2230–2242. 10.1039/D0QO00678E. [DOI] [Google Scholar]
  69. 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]
  70. Peng X.; Rahim A.; Peng W.; Jiang F.; Gu Z.; Wen S. Recent Progress in Cyclic Aryliodonium Chemistry: Syntheses and Applications. Chem. Rev. 2023, 123, 1364–1416. 10.1021/acs.chemrev.2c00591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Cheng H.-C.; Ma J.-L.; Guo P.-H. Cyclic Diaryliodonium Salts: Eco-Friendly and Versatile Building Blocks for Organic Synthesis. Adv. Synth. Catal. 2023, 365, 1112–1139. 10.1002/adsc.202201326. [DOI] [Google Scholar]
  72. Jovanovic D.; Huber S. M. Cyclic Diaryliodonium Species as Synthetic Precursors and Halogen Bonding Organocatalysts. Chem. Rev. 2023, 123, 10527–10529. 10.1021/acs.chemrev.3c00357. [DOI] [PubMed] [Google Scholar]
  73. Montgomery C. A.; Murphy G. K. Exploring the role of halogen bonding in iodonium ylides: insights into unexpected reactivity and reaction control. Beilstein J. Org. Chem. 2023, 19, 1171–1190. 10.3762/bjoc.19.86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Varvoglis A. Chemical transformations induced by hypervalent iodine reagents. Tetrahedron 1997, 53, 1179–1255. 10.1016/S0040-4020(96)00970-2. [DOI] [Google Scholar]
  75. Kitamura T.; Fujiwara Y. Recent progress in the use of hypervalent iodine reagents in organic synthesis. A review. Org. Prep. Proced. Int. 1997, 29, 409–458. 10.1080/00304949709355217. [DOI] [Google Scholar]
  76. Wirth T., Eds. Hypervalent Iodine Chemistry. Top. Curr. Chem., vol 224. Springer, Berlin, Heidelberg, 2003. [Google Scholar]
  77. Kita Y.; Dohi T. Pioneering Metal-Free Oxidative Coupling Strategy of Aromatic Compounds Using Hypervalent Iodine Reagents. Chem. Rec. 2015, 15, 886–906. 10.1002/tcr.201500020. [DOI] [PubMed] [Google Scholar]
  78. Dohi T.; Kita Y.. Oxidative C-C Bond Formation (Couplings, Cyclizations, Cyclopropanation, etc.). In Patai’s Chemistry of Functional Groups; John Wiley & Sons, Ltd, 2018. [Google Scholar]
  79. Tamura Y.; Yakura T.; Haruta J.; Kita Y. Hypervalent iodine oxidation of p-alkoxyphenols and related compounds: a general route to p-benzoquinone monoacetals and spiro lactones. J. Org. Chem. 1987, 52, 3927–3930. 10.1021/jo00226a041. [DOI] [Google Scholar]
  80. Kita Y.; Tohma H.; Inagaki M.; Hatanaka K.; Yakura Y. A novel oxidative azidation of aromatic compounds with hypervalent iodine reagent, phenyliodine(III) bis(trifluoroacetate) (PIFA) and trimethylsilyl azide. Tetrahedron Lett. 1991, 32, 4321–4324. 10.1016/S0040-4039(00)92160-9. [DOI] [Google Scholar]
  81. Kita Y.; Tohma H.; Hatanaka K.; Takada T.; Fujita S.; Mitoh S.; Sakurai H.; Oka S. Hypervalent Iodine-Induced Nucleophilic Substitution of para-Substituted Phenol Ethers. Generation of Cation Radicals as Reactive Intermediates. J. Am. Chem. Soc. 1994, 116, 3684–3691. 10.1021/ja00088a003. [DOI] [Google Scholar]
  82. Kita Y.; Takada T.; Tohma H. Hypervalent iodine reagents in organic synthesis: Nucleophilic substitution of p-substituted phenol ethers. Pure Appl. Chem. 1996, 68, 627–630. 10.1351/pac199668030627. [DOI] [Google Scholar]
  83. Kita Y.; Takada T.; Mihara S.; Tohma H. A Novel and Direct Sulfenylation of Phenol Ethers Using Phenyl Iodine(III) Bis(trifluoroacetate) (PIFA) and Various Thiophenols. Synlett 1995, 1995, 211–212. 10.1055/s-1995-4907. [DOI] [Google Scholar]
  84. Kita Y.; Takada T.; Mihara S.; Whelan B. A.; Tohma H. Novel and Direct Nucleophilic Sulfenylation and Thiocyanation of Phenol Ethers Using a Hypervalent Iodine(III) Reagent. J. Org. Chem. 1995, 60, 7144–7148. 10.1021/jo00127a018. [DOI] [Google Scholar]
  85. Arisawa M.; Ramesh N. G.; Nakajima M.; Tohma H.; Kita Y. Hypervalent Iodine(III)-Induced Intramolecular Cyclization of α-(Aryl)alkyl-β-dicarbonyl Compounds: A Convenient Synthesis of Benzannulated and Spirobenzannulated Compounds. J. Org. Chem. 2001, 66, 59–65. 10.1021/jo000953f. [DOI] [PubMed] [Google Scholar]
  86. Kita Y.; Gyoten M.; Ohtsubo M.; Tohma H.; Takada T. Non-phenolic oxidative coupling of phenol ether derivatives using phenyliodine(III) bis(trifluoroacetate). Chem. Commun. 1996, 1481–1482. 10.1039/cc9960001481. [DOI] [Google Scholar]
  87. Takada T.; Arisawa M.; Gyoten M.; Hamada R.; Tohma H.; Kita Y. Oxidative Biaryl Coupling Reaction of Phenol Ether Derivatives Using a Hypervalent Iodine(III) Reagent. J. Org. Chem. 1998, 63, 7698–7706. 10.1021/jo980704f. [DOI] [Google Scholar]
  88. Hamamoto H.; Anilkumar G.; Tohma H.; Kita Y. A novel and efficient oxidative biaryl coupling reaction of phenol ether derivatives using a combination of hypervalent iodine(iii) reagent and heteropoly acid. Chem. Commun. 2002, 450–451. 10.1039/b111178g. [DOI] [PubMed] [Google Scholar]
  89. Arisawa M.; Utsumi S.; Nakajima M.; Ramesh N. G.; Tohma H.; Kita Y. An unexpected intermolecular chiral biaryl coupling reaction induced by a hypervalent iodine(III) reagent: synthesis of potential precursor for ellagitannin. Chem. Commun. 1999, 469–470. 10.1039/a809680e. [DOI] [Google Scholar]
  90. Tohma H.; Iwata M.; Maegawa T.; Kita Y. Novel and efficient oxidative biaryl coupling reaction of alkylarenes using a hypervalent iodine(III) reagent. Tetrahedron Lett. 2002, 43, 9241–9244. 10.1016/S0040-4039(02)02150-0. [DOI] [Google Scholar]
  91. Dohi T.; Ito M.; Morimoto K.; Iwata M.; Kita Y. Oxidative Cross-Coupling of Arenes Induced by Single-Electron Transfer Leading to Biaryls by Use of Organoiodine(III) Oxidants. Angew. Chem., Int. Ed. 2008, 47, 1301–1304. 10.1002/anie.200704495. [DOI] [PubMed] [Google Scholar]
  92. Dohi T.; Ito M.; Itani I.; Yamaoka N.; Morimoto K.; Fujioka H.; Kita Y. Metal-Free C-H Cross-Coupling toward Oxygenated Naphthalene-Benzene Linked Biaryls. Org. Lett. 2011, 13, 6208–6211. 10.1021/ol202632h. [DOI] [PubMed] [Google Scholar]
  93. Dohi T.; Ito M.; Sekiguchi S.; Ishikado Y.; Kita Y. Hypervalent iodine induced oxidative cross coupling via thiophene cation radical intermediate. Heterocycles 2012, 86, 767–776. 10.3987/COM-12-S(N)39. [DOI] [Google Scholar]
  94. Tohma H.; Iwata M.; Maegawa T.; Kiyono Y.; Maruyama A.; Kita Y. A novel and direct synthesis of alkylated 2,2′-bithiophene derivatives using a combination of hypervalent iodine(iii) reagent and BF3·Et2O. Org. Biomol. Chem. 2003, 1, 1647–1649. 10.1039/B302462H. [DOI] [PubMed] [Google Scholar]
  95. Dohi T.; Morimoto K.; Kiyono Y.; Maruyama A.; Tohma H.; Kita Y. The synthesis of head-to-tail (H-T) dimers of 3-substituted thiophenes by the hypervalent iodine(iii)-induced oxidative biaryl coupling reaction. Chem. Commun. 2005, 2930–2932. 10.1039/b503058g. [DOI] [PubMed] [Google Scholar]
  96. Dohi T.; Morimoto K.; Maruyama A.; Kita Y. Direct Synthesis of Bipyrroles Using Phenyliodine Bis(trifluoroacetate) with Bromotrimethylsilane. Org. Lett. 2006, 8, 2007–2010. 10.1021/ol060333m. [DOI] [PubMed] [Google Scholar]
  97. Dohi T.; Morimoto K.; Ito M.; Kita Y. Regioselective Bipyrrole Coupling of Pyrroles and 3-Substituted Pyrroles Using Phenyliodine(III) Bis(trifluoroacetate). Synthesis 2007, 2007, 2913–2919. 10.1055/s-2007-983798. [DOI] [Google Scholar]
  98. Morimoto K.; Yamaoka N.; Ogawa C.; Nakae T.; Fujioka H.; Dohi T.; Kita Y. Metal-Free Regioselective Oxidative Biaryl Coupling Leading to Head-to-Tail Bithiophenes: Reactivity Switching, a Concept Based on the Iodonium(III) Intermediate. Org. Lett. 2010, 12, 3804–3807. 10.1021/ol101498r. [DOI] [PubMed] [Google Scholar]
  99. Morimoto K.; Nakae T.; Yamaoka N.; Dohi T.; Kita Y. Metal-Free Oxidative Coupling Reactions via σ-Iodonium Intermediates: The Efficient Synthesis of Bithiophenes Using Hypervalent Iodine Reagents. Eur. J. Org. Chem. 2011, 2011, 6326–6334. 10.1002/ejoc.201100969. [DOI] [Google Scholar]
  100. Kita Y.; Morimoto K.; Ito M.; Ogawa C.; Goto A.; Dohi T. Metal-Free Oxidative Cross-Coupling of Unfunctionalized Aromatic Compounds. J. Am. Chem. Soc. 2009, 131, 1668–1669. 10.1021/ja808940n. [DOI] [PubMed] [Google Scholar]
  101. Dohi T.; Ito M.; Yamaoka N.; Morimoto K.; Fujioka H.; Kita Y. Unusual ipso Substitution of Diaryliodonium Bromides Initiated by a Single-Electron-Transfer Oxidizing Process. Angew. Chem., Int. Ed. 2010, 49, 3334–3337. 10.1002/anie.200907281. [DOI] [PubMed] [Google Scholar]
  102. Yamaoka N.; Sumida K.; Itani I.; Kubo H.; Ohnishi Y.; Sekiguchi S.; Dohi T.; Kita Y. Single-Electron-Transfer (SET)-Induced Oxidative Biaryl Coupling by Polyalkoxybenzene-Derived Diaryliodonium(III) Salts. Chem.—Eur. J. 2013, 19, 15004–15011. 10.1002/chem.201301148. [DOI] [PubMed] [Google Scholar]
  103. Beringer F. M.; Drexler M.; Gindler E. M.; Lumpkin C. C. Diaryliodonium salts. I. Synthesis. J. Am. Chem. Soc. 1953, 75, 2705–2708. 10.1021/ja01107a046. [DOI] [Google Scholar]
  104. Beringer F. M.; Brierley A.; Drexler M.; Gindler E. M.; Lumpkin C. C. Diaryliodonium Salts. II. The Phenylation of Organic and Inorganic Bases. J. Am. Chem. Soc. 1953, 75, 2708–2712. 10.1021/ja01107a047. [DOI] [Google Scholar]
  105. Kalyani D.; Deprez N. R.; Desai L. V.; Sanford M. S. Oxidative C-H Activation/C-C Bond Forming Reactions: Synthetic Scope and Mechanistic Insights. J. Am. Chem. Soc. 2005, 127, 7330–7331. 10.1021/ja051402f. [DOI] [PubMed] [Google Scholar]
  106. Deprez N. R.; Kalyani D.; Sanford M. S. Room Temperature Palladium-Catalyzed 2-Arylation of Indoles. J. Am. Chem. Soc. 2006, 128, 4972–4973. 10.1021/ja060809x. [DOI] [PubMed] [Google Scholar]
  107. Deprez N. R.; Sanford M. S. Reactions of Hypervalent Iodine Reagents with Palladium: Mechanisms and Applications in Organic Synthesis. Inorg. Chem. 2007, 46, 1924–1935. 10.1021/ic0620337. [DOI] [PubMed] [Google Scholar]
  108. Deprez N. R.; Sanford M. S. Synthetic and Mechanistic Studies of Pd-Catalyzed C-H Arylation with Diaryliodonium Salts: Evidence for a Bimetallic High Oxidation State Pd Intermediate. J. Am. Chem. Soc. 2009, 131, 11234–11241. 10.1021/ja904116k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Phipps R. J.; Grimster N. P.; Gaunt M. J. Cu(II)-Catalyzed Direct and Site-Selective Arylation of Indoles Under Mild Conditions. J. Am. Chem. Soc. 2008, 130, 8172–8174. 10.1021/ja801767s. [DOI] [PubMed] [Google Scholar]
  110. Phipps R. J.; Gaunt M. J. A Meta-Selective Copper-Catalyzed C-H Bond Arylation. Science 2009, 323, 1593–1597. 10.1126/science.1169975. [DOI] [PubMed] [Google Scholar]
  111. Phipps R. J.; McMurray L.; Ritter S.; Duong H. A.; Gaunt M. J. Copper-Catalyzed Alkene Arylation with Diaryliodonium Salts. J. Am. Chem. Soc. 2012, 134, 10773–10776. 10.1021/ja3039807. [DOI] [PubMed] [Google Scholar]
  112. Suero M. G.; Bayle E. D.; Collins B. S. L.; Gaunt M. J. Copper-Catalyzed Electrophilic Carbofunctionalization of Alkynes to Highly Functionalized Tetrasubstituted Alkenes. J. Am. Chem. Soc. 2013, 135, 5332–5335. 10.1021/ja401840j. [DOI] [PubMed] [Google Scholar]
  113. Walkinshaw A. J.; Xu W.; Suero M. G.; Gaunt M. J. Copper-Catalyzed Carboarylation of Alkynes via Vinyl Cations. J. Am. Chem. Soc. 2013, 135, 12532–12535. 10.1021/ja405972h. [DOI] [PubMed] [Google Scholar]
  114. Aradi K.; Tóth B. L.; Tolnai G. L.; Novák Z. Diaryliodonium Salts in Organic Syntheses: A Useful Compound Class for Novel Arylation Strategies. Synlett 2016, 27, 1456–1485. 10.1055/s-0035-1561369. [DOI] [Google Scholar]
  115. Merritt E.; Olofsson B. Diaryliodonium Salts: A Journey from Obscurity to Fame. Angew. Chem., Int. Ed. 2009, 48, 9052–9070. 10.1002/anie.200904689. [DOI] [PubMed] [Google Scholar]
  116. Olofsson B. Arylation with Diaryliodonium Salts. In: Wirth, T. (eds) Hypervalent Iodine Chemistry. Top. Curr. Chem. 2015, 373, 135–166. 10.1007/128_2015_661. [DOI] [PubMed] [Google Scholar]
  117. Villo P.; Olofsson B.. Arylations Promoted by Hypervalent Iodine Reagents. In Patai’s Chemistry of Functional Groups; John Wiley & Sons, Ltd, 2018. [Google Scholar]
  118. Kikushima K.; Elboray E. E.; Jiménez-Halla J. O. C.; Solorio-Alvarado C. R.; Dohi T. Diaryliodonium(III) salts in one-pot double functionalization of C-I III and ortho C-H bonds. Org. Biomol. Chem. 2022, 20, 3231–3248. 10.1039/D1OB02501E. [DOI] [PubMed] [Google Scholar]
  119. Pan C.; Wang L.; Han J. Diaryliodonium Salts Enabled Arylation, Arylocyclization, and Aryl-Migration. Chem. Rec. 2023, 23, e202300138 10.1002/tcr.202300138. [DOI] [PubMed] [Google Scholar]
  120. Yusubov M. S.; Maskaev A. V.; Zhdankin V. V. Iodonium salts in organic synthesis. Arkivoc 2011, 2011, 370–409. 10.3998/ark.5550190.0012.107. [DOI] [Google Scholar]
  121. Takenaga N.; Kumar R.; Dohi T. Heteroaryliodonium(III) Salts as Highly Reactive Electrophiles. Front. Chem. 2020, 8, 599026. 10.3389/fchem.2020.599026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Lucas H. J.; Kennedy E. R. IODOBENZENE DICHLORIDE. Org. Synth. 1942, 22, 69. 10.15227/orgsyn.022.0069. [DOI] [Google Scholar]
  123. Sharefkin J G.; Saltzman H. IODOSOBENZENE DIACETATE. Org. Synth. 1963, 43, 62. 10.15227/orgsyn.043.0062. [DOI] [Google Scholar]
  124. Kazmierczak P.; Skulski L.; Kraszkiewicz L. Syntheses of (Diacetoxyiodo)arenes or Iodylarenes from Iodoarenes, with Sodium Periodate as the Oxidant. Molecules 2001, 6, 881–891. 10.3390/61100881. [DOI] [Google Scholar]
  125. Zielinska A.; Skulski L. Easy Preparation of (Diacetoxyiodo)arenes from Iodoarenes with Sodium Percarbonate as the Oxidant. Molecules 2002, 7, 806–809. 10.3390/71100806. [DOI] [Google Scholar]
  126. Hossain D.; Kitamura T. Alternative, Easy Preparation of (Diacetoxyiodo)arenes from Iodoarenes Using Potassium Peroxodisulfate as the Oxidant. Synthesis 2005, 2005, 1932–1934. 10.1055/s-2005-869962. [DOI] [Google Scholar]
  127. McKillop A.; Kemp D. Further functional group oxidations using sodium perborate. Tetrahedron 1989, 45, 3299–3306. 10.1016/S0040-4020(01)81008-5. [DOI] [Google Scholar]
  128. Ochiai M.; Takaoka Y.; Masaki Y.; Nagao Y.; Shiro M. Synthesis of chiral hypervalent organoiodinanes, iodo(III)binaphthyls, and evidence for pseudorotation on iodine. J. Am. Chem. Soc. 1990, 112, 5677–5678. 10.1021/ja00170a063. [DOI] [Google Scholar]
  129. Togo H.; Nabana T.; Yamaguchi K. Preparation and Reactivities of Novel (Diacetoxyiodo)arenes Bearing Heteroaromatics. J. Org. Chem. 2000, 65, 8391–8394. 10.1021/jo001186n. [DOI] [PubMed] [Google Scholar]
  130. Tohma H.; Maruyama A.; Maeda A.; Maegawa T.; Dohi T.; Shiro M.; Morita T.; Kita Y. Preparation and Reactivity of 1,3,5,7-Tetrakis[4-(diacetoxyiodo)phenyl]adamantane, a Recyclable Hypervalent Iodine(III) Reagent. Angew. Chem., Int. Ed. 2004, 43, 3595–3598. 10.1002/anie.200454234. [DOI] [PubMed] [Google Scholar]
  131. Moroda A.; Togo H. Biphenyl- and terphenyl-based recyclable organic trivalent iodine reagents. Tetrahedron 2006, 62, 12408–12414. 10.1016/j.tet.2006.09.112. [DOI] [Google Scholar]
  132. Lucas H. J.; Kennedy E. R.; Formo M. W. IODOSOBENZENE. Org. Synth. 1942, 22, 70. 10.15227/orgsyn.022.0070. [DOI] [Google Scholar]
  133. Saltzman H.; Sharefkin J. G. IODOSOBENZENE. Org. Synth. 1963, 43, 60. 10.15227/orgsyn.043.0060. [DOI] [Google Scholar]
  134. Moriarty R. M.; Vaid R. K.; Koser G. F. [Hydroxy(organosulfonyloxy)iodo]arenes in Organic Synthesis. Synlett 1990, 1990, 365–383. 10.1055/s-1990-21097. [DOI] [Google Scholar]
  135. Yamamoto Y.; Togo H. Facile One-Pot Preparation of [Hydroxy(sulfonyloxy)iodo]arenes from-Iodoarenes with MCPBA in the Presence of Sulfonic Acids. Synlett 2005, 2486–2488. 10.1055/s-2005-872697. [DOI] [Google Scholar]
  136. Merritt E. A.; Carneiro V. M. T.; Silva L. F. Jr; Olofsson B. Facile Synthesis of Koser’s Reagent and Derivatives from Iodine or Aryl Iodides. J. Org. Chem. 2010, 75, 7416–7419. 10.1021/jo101227j. [DOI] [PubMed] [Google Scholar]
  137. Kitamura T.; Matsuyuki J.; Nagata K.; Furuki R.; Taniguchi H. A Convenient Preparation of Diaryliodonium Triflates. Synthesis 1992, 1992, 945–946. 10.1055/s-1992-26272. [DOI] [Google Scholar]
  138. Dohi T.; Koseki D.; Sumida K.; Okada K.; Mizuno S.; Kato A.; Morimoto K.; Kita Y. Metal-Free O-Arylation of Carboxylic Acid by Active Diaryliodonium(III) Intermediates Generated in situ from Iodosoarenes. Adv. Synth. Catal. 2017, 359, 3503–3508. 10.1002/adsc.201700843. [DOI] [Google Scholar]
  139. Shah A.; Pike V. W.; Widdowson D. A. Synthesis of functionalised unsymmetrical diaryliodonium salts. J. Chem. Soc., Perkin Trans. 1 1997, 2463–2466. 10.1039/a704062h. [DOI] [Google Scholar]
  140. Kitamura T.; Inoue D.; Wakimoto I.; Nakamura T.; Katsuno R.; Fujiwara Y. Reaction of (diacetoxyiodo)benzene with excess of trifluoromethanesulfonic acid. A convenient route to para-phenylene type hypervalent iodine oligomers. Tetrahedron 2004, 60, 8855–8860. 10.1016/j.tet.2004.07.026. [DOI] [Google Scholar]
  141. Dohi T.; Hayashi T.; Ueda S.; Shoji T.; Komiyama K.; Takeuchi H.; Kita Y. Recyclable synthesis of mesityl iodonium(III) salts. Tetrahedron 2019, 75, 3617–3627. 10.1016/j.tet.2019.05.033. [DOI] [Google Scholar]
  142. Koser G. F.; Wettach R. H. [Hydroxy(tosyloxy)iodo]benzene, a versatile reagent for the mild oxidation of aryl iodides at the iodine atom by ligand transfer. J. Org. Chem. 1980, 45, 1542–1543. 10.1021/jo01296a049. [DOI] [Google Scholar]
  143. Margida A. J.; Koser G. F. Direct condensation of [hydroxy(tosyloxy)iodo]arenes with thiophenes. A convenient, mild synthesis of aryl(2-thienyl)iodonium tosylates. J. Org. Chem. 1984, 49, 3643–3646. 10.1021/jo00193a039. [DOI] [Google Scholar]
  144. Dohi T.; Ito M.; Morimoto K.; Minamitsuji Y.; Takenaga N.; Kita Y. Versatile direct dehydrative approach for diaryliodonium(iii) salts in fluoroalcohol media. Chem. Commun. 2007, 4152–4154. 10.1039/b708802g. [DOI] [PubMed] [Google Scholar]
  145. Dohi T.; Yamaoka N.; Kita Y. Fluoroalcohols: versatile solvents in hypervalent iodine chemistry and syntheses of diaryliodonium(III) salts. Tetrahedron 2010, 66, 5775. 10.1016/j.tet.2010.04.116. [DOI] [Google Scholar]
  146. Ito M.; Ogawa C.; Yamaoka N.; Fujioka H.; Dohi T.; Kita Y. Enhanced Reactivity of [Hydroxy(tosyloxy)iodo]benzene in Fluoroalcohol Media. Efficient Direct Synthesis of Thienyl(aryl)iodonium Salts. Molecules 2010, 15, 1918–1931. 10.3390/molecules15031918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Beringer F. M.; Dehn J. W. Jr; Winicov M. Diaryliodonium Salts. XIV. Reactions of Organometallic Compounds with Iodosobenzene Dichlorides and with Iodonium Salts. J. Am. Chem. Soc. 1960, 82, 2948–2952. 10.1021/ja01496a066. [DOI] [Google Scholar]
  148. Koser G. F.; Wettach R. H.; Smith C. S. New methodology in iodonium salt synthesis. Reactions of [hydroxy(tosyloxy)iodo]arenes with aryltrimethylsilanes. J. Org. Chem. 1980, 45, 1543–1544. 10.1021/jo01296a050. [DOI] [Google Scholar]
  149. 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]
  150. 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]
  151. Pike V. W.; Butt F.; Shah A.; Widdowson D. A. Facile synthesis of substituted diaryliodonium tosylates by treatment of aryltributylstannanes with Koser’s reagent. J. Chem. Soc., Perkin Trans. 1 1999, 245–248. 10.1039/a809349k. [DOI] [Google Scholar]
  152. 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]
  153. Bykowski D.; McDonald R.; Hinkle R. J.; Tykwinski R. R. Structural and Electronic Characteristics of Thienyl(aryl)iodonium Triflates. J. Org. Chem. 2002, 67, 2798–2804. 10.1021/jo015910t. [DOI] [PubMed] [Google Scholar]
  154. Carroll M. A.; Pike V. W.; Widdowson D. A. New synthesis of diaryliodonium sulfonates from arylboronic acids. Tetrahedron Lett. 2000, 41, 5393–5396. 10.1016/S0040-4039(00)00861-3. [DOI] [Google Scholar]
  155. Ochiai M.; Toyonari M.; Sueda T.; Kitagawa Y. Boron-iodine(III) exchange reaction: Direct synthesis of diaryliodonium tetraarylborates from (diacetoxyiodo)arenes by the reaction with alkali metal tetraarylborates in acetic acid. Tetrahedron Lett. 1996, 37, 8421–8424. 10.1016/0040-4039(96)01926-0. [DOI] [Google Scholar]
  156. Seidl T. L.; Sundalam S. K.; McCullough B.; Stuart D. R. Unsymmetrical Aryl(2,4,6-trimethoxyphenyl)iodonium Salts: One-Pot Synthesis, Scope, Stability, and Synthetic Studies. J. Org. Chem. 2016, 81, 1998–2009. 10.1021/acs.joc.5b02833. [DOI] [PubMed] [Google Scholar]
  157. Bielawski M.; Olofsson B. High-yielding one-pot synthesis of diaryliodonium triflates from arenes and iodine or aryl iodides. Chem. Commun. 2007, 2521–2523. 10.1039/b701864a. [DOI] [PubMed] [Google Scholar]
  158. Bielawski M.; Zhu M.; Olofsson B. Efficient and General One-Pot Synthesis of Diaryliodonium Triflates: Optimization, Scope and Limitations. Adv. Synth. Catal. 2007, 349, 2610–2618. 10.1002/adsc.200700373. [DOI] [Google Scholar]
  159. Zhu M.; Jalalian N.; Olofsson B. One-Pot Synthesis of Diaryliodonium Salts Using Toluenesulfonic Acid: A Fast Entry to Electron-Rich Diaryliodonium Tosylates and Triflates Synthesis of Electron-Rich Diaryliodonium Tosylates and Triflates. Synlett 2008, 2008, 0592–0596. 10.1055/s-2008-1032050. [DOI] [Google Scholar]
  160. Bielawski M.; Malmgren J.; Pardo L. M.; Wikmark Y.; Olofsson B. One-Pot Synthesis and Applications of N-Heteroaryl Iodonium Salts. ChemistryOpen 2014, 3, 19–22. 10.1002/open.201300042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Carreras V.; Sandtorv A. H.; Stuart D. R. Synthesis of Aryl(2,4,6-trimethoxyphenyl)iodonium Trifluoroacetate Salts. J. Org. Chem. 2017, 82, 1279–1284. 10.1021/acs.joc.6b02811. [DOI] [PubMed] [Google Scholar]
  162. Lindstedt E.; Reitti M.; Olofsson B. One-Pot Synthesis of Unsymmetric Diaryliodonium Salts from Iodine and Arenes. J. Org. Chem. 2017, 82, 11909–11914. 10.1021/acs.joc.7b01652. [DOI] [PubMed] [Google Scholar]
  163. Hossain Md. D.; Ikegami Y.; Kitamura T. Reaction of Arenes with Iodine in the Presence of Potassium Peroxodisulfate in Trifluoroacetic Acid. Direct and Simple Synthesis of Diaryliodonium Triflates. J. Org. Chem. 2006, 71, 9903–9905. 10.1021/jo061889q. [DOI] [PubMed] [Google Scholar]
  164. Hossain Md. D.; Kitamura T. Reaction of iodoarenes with potassium peroxodisulfate/ trifluoroacetic acid in the presence of aromatics. Direct preparation of diaryliodonium triflates from iodoarenes. Tetrahedron 2006, 62, 6955–6960. 10.1016/j.tet.2006.04.073. [DOI] [Google Scholar]
  165. Hossain Md. D.; Kitamura T. New and Direct Approach to Hypervalent Iodine Compounds from Arenes and Iodine. Straightforward Synthesis of (Diacetoxyiodo)arenes and Diaryliodonium Salts Using Potassium μ-Peroxo-hexaoxodisulfate. Bull. Chem. Soc. Jpn. 2007, 80, 2213–2219. 10.1246/bcsj.80.2213. [DOI] [Google Scholar]
  166. Merritt E. A.; Malmgren J.; Klinke F. J.; Olofsson B. Synthesis of Diaryliodonium Triflates Using Environmentally Benign Oxidizing Agents. Synlett 2009, 2277–2280. 10.1055/s-0029-1217723. [DOI] [Google Scholar]
  167. Soldatova N.; Postnikov P.; Kukurina O.; Zhdankin V. V.; Yoshimura A.; Wirth T.; Yusubov M. S. Facile One-Pot Synthesis of Diaryliodonium Salts from Arenes and Aryl Iodides with Oxone. ChemistryOpen 2017, 6, 18–20. 10.1002/open.201600129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Kryska A.; Skulski L. One-Pot Preparations of Diaryliodonium Bromides from Iodoarenes and Arenes, with Sodium Perborate as the Oxidant. Molecules 2001, 6, 875–880. 10.3390/61100875. [DOI] [Google Scholar]
  169. Soldatova N.; Postnikov P.; Kukurina O.; Zhdankin V. V.; Yoshimura A.; Wirth T.; Yusubov M. S. One-pot synthesis of diaryliodonium salts from arenes and aryl iodides with Oxone-sulfuric acid. Beilstein J. Org. Chem. 2018, 14, 849–855. 10.3762/bjoc.14.70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Dohi T.; Yamaoka N.; Itani I.; Kita Y. One-Pot Syntheses of Diaryliodonium Salts from Aryl Iodides Using Peracetic Acid as Green Oxidant. Aust. J. Chem. 2011, 64, 529–535. 10.1071/CH11057. [DOI] [Google Scholar]
  171. Laudadio G.; Gemoets H. P. L.; Hessel V.; Noel T. Flow Synthesis of Diaryliodonium Triflates. J. Org. Chem. 2017, 82, 11735–11741. 10.1021/acs.joc.7b01346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Gemoets H. P. L.; Laudadio G.; Verstraete K.; Hessel V.; Noel T. A Modular Flow Design for the meta-Selective C-H Arylation of Anilines. Angew. Chem., Int. Ed. 2017, 56, 7161–7165. 10.1002/anie.201703369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Soldatova N. S.; Postnikov P. S.; Yusubov M. S.; Wirth T. Flow Synthesis of Iodonium Trifluoroacetates through Direct Oxidation of Iodoarenes by Oxone. Eur. J. Org. Chem. 2019, 2019, 2081–2088. 10.1002/ejoc.201900220. [DOI] [Google Scholar]
  174. Winterson B.; Bhattacherjee D.; Wirth T. Hypervalent Halogen Compounds in Electrochemical Reactions: Advantages and Prospects. Adv. Synth. Catal. 2023, 365, 2676–2689. 10.1002/adsc.202300412. [DOI] [Google Scholar]
  175. Elsherbini M.; Wirth T. Hypervalent Iodine Reagents by Anodic Oxidation: A Powerful Green Synthesis. Chem.—Eur. J. 2018, 24, 13399–13407. 10.1002/chem.201801232. [DOI] [PubMed] [Google Scholar]
  176. 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]
  177. 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]
  178. Watts K.; Gattrell W.; Wirth T. A practical microreactor for electrochemistry in flow. Beilstein J. Org. Chem. 2011, 7, 1108–1114. 10.3762/bjoc.7.127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Bielawski M.; Aili D.; Olofsson B. Regiospecific One-Pot Synthesis of Diaryliodonium Tetrafluoroborates from Arylboronic Acids and Aryl Iodides. J. Org. Chem. 2008, 73, 4602–4607. 10.1021/jo8004974. [DOI] [PubMed] [Google Scholar]
  180. Linde E.; Mondal S.; Olofsson B. Advancements in the Synthesis of Diaryliodonium Salts: Updated Protocols. Adv. Synth. Catal. 2023, 365, 2751–2756. 10.1002/adsc.202300354. [DOI] [Google Scholar]
  181. Gallagher R. T.; Seidl T. L.; Bader J.; Orella C.; Vickery T.; Stuart D. R. Anion Metathesis of Diaryliodonium Tosylate Salts with a Solid-Phase Column Constructed from Readily Available Laboratory Consumables. Org. Process Res. Dev. 2019, 23, 1269–1274. 10.1021/acs.oprd.9b00163. [DOI] [Google Scholar]
  182. Kumar R.; Dohi T.; Zhdankin V. V. Organohypervalent heterocycles. Chem. Soc. Rev. 2024, 53, 4786–4827. 10.1039/D2CS01055K. [DOI] [PubMed] [Google Scholar]
  183. Singhal R.; Choudhary S. P.; Malik B.; Pilania M. Cyclic diaryliodonium salts: applications and overview. Org. Biomol. Chem. 2023, 21, 4358–4378. 10.1039/D3OB00134B. [DOI] [PubMed] [Google Scholar]
  184. 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]
  185. Spils J.; Wirth T.; Nachtsheim B. J. Two-step continuous-flow synthesis of 6-membered cyclic iodonium salts via anodic oxidation. Beilstein J. Org. Chem. 2023, 19, 27–32. 10.3762/bjoc.19.2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. de Magalhães H. P.; Lüthi H. P.; Togni A. Reductive Eliminations from λ3-Iodanes: Understanding Selectivity and the Crucial Role of the Hypervalent Bond. Org. Lett. 2012, 14, 3830–3833. 10.1021/ol3014039. [DOI] [PubMed] [Google Scholar]
  187. Stridfeldt E.; Lindstedt E.; Reitti M.; Blid J.; Norrby P.-O.; Olofsson B. Competing Pathways in O-Arylations with Diaryliodonium Salts: Mechanistic Insights. Chem.—Eur. J. 2017, 23, 13249–13258. 10.1002/chem.201703057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Kakinuma Y.; Moriyama K.; Togo H. Facile Preparation of Unsymmetrical Diaryl Ethers from Unsymmetrical Diaryliodonium Tosylates and Phenols with High Regioselectivity. Synthesis 2013, 45, 183–188. 10.1055/s-0032-1316824. [DOI] [Google Scholar]
  189. Ross T. L.; Ermert J.; Hocke C.; Coenen H. H. Nucleophilic 18F-Fluorination of Heteroaromatic Iodonium Salts with No-Carrier-Added [18F]Fluoride. J. Am. Chem. Soc. 2007, 129, 8018–8025. 10.1021/ja066850h. [DOI] [PubMed] [Google Scholar]
  190. Malmgren J.; Santoro S.; Jalalian N.; Himo F.; Olofsson B. Arylation with Unsymmetrical Diaryliodonium Salts: A Chemoselectivity Study. Chem.—Eur. J. 2013, 19, 10334–10342. 10.1002/chem.201300860. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Ochiai M.; Kitagawa Y.; Toyonari M. On the mechanism of α-phenylation of β-keto esters with diaryl-λ3- iodanes: evidence for a non-radical pathway. Arkivoc 2003, 2003, 43–48. 10.3998/ark.5550190.0004.606. [DOI] [Google Scholar]
  192. Stuart D. R. Aryl Transfer Selectivity in Metal-Free Reactions of Unsymmetrical Diaryliodonium Salts. Chem.—Eur. J. 2017, 23, 15852–15863. 10.1002/chem.201702732. [DOI] [PubMed] [Google Scholar]
  193. Linde E.; Knippenberg; Olofsson N. B. Synthesis of Cyclic and Acyclic ortho-Aryloxy Diaryliodonium Salts for Chemoselective Functionalizations. Chem.—Eur. J. 2022, 28, e202202453 10.1002/chem.202202453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Di Tommaso E. M.; Walther M.; Staubitz A.; Olofsson B. ortho-Functionalization of azobenzenes via hypervalent iodine reagents. Chem. Commun. 2023, 59, 5047–5050. 10.1039/D3CC01060K. [DOI] [PubMed] [Google Scholar]
  195. Jalalian N.; Ishikawa E. E.; Silva L. F.; Olofsson B. Room Temperature, Metal-Free Synthesis of Diaryl Ethers with Use of Diaryliodonium Salts. Org. Lett. 2011, 13, 1552–1555. 10.1021/ol200265t. [DOI] [PubMed] [Google Scholar]
  196. Lindstedt E.; Ghosh R.; Olofsson B. Metal-Free Synthesis of Aryl Ethers in Water. Org. Lett. 2013, 15, 6070–6073. 10.1021/ol402960f. [DOI] [PubMed] [Google Scholar]
  197. Jalalian N.; Petersen T. B.; Olofsson B. Metal-Free Arylation of Oxygen Nucleophiles with Diaryliodonium Salts. Chem.—Eur. J. 2012, 18, 14140–14149. 10.1002/chem.201201645. [DOI] [PubMed] [Google Scholar]
  198. Chan L.; McNally A.; Toh Q. Y.; Mendoza A.; Gaunt M. J. A counteranion triggered arylation strategy using diaryliodonium fluorides. Chem. Sci. 2015, 6, 1277–1281. 10.1039/C4SC02856B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Ghosh R.; Lindstedt E.; Jalalian N.; Olofsson B. Room Temperature, Metal-Free Arylation of Aliphatic Alcohols. ChemistryOpen 2014, 3, 54–57. 10.1002/open.201402006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Tolnai G. L.; Nilsson U. J.; Olofsson B. Efficient O-Functionalization of Carbohydrates with Electrophilic Reagents. Angew. Chem., Int. Ed. 2016, 55, 11226–11230. 10.1002/anie.201605999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. 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]
  202. Sundalam S. K.; Stuart D. R. Base Mediated Synthesis of Alkyl-aryl Ethers from the Reaction of Aliphatic Alcohols and Unsymmetric Diaryliodonium Salts. J. Org. Chem. 2015, 80, 6456–6466. 10.1021/acs.joc.5b00907. [DOI] [PubMed] [Google Scholar]
  203. Gallagher R. T.; Jha S.; Metze B. E.; Stuart D. R. Synthesis of Alkyl Aryl Ethers by O-Arylation of Alcohols with Diaryliodonium Salts: Scope, Limitations, and Mechanism. Synlett 2024, 35, 889–894. 10.1055/a-2198-3637. [DOI] [Google Scholar]
  204. Wang D.; Ge C.; Yu X.; Wan H.; Xu X. Transition-Metal-Free Direct Arylation and Esterification Reaction of Unprotected Indolylcarboxylic Acid Derivatives: A New Entry to 2-(1H-Indol-2-yl)-5-(phenylthio)-1,3,4-oxadiazoles and Aryl 1H-Indole-2-carboxylates. Synlett 2016, 27, 2616–2620. 10.1055/s-0035-1562526. [DOI] [Google Scholar]
  205. 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]
  206. Ghosh R.; Olofsson B. Metal-Free Synthesis of N-Aryloxyimides and Aryloxyamines. Org. Lett. 2014, 16, 1830–1832. 10.1021/ol500478t. [DOI] [PubMed] [Google Scholar]
  207. Wang Z.-X.; Shi W.-M.; Bi H.-Y.; Li X.-H.; Su G.-F.; Mo D.-L. Synthesis of N-(2-Hydroxyaryl)benzotriazoles via Metal-Free O-Arylation and N-O Bond Cleavage. J. Org. Chem. 2016, 81, 8014–8021. 10.1021/acs.joc.6b01390. [DOI] [PubMed] [Google Scholar]
  208. Shi W.-M.; Ma X.-P.; Pan C.-X.; Su G.-F.; Mo D.-L. Tandem C-O and C-N Bonds Formation Through O-Arylation and [3,3]-Rearrangement by Diaryliodonium Salts: Synthesis of N-Aryl Benzo[1,2,3]triazin-4(1H)-one Derivatives. J. Org. Chem. 2015, 80, 11175–11183. 10.1021/acs.joc.5b01947. [DOI] [PubMed] [Google Scholar]
  209. Xiong W.; Qi C.; Peng Y.; Guo T.; Zhang M.; Jiang H. Base-Promoted Coupling of Carbon Dioxide, Amines, and Diaryliodonium Salts: A Phosgene- and Metal-Free Route to O-Aryl Carbamates. Chem.—Eur. J. 2015, 21, 14314–14318. 10.1002/chem.201502689. [DOI] [PubMed] [Google Scholar]
  210. Xiong B.; Feng X.; Zhu L.; Chen T.; Zhou Y.; Au C.-T.; Yin S.-F. Direct Aerobic Oxidative Esterification and Arylation of P(O)-OH Compounds with Alcohols and Diaryliodonium Triflates. ACS Catal. 2015, 5, 537–543. 10.1021/cs501523g. [DOI] [Google Scholar]
  211. Gao H.; Xu Q.-L.; Keene C.; Kürti L. Scalable, Transition-Metal-Free Direct Oxime O-Arylation: Rapid Access to O-Arylhydroxylamines and Substituted Benzo[b]furans. Chem.—Eur. J. 2014, 20, 8883–8887. 10.1002/chem.201403519. [DOI] [PubMed] [Google Scholar]
  212. Miyagi K.; Moriyama K.; Togo H. One-Pot Preparation of 2-Arylbenzofurans from Oximes with Diaryliodonium Triflate. Heterocycles 2014, 89, 2122–2136. 10.3987/COM-14-13071. [DOI] [Google Scholar]
  213. Ghosh R.; Stridfeldt E.; Olofsson B. Metal-Free One-Pot Synthesis of Benzofurans. Chem.—Eur. J. 2014, 20, 8888–8892. 10.1002/chem.201403523. [DOI] [PubMed] [Google Scholar]
  214. Yang Y.; Wu X.; Han J.; Mao S.; Qian X.; Wang L. Cesium Carbonate Promoted Direct Arylation of Hydroxylamines and Oximes with Diaryliodonium Salts. Eur. J. Org. Chem. 2014, 2014, 6854–6857. 10.1002/ejoc.201402920. [DOI] [Google Scholar]
  215. Misal Castro L.; Chatani N. Potassium tert-Butoxide Mediated O-Arylation of N-Hydroxyphthalimide and Oximes with Diaryliodonium Salts. Synthesis 2014, 46, 2312–2316. 10.1055/s-0033-1340192. [DOI] [Google Scholar]
  216. Zhou Y. Facile and metal-free synthesis of phenols from benzaldoxime and diaryliodonium salts. J. Chem. Res. 2017, 41, 591–593. 10.3184/174751917X15064232103119. [DOI] [Google Scholar]
  217. Reitti M.; Gurubrahamam R.; Walther M.; Lindstedt E.; Olofsson B. Synthesis of Phenols and Aryl Silyl Ethers via Arylation of Complementary Hydroxide Surrogates. Org. Lett. 2018, 20, 1785–1788. 10.1021/acs.orglett.8b00287. [DOI] [PubMed] [Google Scholar]
  218. Katagiri K.; Kuriyama M.; Yamamoto K.; Demizu Y.; Onomura O. Organocatalytic Synthesis of Phenols from Diaryliodonium Salts with Water under Metal-Free Conditions. Org. Lett. 2022, 24 (28), 5149–5154. 10.1021/acs.orglett.2c01989. [DOI] [PubMed] [Google Scholar]
  219. Gallagher R. T.; Basu S.; Stuart D. R. Trimethoxyphenyl (TMP) as a Useful Auxiliary for in situ Formation and Reaction of Aryl(TMP)iodonium Salts: Synthesis of Diaryl Ethers. Adv. Synth. Catal. 2020, 362, 320–325. 10.1002/adsc.201901187. [DOI] [Google Scholar]
  220. Kikushima K.; Miyamoto N.; Watanabe K.; Koseki D.; Kita Y.; Dohi T. Ligand- and Counterion-Assisted Phenol O-Arylation with TMP-Iodonium(III) Acetates. Org. Lett. 2022, 24, 1924–1928. 10.1021/acs.orglett.2c00294. [DOI] [PubMed] [Google Scholar]
  221. China H.; Koseki D.; Samura K.; Kikushima K.; In Y.; Dohi T. Dataset on synthesis and crystallographic structure of phenyl(TMP)iodonium(III) acetate. Data in brief 2019, 25, 104063. 10.1016/j.dib.2019.104063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Milzarek T. M.; Gulder T. A. M. Total Synthesis of the Ambigols: A Cyanobacterial Class of Polyhalogenated Natural Products. Org. Lett. 2021, 23, 102–106. 10.1021/acs.orglett.0c03784. [DOI] [PubMed] [Google Scholar]
  223. Kervefors G.; Pal K. B.; Tolnai G. L.; Mahanti M.; Leffler H.; Nilsson U. J.; Olofsson B. Synthesis and Biological Studies of O3-Aryl Galactosides as Galectin Inhibitors. Helv. Chim. Acta 2021, 104, e2000220 10.1002/hlca.202000220. [DOI] [Google Scholar]
  224. Lucchetti N.; Gilmour R. Reengineering Chemical Glycosylation: Direct, Metal-Free Anomeric O-Arylation of Unactivated Carbohydrates. Chem.—Eur. J. 2018, 24, 16266–16270. 10.1002/chem.201804416. [DOI] [PubMed] [Google Scholar]
  225. Xu H.; Tan T.; Zhang Y.; Wang Y.; Pan K.; Yao Y.; Zhang S.; Gu Y.; Chen W.; Li J.; Dong H.; Meng Y.; Ma P.; Hou W.; Yang G. Metal-Free and Open-Air Arylation Reactions of Diaryliodonium Salts for DNA-Encoded Library Synthesis. Adv. Sci. 2022, 9, 2202790. 10.1002/advs.202202790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. 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]
  227. Zhang J.-W.; Qi L.-W.; Li S.; Xiang S.-H.; Tan B. Direct Construction of NOBINs via Domino Arylation and Sigmatropic Rearrangement Reactions. Chin. J. Chem. 2020, 38, 1503–1514. 10.1002/cjoc.202000358. [DOI] [Google Scholar]
  228. Liu F.; Wang M.; Qu J.; Lu H.; Gao H. Synthesis of non-C2 symmetrical NOBIN-type biaryls through a cascade N-arylation and [3,3]-sigmatropic rearrangement from O-arylhydroxylamines and diaryliodonium salts. Org. Biomol. Chem. 2021, 19, 7246–7251. 10.1039/D1OB00636C. [DOI] [PubMed] [Google Scholar]
  229. Li J.; Liu L. Simple and efficient amination of diaryliodonium salts with aqueous ammonia in water without metal-catalyst. RSC Adv. 2012, 2, 10485–10487. 10.1039/c2ra22046f. [DOI] [Google Scholar]
  230. Reitti M.; Villo P.; Olofsson B. One-Pot C-H Functionalization of Arenes by Diaryliodonium Salts. Angew. Chem., Int. Ed. 2016, 55, 8928–8932. 10.1002/anie.201603175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Yusubov M. S.; Yusubova R. Y.; Nemykin V. N.; Zhdankin V. V. Preparation and X-ray Structural Study of 1-Arylbenziodoxolones. J. Org. Chem. 2013, 78, 3767–3773. 10.1021/jo400212u. [DOI] [PubMed] [Google Scholar]
  232. Yusubov M. S.; Soldatova N. S.; Postnikov P. S.; Valiev R. R.; Svitich D. Y.; Yusubova R. Y.; Yoshimura A.; Wirth T.; Zhdankin V. V. Reactions of 1-Arylbenziodoxolones with Azide Anion: Experimental and Computational Study of Substituent Effects. Eur. J. Org. Chem. 2018, 2018, 640–647. 10.1002/ejoc.201701595. [DOI] [Google Scholar]
  233. Seidl T. L.; Stuart D. R. An Admix Approach To Determine Counter Anion Effects on Metal-Free Arylation Reactions with Diaryliodonium Salts. J. Org. Chem. 2017, 82, 11765–11771. 10.1021/acs.joc.7b01599. [DOI] [PubMed] [Google Scholar]
  234. Carroll M. A.; Wood R. A. Arylation of anilines: formation of diarylamines using diaryliodonium salts. Tetrahedron 2007, 63, 11349–11354. 10.1016/j.tet.2007.08.076. [DOI] [Google Scholar]
  235. Pang X.; Lou Z.; Li M.; Wen L.; Chen C. Tandem Arylation/Friedel-Crafts Reactions of o-Acylanilines with Diaryliodonium Salts: A Modular Synthesis of Acridine Derivatives. Eur. J. Org. Chem. 2015, 2015, 3361–3369. 10.1002/ejoc.201500161. [DOI] [Google Scholar]
  236. Sandtorv A. H.; Stuart D. R. Metal-free Synthesis of Aryl Amines: Beyond Nucleophilic Aromatic Substitution. Angew. Chem., Int. Ed. 2016, 55, 15812–15815. 10.1002/anie.201610086. [DOI] [PubMed] [Google Scholar]
  237. Sundalam S. K.; Nilova A.; Seidl T. L.; Stuart D. R. A Selective C-H Deprotonation Strategy to Access Functionalized Arynes by Using Hypervalent Iodine. Angew. Chem., Int. Ed. 2016, 55, 8431–8434. 10.1002/anie.201603222. [DOI] [PubMed] [Google Scholar]
  238. Li P.; Cheng G.; Zhang H.; Xu X.; Gao J.; Cui X. Copper-Catalyzed One-Pot Synthesis of Unsymmetrical Arylurea Derivatives via Tandem Reaction of Diaryliodonium Salts with N-Arylcyanamide. J. Org. Chem. 2014, 79, 8156–8162. 10.1021/jo501334u. [DOI] [PubMed] [Google Scholar]
  239. Li J.; Zheng X.; Li W.; Zhou W.; Zhu W.; Zhang Y. Transition-metal free N-arylation of cyanamides by diaryliodonium triflates in aqueous media. New J. Chem. 2016, 40, 77–80. 10.1039/C5NJ02153G. [DOI] [Google Scholar]
  240. Lucchetti N.; Scalone M.; Fantasia S.; Muñiz K. Sterically Congested 2,6-Disubstituted Anilines from Direct C-N Bond Formation at an Iodine(III) Center. Angew. Chem., Int. Ed. 2016, 55, 13335–13339. 10.1002/anie.201606599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Basu S.; Sandtorv A. H.; Stuart D. R. Imide arylation with aryl(TMP)iodonium tosylates. Beilstein J. Org. Chem. 2018, 14, 1034–1038. 10.3762/bjoc.14.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Tinnis F.; Stridfeldt E.; Lundberg H.; Adolfsson H.; Olofsson B. Metal-Free N-Arylation of Secondary Amides at Room Temperature. Org. Lett. 2015, 17, 2688–2691. 10.1021/acs.orglett.5b01079. [DOI] [PubMed] [Google Scholar]
  243. Wang M.; Huang Z. Transition metal-free N-arylation of secondary amides through iodonium salts as aryne precursors. Org. Biomol. Chem. 2016, 14, 10185–10188. 10.1039/C6OB01649A. [DOI] [PubMed] [Google Scholar]
  244. Gonda Z.; Novák Z. Transition-Metal-Free N-Arylation of Pyrazoles with Diaryliodonium Salts. Chem.—Eur. J. 2015, 21, 16801–16806. 10.1002/chem.201502995. [DOI] [PubMed] [Google Scholar]
  245. Teskey C. J.; Sohel S. M. A.; Bunting D. L.; Modha S. G.; Greaney M. F. Domino N-/C-Arylation via In Situ Generation of a Directing Group: Atom-Efficient Arylation Using Diaryliodonium Salts. Angew. Chem., Int. Ed. 2017, 56, 5263–5266. 10.1002/anie.201701523. [DOI] [PubMed] [Google Scholar]
  246. Bihari T.; Babinszki B.; Gonda Z.; Kovács S.; Novák Z.; Stirling A. Understanding and Exploitation of Neighboring Heteroatom Effect for the Mild N-Arylation of Heterocycles with Diaryliodonium Salts under Aqueous Conditions: A Theoretical and Experimental Mechanistic Study. J. Org. Chem. 2016, 81, 5417–5422. 10.1021/acs.joc.6b00779. [DOI] [PubMed] [Google Scholar]
  247. Bugaenko D. I.; Yurovskaya M. A.; Karchava A. V. N-Arylation of DABCO with Diaryliodonium Salts: General Synthesis of N-Aryl-DABCO Salts as Precursors for 1,4-Disubstituted Piperazines. Org. Lett. 2018, 20, 6389–6393. 10.1021/acs.orglett.8b02676. [DOI] [PubMed] [Google Scholar]
  248. 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. 2018, 57, 11427–11431. 10.1002/anie.201807001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Kervefors G.; Kersting L.; Olofsson B. Transition Metal-Free N-Arylation of Amino Acid Esters with Diaryliodonium Salts. Chem.—Eur. J. 2021, 27, 5790–5795. 10.1002/chem.202005351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  250. Roshandel S.; Lunn M. J.; Rasul G.; Ravinson D. S. M.; Suri S. C.; Prakash G. K. S. Catalyst-Free Regioselective N2 Arylation of 1,2,3-Triazoles Using Diaryl Iodonium Salts. Org. Lett. 2019, 21, 6255–6258. 10.1021/acs.orglett.9b02140. [DOI] [PubMed] [Google Scholar]
  251. Saikia R. A.; Dutta A.; Sarma B.; Thakur A. J. Metal-Free Regioselective N2 -Arylation of 1H-Tetrazoles with Diaryliodonium Salts. J. Org. Chem. 2022, 87, 9782–9796. 10.1021/acs.joc.2c00848. [DOI] [PubMed] [Google Scholar]
  252. Xu B.; Han J.; Wang L. Metal- and Base-Free Direct N-Arylation of Pyridazinones by Using Diaryliodonium Salts: An Anion Effect. Asian J. Org. Chem. 2018, 7, 1674–1680. 10.1002/ajoc.201800268. [DOI] [Google Scholar]
  253. Kikushima K.; Morita A.; Elboray E. E.; BAE T.; Miyamoto N.; Kita Y.; Dohi T. Transition-Metal-Free N-Arylation of N-Methoxysulfonamides and N,O- Protected Hydroxylamines with Trimethoxyphenyliodonium(III) Acetates. Synthesis 2022, 54, 5192–5202. 10.1055/a-1922-8846. [DOI] [Google Scholar]
  254. Vlasenko Y.; Kuczmera T. J.; Antonkin N. S.; Valiev R. R.; Postnikov P. S.; Nachtsheim B. J. Site Selective Concerted Nucleophilic Aromatic Substitutions of Azole-Ligated Diaryliodonium Salts. Adv. Synth. Catal. 2023, 365, 535–543. 10.1002/adsc.202201001. [DOI] [Google Scholar]
  255. Ma X.-P.; Shi W.-M.; Mo X.-L.; Li X.-H.; Li L.-G.; Pan C.-X.; Chen B.; Su G.-F.; Mo D.-L. Synthesis of α,β-Unsaturated N-Aryl Ketonitrones from Oximes and Diaryliodonium Salts: Observation of a Metal-Free N-Arylation Process. J. Org. Chem. 2015, 80, 10098–10107. 10.1021/acs.joc.5b01716. [DOI] [PubMed] [Google Scholar]
  256. Wu S.-Y.; Ma X.-P.; Liang C.; Mo D.-L. Synthesis of N-Aryl Oxindole Nitrones through a Metal-Free Selective N-Arylation Process. J. Org. Chem. 2017, 82, 3232–3238. 10.1021/acs.joc.6b02774. [DOI] [PubMed] [Google Scholar]
  257. Li X.-H.; Ye A.-H.; Liang C.; Mo D.-L. Substituent Effects of 2-Pyridones on Selective O-Arylation with Diaryliodonium Salts: Synthesis of 2-Aryloxypyridines under Transition-Metal-Free Conditions. Synthesis 2018, 50, 1699–1710. 10.1055/s-0036-1591884. [DOI] [Google Scholar]
  258. Qi C.; Guo T.; Xiong W.; Wang L.; Jiang H. Silver-Promoted Coupling of Carbon Dioxide, o-Alkynylanilines and Diaryliodonium Salts: Straightforward Access to 4-Aryloxy-2-quinolinones. ChemistrySelect 2017, 2, 4691–4695. 10.1002/slct.201701045. [DOI] [Google Scholar]
  259. Mehra M. K.; Tantak M. P.; Arun V.; Kumar I.; Kumar D. Metal-free regioselective formation of C-N and C-O bonds with the utilization of diaryliodonium salts in water: facile synthesis of N-arylquinolones and aryloxyquinolines. Org. Biomol. Chem. 2017, 15, 4956–4961. 10.1039/C7OB00940B. [DOI] [PubMed] [Google Scholar]
  260. Kuriyama M.; Hanazawa N.; Abe Y.; Katagiri K.; Ono S.; Yamamoto O. K.; Onomura N- and O-arylation of pyridin-2-ones with diaryliodonium salts: base-dependent orthogonal selectivity under metal-free conditions. Chem. Sci. 2020, 11, 8295–8300. 10.1039/D0SC02516J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  261. Kuriyama M.; Mochizuki Y.; Miyagi T.; Yamamoto K.; Demizu Y.; Onomura O. Transition-Metal Free O-arylation of Quinoxalin-2-ones with Diaryliodonium Salts. Heterocycles 2021, 103, 502–510. 10.3987/COM-20-S(K)26. [DOI] [Google Scholar]
  262. Elboray E. E.; Bae T.; Kikushima K.; Kita Y.; Dohi T. Transition Metal-Free O-Arylation of N-Alkoxybenzamides Enabled by Aryl(trimethoxyphenyl)iodonium Salts. Adv. Synth. Catal. 2023, 365, 2703–2710. 10.1002/adsc.202300406. [DOI] [Google Scholar]
  263. Chen K.; Koser G. F. Direct and Regiocontrolled Synthesis of α-Phenyl Ketones from Silyl Enol Ethers and Diphenyliodonium Fluoride. J. Org. Chem. 1991, 56, 5764–5767. 10.1021/jo00020a013. [DOI] [Google Scholar]
  264. Iwama T.; Birman V. B.; Kozmin S. A.; Rawal V. H. Regiocontrolled Synthesis of Carbocycle-Fused Indoles via Arylation of Silyl Enol Ethers with o-Nitrophenylphenyliodonium Fluoride. Org. Lett. 1999, 1, 673–676. 10.1021/ol990759j. [DOI] [PubMed] [Google Scholar]
  265. Qian X.; Han J.; Wang L. tert-Butoxide-Mediated Arylation of 2-Substituted Cyanoacetates with Diaryliodonium Salts. Adv. Synth. Catal. 2016, 358, 940–946. 10.1002/adsc.201501013. [DOI] [Google Scholar]
  266. Oh C. H.; Kim J. S.; Jung H. H. Highly Efficient Arylation of Malonates with Diaryliodonium Salts. J. Org. Chem. 1999, 64, 1338–1340. 10.1021/jo981065b. [DOI] [Google Scholar]
  267. Han J.; Qian X.; Xu B.; Wang L. Potassium tert-Butoxide Mediated Arylation of 2-Substituted Malononitriles Using Diaryliodonium Salts. Synlett 2017, 28, 2139–2142. 10.1055/s-0036-1589061. [DOI] [Google Scholar]
  268. Monastyrskyi A.; Namelikonda N. K.; Manetsch R. Metal-Free Arylation of Ethyl Acetoacetate with Hypervalent Diaryliodonium Salts: An Immediate Access to Diverse 3-Aryl-4(1H)-Quinolone. J. Org. Chem. 2015, 80, 2513–2520. 10.1021/jo5023958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Dey C.; Lindstedt E.; Olofsson B. Metal-Free C-Arylation of Nitro Compounds with Diaryliodonium Salts. Org. Lett. 2015, 17, 4554–4557. 10.1021/acs.orglett.5b02270. [DOI] [PubMed] [Google Scholar]
  270. Matsuzaki K.; Okuyama K.; Tokunaga E.; Shiro M.; Shibata N. Sterically Demanding Unsymmetrical Diaryl-λ3-iodanes for Electrophilic Pentafluorophenylation and an Approach to α-Pentafluorophenyl Carbonyl Compounds with an All-Carbon Stereocenter. ChemistryOpen 2014, 3, 233–237. 10.1002/open.201402045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Das P.; Shibata N. Electrophilic Triflyl-arylation and Triflyl-pyridylation by Unsymmetrical Aryl/Pyridyl-λ3 -iodonium Salts: Synthesis of Aryl and Pyridyl Triflones. J. Org. Chem. 2017, 82, 11915–11924. 10.1021/acs.joc.7b01690. [DOI] [PubMed] [Google Scholar]
  272. Matsuzaki K.; Okuyama K.; Tokunaga E.; Saito N.; Shiro M.; Shibata N. Synthesis of Diaryliodonium Salts Having Pentafluorosulfanylarenes and Their Application to Electrophilic Pentafluorosulfanylarylation of C-, O-, N-, and S-Nucleophiles. Org. Lett. 2015, 17, 3038–3041. 10.1021/acs.orglett.5b01323. [DOI] [PubMed] [Google Scholar]
  273. Aggarwal V. K.; Olofsson B. Enantioselective a-Arylation of Cyclohexanones with Diaryl Iodonium Salts: Application to the Synthesis of (−)-Epibatidine. Angew. Chem., Int. Ed. 2005, 44, 5516–5519. 10.1002/anie.200501745. [DOI] [PubMed] [Google Scholar]
  274. Li X.; Ni W.; Mao F.; Wang W.; Li J. A Metal-free Approach to 3-Aryl-3-hydroxy-2-oxindoles by Treatment of 3-Acyloxy-2-oxindoles with Diaryliodonium Salts. Chem.—Asian J. 2016, 11, 226–230. 10.1002/asia.201500854. [DOI] [PubMed] [Google Scholar]
  275. Zhang Y.; Han J.; Liu Z.-J. tert-Butoxide-Mediated Arylation of 1-Acetylindolin-3-ones with Diaryliodonium Salts. Synlett 2015, 26, 2593–2597. 10.1055/s-0035-1560575. [DOI] [Google Scholar]
  276. Savechenkov P. Y.; Zhang X.; Chiara D. C.; Stewart D. S.; Ge R.; Zhou X.; Raines D. E.; Cohen J. B.; Forman S. A.; Miller K. W.; Bruzik K. S. Allyl m-Trifluoromethyldiazirine Mephobarbital: An Unusually Potent Enantioselective and Photoreactive Barbiturate General Anesthetic. J. Med. Chem. 2012, 55, 6554–6565. 10.1021/jm300631e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  277. Chai Z.; Wang B.; Chen J.-N.; Yang G. Arylation of Azlactones using Diaryliodonium Bromides and Silver Carbonate: Facile Access to a-Tetrasubstituted Amino Acid Derivatives. Adv. Synth. Catal. 2014, 356, 2714–2718. 10.1002/adsc.201400035. [DOI] [Google Scholar]
  278. Mao S.; Geng X.; Yang Y.; Qian X.; Wu S.; Han J.; Wang L. Base promoted direct C4-arylation of 4- substituted-pyrazolin-5-ones with diaryliodonium salts. RSC Adv. 2015, 5, 36390–36393. 10.1039/C5RA03819G. [DOI] [Google Scholar]
  279. Norrby P.-O.; Petersen T. B.; Bielawski M.; Olofsson B. α-Arylation by Rearrangement: On the Reaction of Enolates with Diaryliodonium Salts. Chem.—Eur. J. 2010, 16, 8251–8254. 10.1002/chem.201001110. [DOI] [PubMed] [Google Scholar]
  280. An Y.; Zhang X.-M.; Li Z.-Y.; Xiong W.-H.; Yu R.-D.; Zhang F.-M. Transition-metal-free α-arylation of nitroketones with diaryliodonium salts for the synthesis of tertiary α-aryl, α-nitro ketones. Chem. Commun. 2019, 55, 119–122. 10.1039/C8CC08920E. [DOI] [PubMed] [Google Scholar]
  281. 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]
  282. Zaheer M. K.; Vaishanv N. K.; Kumar A.; Mishra S.; Kant R.; Mohanan K. Synthesis 2023, 55, 3382–3392. 10.1055/a-2109-1419. [DOI] [Google Scholar]
  283. Zaheer M. K.; Vaishanv N. K.; Kant R.; Mohanan K. Utilization of Unsymmetric Diaryliodonium Salts in α-Arylation of α-Fluoroacetoacetamides. Chem.—Asian J. 2020, 15, 4297–4301. 10.1002/asia.202001160. [DOI] [PubMed] [Google Scholar]
  284. Knieb A.; Krishnamurti V.; Zhu Z.; Prakash G. K. S. Synthesis of Monofluoromethylarenes: Direct Monofluoromethylation of Diaryliodonium Bromides using Fluorobis(phenylsulfonyl) methane (FBSM). J. Fluor. Chem. 2023, 267, 110095. 10.1016/j.jfluchem.2023.110095. [DOI] [Google Scholar]
  285. Jiang X.; Meyer D.; Baran D.; González M. A. C.; Szabó K. J. Trifluoromethylthiolation, Trifluoromethylation, and Arylation Reactions of Difluoro Enol Silyl Ethers. J. Org. Chem. 2020, 85, 8311–8319. 10.1021/acs.joc.0c01030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Kikushima K.; Yamada K.; Umekawa N.; Yoshio N.; Kita Y.; Dohi T. Decarboxylative arylation with diaryliodonium(III) salts: alternative approach for catalyst-free difluoroenolate coupling to aryldifluoromethyl ketones. Green Chem. 2023, 25, 1790–1796. 10.1039/D2GC04445E. [DOI] [Google Scholar]
  287. Hori M.; Guo J.-D.; Yanagi T.; Nogi K.; Sasamori T.; Yorimitsu H. Sigmatropic Rearrangements of Hypervalent-Iodine-Tethered Intermediates for the Synthesis of Biaryls. Angew. Chem., Int. Ed. 2018, 57, 4663–4667. 10.1002/anie.201801132. [DOI] [PubMed] [Google Scholar]
  288. Ghosh M. K.; Rzymkowski J.; Kalek M. 0 Transition-Metal-Free Aryl-Aryl Cross-Coupling: C-H Arylation of 2-Naphthols with Diaryliodonium Salts. Chem.—Eur. J. 2019, 25, 9619–9623. 10.1002/chem.201902204. [DOI] [PubMed] [Google Scholar]
  289. Ozanne-Beaudenon A.; Quideau S. Regioselective Hypervalent-Iodine(iii)-Mediated Dearomatizing Phenylation of Phenols through Direct Ligand Coupling. Angew. Chem., Int. Ed. 2005, 44, 7065–7069. 10.1002/anie.200501638. [DOI] [PubMed] [Google Scholar]
  290. Wang H.; Greaney M. F. Regiodivergent Arylation of Pyridines via Zincke Intermediates. Angew. Chem., Int. Ed. 2024, 63, e202315418 10.1002/anie.202315418. [DOI] [PubMed] [Google Scholar]
  291. Sandin R. B.; Christiansen R. G.; Brown R. K.; Kirkwood S. The Reaction of Iodonium Salts with Thiol Compounds. J. Am. Chem. Soc. 1947, 69, 1550. 10.1021/ja01198a523. [DOI] [Google Scholar]
  292. Wang D.; Yu X.; Zhao K.; Li L.; Ding Y. Rapid synthesis of aryl sulfides through metal-free C-S coupling of thioalcohols with diaryliodonium salts. Tetrahedron Lett. 2014, 55, 5739–5741. 10.1016/j.tetlet.2014.08.088. [DOI] [Google Scholar]
  293. Krief A.; Dumont W.; Robert M. Arylation of n-Hexylthiol and n-Hexyl Phenyl Sulfide Using Diphenyliodonium Triflate: Synthetic and Mechanistic Aspects - Application to the Transformation of n-Hexylthiol to n-Hexylselenide. Synlett 2006, 2006, 484–4861. 10.1055/s-2006-926235. [DOI] [Google Scholar]
  294. Huang X.; Zhu Q.; Xu Y. Synthesis of Polymeric Diaryliodonium Salts and Its Use in Preparation of Diaryl Sulfides and diaryl Ethers. Synth. Commun. 2001, 31, 2823–2828. 10.1081/SCC-100105332. [DOI] [Google Scholar]
  295. Wagner A. M.; Sanford M. S. Transition-Metal-Free Acid-Mediated Synthesis of Aryl Sulfides from Thiols and Thioethers. J. Org. Chem. 2014, 79, 2263–2267. 10.1021/jo402567b. [DOI] [PubMed] [Google Scholar]
  296. Li X.-H.; Li L.-G.; Mo X.-L.; Mo D.-L. Transition-metal-free synthesis of thiocyanato- or nitro-arenes through diaryliodonium salts. Synth. Commun. 2016, 46, 963–970. 10.1080/00397911.2016.1181764. [DOI] [Google Scholar]
  297. Vaddula B. R.; Varma R. S.; Leazer J. One-Pot Catalyst-Free Synthesis of β- and γ-Hydroxy Sulfides Using Diaryliodonium Salts and Microwave Irradiation. Eur. J. Org. Chem. 2012, 2012, 6852–6855. 10.1002/ejoc.201201313. [DOI] [Google Scholar]
  298. You J.; Chen Z.-C. Hypervalent Iodine in Synthesis; 5. The Reaction Between Diaryliodonium salts and Thiocarboxylic Acid Salts: A Convenient Method for the Synthesis of S-Aryl Thiocarboxylates. Synthesis 1992, 1992, 521–522. 10.1055/s-1992-26150. [DOI] [Google Scholar]
  299. Xia M.; Chen Z.-C. Hypervalent Iodine in Synthesis XXII: A Novel Way for The Preparation of Unsymmetric S-Aryl Thiosulfonates by The Reaction of Potassium Thiosulfontes with Diaryliodonium salts. Synth. Commun. 1997, 27, 1309–1313. 10.1080/00397919708006058. [DOI] [Google Scholar]
  300. Chen Z.; Jin Y.; Stang P. J. Polyvalent Iodine in Synthesis. 2. A New Method for the Preparation of Aryl Esters of Dithiocarbamic Acids. J. Org. Chem. 1987, 52, 4117–4118. 10.1021/jo00227a032. [DOI] [Google Scholar]
  301. Racicot L.; Kasahara T.; Ciufoliní M. A. Arylation of Diorganochalcogen Compounds with Diaryliodonium Triflates: Metal Catalysts Are Unnecessary. Org. Lett. 2014, 16, 6382–6385. 10.1021/ol503177q. [DOI] [PubMed] [Google Scholar]
  302. Takenaga N.; Yoto Y.; Hayashi T.; Miyamoto N.; Nojiri H.; Kumar R.; Dohi T. Catalytic and non-catalytic selective aryl transfer from (mesityl)iodonium(III) salts to diarylsulfide compounds. Arkivoc 2023, 2022, 7–18. 10.24820/ark.5550190.p011.732. [DOI] [Google Scholar]
  303. Kumar D.; Arun V.; Pilania M.; Shekar K. P. C. A Metal-Free and Microwave-Assisted Efficient Synthesis of Diaryl Sulfones. Synlett 2013, 24, 831–836. 10.1055/s-0032-1317803. [DOI] [Google Scholar]
  304. Umierski N.; Manolikakes G. Metal-Free Synthesis of Diaryl Sulfones from Arylsulfinic Acid Salts and Diaryliodonium Salts. Org. Lett. 2013, 15, 188–191. 10.1021/ol303248h. [DOI] [PubMed] [Google Scholar]
  305. Umierski N.; Manolikakes G. Arylation of Lithium Sulfinates with Diaryliodonium Salts: A Direct and Versatile Access to Arylsulfones. Org. Lett. 2013, 15, 4972–4975. 10.1021/ol402235v. [DOI] [PubMed] [Google Scholar]
  306. Margraf N.; Manolikakes G. One-Pot Synthesis of Aryl Sulfones from Organometallic Reagents and Iodonium Salts. J. Org. Chem. 2015, 80, 2582–2600. 10.1021/jo5027518. [DOI] [PubMed] [Google Scholar]
  307. Luo J.; Chen X.; Ding W.; Ma J.; Ni Z.; Xie L.; Xu C. Transition-metal-free glycosyl sulfonation of diaryliodonium salts with sodium glycosyl sulfinate: an efficient approach to access glycosyl aryl sulfones. New J. Chem. 2024, 48, 4218–4223. 10.1039/D3NJ05942A. [DOI] [Google Scholar]
  308. Sarkar S.; Kalek M. Metal-Free S-Arylation of Phosphorothioate Diesters and Related Compounds with Diaryliodonium Salts. Org. Lett. 2023, 25, 671–675. 10.1021/acs.orglett.2c04310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  309. Villo P.; Kervefors G.; Olofsson B. Transition metal-free, chemoselective arylation of thioamides yielding aryl thioimidates or N-aryl thioamides. Chem. Commun. 2018, 54, 8810–8813. 10.1039/C8CC04795B. [DOI] [PubMed] [Google Scholar]
  310. Zhang B.; Kang Y.; Yang L.; Chen X. A S-Arylation of 2-Mercaptoazoles with Diaryliodonium Triflates under Metal-Free and Base-Free Conditions. ChemistrySelect 2016, 1, 1529–1532. 10.1002/slct.201600204. [DOI] [Google Scholar]
  311. Sarkar S.; Wojciechowska N.; Rajkiewicz A. A.; Kalek M. Synthesis of Aryl Sulfides by Metal-Free Arylation of Thiols with Diaryliodonium Salts under Basic Conditions. Eur. J. Org. Chem. 2022, 2022, e202101408 10.1002/ejoc.202101408. [DOI] [Google Scholar]
  312. Saikia R. A.; Hazarik N.; Biswakarma N.; Deka R. C.; Thakur A. J. Metal-free S-arylation of 5-mercaptotetrazoles and 2-mercaptopyridine with unsymmetrical diaryliodonium salts. Org. Biomol. Chem. 2022, 20, 3890–3896. 10.1039/D2OB00406B. [DOI] [PubMed] [Google Scholar]
  313. Zhang Y.; Wang Y.; Wang L.; Han J. Selective S-arylation of thiols with o-OTf-substituted diaryliodonium salts toward diarylsulfides. Org. Biomol. Chem. 2024, 22, 486–490. 10.1039/D3OB01922E. [DOI] [PubMed] [Google Scholar]
  314. Byrne S. A.; Bedding M. J.; Corcilius L.; Ford D. J.; Zhong Y.; Franck C.; Larance M.; Mackay J. P.; Payne R. J. Late-stage modification of peptides and proteins at cysteine with diaryliodonium salts. Chem. Sci. 2021, 12, 14159–14166. 10.1039/D1SC03127A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Wang X.; Luo D.; Wang X.; Zeng X.; Wang X.; Hu Y. N, N′-Disulfonylhydrazines: A novel source of sulfonyl moieties for synthesis of diaryl sulfones. Tetrahedron Lett. 2021, 87, 153540. 10.1016/j.tetlet.2021.153540. [DOI] [Google Scholar]
  316. Gong B.; Zhu H.; Yang L.; Wang H.; Fan Q.; Xie Z.; Le Z. Base-promoted synthesis of diarylsulfones from sulfonyl hydrazines and diaryliodonium salts. Org. Biomol. Chem. 2022, 20, 3501–3505. 10.1039/D2OB00389A. [DOI] [PubMed] [Google Scholar]
  317. Qiu Y.; Yao J.; Xia H.; Zhu C.; Ye S.; Wu J.; Chen Q. An Aqueous Three-Component Reaction of Cyclopropanols, DABCO·(SO2)2 and Diaryliodonium salts. Adv. Synth. Catal. 2023, 365, 3392–3396. 10.1002/adsc.202300839. [DOI] [Google Scholar]
  318. Yu H.; Li Z.; Bolm C. Transition-Metal-Free Arylations of In-Situ Generated Sulfenates with Diaryliodonium Salts. Org. Lett. 2018, 20, 7104–7106. 10.1021/acs.orglett.8b03046. [DOI] [PubMed] [Google Scholar]
  319. Wang L.; Chen M.; Zhang J. Transition metal-free base-promoted arylation of sulfenate anions with diaryliodonium salts. Org. Chem. Front. 2019, 6, 32–35. 10.1039/C8QO00914G. [DOI] [Google Scholar]
  320. Liu W.; Zhang Y.; Xing S.; Lan H.; Chen X.; Bai Y.; Shao X. β-Trifluorosulfinylesters: tuneable reagents for switchable trifluoromethylsulfinylation and C-H trifluoromethylthiolation. Org. Chem. Front. 2023, 10, 2186–2192. 10.1039/D3QO00013C. [DOI] [Google Scholar]
  321. 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]
  322. Bugaenko D. I.; Volkov A. A.; Andreychev V. V.; Karchava A. V. Reaction of Diaryliodonium Salts with Potassium Alkyl Xanthates as an Entry Point to Accessing Organosulfur Compounds. Org. Lett. 2023, 25, 272–276. 10.1021/acs.orglett.2c04143. [DOI] [PubMed] [Google Scholar]
  323. Bugaenko D. I.; Volkov A. A.; Karchava A. V. A Thiol-Free Route to Alkyl Aryl Thioethers. J. Org. Chem. 2023, 88, 9968–9972. 10.1021/acs.joc.3c00734. [DOI] [PubMed] [Google Scholar]
  324. Huang G.; Lu X.; Liang F. Redox-Neutral Strategy for Sulfilimines Synthesis via S-Arylation of Sulfenamides. Org. Lett. 2023, 25, 3179–3183. 10.1021/acs.orglett.3c01077. [DOI] [PubMed] [Google Scholar]
  325. Zhou Q.; Li J.; Wang T.; Yang X. Base-Promoted S-Arylation of Sulfenamides for the Synthesis of Sulfilimines. Org. Lett. 2023, 25, 4335–4339. 10.1021/acs.orglett.3c01436. [DOI] [PubMed] [Google Scholar]
  326. Wu X.; Li Y.; Chen M.; He F.-S.; Wu J. Metal-Free Chemoselective S-Arylation of Sulfenamides To Access Sulfilimines. J. Org. Chem. 2023, 88, 9352–9359. 10.1021/acs.joc.3c00961. [DOI] [PubMed] [Google Scholar]
  327. Liu Z.-D.; Chen Z.-C. Synthesis of Aryl phosphonates by Arylation of Phosphite Anions Using Diaryliodonium salts. Synthesis 1993, 1993, 373–374. 10.1055/s-1993-25865. [DOI] [Google Scholar]
  328. van Druenen M.; Davitt F.; Collins T.; Glynn C.; O’Dwyer C.; Holmes J. D.; Collins G. Covalent Functionalization of Few-Layer Black Phosphorus Using Iodonium Salts and Comparison to Diazonium Modified Black Phosphorus. Chem. Mater. 2018, 30, 4667–4674. 10.1021/acs.chemmater.8b01306. [DOI] [Google Scholar]
  329. Miralles N.; Romero R. M.; Fernández E.; Muñiz K. A mild carbon-boron bond formation from diaryliodonium salts. Chem. Commun. 2015, 51, 14068–14071. 10.1039/C5CC04944J. [DOI] [PubMed] [Google Scholar]
  330. Yoshimura A.; Shea M. T.; Guselnikova O.; Postnikov P. S.; Rohde G. T.; Saito A.; Yusubov M. S.; Nemykine V. N.; Zhdankin V. V. Preparation and structure of phenolic aryliodonium salts. Chem. Commun. 2018, 54, 10363–10366. 10.1039/C8CC06211K. [DOI] [PubMed] [Google Scholar]
  331. 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]
  332. 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]
  333. Sathunuru R.; Biehl E. Facile synthesis of 4H-naphtho[2,3-e] derivatives of 1,3-thiazines and 1,3-selenazines and naphtho[2’,3′:4,5] derivatives of selenolo[2,3-b]pyridines and thieno[2,3-b]pyridines via 2,3- didehydronaphthalene. Arkivoc 2004, 2004, 51–60. 10.3998/ark.5550190.0005.e05. [DOI] [Google Scholar]
  334. Liu Z.-D.; Zeng H.; Chen Z.-C. Hypervalent Iodine in Synthesis XV: A Convenient New Method for The Preparation of Unsymmetrical Diaryl Selenides. Synth. Commun. 1994, 24, 475–479. 10.1080/00397919408011497. [DOI] [Google Scholar]
  335. Liu Z.-D.; Chen D.-C. A Convenient New Method for The Preparation of Diaryl Diselenide. Synth. Commun. 1993, 23, 2673–2676. 10.1080/00397919308013796. [DOI] [Google Scholar]
  336. 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]
  337. Guan Y.; Townsend S. D. Metal-Free Synthesis of Unsymmetrical Organoselenides and Selenoglycosides. Org. Lett. 2017, 19, 5252–5255. 10.1021/acs.orglett.7b02526. [DOI] [PubMed] [Google Scholar]
  338. Dong T.; He J.; Li Z.-H.; Zhang C.-P. Catalyst- and Additive-Free Trifluoromethylselenolation with [Me4N][SeCF3]. ACS Sustainable Chem. Eng. 2018, 6, 1327–1335. 10.1021/acssuschemeng.7b03673. [DOI] [Google Scholar]
  339. Radzhabov A. D.; Soldatova N. S.; Ivanov D. M.; Yusubov M. S.; Kukushkin V. Yu.; Postnikov P. S. Metal-free and atom-efficient protocol for diarylation of selenocyanate by diaryliodonium salts. Org. Biomol. Chem. 2023, 21, 6743–6749. 10.1039/D3OB00833A. [DOI] [PubMed] [Google Scholar]
  340. Motiwala H. F.; Armaly A. M.; Cacioppo J. G.; Coombs T. C.; Koehn K. R. K.; Norwood V. M.; Aubé J. HFIP in Organic Synthesis. Chem. Rev. 2022, 122, 12544–12747. 10.1021/acs.chemrev.1c00749. [DOI] [PubMed] [Google Scholar]
  341. Linstad E. J.; Va̅vere A. L.; Hu B.; Kempinger J. J.; Snyder S. E.; DiMagno S. G. Thermolysis and radiofluorination of diaryliodonium salts derived from anilines. Org. Biomol. Chem. 2017, 15, 2246–2252. 10.1039/C7OB00253J. [DOI] [PubMed] [Google Scholar]
  342. Hu B.; Miller W. H.; Neumann K. D.; Linstad E. J.; DiMagno S. G. An Alternative to the Sandmeyer Approach to Aryl Iodides. Chem.—Eur. J. 2015, 21, 6394–6398. 10.1002/chem.201500151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  343. Moriyama K.; Ishida K.; Togo H. Regioselective Csp2-H dual functionalization of indoles using hypervalent iodine(III): bromo-amination via 1,3-migration of imides on indolyl(phenyl)iodonium imides. Chem. Commun. 2015, 51, 2273–2276. 10.1039/C4CC09077B. [DOI] [PubMed] [Google Scholar]
  344. Gruber S.; Ametamey S. M.; Schibli R. Unexpected reac.tivity of cyclic perfluorinated iodanes with electrophiles. Chem. Commun. 2018, 54, 8999–9002. 10.1039/C8CC04558E. [DOI] [PubMed] [Google Scholar]
  345. Sadek O.; Perrin D. M.; Gras E. Unsymmetrical diaryliodonium phenyltrifluoroborate salts: Synthesis, structure and fluorination. J. Fluor. Chem. 2019, 222–223, 68–74. 10.1016/j.jfluchem.2019.04.004. [DOI] [Google Scholar]
  346. Zhang M.-R.; Kumata K.; Takei M.; Fukumura T.; Suzuki K. How to introduce radioactive chlorine into a benzene ring using [*Cl]Cl?. Appl. Radiat. Isot. 2008, 66, 1341–1345. 10.1016/j.apradiso.2008.04.011. [DOI] [PubMed] [Google Scholar]
  347. Zhou D.; Kim S. H.; Chu W.; Voller T.; Katzenellenbogen J. A. Evaluation of aromatic radiobromination by nucleophilic substitution using diaryliodonium salt precursors. J. Label Compd. Radiopharm. 2017, 60, 450–456. 10.1002/jlcr.3519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  348. Guérard F.; Navarro L.; Lee Y.-S.; Roumesy A.; Alliot C.; Chérel M.; Brechbiel M. W.; Gestin J.-F. Bifunctional aryliodonium salts for highly efficient radioiodination and astatination of antibodies. Bioorg. Med. Chem. 2017, 25, 5975–5980. 10.1016/j.bmc.2017.09.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  349. Navarro L.; Berdal M.; Chérel M.; Pecorari F.; Gestin J.-F.; Guérard F. Prosthetic groups for radioiodination and astatination of peptides and proteins: A comparative study of five potential bioorthogonal labeling strategies. Bioorg. Med. Chem. 2019, 27, 167–174. 10.1016/j.bmc.2018.11.034. [DOI] [PubMed] [Google Scholar]
  350. Pike V. W.; Aigbirhio F. I. Reactions of Cyclotron-produced [18F]Fluoride with Diaryliodonium Salts-a Novel Single-step Route to No-carrier-added [18]Fluoroarenes. J. Chem. Soc. Chem. Commun. 1995, 21, 2215–2216. 10.1039/C39950002215. [DOI] [Google Scholar]
  351. Rotstein B. H.; Stephenson N. A.; Vasdev N.; Liang S. H. Spirocyclic hypervalent iodine(III)-mediated radiofluorination of non-activated and hindered aromatics. Nat. Commun. 2014, 5, 4365. 10.1038/ncomms5365. [DOI] [PubMed] [Google Scholar]
  352. Wang L.; Jacobson O.; Avdic D.; Rotstein B. H.; Weiss I. D.; Collier L.; Chen X.; Vasdev N.; Liang S. H. Ortho-Stabilized 18F-Azido Click Agents and their Application in PET Imaging with Single-Stranded DNA Aptamers. Angew. Chem., Int. Ed. 2015, 54, 12777–12781. 10.1002/anie.201505927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  353. Yuan Z.; Cheng R.; Chen P.; Liu G.; Liang S. H. Efficient Pathway for the Preparation of Aryl(isoquinoline)iodonium- (III) Salts and Synthesis of Radiofluorinated Isoquinolines. Angew. Chem., Int. Ed. 2016, 55, 11882–11886. 10.1002/anie.201606381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  354. Wang J.; Holloway T.; Lisova K.; van Dam R. M. Green and efficient synthesis of the radiopharmaceutical [18F]FDOPA using a microdroplet reactor. React. Chem. Eng. 2020, 5, 320–329. 10.1039/C9RE00354A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  355. Rotstein B. H.; Wang L.; Liu R. Y.; Patteson J.; Kwan E. E.; Vasdev N.; Liang S. H. Mechanistic studies and radiofluorination of structurally diverse pharmaceuticals with spirocyclic iodonium(III) ylides. Chem. Sci. 2016, 7, 4407–4417. 10.1039/C6SC00197A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  356. Yusubov M. S.; Svitich D. Yu.; Larkina M. S.; Zhdankin V. V. Applications of iodonium salts and iodonium ylides as precursors for nucleophilic fluorination in Positron Emission Tomography. Arkivoc 2013, 2013, 364–395. 10.3998/ark.5550190.p008.225. [DOI] [Google Scholar]
  357. Jacobson O.; Kiesewetter D. O.; Chen X. Fluorine-18 Radiochemistry, Labeling Strategies and Synthetic Routes. Bioconjugate Chem. 2015, 26, 1–18. 10.1021/bc500475e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  358. Pike V. W. Hypervalent aryliodine compounds as precursors for radiofluorination. J. Label Compd Radiopharm. 2018, 61, 196–227. 10.1002/jlcr.3570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  359. Krasikova R. N. Nucleophilic Synthesis of 6-L-[18F]FDOPA. Is Copper-Mediated Radiofluorination the Answer?. Molecules 2020, 25, 4365. 10.3390/molecules25194365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  360. Kikushima K.; Kumar R.; Dohi T. Progress in [18F]Fluorination by Using Aryliodonium(III) Compounds and Application for PET Tracer Syntheses. Mini Rev. Org. Chem. 2021, 18, 173–195. 10.2174/1570193X17999200629155733. [DOI] [Google Scholar]
  361. Chassé M.; Pees A.; Lindberg A.; Liang S. H.; Vasdev N. Spirocyclic Iodonium Ylides for Fluorine-18 Radiolabeling of Non-Activated Arenes: From Concept to Clinical Research. Chem. Rec. 2023, 23, e202300072 10.1002/tcr.202300072. [DOI] [PubMed] [Google Scholar]
  362. Matsuoka K.; Obata H.; Nagatsu K.; Kojima M.; Yoshino T.; Ogawa M.; Matsunaga S. Transition-metal-free nucleophilic 211At-astatination of spirocyclic aryliodonium ylides. Org. Biomol. Chem. 2021, 19, 5525–5528. 10.1039/D1OB00789K. [DOI] [PubMed] [Google Scholar]
  363. Maingueneau C.; Berdal M.; Eychenne R.; Gaschet J.; Chérel M.; Gestin J.-F.; Guérard F. 211At and 125I-Labeling of (Hetero)Aryliodonium Ylides: Astatine Wins Again. Chem.—Eur. J. 2022, 28, e202104169 10.1002/chem.202104169. [DOI] [PubMed] [Google Scholar]
  364. Guérard F.; Lee Y.-S.; Baidoo K.; Gestin J.-F.; Brechbiel M. W. Unexpected Behavior of the Heaviest Halogen Astatine in the Nucleophilic Substitution of Aryliodonium Salts. Chem.—Eur. J. 2016, 22, 12332–12339. 10.1002/chem.201600922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  365. Kwon Y.-D.; Son J.; Chun J.-H. Chemoselective Radiosyntheses of Electron-Rich [18F]Fluoroarenes from Aryl(2,4,6-trimethoxyphenyl)iodonium Tosylates. J. Org. Chem. 2019, 84, 3678–3686. 10.1021/acs.joc.9b00019. [DOI] [PubMed] [Google Scholar]
  366. Zhao Q.; Etersque J. M.; Lu S.; Telu S.; Pike V. W. On the Risk of 18F-Regioisomer Formation in the Copper-Free Radiofluorination of Aryliodonium Precursors. Org. Lett. 2023, 25, 8650–8654. 10.1021/acs.orglett.3c03499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  367. Cardinale J.; Ermert J.; Humpert S.; Coenen H. H. Iodonium ylides for one-step, no-carrier-added radiofluorination of electron rich arenes, exemplified with 4-(([18F]fluorophenoxy)- phenylmethyl)piperidine NET and SERT ligands. RSC Adv. 2014, 4, 17293–17299. 10.1039/C4RA00674G. [DOI] [Google Scholar]
  368. Jakobsson J. E.; Riss P. J. Transition metal free, late-stage, regiospecific, aromatic fluorination on a preparative scale using a KF/crypt-222 complex. RSC Adv. 2018, 8, 21288–21291. 10.1039/C8RA03757D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  369. Landge K. P.; Jang K. S.; Lee S. Y.; Chi D. Y. Approach to the Synthesis of Indoline Derivatives from Diaryliodonium Salts. J. Org. Chem. 2012, 77, 5705–5713. 10.1021/jo300874m. [DOI] [PubMed] [Google Scholar]
  370. Shi W.-M.; Li X.-H.; Liang C.; Mo D.-L. Base-Free Selective O-Arylation and Sequential [3,3]-Rearrangement of Amidoximes with Diaryliodonium Salts: Synthesis of 2-Substituted Benzoxazoles. Adv. Synth. Catal. 2017, 359, 4129–4135. 10.1002/adsc.201700906. [DOI] [Google Scholar]
  371. Ma X.-P.; Li K.; Wu S.-Y.; Liang C.; Su G.-F.; Mo D.-L. Construction of 2,3-quaternary fused indolines from alkynyl tethered oximes and diaryliodonium salts through a cascade strategy of N-arylation/ cycloaddition/[3,3]-rearrangement. Green Chem. 2017, 19, 5761–5766. 10.1039/C7GC02844J. [DOI] [Google Scholar]
  372. Lovato K.; Bhakta U.; Ng Y. P.; Kürti L. O-Cyclopropyl hydroxylamines: gram-scale synthesis and utility as precursors for N-heterocycles. Org. Biomol. Chem. 2020, 18, 3281–3287. 10.1039/D0OB00611D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  373. Mao S.; Hua Z.; Wu X.; Yang Y.; Han J.; Wang L. Tandem Arylation/Friedel-Crafts Reaction of O-Hydroxy Bisbenzylic Alcohols with Diaryliodonium Salts: One Pot Synthesis of Unsymmetrical 9-Arylxanthenes. ChemistrySelect 2016, 1, 403–407. 10.1002/slct.201600036. [DOI] [Google Scholar]
  374. 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]
  375. Kvasovs N.; Gevorgyan V. Contemporary methods for generation of aryl radicals. Chem. Soc. Rev. 2021, 50, 2244–2259. 10.1039/D0CS00589D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  376. Wang X.; Studer A. Iodine(III) Reagents in Radical Chemistry. Acc. Chem. Res. 2017, 50, 1712–1724. 10.1021/acs.accounts.7b00148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  377. Bellina F. Real Metal-Free C-H Arylation of (Hetero)arenes: The Radical Way. Synthesis 2021, 53, 2517–2544. 10.1055/a-1437-9761. [DOI] [Google Scholar]
  378. Narayan R.; Manna S.; Antonchick A. P. Hypervalent Iodine(III) in Direct Carbon-Hydrogen Bond Functionalization. Synlett 2015, 26, 1785–1803. 10.1055/s-0034-1379912. [DOI] [Google Scholar]
  379. Besson T.; Fruit C. Recent Advances in Transition-Metal-Free Late-Stage C-H and N-H Arylation of Heteroarenes Using Diaryliodonium Salts. Pharmaceuticals 2021, 14, 661. 10.3390/ph14070661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  380. Ghosh M. K.; Rout N. Aryl-Aryl Cross-Coupling with Hypervalent Iodine Reagents: Aryl Group Transfer Reactions. ChemistrySelect 2020, 5, 13644–13655. 10.1002/slct.202003396. [DOI] [Google Scholar]
  381. Morimoto K. Metal-Free Oxidative Cross-Coupling Reaction of Heteroaromatic and Related Compounds. Chem. Pharm. Bull. 2019, 67, 1259–1270. 10.1248/cpb.c19-00286. [DOI] [PubMed] [Google Scholar]
  382. Rossi R.; Lessi M.; Manzini C.; Marianetti G.; Bellina F. Transition Metal-Free Direct C-H (Hetero)arylation of Heteroarenes: A Sustainable Methodology to Access (Hetero)aryl-Substituted Heteroarenes. Adv. Synth. Catal. 2015, 357, 3777–3814. 10.1002/adsc.201500799. [DOI] [Google Scholar]
  383. Morimoto K.; Dohi T.; Kita Y. Metal-free Oxidative Cross-Coupling Reaction of Aromatic Compounds Containing Heteroatoms. Synlett 2017, 28, 1680–1694. 10.1055/s-0036-1588455. [DOI] [Google Scholar]
  384. Castro S.; Fernández J. J.; Vicente R.; Fañanás F. J.; Rodríguez F. Base- and metal-free C-H direct arylations of naphthalene and other unbiased arenes with diaryliodonium salts. Chem. Commun. 2012, 48, 9089–9091. 10.1039/c2cc34592g. [DOI] [PubMed] [Google Scholar]
  385. Ackermann L.; Dell’Acqua M.; Fenner S.; Vicente R.; Sandmann R. Metal-Free Direct Arylations of Indoles and Pyrroles with Diaryliodonium Salts. Org. Lett. 2011, 13, 2358–2360. 10.1021/ol200601e. [DOI] [PubMed] [Google Scholar]
  386. Zhu Y.; Bauer M.; Ploog J.; Ackermann L. Late-Stage Diversification of Peptides by Metal-Free C-H Arylation. Chem.—Eur. J. 2014, 20, 13099–13102. 10.1002/chem.201404603. [DOI] [PubMed] [Google Scholar]
  387. Wen J.; Zhang R.-Y.; Chen S.-Y.; Zhang J.; Yu X.-Q. Direct Arylation of Arene and N-Heteroarenes with Diaryliodonium Salts without the Use of Transition Metal Catalyst. J. Org. Chem. 2012, 77, 766–771. 10.1021/jo202150t. [DOI] [PubMed] [Google Scholar]
  388. Wang D.; Ge B.; Li L.; Shan J.; Ding Y. Transition Metal-Free Direct C-H Functionalization of Quinones and Naphthoquinones with Diaryliodonium Salts: Synthesis of Aryl Naphthoquinones as β-Secretase Inhibitors. J. Org. Chem. 2014, 79, 8607–8613. 10.1021/jo501467v. [DOI] [PubMed] [Google Scholar]
  389. Dohi T.; Ueda S.; Hirai A.; Kojima Y.; Morimoto K.; Kita Y. Selective Aryl Radical Transfers into N-Hetreoaromatics from Diaryliodonium Salts with Trimethoxybenzene Auxiliary. Heterocycles 2017, 95, 1272–1284. 10.3987/COM-16-S(S)90. [DOI] [Google Scholar]
  390. Kumar R.; Ravi C.; Rawat D.; Adimurthy S. Base-Promoted Transition-Metal-Free Arylation of Imidazo-Fused Heterocycles with Diaryliodonium Salts. Eur. J. Org. Chem. 2018, 2018, 1665–1673. 10.1002/ejoc.201800286. [DOI] [Google Scholar]
  391. Yin K.; Zhang R. Transition-Metal-Free Direct C-H Arylation of Quinoxalin-2(1H)-ones with Diaryliodonium Salts at Room Temperature. Org. Lett. 2017, 19, 1530–1533. 10.1021/acs.orglett.7b00310. [DOI] [PubMed] [Google Scholar]
  392. Zhou X.; Wu Q.; Yu Y.; Yu C.; Hao E.; Wei Y.; Mu X.; Jiao L. Metal-Free Direct α-Selective Arylation of Boron Dipyrromethenes via Base-Mediated C-H Functionalization. Org. Lett. 2016, 18, 736–739. 10.1021/acs.orglett.5b03706. [DOI] [PubMed] [Google Scholar]
  393. Caramenti P.; Nandi R. K.; Waser J. Metal-Free Oxidative Cross Coupling of Indoles with Electron-Rich (Hetero)arenes. Chem.—Eur. J. 2018, 24, 10049–10053. 10.1002/chem.201802142. [DOI] [PubMed] [Google Scholar]
  394. Chen H.; Han J.; Wang L. Intramolecular Aryl Migration of Diaryliodonium Salts: Access to ortho-Iodo Diaryl Ethers. Angew. Chem., Int. Ed. 2018, 57, 12313–12317. 10.1002/anie.201806405. [DOI] [PubMed] [Google Scholar]
  395. Wang Y.; Zhang Y.; Wang L.; Han J. Late-stage Modification of Coumarins via Aryliodonium Intramolecular Aryl Migration. Asian J. Org. Chem. 2022, 11, e202100669 10.1002/ajoc.202100669. [DOI] [Google Scholar]
  396. Chen H.; Wang L.; Han J. Deacetylative Aryl Migration of Diaryliodonium Salts with C(sp2)-N Bond Formation toward ortho-Iodo N-Aryl Sulfonamides. Org. Lett. 2020, 22, 3581–3585. 10.1021/acs.orglett.0c01024. [DOI] [PubMed] [Google Scholar]
  397. Liu X.; Wang L.; Wang H.-Y.; Han J. Diversification of Complex Diaryl Ethers via Diaryliodonium Intramolecular Aryl Rearrangement. J. Org. Chem. 2023, 88, 13089–13101. 10.1021/acs.joc.3c01293. [DOI] [PubMed] [Google Scholar]
  398. 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]
  399. 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]
  400. Lee J. B.; Kim G. H.; Jeon J. H.; Jeong S. Y.; Lee S.; Park J.; Lee D.; Kwon Y.; Seo J. K.; Chun J.-H.; Kang S. J.; Choe W.; Rohde J.-U.; Hong S. Y. Rapid access to polycyclic N-heteroarenes from unactivated, simple azines via a base-promoted Minisci-type annulation. Nat. Commun. 2022, 13, 2421. 10.1038/s41467-022-30086-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  401. 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]
  402. 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]
  403. Li Y.; Studer A. Transition-Metal-Free Trifluoromethylaminoxylation of Alkenes. Angew. Chem., Int. Ed. 2012, 51, 8221–8224. 10.1002/anie.201202623. [DOI] [PubMed] [Google Scholar]
  404. Zhang B.; Studer A. Stereoselective Radical Azidooxygenation of Alkenes. Org. Lett. 2013, 15, 4548–4551. 10.1021/ol402106x. [DOI] [PubMed] [Google Scholar]
  405. Hartmann M.; Li Y.; Mück-Lichtenfeld C.; Studer A. Generation of Aryl Radicals through Reduction of Hypervalent Iodine(III) Compounds with TEMPONa: Radical Alkene Oxyarylation. Chem.—Eur. J. 2016, 22, 3485–3490. 10.1002/chem.201504852. [DOI] [PubMed] [Google Scholar]
  406. Wu C.; Zhao C.; Zhou J.; Hu H.-S.; Li J.; Wu P.; Chen C. Wet carbonate-promoted radical arylation of vinyl pinacolboronates with diaryliodonium salts yields substituted olefins. Commun. Chem. 2020, 3, 92. 10.1038/s42004-020-00343-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  407. Satkar Y.; Wrobel K.; Trujillo-González D. E.; Ortiz-Alvarado R.; Jiménez-Halla J. O. C.; Solorio-Alvarado C. R. The Diaryliodonium(III) Salts Reaction With Free-Radicals Enables One-Pot Double Arylation of Naphthols. Front. Chem. 2020, 8, 563470. 10.3389/fchem.2020.563470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  408. Nie S.-M.; Zhang X.; Wang Z.-X.; Su G.-F.; Pan C.-X.; Mo D.-L. Synthesis of 3-(2-Hydroxyaryl)indazole Derivatives through Radical O-Arylation and [3,3]-Rearrangement from N-Hydroxyindazoles and Diaryliodonium Salts. Adv. Synth. Catal. 2022, 364, 3782–3788. 10.1002/adsc.202200569. [DOI] [Google Scholar]
  409. Jiang J.; Li J. Mechanically Induced N-arylation of Amines with Diaryliodonium Salts. ChemistrySelct 2020, 5, 542–548. 10.1002/slct.201904188. [DOI] [Google Scholar]
  410. Yoshimura A.; Saito A.; Zhdankin V. V. Iodonium Salts as Benzyne Precursors. Chem.—Eur. J. 2018, 24, 15156–15166. 10.1002/chem.201802111. [DOI] [PubMed] [Google Scholar]
  411. Shi J.; Li L.; Li Y. o-Silylaryl Triflates: A Journey of Kobayashi Aryne Precursors. Chem. Rev. 2021, 121, 3892–4044. 10.1021/acs.chemrev.0c01011. [DOI] [PubMed] [Google Scholar]
  412. Miyabe H. Regiocontrol by Halogen Substituent on Arynes: Generation of 3-Haloarynes and Their Synthetic Reactions. Synthesis 2022, 54, 5003–5016. 10.1055/a-1814-9853. [DOI] [Google Scholar]
  413. Beringer F. M.; Huang S. J. Rearrangement and Cleavage of 2-Aryliodoniobenzoates. Trapping Agents for Benzyne. J. Org. Chem. 1964, 29, 445–448. 10.1021/jo01025a049. [DOI] [Google Scholar]
  414. Akiyama T.; Imasaki Y.; Kawmisi M. Arylation of Tetrazolide with Diaryliodonium halides: Evidence for Intermediacy of Benzyne. Chem. Lett. 1974, 3, 229–230. 10.1246/cl.1974.229. [DOI] [Google Scholar]
  415. Kitamura T.; Yamane M. (Phenyl)[o-(Trimethylsilyl)phenyl]iodonium Triflate. A New and Efficient Benzyne. J. Chem. Soc. Chem. Commun. 1995, 983–984. 10.1039/c39950000983. [DOI] [Google Scholar]
  416. Kitamura T.; Yamane M.; Inoue K.; Todaka M.; Fukatsu N.; Meng Z.; Fujiwara Y. A New and Efficient Hypervalent Iodine-Benzyne Precursor, (Phenyl)[o-(trimethylsilyl)phenyl]iodonium Triflate: Generation, Trapping Reaction, and Nature of Benzyne. J. Am. Chem. Soc. 1999, 121, 11674–11679. 10.1021/ja992324x. [DOI] [Google Scholar]
  417. Kitamura T.; Gondo K.; Oyamada J. Hypervalent Iodine/Triflate Hybrid Benzdiyne Equivalents: Access to Controlled Synthesis of Polycyclic Aromatic Compounds. J. Am. Chem. Soc. 2017, 139, 8416–8419. 10.1021/jacs.7b04483. [DOI] [PubMed] [Google Scholar]
  418. Yoshimura A.; Fuchs J. M.; Middleton K. R.; Maskaev A. V.; Rohde G. T.; Saito A.; Postnikov P. S.; Yusubov M. S.; Nemykin V. N.; Zhdankin V. V. Pseudocyclic Arylbenziodoxaboroles: Efficient Benzyne Precursors Triggered by Water at Room Temperature. Chem.—Eur. J. 2017, 23, 16738–16742. 10.1002/chem.201704393. [DOI] [PubMed] [Google Scholar]
  419. Sviridova E.; Li M.; Barras A.; Addad A.; Yusubov M. S.; Zhdankin V. V.; Yoshimura A.; Szunerits S.; Postnikov P. S.; Boukherroub R. Aryne cycloaddition reaction as a facile and mild modification method for design of electrode materials for high-performance symmetric supercapacitor. Electrochim. Acta 2021, 369, 137667. 10.1016/j.electacta.2020.137667. [DOI] [Google Scholar]
  420. Gillespie D. G.; Walker B. J.; Stevens D. Phosphides and Arsenides as Metal-Halogen Exchange Reagents. Part 2. Reactions with Aromatic Dihalides. J. Chem. Soc. Perkin Trans. I 1983, 1697–1703. 10.1039/P19830001697. [DOI] [Google Scholar]
  421. Yoshida S.; Hosoya T. Synthesis of Diverse Aromatic Oxophosphorus Compounds by the Michaelis Arbuzov-type Reaction of Arynes. Chem. Lett. 2013, 42, 583–585. 10.1246/cl.130116. [DOI] [Google Scholar]
  422. Yoshimura A.; Ngo K.; Mironova I. A.; Gardner Z. S.; Rohde G. T.; Ogura N.; Ueki A.; Yusubov M. S.; Saito A.; Zhdankin V. V. Pseudocyclic Arylbenziodoxaboroles as Water-Triggered Aryne Precursors in Reactions with Organic Sulfides. Org. Lett. 2024, 26, 1891–1895. 10.1021/acs.orglett.4c00197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  423. Stuart D. R. Unsymmetrical Diaryliodonium Salts as Aryne Synthons: Renaissance of a C-H Deprotonative Approach to Arynes. Synlett 2017, 28, 275–279. 10.1055/s-0036-1588683. [DOI] [Google Scholar]
  424. Chen H.; Han J.; Wang L. Diels-Alder cycloadditions of N-arylpyrroles via aryne intermediates using diaryliodonium salts. Beilstein J. Org. Chem. 2018, 14, 354–363. 10.3762/bjoc.14.23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  425. Zhang Z.; Wu X.; Han J.; Wu W.; Wang L. Direct arylation of tertiary amines via aryne intermediates using diaryliodonium salts. Tetrahedron Lett. 2018, 59, 1737–1741. 10.1016/j.tetlet.2018.03.067. [DOI] [Google Scholar]
  426. Nilova A.; Sibbald P. A.; Valente E. J.; González-Montiel G. A.; Richardson H. C.; Brown K. S.; Cheong P. H.-Y.; Stuart D. R. Regioselective Synthesis of 1,2,3,4-Tetrasubstituted Arenes by Vicinal Functionalization of Arynes Derived from Aryl(Mes)iodonium Salts. Chem.—Eur. J. 2021, 27, 7168–7175. 10.1002/chem.202100201. [DOI] [PubMed] [Google Scholar]
  427. Nilova A.; Metze B.; Stuart D. R. Aryl(TMP)iodonium Tosylate Reagents as a Strategic Entry Point to Diverse Aryl Intermediates: Selective Access to Arynes. Org. Lett. 2021, 23, 4813–4817. 10.1021/acs.orglett.1c01534. [DOI] [PubMed] [Google Scholar]
  428. Metze B.; Bhattacharjee A.; McCormick T. M.; Stuart D. R. Parameterization of Arynophiles: Experimental Investigations Towards a Quantitative Understanding of Aryne Trapping Reactions. Synthesis 2022, 54, 4989–4996. 10.1055/a-1845-3066. [DOI] [Google Scholar]
  429. Metze B. E.; Roberts R. A.; Nilova A.; Stuart D. R. An efficient and chemoselective method to generate arynes. Chem. Sci. 2023, 14, 13885–13892. 10.1039/D3SC05429B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  430. Li X.; Chen H.; Liu X.; Wang L.; Han J. Meta-OTf-Substituted Diaryliodonium Salts Enabled Diels-Alder Cycloadditions of Furans via Aryne Intermediates. ChemistrySelect 2023, 8, e202301890 10.1002/slct.202301890. [DOI] [Google Scholar]
  431. Takenaga N.; Hayashi T.; Ueda S.; Satake; Takenaga N.; Ueda S.; Hayashi T.; Dohi T.; Kitagaki S. Vicinal Functionalization of Uracil Heterocycles with Base Activation of Iodonium(III) Salts. Heterocycles 2019, 99, 865–874. 10.3987/COM-18-S(F)93. [DOI] [Google Scholar]; Takenaga N.; Hayashi T.; Ueda S.; Satake H.; Yamada Y.; Kodama T.; Dohi T. Synthesis of Uracil-Iodonium(III) Salts for Practical Utilization as Nucleobase Synthetic Modules. Molecules 2019, 24, 3034. 10.3390/molecules24173034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  432. Takenaga N.; Ueda S.; Hayashi T.; Dohi T.; Kitagaki S. Vicinal Functionalization of Uracil Heterocycles with Base Activation of Iodonium(III) Salts. Heterocycles 2019, 99, 865–874. 10.3987/COM-18-S(F)93. [DOI] [Google Scholar]
  433. Yuan H.; Yin W.; Hu J.; Li Y. 3-sulfonyloxyaryl(mesityl)iodonium triflates as 1,2-benzdiyne precursors with activation via ortho-deprotonative elimination strategy. Nat. Commun. 2023, 14, 1841. 10.1038/s41467-023-37196-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  434. 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]
  435. Li S.; Lv H.; Yu Y.; Ye X.; Li B.; Yang S.; Mo Y.; Kong X. Domino N-/C- or N-/N-/C-arylation of imidazoles to yield polyaryl imidazolium salts via atom-economical use of diaryliodonium salts. Chem. Commun. 2019, 55, 11267–11270. 10.1039/C9CC05237B. [DOI] [PubMed] [Google Scholar]
  436. 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]
  437. 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]
  438. Mathew B. P.; Yang H. J.; Kim J.; Lee J. B.; Kim Y.-T.; Lee S.; Lee C. Y.; Choe W.; Myung K.; Park J.-U.; Hong S. Y. An Annulative Synthetic Strategy for Building Triphenylene Frameworks by Multiple C-H Bond Activations. Angew. Chem., Int. Ed. 2017, 56, 5007–5011. 10.1002/anie.201700405. [DOI] [PubMed] [Google Scholar]
  439. 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]
  440. 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]
  441. Wang Y.; Pan W.; Zhang Y.; Wang L.; Han J. Truce-Smiles Rearrangement of Diaryliodonium Salts in Ionic Liquids. Angew. Chem., Int. Ed. 2023, 62, e202304897 10.1002/anie.202304897. [DOI] [PubMed] [Google Scholar]
  442. Linde E.; Bulfield D.; Kervefors G.; Purkait N.; Olofsson B. Diarylation of N- and O-nucleophiles through a metal-free cascade reaction. Chem. 2022, 8, 850–865. 10.1016/j.chempr.2022.01.009. [DOI] [Google Scholar]
  443. Wu T.; Greaney M. F. Chemoselective cascade arylation. Chem. 2022, 8, 609–611. 10.1016/j.chempr.2022.02.020. [DOI] [Google Scholar]
  444. Mondal S.; Di Tommaso E. M.; Olofsson B. Transition-Metal-Free Difunctionalization of Sulfur Nucleophiles. Angew. Chem., Int. Ed. 2023, 62, e202216296 10.1002/anie.202216296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  445. Linde E.; Olofsson B. Synthesis of Complex Diarylamines through a Ring-Opening Difunctionalization Strategy. Angew. Chem., Int. Ed. 2023, 62, e202310921 10.1002/anie.202310921. [DOI] [PubMed] [Google Scholar]
  446. Liu M.; Jiang H.; Tang J.; Ye Z.; Zhang F.; Wu Y. meta-Selective O-Arylation of Cyclic Diaryliodonium Salts with Phenols via Aryne Intermediates. Org. Lett. 2023, 25, 2777–2781. 10.1021/acs.orglett.3c00600. [DOI] [PubMed] [Google Scholar]
  447. Jiang H.; Liu M.; Ye Z.; Song Z.; Wu Y.; Zhang F. Metal-free meta-halogenation of cyclic diaryliodonium salts via aryne intermediates. Org. Chem. Front. 2023, 10, 3856–3860. 10.1039/D3QO00570D. [DOI] [Google Scholar]
  448. Rimi R.; Soni S.; Uttam B.; China H.; Dohi T.; Zhdankin V. V.; Kumar R. Recyclable Hypervalent Iodine Reagents in Modern Organic Synthesis. Synthesis 2022, 54, 2731–2748. 10.1055/s-0041-1737909. [DOI] [Google Scholar]
  449. Roy S.; Panja S.; Sahoo S. R.; Chatterjee S.; Maiti D. Enroute sustainability: metal free C-H bond functionalization. Chem. Soc. Rev. 2023, 52, 2391–2479. 10.1039/D0CS01466D. [DOI] [PubMed] [Google Scholar]
  450. Bellotti P.; Huang H.-M.; Faber T.; Glorius F. Photocatalytic Late-Stage C-H Functionalization. Chem. Rev. 2023, 123, 4237–4352. 10.1021/acs.chemrev.2c00478. [DOI] [PubMed] [Google Scholar]
  451. Morimoto K.; Nakamura A.; Dohi T.; Kita Y. Metal-Free Oxidative Cross-Coupling Reaction of Thiophene Iodonium Salts with Pyrroles. Eur. J. Org. Chem. 2016, 2016, 4294–4297. 10.1002/ejoc.201600791. [DOI] [PubMed] [Google Scholar]
  452. Morimoto K.; Ohnishi Y.; Koseki D.; Nakamura A.; Dohi T.; Kita Y. Stabilized pyrrolyl iodonium salts and metal-free oxidative cross-coupling. Org. Biomol. Chem. 2016, 14, 8947–8951. 10.1039/C6OB01764A. [DOI] [PubMed] [Google Scholar]
  453. Morimoto K.; Ohnishi Y.; Nakamura A.; Sakamoto K.; Dohi T.; Kita Y. N1-Selective Oxidative C-N Coupling of Azoles with Pyrroles Using a Hypervalent Iodine Reagent. Asian J. Org. Chem. 2014, 3, 382–386. 10.1002/ajoc.201400027. [DOI] [Google Scholar]
  454. Dohi T.; Yamaoka N.; Nakamura S.; Sumida K.; Morimoto K.; Kita Y. Efficient Synthesis of a Regioregular Oligothiophene Photovoltaic Dye Molecule, MK-2, and Related Compounds: A Cooperative Hypervalent Iodine and Metal-Catalyzed Synthetic Route. Chem.—Eur. J. 2013, 19, 2067–2075. 10.1002/chem.201203503. [DOI] [PubMed] [Google Scholar]
  455. Hong F.; Chen Y.; Lu B.; Cheng J. One-Pot Assembly of Fused Heterocycles via Oxidative Palladium-Catalyzed Cyclization of Arylols and Iodoarenes. Adv. Synth. Catal. 2016, 358, 353–357. 10.1002/adsc.201500925. [DOI] [Google Scholar]
  456. Cheng J.; Huang L.; Xiao H.; Jiang S. One-Pot Transformation of Hypervalent Iodines into Diversified Phenoxazine Analogues as Promising Photocatalysts. J. Org. Chem. 2021, 86, 15792–15799. 10.1021/acs.joc.1c01883. [DOI] [PubMed] [Google Scholar]
  457. Lu N.; Huang L.; Xie L.; Cheng J. Transition-Metal-Free Selective Iodoarylation of Pyrazoles via Heterocyclic Aryliodonium Ylides. Eur. J. Org. Chem. 2018, 2018, 3437–3443. 10.1002/ejoc.201800416. [DOI] [Google Scholar]
  458. 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]
  459. Karandikar S. S.; Metze B. E.; Roberts R. A.; Stuart D. R. Oxidative Cycloaddition Reactions of Arylboron Reagents via a One-pot Formal Dehydroboration Sequence. Org. Lett. 2023, 25, 6374–6379. 10.1021/acs.orglett.3c02379. [DOI] [PubMed] [Google Scholar]
  460. 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. 2022, 134, e202201240 10.1002/ange.202201240. [DOI] [PubMed] [Google Scholar]
  461. Zhou J.; Bao Z.; Wu P.; Chen C. The Preparation and Application of Diaryliodonium Salts Derived from Gemfibrozil and Gemfibrozil Methyl Ester. Synthesis 2022, 54, 1388–1394. 10.1055/a-1679-7753. [DOI] [Google Scholar]
  462. Zhou J.; Bao Z.; Wu P.; Chen C. Preparation and Synthetic Application of Naproxen-Containing Diaryliodonium Salts. Molecules 2021, 26, 3240. 10.3390/molecules26113240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  463. Sun C.-L.; Shi Z.-J. Transition-Metal-Free Coupling Reactions. Chem. Rev. 2014, 114, 9219–9280. 10.1021/cr400274j. [DOI] [PubMed] [Google Scholar]
  464. Lekkala R.; Lekkala R.; Moku B.; Rakesh K. P.; Qin H.-L. Recent Developments in Radical-Mediated Transformations of Organohalides. Eur. J. Org. Chem. 2019, 2019, 2769–2806. 10.1002/ejoc.201900098. [DOI] [Google Scholar]
  465. Madasu J.; Shinde S.; Das R.; Patel S.; Shard A. Potassium tert-butoxide mediated C-C, C-N, C-O and C-S bond forming reactions. Org. Biomol. Chem. 2020, 18, 8346–8365. 10.1039/D0OB01382J. [DOI] [PubMed] [Google Scholar]
  466. Koike T.; Akita M. Modern Synthetic Strategies for One-Electron Injection. Trends Chem. 2021, 3, 416–427. 10.1016/j.trechm.2021.02.006. [DOI] [Google Scholar]
  467. Koefoed L.; Vase K. H.; Stenlid J. H.; Brinck T.; Yoshimura Y.; Lund H.; Pedersen S. U.; Daasbjerg K. On the Kinetic and Thermodynamic Properties of Aryl Radicals Using Electrochemical and Theoretical Approaches. ChemElectroChem. 2017, 4, 3212–3221. 10.1002/celc.201700772. [DOI] [Google Scholar]
  468. Romanczyk P. P.; Kurek S. S. Reliable reduction potentials of diaryliodonium cations and aryl radicals in acetonitrile from high-level ab initio computations. Electrochim. Acta 2020, 351, 136404. 10.1016/j.electacta.2020.136404. [DOI] [Google Scholar]
  469. Nocera G.; Murphy J. A. Ground State Cross-Coupling of Haloarenes with Arenes Initiated by Organic Electron Donors, Formed in situ: An Overview. Synthesis 2020, 52, 327–336. 10.1055/s-0039-1690614. [DOI] [Google Scholar]
  470. Syroeshkin M. A.; Kuriakose F.; Saverina E. A.; Timofeeva V. A.; Egorov M. P.; Alabugin I. V. Upconversion of Reductants. Angew. Chem., Int. Ed. 2019, 58, 5532–5550. 10.1002/anie.201807247. [DOI] [PubMed] [Google Scholar]
  471. Wang C.-S.; Dixneuf P. H.; Soulé J.-F. Photoredox Catalysis for Building C-C Bonds from C(sp2)-H Bonds. Chem. Rev. 2018, 118, 7532–7585. 10.1021/acs.chemrev.8b00077. [DOI] [PubMed] [Google Scholar]
  472. Weick F.; Hagmeyer N.; Giraud M.; Dietzek-Ivanšić B.; Wagenknecht H.-A. Reductive Activation of Aryl Chlorides by Tuning the Radical Cation Properties of N-Phenylphenothiazines as Organophotoredox Catalysts. Chem.—Eur. J. 2023, 29, e202302347 10.1002/chem.202302347. [DOI] [PubMed] [Google Scholar]
  473. Pause L.; Robert M.; Savéant J.-M. Can Single-Electron Transfer Break an Aromatic Carbon-Heteroatom Bond in One Step? A Novel Example of Transition between Stepwise and Concerted Mechanisms in the Reduction of Aromatic Iodides. J. Am. Chem. Soc. 1999, 121, 7158–7159. 10.1021/ja991365q. [DOI] [Google Scholar]
  474. Xu Z.; Gao L.; Wang L.; Gong M.; Wang W.; Yuan R. Visible Light Photoredox Catalyzed Biaryl Synthesis Using Nitrogen Heterocycles as Promoter. ACS Catal. 2015, 5, 45–50. 10.1021/cs5011037. [DOI] [Google Scholar]
  475. Wang L.; Byun J.; Li R.; Huang W.; Zhang K. A. I. Molecular Design of Donor-Acceptor-Type Organic Photocatalysts for Metal-free Aromatic C-C Bond Formations under Visible Light. Adv. Synth. Catal. 2018, 360, 4312–4318. 10.1002/adsc.201800950. [DOI] [Google Scholar]
  476. Barange S. H.; Bhagat P. R. A Metal/Solvent/Additive Free Compliant Route to Ullmann-Type C-N Coupling using Ionic Liquid Entangled Porphyrin Heterogeneous Photocatalyst. ChemistrySelect 2022, 7, e202201177 10.1002/slct.202201177. [DOI] [Google Scholar]
  477. Pavlovska T.; Lesný D. K.; Svobodová E.; Hoskovcová I.; Archipowa N.; Kutta R. J.; Cibulka R. Tuning Deazaflavins Towards Highly Potent Reducing Photocatalysts Guided by Mechanistic Understanding - Enhancement of the Key Step by the Internal Heavy Atom Effect. Chem.—Eur. J. 2022, 28, e202200768 10.1002/chem.202200768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  478. Weinhold T. D.; Reece N. A.; Ribeiro K.; Ocasio M. L.; Watson N.; Hanson K.; Longstreet A. R. Assessing Carbazole Derivatives as Single-Electron Photoreductants. J. Org. Chem. 2022, 87, 16928–16936. 10.1021/acs.joc.2c02312. [DOI] [PubMed] [Google Scholar]
  479. Kwon Y.; Lee J.; Noh Y.; Kim D.; Lee Y.; Yu C.; J; Roldao C.; Feng S.; Gierschner J.; Wannemacher R.; Kwon M. S. Formation and degradation of strongly reducing cyanoarene-based radical anions towards efficient radical anion-mediated photoredox catalysis. Nat. Commun. 2023, 14, 92. 10.1038/s41467-022-35774-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  480. Tintori G.; Fall A.; Assani N.; Zhao Y.; Bergé-Lefranc D.; Redon S.; Vanelle P.; Broggi J. Generation of powerful organic electron donors by water-assisted decarboxylation of benzimidazolium carboxylates. Org. Chem. Front. 2021, 8, 1197–1205. 10.1039/D0QO01488E. [DOI] [Google Scholar]
  481. Hasegawa E.; Izumiya N.; Miura T.; Ikoma T.; Iwamoto H.; Takizawa S.-y.; Murata S. Benzimidazolium Naphthoxide Betaine Is a Visible Light Promoted Organic Photoredox Catalyst. J. Org. Chem. 2018, 83, 3921–3927. 10.1021/acs.joc.8b00282. [DOI] [PubMed] [Google Scholar]
  482. Ghasimi S.; Bretschneider S. A.; Huang W.; Landfester K.; Zhang K. A. I. A Conjugated Microporous Polymer for Palladium-Free, Visible Light-Promoted Photocatalytic Stille-Type Coupling Reactions. Adv. Sci. 2017, 4, 1700101. 10.1002/advs.201700101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  483. López-Calixto C. G.; Liras M.; de la Peña O’Shea V. A.; Pérez-Ruiz R. Synchronized biphotonic process triggering C-C coupling catalytic reactions. Appl. Catal. 2018, 237, 18–23. 10.1016/j.apcatb.2018.05.062. [DOI] [Google Scholar]
  484. Xu J.; Cao J.; Wu X.; Wang H.; Yang X.; Tang X.; Toh R. W.; Zhou R.; Yeow E. K. L.; Wu J. Unveiling Extreme Photoreduction Potentials of Donor-Acceptor Cyanoarenes to Access Aryl Radicals from Aryl Chlorides. J. Am. Chem. Soc. 2021, 143, 13266–13273. 10.1021/jacs.1c05994. [DOI] [PubMed] [Google Scholar]
  485. Enders P.; Májek M.; Lam C. M.; Little R. D.; Francke R. How to Harness Electrochemical Mediators for Photocatalysis - A Systematic Approach Using the Phenanthro[9,10-d]imidazole Framework as a Test Case. ChemCatChem. 2023, 15, e202200830 10.1002/cctc.202200830. [DOI] [Google Scholar]
  486. Nocera G.; Young A.; Palumbo F.; Emery K. J.; Coulthard G.; McGuire T.; Tuttle T.; Murphy J. A. Electron Transfer Reactions: KOtBu (but not NaOtBu) Photoreduces Benzophenone under Activation by Visible Light. J. Am. Chem. Soc. 2018, 140, 9751–9757. 10.1021/jacs.8b06089. [DOI] [PubMed] [Google Scholar]
  487. Bardagi J. I.; Ghosh I.; Schmalzbauer M.; Ghosh T.; König B. Anthraquinones as Photoredox Catalysts for the Reductive Activation of Aryl Halides. Eur. J. Org. Chem. 2018, 2018, 34–40. 10.1002/ejoc.201701461. [DOI] [Google Scholar]
  488. Pezzetta C.; Folli A.; Matuszewska O.; Murphy D.; Davidson R. W. M.; Bonifazi D. peri-Xanthenoxanthene (PXX): a Versatile Organic Photocatalyst in Organic Synthesis. Adv. Synth. Catal. 2021, 363, 4740–4753. 10.1002/adsc.202100030. [DOI] [Google Scholar]
  489. Jeong D. Y.; Lee D. S.; Lee H. L.; Nah S.; Lee J. Y.; Cho E. J.; You Y. Evidence and Governing Factors of the Radical-Ion Photoredox Catalysis. ACS Catal. 2022, 12, 6047–6059. 10.1021/acscatal.2c00763. [DOI] [Google Scholar]
  490. Glaser F.; Wenger O. S. Red Light-Based Dual Photoredox Strategy Resembling the Z-Scheme of Natural Photosynthesis. JACS Au 2022, 2, 1488–1503. 10.1021/jacsau.2c00265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  491. Singh V.; Singh R.; Hazari A. S.; Adhikari D. Unexplored Facet of Pincer Ligands: Super-Reductant Behavior Applied to Transition-Metal-Free Catalysis. JACS Au 2023, 3, 1213–1220. 10.1021/jacsau.3c00077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  492. Studer A.; Curran D. P. Organocatalysis and C–H Activation Meet Radical- and Electron-Transfer Reactions. Angew. Chem., Int. Ed. 2011, 50, 5018–5022. 10.1002/anie.201101597. [DOI] [PubMed] [Google Scholar]
  493. Studer A.; Curran D. P. The electron is a catalyst. Nat. Chem. 2014, 6, 765–273. 10.1038/nchem.2031. [DOI] [PubMed] [Google Scholar]
  494. Yanagisawa S.; Ueda K.; Taniguchi T.; Itami K. Potassium t-Butoxide Alone Can Promote the Biaryl Coupling of Electron-Deficient Nitrogen Heterocycles and Haloarenes. Org. Lett. 2008, 10, 4673–4676. 10.1021/ol8019764. [DOI] [PubMed] [Google Scholar]
  495. Yanagisawa S.; Itami K. tert-Butoxide-Mediated C-H Bond Arylation of Aromatic Compounds with Haloarenes. ChemCatChem. 2011, 3, 827–829. 10.1002/cctc.201000431. [DOI] [Google Scholar]
  496. Barham J. P.; Coulthard G.; Kane R. G.; Delgado N.; John M. P.; Murphy J. A. Double Deprotonation of Pyridinols Generates Potent Organic Electron-Donor Initiators for Haloarene-Arene Coupling. Angew. Chem., Int. Ed. 2016, 55, 4492–4496. 10.1002/anie.201511847. [DOI] [PubMed] [Google Scholar]
  497. Zhou S.; Anderson G. M.; Mondal B.; Doni E.; Ironmonger V.; Kranz M.; Tuttle T.; Murphy J. A. Organic super-electron-donors: initiators in transition metal-free haloarene-arene coupling. Chem. Sci. 2014, 5, 476–482. 10.1039/C3SC52315B. [DOI] [Google Scholar]
  498. Anderson G. M.; Cameron I.; Murphy J. A.; Tuttle T. Predicting the reducing power of organic super electron donors. RSC Adv. 2016, 6, 11335–11343. 10.1039/C5RA26483A. [DOI] [Google Scholar]
  499. Zhou S.; Doni E.; Anderson G. M.; Kane R. G.; MacDougall S. W.; Ironmonger V. M.; Tuttle T.; Murphy J. A. Identifying the Roles of Amino Acids, Alcohols and 1,2-Diamines as Mediators in Coupling of Haloarenes to Arenes. J. Am. Chem. Soc. 2014, 136, 17818–17826. 10.1021/ja5101036. [DOI] [PubMed] [Google Scholar]
  500. Smith A. J.; Poole D. L.; Murphy J. A. The role of organic electron donors in the initiation of BHAS base-induced coupling reactions between haloarenes and arenes. Sci. China Chem. 2019, 62, 1425–1438. 10.1007/s11426-019-9611-x. [DOI] [Google Scholar]
  501. Patil M. Mechanistic Insights into the Initiation Step of the Base Promoted Direct C-H Arylation of Benzene in the Presence of Additive. J. Org. Chem. 2016, 81, 632–639. 10.1021/acs.joc.5b02438. [DOI] [PubMed] [Google Scholar]
  502. Ahmed J.; Mandal S. K. Phenalenyl Radical: Smallest Polycyclic Odd Alternant Hydrocarbon Present in the Graphene Sheet. Chem. Rev. 2022, 122, 11369–11431. 10.1021/acs.chemrev.1c00963. [DOI] [PubMed] [Google Scholar]
  503. Ahmed J.; Datta P.; Das A.; Jomy S.; Mandal S. K. Switching between mono and doubly reduced odd alternant hydrocarbon: designing a redox catalyst. Chem. Sci. 2021, 12, 3039–3049. 10.1039/D0SC05972B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  504. Raviola C.; Protti S. Leaving Groups in Metal-Free Arylations: Make Your Choice!. Eur. J. Org. Chem. 2020, 2020, 5292–5304. 10.1002/ejoc.202000143. [DOI] [Google Scholar]
  505. Liu W.; Cao H.; Zhang H.; Zhang H.; Chung K. H.; He C.; Wang H.; Kwong F. Y.; Lei A. Organocatalysis in Cross-Coupling: DMEDA-Catalyzed Direct C-H Arylation of Unactivated Benzene. J. Am. Chem. Soc. 2010, 132, 16737–16740. 10.1021/ja103050x. [DOI] [PubMed] [Google Scholar]
  506. Zhang L.; Yang H.; Jiao L. Revisiting the Radical Initiation Mechanism of the Diamine-Promoted Transition-Metal-Free Cross-Coupling Reaction. J. Am. Chem. Soc. 2016, 138, 7151–7160. 10.1021/jacs.6b03442. [DOI] [PubMed] [Google Scholar]
  507. Patil M. Mechanism of the t-BuOM (M = K, Na, Li)/DMEDA-Mediated Direct C-H Arylation of Benzene: A Computational Study. Synthesis 2020, 52, 2883–2891. 10.1055/s-0040-1707882. [DOI] [Google Scholar]
  508. Shirakawa E.; Itoh K.-i.; Higashino T.; Hayashi T. tert-Butoxide-Mediated Arylation of Benzene with Aryl Halides in the Presence of a Catalytic 1,10-Phenanthroline Derivative. J. Am. Chem. Soc. 2010, 132, 15537–15539. 10.1021/ja1080822. [DOI] [PubMed] [Google Scholar]
  509. Sun C.-L.; Li H.; Yu D.-G.; Yu M.; Zhou X.; Lu X.-Y.; Huang K.; Zheng S.-F.; Li B.-J.; Shi Z.-J. An efficient organocatalytic method for constructing biaryls through aromatic C-H activation. Nat. Chem. 2010, 2, 1044–1049. 10.1038/nchem.862. [DOI] [PubMed] [Google Scholar]
  510. Yang J.; Ren Y.; Wang J.; Li T.; Xiao T.; Jiang Y. Phenanthroline-tBuONa Promoted Intramolecular C-H Arylation of 1,5-Diaryl-1,2,3-Triazoles for Efficient Synthesis of Triazolophenanthridines. ChemistrySelect 2019, 4, 6272–6276. 10.1002/slct.201901710. [DOI] [Google Scholar]
  511. Shan X.-H.; Yang B.; Zheng H.-X.; Qu J.-P.; Kang Y.-B. Phenanthroline-tBuOK Promoted Intramolecular C-H Arylation of Indoles with ArI under Transition-Metal-Free Conditions. Org. Lett. 2018, 20, 7898–7901. 10.1021/acs.orglett.8b03449. [DOI] [PubMed] [Google Scholar]
  512. Meli A.; Ebenhoch B.; Kutonova K.; Bihlmeier A.; Feyrer A.; Deck E.; Breher F.; Nieger M.; Colsmann A.; Bräse S. Star-shaped triarylamines-One-step metal-free synthesis and optoelectronic properties. Synth. Met. 2019, 256, 116138. 10.1016/j.synthmet.2019.116138. [DOI] [Google Scholar]
  513. Nozawa-Kumada K.; Iwakawa Y.; Onuma S.; Shigeno M.; Kondo Y. NaH-mediated direct C-H arylation in the presence of 1,10-phenanthroline. Chem. Commun. 2020, 56, 7773–7776. 10.1039/D0CC00730G. [DOI] [PubMed] [Google Scholar]
  514. Dewanji A.; Murarka S.; Curran D. P.; Studer A. Phenyl Hydrazine as Initiator for Direct Arene C-H Arylation via Base Promoted Homolytic Aromatic Substitution. Org. Lett. 2013, 15, 6102–6105. 10.1021/ol402995e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  515. Zhu Y.; Wu Z.; Sun H.; Ding J. Photo-Induced, Phenylhydrazine-Promoted Transition-Metal-Free Dehalogenation of Aryl Fluorides, Chlorides, Bromides, and Iodides. Molecules 2023, 28, 6915. 10.3390/molecules28196915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  516. Tanimoro K.; Ueno M.; Takeda K.; Kirihata M.; Tanimori S. Proline Catalyzes Direct C-H Arylations of Unactivated Arenes. J. Org. Chem. 2012, 77, 7844–7849. 10.1021/jo3008594. [DOI] [PubMed] [Google Scholar]
  517. Liu W.; Tian F.; Wang X.; Yu H.; Bi Y. Simple alcohols promoted direct C-H arylation of unactivated arenes with aryl halides. Chem. Commun. 2013, 49, 2983–2985. 10.1039/c3cc40695d. [DOI] [PubMed] [Google Scholar]
  518. Zhu Y.; Wu Z.; Sun H.; Ding J. Visible-light-induced, base-promoted transition-metal-free hydrogenation of aryl halides with n-butanol. New J. Chem. 2023, 47, 15852–15855. 10.1039/D3NJ03228K. [DOI] [Google Scholar]
  519. Zhao H.; Xu X.; Wu W.; Zhang W.; Zhang Y. Urea-based organocatalyst catalyzed direct C-H bond arylations of unactivated arenes. Catal. Commun. 2018, 111, 95–99. 10.1016/j.catcom.2018.04.008. [DOI] [Google Scholar]
  520. Yang H.; Zhang L.; Jiao L. N-Methylanilines as Simple and Efficient Promoters for Radical-Type Cross-Coupling Reactions of Aryl Iodides. Chem.—Eur. J. 2017, 23, 65–69. 10.1002/chem.201604602. [DOI] [PubMed] [Google Scholar]
  521. Yang H.; Chu D.-Z.; Jiao L. Aromatization modulates the activity of small organic molecules as promoters for carbon- halogen bond activation. Chem. Sci. 2018, 9, 1534–1539. 10.1039/C7SC04450J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  522. Zheng X.; Wu X.-N.; Chen J.-Y.; Luo H.-B.; Wu D.; Wu Y. Transition-Metal-Free C-H Arylation of Unactivated Arenes with 8-Hydroxyquinoline as a Promoter. Synthesis 2018, 50, 1721–1727. 10.1055/s-0036-1591874. [DOI] [Google Scholar]
  523. Wu Y.; Choy P. Y.; Kwong F. Y. Direct intermolecular C-H arylation of unactivated arenes with aryl bromides catalysed by 2-pyridyl carbinol. Org. Biomol. Chem. 2014, 12, 6820–6823. 10.1039/C4OB01211A. [DOI] [PubMed] [Google Scholar]
  524. Yadav L.; Tiwari M. K.; Shyamlal B. R. K.; Chaudhary S. Organocatalyst in Direct C(sp2)-H Arylation of Unactivated Arenes: [1-(2-Hydroxyethyl)-piperazine]-Catalyzed Inter-/Intra-molecular C-H Bond Activation. J. Org. Chem. 2020, 85, 8121–8141. 10.1021/acs.joc.0c01019. [DOI] [PubMed] [Google Scholar]
  525. Liu Z.; Wang P.; Chen Y.; Yan Z.; Chen S.; Chen W.; Mu T. Small organic molecules with tailored structures: initiators in the transition-metal-free C-H arylation of unactivated arenes. RSC Adv. 2020, 10, 14500–14509. 10.1039/D0RA01845G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  526. Song Q.; Zhang D.; Zhu Q.; Xu Y. p-Toluenesulfonohydrazide as Highly Efficient Initiator for Direct C-H Arylation of Unactivated Arenes. Org. Lett. 2014, 16, 5272–5274. 10.1021/ol502370r. [DOI] [PubMed] [Google Scholar]
  527. Qiu Y.; Liu Y.; Yang K.; Hong W.; Li Z.; Wang Z.; Yao Z.; Jiang S. New Ligands That Promote Cross-Coupling Reactions between Aryl Halides and Unactivated Arenes. Org. Lett. 2011, 13, 3556–3559. 10.1021/ol2009208. [DOI] [PubMed] [Google Scholar]
  528. Cumine F.; Zhou S.; Tuttle T.; Murphy J. A. A study of diketopiperazines as electron-donor initiators in transition metal-free haloarene-arene coupling. Org. Biomol. Chem. 2017, 15, 3324–3336. 10.1039/C7OB00036G. [DOI] [PubMed] [Google Scholar]
  529. Emery K. J.; Tuttle T.; Murphy J. A. Evidence of single electron transfer from the enolate anion of an N, N′-dialkyldiketopiperazine additive in BHAS coupling reactions. Org. Biomol. Chem. 2017, 15, 8810–8819. 10.1039/C7OB02209C. [DOI] [PubMed] [Google Scholar]
  530. Emery K. J.; Tuttle T.; Kennedy A. R.; Murphy J. A. C-C bond-forming reactions of ground-state aryl halides under reductive activation. Tetrahedron 2016, 72, 7875–7887. 10.1016/j.tet.2016.05.083. [DOI] [Google Scholar]
  531. Doni E.; Zhou S.; Murphy J. A. Electron Transfer-Induced Coupling of Haloarenes to Styrenes and 1,1-Diphenylethenes Triggered by Diketopiperazines and Potassium tert-Butoxide. Molecules 2015, 20, 1755–1774. 10.3390/molecules20021755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  532. Paira R.; Singh B.; Hota P. K.; Ahmed J.; Sau S. C.; Johnpeter J. P.; Mandal S. K. Open-Shell Phenalenyl in Transition Metal-Free Catalytic C-H Functionalization. J. Org. Chem. 2016, 81, 2432–2441. 10.1021/acs.joc.6b00002. [DOI] [PubMed] [Google Scholar]
  533. Sigen A.; Liu X.; Li H.; He C.; Mu Y. Direct C-H Arylation of Unactivated Arenes with Aryl Halides Promoted by Bis(imino)pyridine Derivatives. Asian J. Org. Chem. 2013, 2, 857–861. 10.1002/ajoc.201300129. [DOI] [Google Scholar]
  534. Babaguchi K.; Bedi D.; Chacon-Teran M. A.; Findlater M. α-Diimine-mediated C-H functionalization of arenes for aryl-aryl cross-coupling reactions. Org. Biomol. Chem. 2024, 22, 2389–2394. 10.1039/D3OB01992F. [DOI] [PubMed] [Google Scholar]
  535. Ghosh D.; Lee J.-Y.; Liu C.-Y.; Chiang Y.-H.; Lee H. M. Direct CH Arylations of Unactivated Arenes Catalyzed by Amido-Functionalized Imidazolium Salts. Adv. Synth. Catal. 2014, 356, 406–410. 10.1002/adsc.201300877. [DOI] [Google Scholar]
  536. Chen W.-C.; Hsu Y.-C.; Shih W.-C.; Lee C.-Y.; Chuang W.-H.; Tsai Y.-F.; Chen P. P.-Y.; Ong T.-G. Metal-free arylation of benzene and pyridine promoted by amino-linked nitrogen heterocyclic carbenes. Chem. Commun. 2012, 48, 6702–6704. 10.1039/c2cc32519e. [DOI] [PubMed] [Google Scholar]
  537. Ng Y. S.; Chan C. S.; Chan K. S. Free porphyrin catalyzed direct C-H arylation of benzene with aryl halides. Tetrahedron Lett. 2012, 53, 3911–3914. 10.1016/j.tetlet.2012.05.073. [DOI] [Google Scholar]
  538. Zhao H.; Shen J.; Guo J.; Ye R.; Zeng H. A macrocyclic aromatic pyridone pentamer as a highly efficient organocatalyst for the direct arylations of unactivated arenes. Chem. Commun. 2013, 49, 2323–2325. 10.1039/c3cc00019b. [DOI] [PubMed] [Google Scholar]
  539. Zhao H.; Shen J.; Ren C.; Zeng W.; Zeng H. A Foldamer-Based Organocatalyst for Direct Arylations of Unactivated Arenes. Org. Lett. 2017, 19, 2190–2193. 10.1021/acs.orglett.7b00921. [DOI] [PubMed] [Google Scholar]
  540. Ghonchepour E.; Islami M. R.; Tikdari A. M. Methyl Red as Organocatalyst for Arylation of Unactivated Benzene Derivatives with Aryl Halides. ChemistrySelect 2018, 3, 11517–11521. 10.1002/slct.201802643. [DOI] [Google Scholar]
  541. Ghonchepour E.; Islami M. R.; Mostafavi H.; Tikdari A. M.; Sheikhshoaie I. Transition metal-free and base-mediated transformation arylation of unactivated benzene with aryl halides in presence of N,N′-bis(salicylidene) ethylenediamine as organocatalyst. Catal. Commun. 2018, 107, 87–91. 10.1016/j.catcom.2018.01.007. [DOI] [Google Scholar]
  542. Liang S.; Shi S.; Ding S.; Xiao W.; Wang H.; Wang S.; Zeng R.; Chen C.; Song W. Construction of a transition-metal-free mesoporous organic phenanthroline-based polymeric catalyst for boosting direct activation of aromatic C-H bonds. Catal. Sci. Technol. 2022, 12, 6650–6654. 10.1039/D2CY01309F. [DOI] [Google Scholar]
  543. Banik A.; Paira R.; Shaw B. K.; Vijaykumar G.; Mandal S. K. Accessing Heterobiaryls through Transition-Metal-Free C-H Functionalization. J. Org. Chem. 2018, 83, 3236–3244. 10.1021/acs.joc.8b00140. [DOI] [PubMed] [Google Scholar]
  544. Kiriyama K.; Shirakawa E. tert-Butoxide-promoted Coupling of Aryl Iodides with Arenes Using Di-tert-butyl Hyponitrite as an Initiator. Chem. Lett. 2017, 46, 1757–1759. 10.1246/cl.170814. [DOI] [Google Scholar]
  545. Yong G.-P.; She W.-L.; Zhang Y.-M.; Li Y.-Z. Direct arylation of unactivated aromatic C-H bonds catalyzed by a stable organic radical. Chem. Commun. 2011, 47, 11766–11768. 10.1039/c1cc14420k. [DOI] [PubMed] [Google Scholar]
  546. Sharma S.; Kumar M.; Sharma S.; Nayal O. S.; Kumar N.; Singh B.; Sharma U. Microwave assisted synthesis of phenanthridinones and dihydrophenanthridines by vasicine/KOtBu promoted intramolecular C-H arylation. Org. Biomol. Chem. 2016, 14, 8536–8544. 10.1039/C6OB01362G. [DOI] [PubMed] [Google Scholar]
  547. Gao Y.; Tang P.; Zhou H.; Zhang W.; Yang H.; Yan N.; Hu G.; Mei D.; Wang J.; Ma D. Graphene Oxide Catalyzed C-H Bond Activation: The Importance of Oxygen Functional Groups for Biaryl Construction. Angew. Chem., Int. Ed. 2016, 55, 3124–3128. 10.1002/anie.201510081. [DOI] [PubMed] [Google Scholar]
  548. Yang Z.; Liu Z.; Zhang H.; Yu B.; Zhao Y.; Wang H.; Ji G.; Chen Y.; Liu X.; Liu Z. N-Doped porous carbon nanotubes: synthesis and application in catalysis. Chem. Commun. 2017, 53, 929–932. 10.1039/C6CC09374D. [DOI] [PubMed] [Google Scholar]
  549. Jori P. K.; Jadhav V. H. Transition-Metal-Free Glucose-Derived Carbonaceous Catalyst Catalyzes Direct C-H Arylation of Unactivated Arenes with Aryl Halides. ChemistrySelect 2019, 4, 4848–4853. 10.1002/slct.201900013. [DOI] [Google Scholar]
  550. Tang Y.; Chen F.; Wang S.; Sun Q.; Meng X.; Xiao F.-S. Porous Organic Phenanthroline-Based Polymer as an Efficient Transition-Metal-Free Heterogeneous Catalyst for Direct Aromatic C-H Activation. Chem.—Eur. J. 2021, 27, 8684–8688. 10.1002/chem.202100288. [DOI] [PubMed] [Google Scholar]
  551. Li Z.; Shen J.; Jiang Q.; Shen S.; Zhou J.; Zeng H. Directionally Aligned Crown Ethers as Superactive Organocatalysts for Transition-metal Ion-free Arylation of Unactivated Arenes. Chem. Asian J. 2022, 17, e202200303 10.1002/asia.202200303. [DOI] [PubMed] [Google Scholar]
  552. Tiwari M. K.; Yadav L.; Chaudhary S. [2,3-Bis-(2-pyridyl)pyrazine] as an Efficient Organocatalyst for the Direct C(sp2)-H Arylation of Unactivated Arenes/Heteroarenes via C-H Bond Activation. ChemistrySelect 2020, 5, 11968–11975. 10.1002/slct.202003140. [DOI] [Google Scholar]
  553. Maiti D.; Halder A.; De Sarkar S. Base-Promoted Aerobic Oxidation/Homolytic Aromatic Substitution Cascade toward the Synthesis of Phenanthridines. Adv. Synth. Catal. 2019, 361, 4941–4948. 10.1002/adsc.201900995. [DOI] [Google Scholar]
  554. Jia F.-C.; Xu C.; Zhou Z.-W.; Cai Q.; Wu Y.-D.; Wu A.-X. Substrates as Electron-Donor Precursors: Synthesis of Naphtho-Fused Oxindoles via Benzannulation of 2-Halobenzaldehydes and Indolin-2-ones. Org. Lett. 2016, 18, 5232–5235. 10.1021/acs.orglett.6b02515. [DOI] [PubMed] [Google Scholar]
  555. Doni E.; Murphy J. A. Evolution of neutral organic super-electron donors and their applications. Chem. Commun. 2014, 50, 6073–6087. 10.1039/C3CC48969H. [DOI] [PubMed] [Google Scholar]
  556. Murphy J. A. Discovery and Development of Organic Super-Electron-Donors. J. Org. Chem. 2014, 79, 3731–3746. 10.1021/jo500071u. [DOI] [PMC free article] [PubMed] [Google Scholar]
  557. Broggi J.; Terme T.; Vanelle P. Organic Electron Donors as Powerful Single-Electron Reducing Agents in Organic Synthesis. Angew. Chem., Int. Ed. 2014, 53, 384–413. 10.1002/anie.201209060. [DOI] [PubMed] [Google Scholar]
  558. Murphy J. A.; Khan T. A.; Zhou S.-Z.; Thomson D. W.; Mahesh M. Highly Efficient Reduction of Unactivated Aryl and Alkyl Iodides by a Ground-State Neutral Organic Electron Donor. Angew. Chem., Int. Ed. 2005, 44, 1356–1360. 10.1002/anie.200462038. [DOI] [PubMed] [Google Scholar]
  559. Murphy J. A.; Zhou S.-Z.; Thomson D. W.; Schoenebeck F.; Mahesh M.; Park S. R.; Tuttle T.; Berlouis L. E. A. The Generation of Aryl Anions by Double Electron Transfer to Aryl Iodides from a Neutral Ground-State Organic Super-Electron Donor. Angew. Chem., Int. Ed. 2007, 46, 5178–5183. 10.1002/anie.200700554. [DOI] [PubMed] [Google Scholar]
  560. Murphy J. A.; Garnier J.; Park S. R.; Schoenebeck F.; Zhou S.-Z.; Turner A. T. Super-Electron Donors: Bis-pyridinylidene Formation by Base Treatment of Pyridinium Salts. Org. Lett. 2008, 10, 1227–1230. 10.1021/ol800134g. [DOI] [PubMed] [Google Scholar]
  561. Hanson S. S.; Doni E.; Traboulsee K. T.; Coulthard G.; Murphy J. A.; Dyker C. A. Pushing the Limits of Neutral Organic Electron Donors: A Tetra(iminophosphorano)-Substituted Bispyridinylidene. Angew. Chem., Int. Ed. 2015, 54, 11236–11239. 10.1002/anie.201505378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  562. Martin J. D.; Dyker C. A. Facile preparation and isolation of neutral organic electron donors based on 4-dimethylaminopyridine. Can. J. Chem. 2018, 96, 522–525. 10.1139/cjc-2017-0526. [DOI] [Google Scholar]
  563. Hanson S. S.; Richard N. A.; Dyker C. A. Powerful Bispyridinylidene Organic Reducing Agents with Iminophosphorano π-Donor Substituents. Chem.—Eur. J. 2015, 21, 8052–8055. 10.1002/chem.201500809. [DOI] [PubMed] [Google Scholar]
  564. Frenette B. L.; Arsenault N.; Walker S. L.; Decken A.; Dyker C. A. Bis(Iminophosphorano)-Substituted Pyridinium Ions and their Corresponding Bispyridinylidene Organic Electron Donors. Chem.—Eur. J. 2021, 27, 8528–8536. 10.1002/chem.202100318. [DOI] [PubMed] [Google Scholar]
  565. Rohrbach S.; Shah R. S.; Tuttle T.; Murphy J. A. Neutral Organic Super Electron Donors Made Catalytic. Angew. Chem., Int. Ed. 2019, 58, 11454–11458. 10.1002/anie.201905814. [DOI] [PubMed] [Google Scholar]
  566. Nozawa-Kumada K.; Onuma S.; Ono; Kumagai K. T.; Iwakawa Y.; Sato K.; Shigeno M.; Kondo Y. Transition-Metal-Free Intermolecular Hydrocarbonation of Styrenes Mediated by NaH/1,10-Phenanthroline. Chem.—Eur. J. 2023, 29, e202203143 10.1002/chem.202203143. [DOI] [PubMed] [Google Scholar]
  567. Chung H.; Kim J.; González-Montiel G. A.; Cheong P. H.-Y.; Lee H. G. Modular Counter-Fischer-Indole Synthesis through Radical-Enolate Coupling. Org. Lett. 2021, 23, 1096–1102. 10.1021/acs.orglett.1c00003. [DOI] [PubMed] [Google Scholar]
  568. Ohmiya H. N-Heterocyclic Carbene-Based Catalysis Enabling Cross-Coupling Reactions. ACS Catal. 2020, 10, 6862–6869. 10.1021/acscatal.0c01795. [DOI] [Google Scholar]
  569. Li Q.-Z.; Kou X.-X.; Qi T.; Li J.-L. Merging N-Heterocyclic Carbene Organocatalysis with Hydrogen Atom Transfer Strategy. ChemCatChem. 2023, 15, e202201320 10.1002/cctc.202201320. [DOI] [Google Scholar]
  570. Delfau L.; Nichilo S.; Molton F.; Broggi J.; Tomás-Mendivil E.; Martin D. Critical Assessment of the Reducing Ability of Breslow-type Derivatives and Implications for Carbene-Catalyzed Radical Reactions. Angew. Chem., Int. Ed. 2021, 60, 26783–26789. 10.1002/anie.202111988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  571. Liu W.; Vianna A.; Zhang Z.; Huang S.; Huang L.; Melaimi M.; Bertrand G.; Yan X. Mesoionic carbene-Breslow intermediates as super electron donors: Application to the metal-free arylacylation of alkenes. Chem. Catal. 2021, 1, 196–206. 10.1016/j.checat.2021.03.004. [DOI] [Google Scholar]
  572. Matsuki Y.; Ohnishi N.; Kakeno Y.; Takemoto S.; Ishii T.; Nagao K.; Ohmiya H. Aryl radical-mediated N-heterocyclic carbene catalysis. Nat. Commun. 2021, 12, 3848. 10.1038/s41467-021-24144-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  573. Biswas A.; Bhunia A.; Mandal S. K. Mechanochemical solid state single electron transfer from reduced organic hydrocarbon for catalytic aryl-halide bond activation. Chem. Sci. 2023, 14, 2606–2615. 10.1039/D2SC06119H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  574. Banik A.; Mandal S. K. Tuning Redox States of Phenalenyl-Based Molecules by Consecutive Reduction toward Transition Metal-Free Heck-Type C-C Cross-Coupling. ACS Catal. 2022, 12, 5000–5012. 10.1021/acscatal.2c00173. [DOI] [Google Scholar]
  575. Sil S.; Bhaskaran A. S.; Chakraborty S.; Singh B.; Kuniyil R.; Mandal S. K. Reduced-Phenalenyl-Based Molecule as a Super Electron Donor for Radical-Mediated C-N Coupling Catalysis at Room Temperature. J. Am. Chem. Soc. 2022, 144 (49), 22611–22621. 10.1021/jacs.2c09225. [DOI] [PubMed] [Google Scholar]
  576. Singh B.; Ahmed J.; Biswas A.; Paira R.; Mandal S. K. Reduced Phenalenyl in Catalytic Dehalogenative Deuteration and Hydrodehalogenation of Aryl Halides. J. Org. Chem. 2021, 86, 7242–7255. 10.1021/acs.joc.1c00573. [DOI] [PubMed] [Google Scholar]
  577. Zhang J.; Wu H.-H.; Zhang J. Cesium Carbonate Mediated Borylation of Aryl Iodides with Diboron in Methanol. Eur. J. Org. Chem. 2013, 2013, 6263–6266. 10.1002/ejoc.201300829. [DOI] [Google Scholar]
  578. Niu Y.-J.; Sui G.-H.; Zheng H.-X.; Shan X.-H.; Tie L.; Fu J.-L.; Qu J.-P.; Kang Y.-B. Competing Dehalogenation versus Borylation of Aryl Iodides and Bromides under Transition-Metal-Free Basic Conditions. J. Org. Chem. 2019, 84, 10805–10813. 10.1021/acs.joc.9b01350. [DOI] [PubMed] [Google Scholar]
  579. Waghamare A. B.; Raut R. K.; Patel N.; Majumdar M. Role of N, N′-Diboryl-4,4′-bipyridinylidene in the Transition-Metal-Free Borylation of Aryl Halides and Direct C-H Arylation of Unactivated Benzene. Eur. J. Inorg. Chem. 2022, 2022, e202200089 10.1002/ejic.202200089. [DOI] [Google Scholar]
  580. Zhang L.; Jiao L. Pyridine-Catalyzed Radical Borylation of Aryl Halides. J. Am. Chem. Soc. 2017, 139, 607–610. 10.1021/jacs.6b11813. [DOI] [PubMed] [Google Scholar]
  581. Zhang L.; Jiao L. Visible-Light-Induced Organocatalytic Borylation of Aryl Chlorides. J. Am. Chem. Soc. 2019, 141, 9124–9128. 10.1021/jacs.9b00917. [DOI] [PubMed] [Google Scholar]
  582. Zhang L.; Jiao L. Super electron donors derived from diboron. Chem. Sci. 2018, 9, 2711–2722. 10.1039/C8SC00008E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  583. Li M.; Liu S.; Bao H.; Li Q.; Deng Y.-H.; Sun T.-Y.; Wang L. Photoinduced metal-free borylation of aryl halides catalysed by an in situ formed donor-acceptor complex. Chem. Sci. 2022, 13, 4909–4914. 10.1039/D2SC00552B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  584. Pinet S.; Liautard V.; Debiais M.; Pucheault M. Radical Metal-Free Borylation of Aryl Iodides. Synthesis 2017, 49, 4759–4768. 10.1055/s-0036-1588431. [DOI] [Google Scholar]
  585. Franco M.; Vargas E. L.; Tortosa M.; Cid M. B. Coupling of thiols and aromatic halides promoted by diboron derived super electron donors. Chem. Commun. 2021, 57, 11653–11656. 10.1039/D1CC05294B. [DOI] [PubMed] [Google Scholar]
  586. Roman D. S.; Takahashi Y.; Charette A. B. Potassium tert-Butoxide Promoted Intramolecular Arylation via a Radical Pathway. Org. Lett. 2011, 13, 3242–3245. 10.1021/ol201160s. [DOI] [PubMed] [Google Scholar]
  587. Cuthbertson J.; Gray V. J.; Wilden J. D. Observations on transition metal free biaryl coupling: potassium tert-butoxide alone promotes the reaction without diamine or phenanthroline catalysts. Chem. Commun. 2014, 50, 2575–2578. 10.1039/C3CC49019J. [DOI] [PubMed] [Google Scholar]
  588. Liu W.; Xu L.; Bi Y. Alkoxy base-mediated transition-metal-free cross-coupling reactions of benzene with aryl halides. RSC Adv. 2014, 4, 44943–44947. 10.1039/C4RA07688E. [DOI] [Google Scholar]
  589. De S.; Ghosh S.; Bhunia S.; Sheikh J. A.; Bisai A. Intramolecular Direct Dehydrohalide Coupling Promoted by KOtBu: Total Synthesis of Amaryllidaceae Alkaloids Anhydrolycorinone and Oxoassoanine. Org. Lett. 2012, 14, 4466–4469. 10.1021/ol3019677. [DOI] [PubMed] [Google Scholar]
  590. De S.; Mishra S.; Kakde B. N.; Dey D.; Bisai A. Expeditious Approach to Pyrrolophenanthridones, Phenanthridines, and Benzo[c]phenanthridines via Organocatalytic Direct Biaryl-Coupling Promoted by Potassium tert-Butoxide. J. Org. Chem. 2013, 78, 7823–7844. 10.1021/jo400890k. [DOI] [PubMed] [Google Scholar]
  591. Barham J. P.; Coulthard G.; Emery K. J.; Doni E.; Cumine F.; Nocera G.; John M. P.; L. Berlouis E. A.; McGuire T.; Tuttle T.; Murphy J. A. KOtBu: A Privileged Reagent for Electron Transfer Reactions?. J. Am. Chem. Soc. 2016, 138, 7402–7410. 10.1021/jacs.6b03282. [DOI] [PubMed] [Google Scholar]
  592. Barham J. P.; Dalton S. E.; Allison M.; Nocera G.; Young A.; John M. P.; McGuire T.; Campos S.; Tuttle T.; Murphy J. A. Dual Roles for Potassium Hydride in Haloarene Reduction: CSNAr and Single Electron Transfer Reduction via Organic Electron Donors Formed in Benzene. J. Am. Chem. Soc. 2018, 140, 11510–11518. 10.1021/jacs.8b07632. [DOI] [PubMed] [Google Scholar]
  593. Shigeno M.; Kai Y.; Yamada T.; Hayashi K.; Nozawa-Kumada K.; Denneval C.; Kondo Y. Construction of Biaryl Scaffolds from Iodoarenes and C. Asian J. Org. Chem. 2018, 7, 2082–2086. 10.1002/ajoc.201800438. [DOI] [Google Scholar]
  594. Nozawa-Kumada K.; Nakamura K.; Kurosu S.; Iwakawa Y.; Denneval C.; Shigeno M.; Kondo Y. Tetramethylammonium Fluoride Tetrahydrate-Mediated Transition Metal-Free Coupling of Aryl Iodides with Unactivated Arenes in Air. Chem. Pharm. Bull. 2019, 67, 1042–1045. 10.1248/cpb.c19-00452. [DOI] [PubMed] [Google Scholar]
  595. Iqbal M. A.; Lu L.; Mehmood H.; Hua R. Biaryl Formation via Base-Promoted Direct Coupling Reactions of Arenes with Aryl Halides. ACS Omega 2021, 6, 15981–15987. 10.1021/acsomega.1c01736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  596. Fukuoka S. Transition Metal Catalyst-Free and Radical Initiator-Free Carbonylation of Aryl Iodides. Ind. Eng. Chem. Res. 2016, 55, 4830–4835. 10.1021/acs.iecr.6b00606. [DOI] [Google Scholar]
  597. Kawamoto T.; Sato A.; Ryu I. Photoinduced Aminocarbonylation of Aryl Iodides. Chem.—Eur. J. 2015, 21, 14764–14767. 10.1002/chem.201503164. [DOI] [PubMed] [Google Scholar]
  598. Fukuyama T.; Bando T.; Ryu I. Electron-Transfer-Induced Intramolecular Heck Carbonylation Reactions Leading to Benzolactones and Benzolactams. Synthesis 2018, 50, 3015–3021. 10.1055/s-0037-1609964. [DOI] [Google Scholar]
  599. Kawamoto T.; Fukuyama T.; Picard B.; Ryu I. New directions in radical carbonylation chemistry: combination with electron catalysis, photocatalysis and ring-opening. Chem. Commun. 2022, 58, 7608–7617. 10.1039/D2CC02700C. [DOI] [PubMed] [Google Scholar]
  600. Jin F.; Han W. Transition-metal-free, ambient-pressure carbonylative cross-coupling reactions of aryl halides with potassium aryltrifluoroborates. Chem. Commun. 2015, 51, 9133–9136. 10.1039/C5CC01968K. [DOI] [PubMed] [Google Scholar]
  601. Xu F.; Li D.; Han W. Transition-metal-free carbonylation of aryl halides with arylboronic acids by utilizing stoichiometric CHCl3 as the carbon monoxide-precursor. Green Chem. 2019, 21, 2911–2915. 10.1039/C9GC00598F. [DOI] [Google Scholar]
  602. Yu D.; Xu F.; Li D.; Han W. Transition-Metal-Free Carbonylative Suzuki-Miyaura Reactions of Aryl Iodides with Arylboronic Acids Using N-Formylsaccharin as CO Surrogate. Adv. Synth. Catal. 2019, 361, 3102–3107. 10.1002/adsc.201900306. [DOI] [Google Scholar]
  603. Pichette Drapeau M.; Fabre I.; Grimaud L.; Ciofini I.; Ollevier T.; Taillefer M. Transition-Metal-Free a-Arylation of Enolizable Aryl Ketones and Mechanistic Evidence for a Radical Process. Angew. Chem., Int. Ed. 2015, 54, 10587–10591. 10.1002/anie.201502332. [DOI] [PubMed] [Google Scholar]
  604. Nandy A.; Sekar G. KOtBu-Promoted Halogen-Bond-Assisted Intramolecular C-S Cross-Coupling of o-Iodothioanilides for the Synthesis of 2-Substituted Benzothiazoles. J. Org. Chem. 2021, 86, 15825–15834. 10.1021/acs.joc.1c02023. [DOI] [PubMed] [Google Scholar]
  605. Panetti G. B.; Carroll P. J.; Gau M. R.; Manor B. C.; Schelter E. J.; Walsh P. J. Synthesis of an elusive, stable 2-azaallyl radical guided by electrochemical and reactivity studies of 2-azaallyl anions. Chem. Sci. 2021, 12, 4405–4410. 10.1039/D0SC04822D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  606. Zou D.; Wang W.; Ying J. Recent Advances in Radical Coupling Strategies Enabled by 2- Azaallyl Anions as Super-Electron-Donors. Adv. Synth. Catal. 2024, 366, 1450–1466. 10.1002/adsc.202301436. [DOI] [Google Scholar]
  607. Li M.; Berritt S.; Matuszewski L.; Deng G.; Pascual-Escudero A.; Panetti G. B.; Poznik M.; Yang X.; Chruma J. J.; Walsh P. J. Transition-Metal-Free Radical C(sp3)-C(sp2) and C(sp3)-C(sp3) Coupling Enabled by 2-Azaallyls as Super-Electron-Donors and Coupling-Partners. J. Am. Chem. Soc. 2017, 139, 16327–16333. 10.1021/jacs.7b09394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  608. Deng G.; Li M.; Yu K.; Liu C.; Liu Z.; Duan S.; Chen W.; Yang X.; Zhang H.; Walsh P. J. Synthesis of Benzofuran Derivatives through Cascade Radical Cyclization/Intermolecular Coupling of 2-Azaallyls. Angew. Chem., Int. Ed. 2019, 58, 2826–2830. 10.1002/anie.201812369. [DOI] [PubMed] [Google Scholar]
  609. Yu K.; Li M.; Deng G.; Liu C.; Wang J.; Liu Z.; Zhang H.; Yang X.; Walsh P. J. An Efficient Route to Isochromene Derivatives via Cascade Radical Cyclization and Radical-Radical Coupling. Adv. Synth. Catal. 2019, 361, 4354–4359. 10.1002/adsc.201900497. [DOI] [Google Scholar]
  610. Zi Q.; Li M.; Cong J.; Deng G.; Duan S.; Yin M.; Chen W.; Jing H.; Yang X.; Walsh P. J. Super-Electron-Donor 2-Azaallyl Anions Enable Construction of Isoquinolines. Org. Lett. 2022, 24, 1786–1790. 10.1021/acs.orglett.2c00140. [DOI] [PubMed] [Google Scholar]
  611. Wang B.; Li M.; Gao G.; Sanz-Vidal A.; Zheng B.; Walsh P. J. Synthesis of Tryptamines from Radical Cyclization of 2-Iodoaryl Allenyl Amines and Coupling with 2-Azallyls. J. Org. Chem. 2022, 87, 8099–8103. 10.1021/acs.joc.2c00767. [DOI] [PubMed] [Google Scholar]
  612. Greener A. J.; Ubysz P.; Owens-Ward W.; Smith G.; Ocana I.; Whitwood A. C.; Chechik V.; James M. J. Radical-anion coupling through reagent design: hydroxylation of aryl halides. Chem. Sci. 2021, 12, 14641–14646. 10.1039/D1SC04748E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  613. Yu F.; Mao R.; Yu M.; Gu X.; Wang Y. Generation of Aryl Radicals from Aryl Halides: Rongalite-Promoted Transition-Metal-Free Arylation. J. Org. Chem. 2019, 84, 9946–9956. 10.1021/acs.joc.9b01113. [DOI] [PubMed] [Google Scholar]
  614. Laha J. K.; Gupta P. Sulfoxylate Anion Radical-Induced Aryl Radical Generation and Intramolecular Arylation for the Synthesis of Biarylsultams. J. Org. Chem. 2022, 87, 4204–4214. 10.1021/acs.joc.1c03031. [DOI] [PubMed] [Google Scholar]
  615. Yamamoto E.; Izumi K.; Horita Y.; Ito H. Anomalous Reactivity of Silylborane: Transition-Metal-Free Boryl Substitution of Aryl, Alkenyl, and Alkyl Halides with Silylborane/Alkoxy Base Systems. J. Am. Chem. Soc. 2012, 134, 19997–20000. 10.1021/ja309578k. [DOI] [PubMed] [Google Scholar]
  616. Wang X.; Zhu M.-H.; Schuman D. P.; Zhong D.; Wang W.-Y.; Wu L.-Y.; Liu W.; Stoltz B. M.; Liu W.-B. General and Practical Potassium Methoxide/Disilane-Mediated Dehalogenative Deuteration of (Hetero)Arylhalides. J. Am. Chem. Soc. 2018, 140, 10970–10974. 10.1021/jacs.8b07597. [DOI] [PubMed] [Google Scholar]
  617. Wang H.; Tong X.; Huo Y.; Tang J.; Xia C. Potassium methoxide/disilane-mediated formylation of aryl iodides with DMF at room temperature. Org. Chem. Front. 2020, 7, 4074–4079. 10.1039/D0QO00974A. [DOI] [Google Scholar]
  618. Lee Y.; Baek S.-y.; Park J.; Kim S.-T.; Tussupbayev S.; Kim J.; Baik M.-H.; Cho S. H. Chemoselective Coupling of 1,1-Bis[(pinacolato)boryl]alkanes for the Transition-Metal-Free Borylation of Aryl and Vinyl Halides: A Combined Experimental and Theoretical Investigation. J. Am. Chem. Soc. 2017, 139, 976–984. 10.1021/jacs.6b11757. [DOI] [PubMed] [Google Scholar]
  619. Nelson R. B.; Gribble G. W. Reduction of Aryl Iodides with Sodium Hydride. J. Org. Chem. 1974, 39, 1425–1426. 10.1021/jo00926a023. [DOI] [Google Scholar]
  620. Ong D. Y.; Tejo C.; Xu K.; Hirao H.; Chiba S. Hydrodehalogenation of Haloarenes by a Sodium Hydride-Iodide Composite. Angew. Chem., Int. Ed. 2017, 56, 1840–1844. 10.1002/anie.201611495. [DOI] [PubMed] [Google Scholar]
  621. Luo F.; Li C.-L.; Ji P.; Zhou Y.; Gui J.; Chen L.; Yin Y.; Zhang X.; Hu Y.; Chen X.; Liu X.; Chen X.; Yu Z.-X.; Wang W.; Zhang S.-L. Direct insertion into the C-C bond of unactivated ketones with NaH-mediated aryne chemistry. Chem. 2023, 9, 2620–2636. 10.1016/j.chempr.2023.05.032. [DOI] [Google Scholar]
  622. Jiang Y.; Zhu W.; Huang J.; Luo F.; Chen X.; Fang C.; Chen X.; Liu S.; Hu Y.; Zhang S. A simple method for N-arylation of secondary amides/amines through a NaH-initiated aryne generation strategy. Org. Chem. Front. 2023, 11, 12–20. 10.1039/D3QO01109G. [DOI] [Google Scholar]
  623. Ghosh I.; Marzo L.; Das A.; Shaikh R.; König B. Visible Light Mediated Photoredox Catalytic Arylation Reactions. Acc. Chem. Res. 2016, 49, 1566–1577. 10.1021/acs.accounts.6b00229. [DOI] [PubMed] [Google Scholar]
  624. Tay N. E. S.; Lehnherr D.; Rovis T. Photons or Electrons? A Critical Comparison of Electrochemistry and Photoredox Catalysis for Organic Synthesis. Chem. Rev. 2022, 122, 2487–2649. 10.1021/acs.chemrev.1c00384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  625. Liu W.; Li J.; Huang C.-Y.; Li C.-J. Aromatic Chemistry in the Excited State: Facilitating Metal-Free Substitutions and Cross-Couplings. Angew. Chem., Int. Ed. 2020, 59, 1786–1796. 10.1002/anie.201909138. [DOI] [PubMed] [Google Scholar]
  626. Schmalzbauer M.; Marcon M.; König B. Excited State Anions in Organic Transformations. Angew. Chem., Int. Ed. 2021, 60, 6270–6292. 10.1002/anie.202009288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  627. Tian X.; Liu Y.; Yakubov S.; Schütte J.; Chiba S.; Barham J. P. Photo- and electro-chemical strategies for the activations of strong chemical bonds. Chem. Soc. Rev. 2024, 53, 263–316. 10.1039/D2CS00581F. [DOI] [PubMed] [Google Scholar]
  628. Reischauer S.; Pieber B. Emerging concepts in photocatalytic organic synthesis. iScience 2021, 24, 102209. 10.1016/j.isci.2021.102209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  629. Lang Y.; Li C.-J.; Zeng H. Photo-induced transition-metal and external photosensitizer-free organic reactions. Org. Chem. Front. 2021, 8, 3594–3613. 10.1039/D1QO00359C. [DOI] [Google Scholar]
  630. Amos S. G. E.; Garreau M.; Buzzetti L.; Waser J. Photocatalysis with organic dyes: facile access to reactive intermediates for synthesis. Beilstein J. Org. Chem. 2020, 16, 1163–1187. 10.3762/bjoc.16.103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  631. Rigotti T.; Alemán J. Visible light photocatalysis - from racemic to asymmetric activation strategies. Chem. Commun. 2020, 56, 11169–11190. 10.1039/D0CC03738A. [DOI] [PubMed] [Google Scholar]
  632. Cavedon C.; Seeberger P. H.; Pieber B. Photochemical Strategies for Carbon-Heteroatom Bond Formation. Eur. J. Org. Chem. 2020, 2020, 1379–1392. 10.1002/ejoc.201901173. [DOI] [Google Scholar]
  633. Bryden M. A.; Zysman-Colman E. Organic thermally activated delayed fluorescence (TADF) compounds used in photocatalysis. Chem. Soc. Rev. 2021, 50, 7587–7680. 10.1039/D1CS00198A. [DOI] [PubMed] [Google Scholar]
  634. Markushyna Y.; Savateev A. Light as a Tool in Organic Photocatalysis: Multi-Photon Excitation and Chromoselective Reactions. Eur. J. Org. Chem. 2022, 2022, e202200026 10.1002/ejoc.202200026. [DOI] [Google Scholar]
  635. Ravelli D.; Protti S.; Fagnoni M. Carbon-Carbon Bond Forming Reactions via Photogenerated Intermediates. Chem. Rev. 2016, 116, 9850–9913. 10.1021/acs.chemrev.5b00662. [DOI] [PubMed] [Google Scholar]
  636. Lan J.; Chen R.; Duo F.; Hu M.; Lu X. Visible-Light Photocatalytic Reduction of Aryl Halides as a Source of Aryl Radicals. Molecules 2022, 27, 5364. 10.3390/molecules27175364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  637. Gurry M.; Aldabbagh F. A new era for homolytic aromatic substitution: replacing Bu3SnH with efficient light-induced chain reactions. Org. Biomol. Chem. 2016, 14, 3849–3862. 10.1039/C6OB00370B. [DOI] [PubMed] [Google Scholar]
  638. Dam B.; Das B.; Patel B. K. Graphitic carbon nitride materials in dual metallophotocatalysis: a promising concept in organic synthesis. Green Chem. 2023, 25, 3374–3397. 10.1039/D3GC00669G. [DOI] [Google Scholar]
  639. Qiu G.; Li Y.; Wu J. Recent developments for the photoinduced Ar-X bond dissociation reaction. Org. Chem. Front. 2016, 3, 1011–1027. 10.1039/C6QO00103C. [DOI] [Google Scholar]
  640. Bardagi J. I.; Ghosh I.. Photoredox Catalytic Activation of Carbon—Halogen Bonds: C—H Functionalization Reactions under Visible Light. In Visible-Light-Active Photocatalysis: Nanostructured Catalyst Design, Mechanisms, and Applications, First ed.; Ghosh S., Ed.; 2018; pp 75–114. [Google Scholar]
  641. Halder S.; Mandal S.; Kundu A.; Mandal B.; Adhikari D. Super-Reducing Behavior of Benzo[b]phenothiazine Anion Under Visible-Light Photoredox Condition. J. Am. Chem. Soc. 2023, 145, 22403–22412. 10.1021/jacs.3c05787. [DOI] [PubMed] [Google Scholar]
  642. Zeng L.; Huang L.; Lin W.; Jiang L.-H.; Han G. Red light-driven electron sacrificial agentsfree photoreduction of inert aryl halides via triplet-triplet annihilation. Nat. Commun. 2023, 14, 1102. 10.1038/s41467-023-36679-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  643. Dey D.; Kundu A.; Roy M.; Singh V.; Maji S.; Adhikari D. Single electron transfer catalysis by diphenylthiourea under visible light photoredox conditions. Org. Chem. Front. 2023, 10, 5248–5253. 10.1039/D3QO01189E. [DOI] [Google Scholar]
  644. Wang Y.; Wei Y.; Song W.; Chen C.; Zhao J. Photocatalytic Hydrodehalogenation for the Removal of Halogenated Aromatic Contaminants. ChemCatChem. 2019, 11, 258–268. 10.1002/cctc.201801222. [DOI] [Google Scholar]
  645. Juliá F.; Constantin T.; Leonori D. Applications of Halogen-Atom Transfer (XAT) for the Generation of Carbon Radicals in Synthetic Photochemistry and Photocatalysis. Chem. Rev. 2022, 122, 2292–2352. 10.1021/acs.chemrev.1c00558. [DOI] [PubMed] [Google Scholar]
  646. Sumida Y.; Ohmiya H. Direct excitation strategy for radical generation in organic synthesis. Chem. Soc. Rev. 2021, 50, 6320–6332. 10.1039/D1CS00262G. [DOI] [PubMed] [Google Scholar]
  647. Zheng L.; Cai L.; Tao K.; Xie Z.; Lai Y.-L.; Guo W. Progress in Photoinduced Radical Reactions using Electron Donor-Acceptor Complexe. Asian J. Org. Chem. 2021, 10, 711–748. 10.1002/ajoc.202100009. [DOI] [Google Scholar]
  648. Yang Z.; Liu Y.; Cao K.; Zhang X.; Jiang H.; Li J. Synthetic reactions driven by electron-donor-acceptor (EDA) complexes. Beilstein J. Org. Chem. 2021, 17, 771–799. 10.3762/bjoc.17.67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  649. Crisenza G. E. M.; Mazzarella D.; Melchiorre P. Synthetic Methods Driven by the Photoactivity of Electron Donor- Acceptor Complexes. J. Am. Chem. Soc. 2020, 142, 5461–5476. 10.1021/jacs.0c01416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  650. Lima C. G. S.; Lima T. de M.; Duarte M.; Jurberg I. D.; Paixão M. W. Organic Synthesis Enabled by Light-Irradiation of EDA Complexes: Theoretical Background and Synthetic Applications. ACS Catal. 2016, 6, 1389–1407. 10.1021/acscatal.5b02386. [DOI] [Google Scholar]
  651. Tavakolian M.; Hosseini-Sarvari M. Catalyst-Free Organic Transformations under Visible-Light. ACS Sustainable Chem. Eng. 2021, 9, 4296–4323. 10.1021/acssuschemeng.0c06657. [DOI] [Google Scholar]
  652. Volkov A. A.; Bugaenko D. I.; Karchava A. V. Transition Metal and Photocatalyst Free Arylation via Photoexcitable Electron Donor Acceptor Complexes: Mediation and Catalysis. ChemCatChem. 2024, 16, e202301526 10.1002/cctc.202301526. [DOI] [Google Scholar]
  653. Li H.; Chiba S. Thiolate photocatalysis enables radical borylation of reductively inert aryl electrophiles. Chem. 2021, 7, 1414–1416. 10.1016/j.chempr.2021.05.015. [DOI] [Google Scholar]
  654. Wei Y.; Zhou Q.-Q.; Tan F.; Lu L.-Q.; Xiao W.-J. Visible-Light-Driven Organic Photochemical Reactions in the Absence of External Photocatalysts. Synthesis 2019, 51, 3021–3054. 10.1055/s-0037-1611812. [DOI] [Google Scholar]
  655. Yuan Y.-q.; Majumder S.; Yang M.-h.; Guo S.-r. Recent advances in catalyst-free photochemical reactions via electrondonor-acceptor (EDA) complex process. Tetrahedron Lett. 2020, 61, 151506. 10.1016/j.tetlet.2019.151506. [DOI] [Google Scholar]
  656. van der Zee L. J. C.; Pahar S.; Richards E.; Melen R. L.; Slootweg J. C. Insights into Single-Electron-Transfer Processes in Frustrated Lewis Pair Chemistry and Related Donor-Acceptor Systems in Main Group Chemistry. Chem. Rev. 2023, 123, 9653–9675. 10.1021/acs.chemrev.3c00217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  657. Wortman A. K.; Stephenson C. R.J. EDA photochemistry: Mechanistic investigations and future opportunities. Chem. 2023, 9, 2390–2415. 10.1016/j.chempr.2023.06.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  658. Wang R.; Wang X.; Mao S.; Zhao Y.; Yuan B.; Yang X.-Y.; Li J.; Chen Z. Metal-Free Photochemical C-Se Cross-Coupling of Aryl Halides with Diselenides. Adv. Synth. Catal. 2022, 364, 1607–1612. 10.1002/adsc.202200042. [DOI] [Google Scholar]
  659. Cao D.; Pan P.; Li C.-J.; Zeng H. Photo-induced transition-metal and photosensitizer free cross-coupling of aryl halides with disulfides. Green Synth. Catal. 2021, 2, 303–306. 10.1016/j.gresc.2021.04.006. [DOI] [Google Scholar]
  660. Mao S.; Zhao Y.; Luo Z.; Wang R.; Yuan B.; Hu J.; Hu L.; Zhang S.-Q.; Ye X.; Wang M.; Chen Z. Metal-free photo-induced sulfidation of aryl iodide and other chalcogenation. Front. Chem. 2022, 10, 941016. 10.3389/fchem.2022.941016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  661. Pan L.; Cooke M. V.; Spencer A.; Laulhé S. Dimsyl Anion Enables Visible-Light-Promoted Charge Transfer in Cross-Coupling Reactions of Aryl Halides. Adv. Synth. Catal. 2022, 364, 420–425. 10.1002/adsc.202101052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  662. Ye S.; Li X.; Xie W.; Wu J. Photoinduced Sulfonylation Reactions through the Insertion of Sulfur Dioxide. Eur. J. Org. Chem. 2020, 2020, 1274–1287. 10.1002/ejoc.201900396. [DOI] [Google Scholar]
  663. Zhang J.; Zhou K.; Qiu G.; Wu J. Photoinduced synthesis of allylic sulfones using potassium metabisulfite as the source of sulfur dioxide. Org. Chem. Front. 2019, 6, 36–40. 10.1039/C8QO01048J. [DOI] [Google Scholar]
  664. Zhou K.; Liu J.-B.; Xie W.; Ye S.; Wu J. Photoinduced synthesis of 2-sulfonylacetonitriles with the insertion of sulfur dioxide under ultraviolet irradiation. Chem. Commun. 2020, 56, 2554–2557. 10.1039/D0CC00351D. [DOI] [PubMed] [Google Scholar]
  665. Bartolomei B.; Gentile G.; Rosso C.; Filippini G.; Prato M. Turning the Light on Phenols: New Opportunities in Organic Synthesis. Chem.—Eur. J. 2021, 27, 16062–16070. 10.1002/chem.202102276. [DOI] [PubMed] [Google Scholar]
  666. Filippini G.; Dosso J.; Prato M. Phenols as Novel Photocatalytic Platforms for Organic Synthesis. Helv. Chim. Acta 2023, 106, e202300059 10.1002/hlca.202300059. [DOI] [Google Scholar]
  667. Zhu D.-L.; Jiang S.; Young D. J.; Wu Q.; Li H.-Y.; Li H.-X. Visible-light-driven C(sp2)-H arylation of phenols with arylbromides enabled by electron donor- acceptor excitation. Chem. Commun. 2022, 58, 3637–3640. 10.1039/D1CC07127K. [DOI] [PubMed] [Google Scholar]
  668. Yang Q.-Q.; Liu N.; Yan J.-Y.; Ren Z.-L.; Wang L. Visible Light- and Heat-Promoted C. Asian J. Org. Chem. 2020, 9, 116–120. 10.1002/ajoc.201900666. [DOI] [Google Scholar]
  669. Li H.; Liu Y.; Chiba S. Leveraging of Sulfur Anions in Photoinduced Molecular Transformations. JACS Au 2021, 1, 2121–2129. 10.1021/jacsau.1c00363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  670. Wang S.; Wang H.; König B. Light-Induced Single-Electron Transfer Processes involving Sulfur Anions as Catalysts. J. Am. Chem. Soc. 2021, 143, 15530–15537. 10.1021/jacs.1c07785. [DOI] [PubMed] [Google Scholar]
  671. Wang S.; Wang H.; König B. Photo-induced thiolate catalytic activation of inert Caryl-hetero bonds for radical borylation. Chem. 2021, 7, 1653–1665. 10.1016/j.chempr.2021.04.016. [DOI] [Google Scholar]
  672. Liu B.; Lim C.-H.; Miyake G. M. Visible-Light-Promoted C-S Cross-Coupling via Intermolecular Charge Transfer. J. Am. Chem. Soc. 2017, 139, 13616–13619. 10.1021/jacs.7b07390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  673. Zhu C.; Zhumagazy S.; Yue H.; Rueping M. Metal-free C-Se cross-coupling enabled by photoinduced inter-molecular charge transfer. Chem. Commun. 2021, 58, 96–99. 10.1039/D1CC06152F. [DOI] [PubMed] [Google Scholar]
  674. Piedra H. F.; Valdés C.; Plaza M. Shining light on halogen-bonding complexes: a catalyst-free activation mode of carbon-halogen bonds for the generation of carbon-centered radicals. Chem. Sci. 2023, 14, 5545–5568. 10.1039/D3SC01724A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  675. Nandy A.; Kazi I.; Guha S.; Sekar G. Visible-Light-Driven Halogen-Bond-Assisted Direct Synthesis of Heteroaryl Thioethers Using Transition-Metal-Free One-Pot C-I Bond Formation/C-S Cross-Coupling Reaction. J. Org. Chem. 2021, 86, 2570–2581. 10.1021/acs.joc.0c02672. [DOI] [PubMed] [Google Scholar]
  676. Sundaravelu N.; Nandy A.; Sekar G. Visible Light Mediated Photocatalyst Free C-S Cross Coupling: Domino Synthesis of Thiochromane Derivatives via Photoinduced Electron Transfer. Org. Lett. 2021, 23, 3115–3119. 10.1021/acs.orglett.1c00806. [DOI] [PubMed] [Google Scholar]
  677. Xu Z.-M.; Li H.-X.; Young D. J.; Zhu D.-L.; Li H.-Y.; Lang J.-P. Exogenous Photosensitizer-, Metal-, and Base-Free Visible-Light-Promoted C-H Thiolation via Reverse Hydrogen Atom Transfer. Org. Lett. 2019, 21, 237–241. 10.1021/acs.orglett.8b03679. [DOI] [PubMed] [Google Scholar]
  678. Wang H.; Wu Q.; Zhang J.-D.; Li H.-Y.; Li H.-X. Photocatalyst- and Transition-Metal-Free Visible-Light-Promoted Intramolecular C(sp2)-S Formation. Org. Lett. 2021, 23, 2078–2083. 10.1021/acs.orglett.1c00235. [DOI] [PubMed] [Google Scholar]
  679. Yan K.; Liu M.; Wen J.; Liu X.; Wang X.; Chen X.; Li J.; Wang S.; Wang X.; Wang H. Visible-light-promoted cascade cyclization towards benzo[d]imidazo[5,1-b]thiazoles under metal- and photocatalyst-free conditions. Green Chem. 2021, 23, 1286–1291. 10.1039/D0GC04135A. [DOI] [Google Scholar]
  680. Chen L.; Liang J.; Chen Z.-Y.; Chen J.; Yan M.; Zhang X.-J. A Convenient Synthesis of Sulfones via Light Promoted Coupling of Sodium Sulfinates and Aryl Halides. Adv. Synth. Catal. 2019, 361, 956–960. 10.1002/adsc.201801390. [DOI] [Google Scholar]
  681. Lin B.; Lu W.; Chen Z.-y.; Zhang Y.; Duan Y.-z.; Lu X.; Yan M.; Zhang X.-j. Enhancing the Potential of Miniature-Scale DNA-Compatible Radical Reactions via an Electron Donor-Acceptor Complex and a Reversible Adsorption to Solid Support Strategy. Org. Lett. 2021, 23, 7381–7385. 10.1021/acs.orglett.1c02562. [DOI] [PubMed] [Google Scholar]
  682. Li F.; Han X.; Xu Z.; Zhang C.-P. Photoinduced Trifluoromethylselenolation of Aryl Halides with [Me4N][SeCF3]. Org. Lett. 2023, 25, 7884–7889. 10.1021/acs.orglett.3c03116. [DOI] [PubMed] [Google Scholar]
  683. Liu Y.; Li H.; Chiba S. Photoinduced Cross-Coupling of Aryl Iodides with Alkenes. Org. Lett. 2021, 23, 427–432. 10.1021/acs.orglett.0c03935. [DOI] [PubMed] [Google Scholar]
  684. Li Y.; Wise D. E.; Mitchell J. K.; Parasram M. Cascade Synthesis of Phenanthrenes under Photoirradiation. J. Org. Chem. 2023, 88, 717–721. 10.1021/acs.joc.2c02202. [DOI] [PubMed] [Google Scholar]
  685. Liang K.; Liu Q.; Shen L.; Li X.; Wei D.; Zheng L.; Xia C. Intermolecular oxyarylation of olefins with aryl halides and TEMPOH catalyzed by the phenolate anion under visible light. Chem. Sci. 2020, 11, 6996–7002. 10.1039/D0SC02160A. [DOI] [Google Scholar]
  686. Zhang L.; Hu F.; Shen L.; Gao L.; Yang Y.; Pan Z.; Xia C. Redox-Neutral Intramolecular Dearomative Spirocyclization of Phenols Induced by Visible Light. Org. Lett. 2023, 25, 3168–3172. 10.1021/acs.orglett.3c01257. [DOI] [PubMed] [Google Scholar]
  687. Liang K.; Li T.; Li N.; Zhang Y.; Shen L.; Ma Z.; Xia C. Redox-neutral photochemical Heck-type arylation of vinylphenols activated by visible light. Chem. Sci. 2020, 11, 2130–2135. 10.1039/C9SC06184C. [DOI] [PMC free article] [PubMed] [Google Scholar]
  688. Ye S.; Zhou K.; Rojsitthisak P.; Wu J. Metal-free insertion of sulfur dioxide with aryl iodides under ultraviolet irradiation: direct access to sulfonated cyclic compounds. Org. Chem. Front. 2020, 7, 14–18. 10.1039/C9QO01274E. [DOI] [Google Scholar]
  689. da Silva G. P.; Ali A.; da Silva R. C.; Jiang H.; Paixão M. W. Tris(trimethylsilyl)silane and visible-light irradiation: a new metal- and additive-free photochemical process for the synthesis of indoles and oxindoles. Chem. Commun. 2015, 51, 15110–15113. 10.1039/C5CC06329A. [DOI] [PubMed] [Google Scholar]
  690. Caiuby C.A. D.; Ali A.; Santana V. T.; Lucas F. W. de S.; Santos M. S.; Correa A. G.; Nascimento O. R.; Jiang H.; Paixão M. W. Intramolecular radical cyclization approach to access highly substituted indolines and 2,3- dihydrobenzofurans under visible-light. RSC Adv. 2018, 8, 12879–12886. 10.1039/C8RA01787E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  691. Hendy C. M.; Smith G. C.; Xu Z.; Lian T.; Jui N. T. Radical Chain Reduction via Carbon Dioxide Radical Anion (CO2•–). J. Am. Chem. Soc. 2021, 143, 8987–8992. 10.1021/jacs.1c04427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  692. Chmiel A. F.; Williams O. P.; Chernowsky C. P.; Yeung C. S.; Wickens Z. K. Non-innocent Radical Ion Intermediates in Photoredox Catalysis: Parallel Reduction Modes Enable Coupling of Diverse Aryl Chlorides. J. Am. Chem. Soc. 2021, 143, 10882–10889. 10.1021/jacs.1c05988. [DOI] [PubMed] [Google Scholar]
  693. Xiao W.; Zhang J.; Wu J. Recent Advances in Reactions Involving Carbon Dioxide Radical Anion. ACS Catal. 2023, 13, 15991–16011. 10.1021/acscatal.3c04125. [DOI] [Google Scholar]
  694. Hou J.; Hua L.-L.; Huang Y.; Zhan L.-W.; Li B.-D. Visible-Light-Promoted Catalyst-Free Oxyarylation and Hydroarylation of Alkenes with Carbon Dioxide Radical Anion. Chem.—Asian J. 2023, 18, e202201092 10.1002/asia.202201092. [DOI] [PubMed] [Google Scholar]
  695. Yao Q.; Liu W.; Liu P.; Ren L.; Fang X.; Li C.-J. Metal-Free Photoinduced Transformation of Aryl Halides and Diketones into Aryl Ketones. Eur. J. Org. Chem. 2019, 2019, 2721–2724. 10.1002/ejoc.201801650. [DOI] [Google Scholar]
  696. Pan P.; Liu S.; Lan Y.; Zeng H.; Li C.-J. Visible-light-induced cross-coupling of aryl iodides with hydrazones via an EDA-complex. Chem. Sci. 2022, 13, 7165–7171. 10.1039/D2SC01909D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  697. Liang K.; Li N.; Zhang Y.; Li T.; Xia C. Transition-metal-free a-arylation of oxindoles via visible-light-promoted electron transfer. Chem. Sci. 2019, 10, 3049–3053. 10.1039/C8SC05170D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  698. Liu Y.; Li S.; Chen X.; Jiang H. Visible-Light-Mediated Synthesis of α-Aryl Ester Derivatives via an EDA Complex. J. Org. Chem. 2023, 88, 12474–12480. 10.1021/acs.joc.3c01201. [DOI] [PubMed] [Google Scholar]
  699. Volkov A. A.; Bugaenko D. I.; Karchava A. V. Visible Light instead of Transition Metal: Electron Donor Acceptor Complex Enabled Cross-Coupling of Aryl Halides with Active Methylene Compounds. Adv. Synth. Catal. 2024, 366, 457–464. 10.1002/adsc.202301192. [DOI] [Google Scholar]
  700. Hayashi H.; Wang B.; Wu X.; Teo S. J.; Kaga A.; Watanabe K.; Takita R.; Yeow E. K. L.; Chiba S. Biaryl Cross-Coupling Enabled by Photo-Induced Electron Transfer. Adv. Synth. Catal. 2020, 362, 2223–2231. 10.1002/adsc.201901601. [DOI] [Google Scholar]
  701. Miao R.; Wang D.; Xiao J.; Ma J.; Xue D.; Liu F.; Fang Y. Halogen bonding matters: visible light-induced photoredox catalyst-free aryl radical formation and its applications. Phys. Chem. Chem. Phys. 2020, 22, 10212–10218. 10.1039/D0CP00946F. [DOI] [PubMed] [Google Scholar]
  702. Walia P. K.; Kumar M.; Bhalla V. Empowering Transition-Metal-Free Cascade Protocol for the Green Synthesis of Biaryls and Alkynes. ACS Omega 2018, 3, 1983–1990. 10.1021/acsomega.7b02014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  703. Kazi I.; Nandy A.; Selvam R.; Sekar G. Halogen Bond-Activated Visible-Light-Mediated Regioselective C-H Arylation of 2-Phenylimidazo-[1,2-a]pyridines. J. Org. Chem. 2022, 87, 12323–12333. 10.1021/acs.joc.2c01548. [DOI] [PubMed] [Google Scholar]
  704. Heredia M. D.; Puiatti M.; Rossi R. A.; Budén M. E. Visible light mediated synthesis of 6H-benzo[c]chromenes: transition-metal-free intramolecular direct C-H arylation. Org. Biomol. Chem. 2021, 20, 228–239. 10.1039/D1OB01673C. [DOI] [PubMed] [Google Scholar]
  705. Tang W.-X.; Chen K.-Q.; Sun D.-Q.; Chen X.-Y. Photoinduced halogen-bonding enabled synthesis of oxindoles and isoindolinones from aryl iodides. Org. Biomol. Chem. 2023, 21, 715–718. 10.1039/D2OB01818G. [DOI] [PubMed] [Google Scholar]
  706. Yang Z.-S.; Tang W.-X.; Zhang B.-B.; Sun D.-Q.; Chen K.-Q.; Chen X.-Y. Electron donor-acceptor complex enabled cascade reaction of unprotected o-anilide aryl chlorides for heterocycle synthesis. Org. Chem. Front. 2023, 10, 1219–1223. 10.1039/D2QO01863B. [DOI] [Google Scholar]
  707. Alonso F.; Beletskaya I. P.; Yus M. Metal-Mediated Reductive Hydrodehalogenation of Organic Halides. Chem. Rev. 2002, 102, 4009–4091. 10.1021/cr0102967. [DOI] [PubMed] [Google Scholar]
  708. Kopf S.; Bourriquen F.; Li W.; Neumann H.; Junge K.; Beller M. Recent Developments for the Deuterium and Tritium Labeling of Organic Molecules. Chem. Rev. 2022, 122, 6634–6718. 10.1021/acs.chemrev.1c00795. [DOI] [PubMed] [Google Scholar]
  709. Li N.; Li Y.; Wu X.; Zhu C.; Xie J. Radical deuteration. Chem. Soc. Rev. 2022, 51, 6291–6306. 10.1039/D1CS00907A. [DOI] [PubMed] [Google Scholar]
  710. Pirali T.; Serafini M.; Cargnin S.; Genazzani A. A. Applications of Deuterium in Medicinal Chemistry. J. Med. Chem. 2019, 62, 5276–5297. 10.1021/acs.jmedchem.8b01808. [DOI] [PubMed] [Google Scholar]
  711. Levernier E.; Tatoueix K.; Garcia-Argote S.; Pfeifer V.; Kiesling R.; Gravel E.; Feuillastre S.; Pieters G. Easy-to-Implement Hydrogen Isotope Exchange for the Labeling of N-Heterocycles, Alkylkamines, Benzylic Scaffolds, and Pharmaceuticals. JACS Au 2022, 2, 801–808. 10.1021/jacsau.1c00503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  712. Zhou R.; Ma L.; Yang X.; Cao J. Recent advances in visible-light photocatalytic deuteration reactions. Org. Chem. Front. 2021, 8, 426–444. 10.1039/D0QO01299H. [DOI] [Google Scholar]
  713. Budén M. E.; Bardagí J. I.; Puiatti M.; Rossi R. A. Initiation in Photoredox C-H Functionalization Reactions. Is Dimsyl Anion a Key Ingredient?. J. Org. Chem. 2017, 82, 8325–8333. 10.1021/acs.joc.7b00822. [DOI] [PubMed] [Google Scholar]
  714. Elhage A.; Costa P.; Nasim A.; Lanterna A. E.; Scaiano J. C. Photochemical Dehalogenation of Aryl Halides: Importance of Halogen Bonding. J. Phys. Chem. A 2019, 123, 10224–10229. 10.1021/acs.jpca.9b06716. [DOI] [PubMed] [Google Scholar]
  715. Fukuyama T.; Fujita Y.; Miyoshi H.; Ryu I.; Kao S.-C.; Wu Y.-K. Electron transfer-induced reduction of organic halides with amines. Chem. Commun. 2018, 54, 5582–5585. 10.1039/C8CC02445F. [DOI] [PubMed] [Google Scholar]
  716. Cao D.; Yan C.; Zhou P.; Zeng H.; Li C.-J. Hydrogen bonding promoted simple and clean photo-induced reduction of C-X bond with isopropanol. Chem. Commun. 2019, 55, 767–770. 10.1039/C8CC08942F. [DOI] [PubMed] [Google Scholar]
  717. Yan B.; Zhou Y.; Wu J.; Ran M.; Li H.; Yao Q. Catalyst-free reductive hydrogenation or deuteration of aryl-heteroatom bonds induced by light. Org. Chem. Front. 2021, 8, 5244–5249. 10.1039/D1QO00978H. [DOI] [Google Scholar]
  718. Ding T.-H.; Qu J.-P.; Kang Y.-B. Visible-Light-Induced, Base-Promoted Transition-Metal-Free Dehalogenation of Aryl Fluorides, Chlorides, Bromides, and Iodides. Org. Lett. 2020, 22, 3084–3088. 10.1021/acs.orglett.0c00827. [DOI] [PubMed] [Google Scholar]
  719. Sarkar S.; Wagulde S.; Jia X.; Gevorgyan V. General and selective metal-free radical a-C-H borylation of aliphatic amines. Chem. 2022, 8, 3096–3108. 10.1016/j.chempr.2022.07.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  720. Guo R.; Shu C.; Zhang G. Photoinduced metal-free α-C-H borylation of aliphatic amines. Chem. 2022, 8, 2902–2904. 10.1016/j.chempr.2022.10.004. [DOI] [Google Scholar]
  721. Cai Y.-M.; Xu Y.-T.; Zhang X.; Gao W.-X.; Huang X.-B.; Zhou Y.-B.; Liu M.-C.; Wu H.-Y. Photoinduced Hydroxylation of Organic Halides under Mild Conditions. Org. Lett. 2019, 21, 8479–8484. 10.1021/acs.orglett.9b03317. [DOI] [PubMed] [Google Scholar]
  722. Li L.; Liu W.; Zeng H.; Mu X.; Cosa G.; Mi Z.; Li C.-J. Photo-induced Metal-Catalyst-Free Aromatic Finkelstein Reaction. J. Am. Chem. Soc. 2015, 137, 8328–8331. 10.1021/jacs.5b03220. [DOI] [PubMed] [Google Scholar]
  723. Yang X.; Liu W.; Li L.; Wei W.; Li C.-J. Photo-induced Carboiodination: A Simple Way to Synthesize Functionalized Dihydrobenzofurans and Indolines. Chem.—Eur. J. 2016, 22, 15252–15256. 10.1002/chem.201603608. [DOI] [PubMed] [Google Scholar]
  724. Mfuh A. M.; Doyle J. D.; Chhetri B.; Arman H. D.; Larionov O. V. Scalable, Metal- and Additive-Free, Photoinduced Borylation of Haloarenes and Quaternary Arylammonium Salts. J. Am. Chem. Soc. 2016, 138, 2985–2988. 10.1021/jacs.6b01376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  725. Mfuh A. M.; Nguyen V. T.; Chhetri B.; Burch J. E.; Doyle J. D.; Nesterov V. N.; Arman H. D.; Larionov O. V. Additive- and Metal-Free, Predictably 1,2- and 1,3-Regioselective, Photoinduced Dual C-H/C-X Borylation of Haloarenes. J. Am. Chem. Soc. 2016, 138, 8408–8411. 10.1021/jacs.6b05436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  726. Chen K.; Zhang S.; He P.; Li P. Efficient metal-free photochemical borylation of aryl halides under batch and continuous-flow conditions. Chem. Sci. 2016, 7, 3676–3680. 10.1039/C5SC04521E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  727. Chen K.; Cheung M. S.; Lin Z.; Li P. Metal-free borylation of electron-rich aryl (pseudo)halides under continuous-flow photolytic conditions. Org. Chem. Front. 2016, 3, 875–879. 10.1039/C6QO00109B. [DOI] [Google Scholar]
  728. Cheng Y.; Muck-Lichtenfeld C.; Studer A. Metal-Free Radical Borylation of Alkyl and Aryl Iodides. Angew. Chem., Int. Ed. 2018, 57, 16832–16836. 10.1002/anie.201810782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  729. Friese F. W.; Studer A. New avenues for C-B bond formation via radical intermediates. Chem. Sci. 2019, 10, 8503–8518. 10.1039/C9SC03765A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  730. Matsuo K.; Yamaguchi E.; Itoh A. Halogen-Bonding-Promoted Photoinduced C-X Borylation of Aryl Halide Using Phenol Derivatives. J. Org. Chem. 2023, 88, 6176–6181. 10.1021/acs.joc.3c00201. [DOI] [PubMed] [Google Scholar]
  731. Yuan J.; To W.-P.; Zhang Z.-Y.; Yue C.-D.; Meng S.; Chen J.; Liu Y.; Yu G.-A.; Che C.-M. Visible-Light-Promoted Transition-Metal-Free Phosphinylation of Heteroaryl Halides in the Presence of Potassium tert-Butoxide. Org. Lett. 2018, 20, 7816–7820. 10.1021/acs.orglett.8b03265. [DOI] [PubMed] [Google Scholar]
  732. Zeng H.; Dou Q.; Li C.-J. Photoinduced Transition-Metal-Free Cross-Coupling of Aryl Halides with H-Phosphonates. Org. Lett. 2019, 21, 1301–1305. 10.1021/acs.orglett.8b04081. [DOI] [PubMed] [Google Scholar]
  733. Oßwald S.; Zippel C.; Hassan Z.; Nieger M.; Bräse S. C-P bond formation of cyclophanyl-, and aryl halides via a UV-induced photo Arbuzov reaction: a versatile portal to phosphonate-grafted scaffolds. RSC Adv. 2022, 12, 3309–3312. 10.1039/D2RA00094F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  734. Pan L.; Deckert M. M.; Cooke M. V.; Bleeke A. R.; Laulhé S. Solvent Anions Enable Photoinduced Borylation and Phosphonation of Aryl Halides via EDA Complexes. Org. Lett. 2022, 24, 6466–6471. 10.1021/acs.orglett.2c02631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  735. Roy V. J.; Raha Roy S. Light-Induced Activation of C-X Bond via Carbonate-Assisted Anion-π Interactions: Applications to C-P and C-B Bond Formation. Org. Lett. 2023, 25, 923–927. 10.1021/acs.orglett.2c04208. [DOI] [PubMed] [Google Scholar]
  736. Murray P. R. D.; Cox J. H.; Chiappini N. D.; Roos C. B.; McLoughlin E. A.; Hejna B. G.; Nguyen S. T.; Ripberger H. H.; Ganley J. M.; Tsui E.; Shin N. Y.; Koronkiewicz B.; Qiu G.; Knowles R. R. Photochemical and Electrochemical Applications of Proton-Coupled Electron Transfer in Organic Synthesis. Chem. Rev. 2022, 122, 2017–2291. 10.1021/acs.chemrev.1c00374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  737. Cybularczyk-Cecotka M.; Szczepanik J.; Giedyk M. Photocatalytic strategies for the activation of organic chlorides. Nat. Catal. 2020, 3, 872–886. 10.1038/s41929-020-00515-8. [DOI] [Google Scholar]
  738. McAtee R. C.; McClain E. J.; Stephenson C. R. J. Illuminating Photoredox Catalysis. Trends in Chem. 2019, 1, 111–125. 10.1016/j.trechm.2019.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  739. Prier C. K.; Rankic D. A.; MacMillan D. W. C. Visible Light Photoredox Catalysis with Transition Metal Complexes: Applications in Organic Synthesis. Chem. Rev. 2013, 113, 5322–5363. 10.1021/cr300503r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  740. Ghosh S.; Majumder S.; Ghosh D.; Hajra A. Redox-neutral carbon-heteroatom bond formation under photoredox catalysis. Chem. Commun. 2023, 59, 7004–7027. 10.1039/D3CC01873C. [DOI] [PubMed] [Google Scholar]
  741. Corbin D. A.; Miyake G. M. Photoinduced Organocatalyzed Atom Transfer Radical Polymerization (O-ATRP): Precision Polymer Synthesis Using Organic Photoredox Catalysis. Chem. Rev. 2022, 122, 1830–1874. 10.1021/acs.chemrev.1c00603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  742. Liu Q.; Huo C.; Fu Y.; Du Z. Recent progress in organophotoredox reaction. Org. Biomol. Chem. 2022, 20, 6721–6740. 10.1039/D2OB00807F. [DOI] [PubMed] [Google Scholar]
  743. Lee Y.; Kwon M. S. Emerging Organic Photoredox Catalysts for Organic Transformations. Eur. J. Org. Chem. 2020, 2020, 6028–6043. 10.1002/ejoc.202000720. [DOI] [Google Scholar]
  744. Cannalire R.; Pelliccia S.; Sancineto L.; Novellino E.; Tron G. C.; Giustiniano M. Visible light photocatalysis in the late-stage functionalization of pharmaceutically relevant compounds. Chem. Soc. Rev. 2021, 50, 766–897. 10.1039/D0CS00493F. [DOI] [PubMed] [Google Scholar]
  745. Xu G.-Q.; Xu P.-F. Visible light organic photoredox catalytic cascade reactions. Chem. Commun. 2021, 57, 12914–12935. 10.1039/D1CC04883J. [DOI] [PubMed] [Google Scholar]
  746. Glaser F.; Kerzig C.; Wenger O. S. Multi-Photon Excitation in Photoredox Catalysis: Concepts, Applications, Methods. Angew. Chem., Int. Ed. 2020, 59, 10266–10284. 10.1002/anie.201915762. [DOI] [PubMed] [Google Scholar]
  747. Romero N. A.; Nicewicz D. A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116, 10075–10166. 10.1021/acs.chemrev.6b00057. [DOI] [PubMed] [Google Scholar]
  748. Du Y.; Pearson R. M.; Lim C.-H.; Sartor S. M.; Ryan M. D.; Yang H.; Damrauer N. H.; Miyake G. M. Strongly Reducing, Visible-Light Organic Photoredox Catalysts as Sustainable Alternatives to Precious Metals. Chem.—Eur. J. 2017, 23, 10962–10968. 10.1002/chem.201702926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  749. Mandigma M. J. P.; Kaur J.; Barham J. P. Organophotocatalytic Mechanisms: Simplicity or Naïvety? Diverting Reactive Pathways by Modifications of Catalyst Structure, Redox States and Substrate Preassemblies. ChemCatChem. 2023, 15, e202201542 10.1002/cctc.202201542. [DOI] [Google Scholar]
  750. Sachidanandan K.; Niu B.; Laulhé S. An Overview of α-Aminoalkyl Radical Mediated HalogenAtom Transfer. ChemCatChem. 2023, 15, e202300860 10.1002/cctc.202300860. [DOI] [Google Scholar]
  751. Juliá F.New synthetic strategies based on photoinduced halogen-atom transfer (XAT). In Photochemistry Vol. 51; Crespi S.; Protti S., Eds.; Royal Society of Chemistry, 2023; pp 361–383. [Google Scholar]
  752. Li H.; Tang X.; Pang J. H.; Wu X.; Yeow E. K. L.; Wu J.; Chiba S. Polysulfide Anions as Visible Light Photoredox Catalysts for Aryl Cross-Couplings. J. Am. Chem. Soc. 2021, 143, 481–487. 10.1021/jacs.0c11968. [DOI] [PubMed] [Google Scholar]
  753. Li H.; Liu Y.; Chiba S. Anti-Markovnikov hydroarylation of alkenes via polysulfide anion photocatalysis. Chem. Commun. 2021, 57, 6264–6267. 10.1039/D1CC02185K. [DOI] [PubMed] [Google Scholar]
  754. Das T. K.; Kundu M.; Mondal B.; Ghosh P.; Das S. Organocatalytic synthesis of (Het)biaryl scaffolds via photoinduced intra/intermolecular C(sp2)-H arylation by 2-pyridone derivatives. Org. Biomol. Chem. 2021, 20, 208–218. 10.1039/D1OB01798E. [DOI] [PubMed] [Google Scholar]
  755. Poelma S. O.; Burnett G. L.; Discekici E. H.; Mattson K. M.; Treat N. J.; Luo Y.; Hudson Z. M.; Shankel S. L.; Clark P. G.; Kramer J. W.; Hawker C. J.; de Alaniz J. R. Chemoselective Radical Dehalogenation and C-C Bond Formation on Aryl Halide Substrates Using Organic Photoredox Catalysts. J. Org. Chem. 2016, 81, 7155–7160. 10.1021/acs.joc.6b01034. [DOI] [PubMed] [Google Scholar]
  756. Jin S.; Dang H. T.; Haug G. C.; He R.; Nguyen V. D.; Nguyen V. T.; Arman H. D.; Schanze K. S.; Larionov O. V. Visible Light-Induced Borylation of C-O, C-N, and C-X Bonds. J. Am. Chem. Soc. 2020, 142, 1603–1613. 10.1021/jacs.9b12519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  757. Pan L.; Kelley A. S.; Cooke M. V.; Deckert M. M.; Laulhé S. Transition-Metal-Free Photoredox Phosphonation of Aryl C-N and C-X Bonds in Aqueous Solvent Mixtures. ACS Sustainable Chem. Eng. 2022, 10, 691–695. 10.1021/acssuschemeng.1c07394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  758. Garrido-Castro A. F.; Salaverri N.; Maestro M. C.; Alemán J. Intramolecular Homolytic Substitution Enabled by Photoredox Catalysis: Sulfur, Phosphorus, and Silicon Heterocycle Synthesis from Aryl Halides. Org. Lett. 2019, 21, 5295–5300. 10.1021/acs.orglett.9b01911. [DOI] [PubMed] [Google Scholar]
  759. Heredia M. D.; Guerra W. D.; Barolo S. M.; Fornasier S. J.; Rossi R. A.; Budén M. E. Transition-Metal-Free and Visible-Light-Mediated Desulfonylation and Dehalogenation Reactions: Hantzsch Ester Anion as Electron and Hydrogen Atom Donor. J. Org. Chem. 2020, 85, 13481–13494. 10.1021/acs.joc.0c01523. [DOI] [PubMed] [Google Scholar]
  760. Zhu D.-L.; Wu Q.; Li H.-Y.; Li H.-X.; Lang J.-P. Hantzsch Ester as a Visible-Light Photoredox Catalyst for Transition-Metal-Free Coupling of Arylhalides and Arylsulfinates. Chem.—Eur. J. 2020, 26, 3484–3488. 10.1002/chem.201905281. [DOI] [PubMed] [Google Scholar]
  761. Constantin T.; Zanini M.; Regni A.; Sheikh N. S.; Juliá F.; Leonori D. Aminoalkyl radicals as halogen-atom transfer agents for activation of alkyl and aryl halides. Science 2020, 367, 1021–1026. 10.1126/science.aba2419. [DOI] [PubMed] [Google Scholar]
  762. Constantin T.; Juliá F.; Sheikh N. S.; Leonori D. A case of chain propagation: a-aminoalkyl radicals as initiators for aryl radical chemistry. Chem. Sci. 2020, 11, 12822–12828. 10.1039/D0SC04387G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  763. Mondal K.; Mallik S.; Sardana S.; Baidya M. A Visible-Light-Induced α-Aminoalkyl-Radical-Mediated Halogen-Atom Transfer Process: Modular Synthesis of Phenanthridinone Alkaloids. Org. Lett. 2023, 25, 1689–1694. 10.1021/acs.orglett.3c00358. [DOI] [PubMed] [Google Scholar]
  764. MacKenzie I. A.; Wang L.; Onuska N. P. R.; Williams O. F.; Begam K.; Moran A. M.; Dunietz B. D.; Nicewicz D. A. Discovery and characterization of an acridine radical photoreductant. Nature 2020, 580, 76–81. 10.1038/s41586-020-2131-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  765. Chen W.; Wang H.; Tay N. E. S.; Pistritto V. A.; Li K.-P.; Zhang T.; Wu Z.; Nicewicz D. A.; Li Z. Arene radiofluorination enabled by photoredox-mediated halide interconversion. Nat. Chem. 2022, 14, 216–223. 10.1038/s41557-021-00835-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  766. Chen W.; Huang Z.; Tay N. E. S.; Giglio B.; Wang M.; Wang H.; Wu Z.; Nicewicz D. A.; Li Z. Direct arene C-H fluorination with 18F via organic photoredox catalysis. Science 2019, 364, 1170–1174. 10.1126/science.aav7019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  767. Tay N. E. S.; Chen W.; Levens A.; Pistritto V. A.; Huang Z.; Wu Z.; Li Z.; Nicewicz D. A. 19F- and 18F-arene deoxyfluorination via organic photoredox-catalysed polarity-reversed nucleophilic aromatic substitution. Nat. Catal. 2020, 3, 734–742. 10.1038/s41929-020-0495-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  768. Hossain M. M.; Shaikh A. C.; Moutet J.; Gianetti T. L. Photocatalytic α-arylation of cyclic ketones. Nat. Synth. 2022, 1, 147–157. 10.1038/s44160-021-00021-0. [DOI] [Google Scholar]
  769. Lennert U.; Arockiam P. B.; Streitferdt V.; Scott D. J.; Rödl C.; Gschwind R. M.; Wolf R. Direct catalytic transformation of white phosphorus into arylphosphines and phosphonium salts. Nat. Catal. 2019, 2, 1101–1106. 10.1038/s41929-019-0378-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  770. Arockiam P. B.; Lennert U.; Graf C.; Rothfelder R.; Scott D. J.; Fischer T. G.; Zeitler K.; Wolf R. Versatile Visible-Light-Driven Synthesis of Asymmetrical Phosphines and Phosphonium Salts. Chem.—Eur. J. 2020, 26, 16374–16382. 10.1002/chem.202002646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  771. Liu C.; Shen N.; Shang R. Photocatalytic defluoroalkylation and hydrodefluorination of trifluoromethyls using o-phosphinophenolate. Nat. Commun. 2022, 13, 354. 10.1038/s41467-022-28007-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  772. Shen N.; Liu; Zhang C.; X; Shang R. o-Phosphinodiarylamides as Reductive Photocatalysts for Dehalogenative and Deaminative Cross-Couplings. ACS Catal. 2023, 13, 11753–11761. 10.1021/acscatal.3c03569. [DOI] [Google Scholar]
  773. Shen N.; Li R.; Liu C.; Shen X.; Guan W.; Shang R. Photocatalytic Cross-Couplings of Aryl Halides Enabled by o-Phosphinophenolate and o-Phosphinothiophenolate. ACS Catal. 2022, 12, 2788–2795. 10.1021/acscatal.1c05941. [DOI] [Google Scholar]
  774. Majhi J.; Granados A.; Matsuo B.; Ciccone V.; Dhungana R. K.; Sharique M.; Molander G. A. Practical, scalable, and transition metal-free visible light-induced heteroarylation route to substituted oxindoles. Chem. Sci. 2023, 14, 897–902. 10.1039/D2SC05918E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  775. Chen K.-Q.; Zhang B.-B.; Wang Z.-X.; Chen X.-Y. N-Heterocyclic Nitreniums Can Be Employed as Photoredox Catalysts for the Single-Electron Reduction of Aryl Halides. Org. Lett. 2022, 24, 4598–4602. 10.1021/acs.orglett.2c01702. [DOI] [PubMed] [Google Scholar]
  776. Huchenski B. S. N.; Robertson K. N.; Speed A. W. H. Functionalization of Bis-Diazaphospholene P-P Bonds with Diverse Electrophiles. Eur. J. Org. Chem. 2020, 2020, 5140–5144. 10.1002/ejoc.202000880. [DOI] [Google Scholar]
  777. Zhang J.; Yang J.-D.; Cheng J.-P. Diazaphosphinanes as hydride, hydrogen atom, proton or electron donors under transition-metal free conditions: thermodynamics, kinetics, and synthetic applications. Chem. Sci. 2020, 11, 3672–3679. 10.1039/C9SC05883D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  778. Zhang J.; Yang J.-D.; Cheng J.-P. Exploiting the radical reactivity of diazaphosphinanes in hydrodehalogenations and cascade cyclizations. Chem. Sci. 2020, 11, 4786–4790. 10.1039/D0SC01352H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  779. Riley R. D.; Huchenski B. S. N.; Bamford K. L.; Speed A. W. H. Diazaphospholene-Catalyzed Radical Reactions from Aryl Halides. Angew. Chem., Int. Ed. 2022, 61, e202204088 10.1002/anie.202204088. [DOI] [PubMed] [Google Scholar]
  780. Klett J.; Woźniak Ł.; Cramer N. 1,3,2-Diazaphospholene-Catalyzed Reductive Cyclizations of Organohalides. Angew. Chem., Int. Ed. 2022, 61, e202202306 10.1002/anie.202202306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  781. Liu Y.; Li H.; Tan E. Y. K.; Santiko E. B.; Chitose Y.; Abe M.; Chiba S. Pyrrolo[2,1-a]isoquinolines as multitasking organophotocatalysts in chemical synthesis. Chem. Catal. 2022, 2, 2726–2749. 10.1016/j.checat.2022.08.013. [DOI] [Google Scholar]
  782. Fu Y.; Liu X.; Xia Y.; Guo X.; Guo J.; Zhang J.; Zhao W.; Wu Y.; Wang J.; Zhong F. Whole-cell-catalyzed hydrogenation/deuteration of aryl halides with a genetically repurposed photodehalogenase. Chem. 2023, 9, 1897–1909. 10.1016/j.chempr.2023.03.006. [DOI] [Google Scholar]
  783. Dey D.; Kundu A.; Mandal B.; Roy M.; Adhikari D. Deciphering the single electron transfer ability of fluorene under photoredox conditions. Catal. Sci. Technol. 2022, 12, 7322–7327. 10.1039/D2CY01460B. [DOI] [Google Scholar]
  784. Pathania V.; Roy V. J.; Roy S. R. Transforming Non-innocent Phenalenyl to a Potent Photoreductant: Captivating Reductive Functionalization of Aryl Halides through Visible-Light-Induced Electron Transfer Processes. J. Org. Chem. 2022, 87, 16550–16566. 10.1021/acs.joc.2c02241. [DOI] [PubMed] [Google Scholar]
  785. Maust M. C.; Blakey S. B. Photoredox-Driven Three-Component Coupling of Aryl Halides, Olefins, and O2. ACS Catal. 2024, 14, 2582–2587. 10.1021/acscatal.3c05988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  786. Yang F.; He G.-C.; Sun S.-H.; Song T.-T.; Min X.-T.; Ji D.-W.; Guo S.-Y.; Chen Q.-A. Selective C-S Bond Constructions Using Inorganic Sulfurs via Photoinduced Electron Donor-Acceptor Activation. J. Org. Chem. 2022, 87, 14241–14249. 10.1021/acs.joc.2c01750. [DOI] [PubMed] [Google Scholar]
  787. König B. Photocatalysis in Organic Synthesis - Past, Present, and Future. Eur. J. Org. Chem. 2017, 2017, 1979–1981. 10.1002/ejoc.201700420. [DOI] [Google Scholar]
  788. Srivastava V.; Singh P. K.; Singh P. P. Recent advances of visible-light photocatalysis in the functionalization of organic compounds. J. Photochem. Photobiol. C: Photochem. Rev. 2022, 50, 100488. 10.1016/j.jphotochemrev.2022.100488. [DOI] [Google Scholar]
  789. Albano G.; Punzi A.; Capozzi M. A. M.; Farinola G. M. Sustainable protocols for direct C-H bond arylation of (hetero)arenes. Green Chem. 2022, 24, 1809–1894. 10.1039/D1GC03168F. [DOI] [Google Scholar]
  790. Torregrosa-Chinillach A.; Chinchilla R. Visible Light-Induced Aerobic Oxidative Dehydrogenation of C-N/C-O to C = N/C = O Bonds Using Metal-Free Photocatalysts: Recent Developments. Molecules 2022, 27, 497. 10.3390/molecules27020497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  791. Buglioni L.; Raymenants F.; Slattery A.; Zondag S. D. A.; Noël T. Technological Innovations in Photochemistry for Organic Synthesis: Flow Chemistry, High-Throughput Experimentation, Scale-up, and Photoelectrochemistry. Chem. Rev. 2022, 122, 2752–2906. 10.1021/acs.chemrev.1c00332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  792. Braga F. C.; Martins G. M.; Franco M. S.; Belli L. P. A.; de Oliveira K. T.; de Assis F. F. Catalyst-free photoarylation reactions promoted by visible light. Org. Chem. Front. 2024, 11, 1251–1275. 10.1039/D3QO01738A. [DOI] [Google Scholar]
  793. Singh F. V.; Wirth T. Hypervalent iodine chemistry and light: photochemical reactions involving hypervalent iodine Chemistry. Arkivoc 2022, 2021, 12–47. 10.24820/ark.5550190.p011.483. [DOI] [Google Scholar]
  794. Wang L.; Liu J. Synthetic Applications of Hypervalent Iodine(III) Reagents Enabled by Visible Light Photoredox Catalysis. Eur. J. Org. Chem. 2016, 2016, 1813–1824. 10.1002/ejoc.201501490. [DOI] [Google Scholar]
  795. Sun Y.; Song J.; Qin Q.; Zhang E.; Han Q.; Yang S.; Wang Z.; Yue S.; Dong D. Chin. J. Org. Chem. 2021, 41, 4651–4660. 10.6023/cjoc202106006. [DOI] [Google Scholar]
  796. Narobe R.; König B. Transformations based on direct excitation of hypervalent iodine(III) reagents. Org. Chem. Front. 2023, 10, 1577–1586. 10.1039/D3QO00039G. [DOI] [Google Scholar]
  797. Crivello J. V.Diaryliodonium Salt Photoacid Generators. In Iodine Chemistry and Applications; Kaiho T., Ed.; 2014. [Google Scholar]
  798. Vaish A.; Tsarevsky N. V.. Hypervalent Iodine Compounds in Polymer Science and Technology. In Main Group Strategies towards Functional Hybrid Materials; Baumgartner T.; Jäkle F., Eds.; 2017. [Google Scholar]
  799. Topa M.; Ortyl J. Moving Towards a Finer Way of Light-Cured Resin-Based Restorative Dental Materials: Recent Advances in Photoinitiating Systems Based on Iodonium Salts. Materials 2020, 13, 4093. 10.3390/ma13184093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  800. Zhang J.; Xiao P.; Dietlin C.; Campolo D.; Dumur F.; Gigmes D.; Morlet-Savary F.; Fouassier J.-P.; Lalevée J. Cationic Photoinitiators for Near UV and Visible LEDs: A Particular Insight into One-Component Systems. Macromol. Chem. Phys. 2016, 217, 1214–1227. 10.1002/macp.201500546. [DOI] [Google Scholar]
  801. Crivello J. V.; Reichmanis E. Photopolymer Materials and Processes for Advanced Technologies. Chem. Mater. 2014, 26, 533–548. 10.1021/cm402262g. [DOI] [Google Scholar]
  802. Giacoletto N.; Ibrahim-Ouali M.; Dumur F. Recent advances on squaraine-based photoinitiators of polymerization. Eur. Polym. J. 2021, 150, 110427. 10.1016/j.eurpolymj.2021.110427. [DOI] [Google Scholar]
  803. Malik M. S.; Schlögl S.; Wolfahrt M.; Sangermano M. Review on UV-Induced Cationic Frontal Polymerization of Epoxy Monomers. Polymers 2020, 12, 2146. 10.3390/polym12092146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  804. Dumur F. Recent advances on diaryliodonium-based monocomponent photoinitiating systems. Eur. Polym. J. 2023, 195, 112193. 10.1016/j.eurpolymj.2023.112193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  805. 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]
  806. Chandler C.; Porcincula D. H.; Ford M. J.; Hook C.; Zhang X.; Brodsky J.; Sellinger A. Plastic Scintillators via Rapid Photoinitiated Cationic Polymerizationof Vinyltoluene. ACS Appl. Polym. Mater. 2022, 4, 4069–4074. 10.1021/acsapm.2c00316. [DOI] [Google Scholar]
  807. Petko F.; Galek M.; Hola E.; Topa-Skwarczyńska M.; Tomal W.; Jankowska M.; Pilch M.; Popielarz R.; Graff B.; Morlet-Savary F.; Lalevee J.; Ortyl J. Symmetric Iodonium Salts Based on Benzylidene as One Component Photoinitiators for Applications in 3D Printing. Chem. Mater. 2022, 34, 10077–10092. 10.1021/acs.chemmater.2c02796. [DOI] [Google Scholar]
  808. Wang Q.; Popov S.; Feilen A.; Strehmel V.; Strehmel B. Rational Selection of Cyanines to Generate Conjugate Acid and Free Radicals for Photopolymerization upon Exposure at 860nm. Angew. Chem., Int. Ed. 2021, 60, 26855–26865. 10.1002/anie.202108713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  809. Radchenko A. V.; Duchet-Rumeau J.; Gérard J.-F.; Baudoux J.; Livi S. Cycloaliphatic epoxidized ionic liquids as new versatile monomers for the development of shape memory PIL networks by 3D printing. Polym. Chem. 2020, 11, 5475–5483. 10.1039/D0PY00704H. [DOI] [Google Scholar]
  810. Tobisu M.; Furukawa T.; Chatani N. Visible Light-mediated Direct Arylation of Arenes and Heteroarenes Using Diaryliodonium Salts in the Presence and Absence of a Photocatalyst. Chem. Lett. 2013, 42, 1203–1205. 10.1246/cl.130547. [DOI] [Google Scholar]
  811. Shaw R.; Sihag N.; Jain S.; Sharma R.; Yadav M. R. Photoinduced Alkyl/Aryl Radical Cascade for the Synthesis of Quaternary CF3-Containing Oxindoles and Indoline Alkaloids. J. Org. Chem. 2023, 88, 5652–5660. 10.1021/acs.joc.3c00141. [DOI] [PubMed] [Google Scholar]
  812. Chen Z.; Liu N.-W.; Bolte M.; Ren H.; Manolikakes G. Visible-light mediated 3-component synthesis of sulfonylated coumarins from sulfur dioxide. Green Chem. 2018, 20, 3059–3070. 10.1039/C8GC00838H. [DOI] [Google Scholar]
  813. Liu N.-W.; Chen Z.; Herbert A.; Ren H.; Manolikakes G. Visible-Light-Induced 3-Component Synthesis of Sulfonylated Oxindoles by Fixation of Sulfur Dioxide. Eur. J. Org. Chem. 2018, 2018, 5725–5734. 10.1002/ejoc.201801128. [DOI] [Google Scholar]
  814. Nair A. M.; Halder I.; Khan S.; Volla C. M. R. Metal Free Sulfonylative Spirocyclization of Alkenyl and Alkynyl Amides via Insertion of Sulfur Dioxide. Adv. Synth. Catal. 2020, 362, 224–229. 10.1002/adsc.201901321. [DOI] [Google Scholar]
  815. Nair A. M.; Halder I.; Volla C. M. R. A metal-free four-component sulfonylation, Giese cyclization, selenylation cascade via insertion of sulfur dioxide. Chem. Commun. 2022, 58, 6950–6953. 10.1039/D2CC02315F. [DOI] [PubMed] [Google Scholar]
  816. Bugaenko D. I.; Volkov A. A.; Livantsov M. V.; Yurovskaya M. A.; Karchava A. V. Catalyst-Free Arylation of Tertiary Phosphines with Diaryliodonium Salts Enabled by Visible Light. Chem.—Eur. J. 2019, 25, 12502–12506. 10.1002/chem.201902955. [DOI] [PubMed] [Google Scholar]
  817. Bugaenko D. I.; Karchava A. V. Electron Donor-Acceptor Complex Initiated Photochemical Phosphorus Arylation with Diaryliodonium Salts toward the Synthesis of Phosphine Oxides. Adv. Synth. Catal. 2023, 365, 1893–1900. 10.1002/adsc.202300351. [DOI] [Google Scholar]
  818. Bugaenko D. I.; Karchava A. V. Catalyst-Free Visible Light Mediated Synthesis of Unsymmetrical Tertiary Arylphosphines. Adv. Synth. Catal. 2022, 364, 2248–2253. 10.1002/adsc.202200309. [DOI] [Google Scholar]
  819. Lecroq W.; Bazille P.; Morlet-Savary F.; Breugst M.; Lalevée J.; Gaumont A.-C.; Lakhdar S. Visible-Light-Mediated Metal-Free Synthesis of Aryl Phosphonates: Synthetic and Mechanistic Investigations. Org. Lett. 2018, 20, 4164–4167. 10.1021/acs.orglett.8b01379. [DOI] [PubMed] [Google Scholar]
  820. Liu N.-W.; Liang S.; Manolikakes G. Visible-Light Photoredox-Catalyzed Aminosulfonylation of Diaryliodonium Salts with Sulfur Dioxide and Hydrazines. Adv. Synth. Catal. 2017, 359, 1308–1319. 10.1002/adsc.201601341. [DOI] [Google Scholar]
  821. Sun D.; Yin K.; Zhang R. Visible-light-induced multicomponent cascade cycloaddition involving N-propargyl aromatic amines, diaryliodonium salts and sulfur dioxide: rapid access to 3-arylsulfonylquinolines. Chem. Commun. 2018, 54, 1335–1338. 10.1039/C7CC09410H. [DOI] [PubMed] [Google Scholar]
  822. Breton-Patient C.; Naud-Martin D.; Mahuteau-Betzer F.; Piguel S. Three-Component C-H Bond Sulfonylation of Imidazoheterocycles by Visible-Light Organophotoredox Catalysis. Eur. J. Org. Chem. 2020, 2020, 6653–6660. 10.1002/ejoc.202001219. [DOI] [Google Scholar]
  823. Ma Y.; Pan Q.; Ou C.; Cai Y.; Ma X.; Liu C. Aryl sulfonyl fluoride synthesis via organophotocatalytic fluorosulfonylation of diaryliodonium salts. Org. Biomol. Chem. 2023, 21, 7597–7601. 10.1039/D3OB01200J. [DOI] [PubMed] [Google Scholar]
  824. Senapati S.; Parida S. K.; Karandikar S. S.; Murarka S. Organophotoredox-Catalyzed Arylation and Aryl Sulfonylation of Morita-Baylis-Hillman Acetates with Diaryliodonium Reagents. Org. Lett. 2023, 25, 7900–7905. 10.1021/acs.orglett.3c03146. [DOI] [PubMed] [Google Scholar]
  825. Li D.; Liang C.; Jiang Z.; Zhang J.; Zhuo W.-T.; Zou F.-Y.; Wang W.-P.; Gao G.-L.; Song J. Visible-Light-Promoted C2 Selective Arylation of Quinoline and Pyridine N-Oxides with Diaryliodonium Tetrafluoroborate. J. Org. Chem. 2020, 85, 2733–2742. 10.1021/acs.joc.9b02933. [DOI] [PubMed] [Google Scholar]
  826. Camacho M. B.; Clark A. E.; Liebrecht T. A.; DeLuca J. P. A Phenyliodonium Ylide as a Precursor for Dicarboethoxycarbene: Demonstration of a Strategy for Carbene Generation. J. Am. Chem. Soc. 2000, 122, 5210–5211. 10.1021/ja000334o. [DOI] [Google Scholar]
  827. Chidley T.; Jameel I.; Rizwan S.; Peixoto P. A.; Pouysegu L.; Quideau S.; Hopkins W. S.; Murphy G. K. Blue LED Irradiation of Iodonium Ylides Gives Diradical Intermediates for Efficient Metal-free Cyclopropanation with Alkenes. Angew. Chem., Int. Ed. 2019, 58, 16959–16965. 10.1002/anie.201908994. [DOI] [PubMed] [Google Scholar]
  828. Mi X.; Pi C.; Feng W.; Cui X. Recent progress in the application of iodonium ylides in organic synthesis. Org. Chem. Front. 2022, 9, 6999–7015. 10.1039/D2QO01332K. [DOI] [Google Scholar]
  829. Ren J.; Pi C.; Cui X.; Wu Y. Divergent C(sp2)-H arylation of heterocycles via organic photoredox catalysis. Green Chem. 2022, 24, 3017–3022. 10.1039/D1GC04825B. [DOI] [Google Scholar]
  830. Chen J.-Y.; Wu H.-Y.; Song H.-Y.; Li H.-X.; Jiang J.; Yang T.-B.; He W.-M. Visible-Light-Induced Annulation of Iodonium Ylides and 2-Isocyanobiaryls to Access 6-Arylated Phenanthridines. J. Org. Chem. 2023, 88, 8360–8368. 10.1021/acs.joc.3c00380. [DOI] [PubMed] [Google Scholar]
  831. Meher P.; Panda S. P.; Mahapatra S. K.; Thombare K. R.; Roy L.; Murarka S. A General Electron Donor-Acceptor Photoactivation Platform of Diaryliodonium Reagents: Arylation of Heterocycles. Org. Lett. 2023, 25, 8290–8295. 10.1021/acs.orglett.3c03365. [DOI] [PubMed] [Google Scholar]
  832. Samanta R. K.; Meher P.; Murarka S. Visible Light Photoredox-Catalyzed Direct C-H Arylation of Quinoxalin-2(1H)-ones with Diaryliodonium Salts. J. Org. Chem. 2022, 87, 10947–10957. 10.1021/acs.joc.2c01234. [DOI] [PubMed] [Google Scholar]
  833. Pham P. H.; Petersen H. A.; Katsirubas J. L.; Luca O. R. Recent synthetic methods involving carbon radicals generated by electrochemical catalysis. Org. Biomol. Chem. 2022, 20, 5907–5932. 10.1039/D2OB00424K. [DOI] [PubMed] [Google Scholar]
  834. Meyer T. H.; Choi I.; Tian C.; Ackermann L. Powering the Future: How Can Electrochemistry Make a Difference in Organic Synthesis?. Chem. 2020, 6, 2484–2496. 10.1016/j.chempr.2020.08.025. [DOI] [Google Scholar]
  835. Jiang Y.; Xu K.; Zeng C. Use of Electrochemistry in the Synthesis of Heterocyclic Structures. Chem. Rev. 2018, 118, 4485–4540. 10.1021/acs.chemrev.7b00271. [DOI] [PubMed] [Google Scholar]
  836. Moniruzzaman M.; Afrin S.; Ali Md. K. Progress in the Electrochemical Synthesis of Organoboron Compounds. Asian J. Org. Chem. 2023, 12, e202300090 10.1002/ajoc.202300090. [DOI] [Google Scholar]
  837. Xiong P.; Xu H.-C. Chemistry with Electrochemically Generated N-Centered Radicals. Acc. Chem. Res. 2019, 52, 3339–3350. 10.1021/acs.accounts.9b00472. [DOI] [PubMed] [Google Scholar]
  838. Francke R.; Little R. D. Redox catalysis in organic electrosynthesis: basic principles and recent developments. Chem. Soc. Rev. 2014, 43, 2492–2521. 10.1039/c3cs60464k. [DOI] [PubMed] [Google Scholar]
  839. Feng R.; Smith J. A.; Moeller K. D. Anodic Cyclization Reactions and the Mechanistic Strategies That Enable Optimization. Acc. Chem. Res. 2017, 50, 2346–2352. 10.1021/acs.accounts.7b00287. [DOI] [PubMed] [Google Scholar]
  840. Ma C.; Fang P.; Mei T.-S. Recent Advances in C-H Functionalization Using Electrochemical Transition Metal Catalysis. ACS Catal. 2018, 8, 7179–7189. 10.1021/acscatal.8b01697. [DOI] [Google Scholar]
  841. Yoshida J.-i.; Shimizu A.; Hayashi R. Electrogenerated Cationic Reactive Intermediates: The Pool Method and Further Advances. Chem. Rev. 2018, 118, 4702–4730. 10.1021/acs.chemrev.7b00475. [DOI] [PubMed] [Google Scholar]
  842. Wiebe A.; Gieshoff T.; Möhle S.; Rodrigo E.; Zirbes M.; Waldvogel S. R. Electrifying Organic Synthesis. Angew. Chem., Int. Ed. 2018, 57, 5594–5619. 10.1002/anie.201711060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  843. Xiang H.; He J.; Qian W.; Qiu M.; Xu H.; Duan W.; Ouyang Y.; Wang Y.; Zhu C. Electroreductively Induced Radicals for Organic Synthesis. Molecules 2023, 28, 857. 10.3390/molecules28020857. [DOI] [PMC free article] [PubMed] [Google Scholar]
  844. Mahanty K.; Halder A.; Maiti D.; De Sarkar S. Recent Developments in the Electroreductive Functionalization of Carbon-Halogen Bonds. Synthesis 2023, 55, 400–416. 10.1055/a-1944-9494. [DOI] [Google Scholar]
  845. Huang B.; Sun Z.; Sun G. Recent progress in cathodic reduction-enabled organic electrosynthesis: Trends, challenges, and opportunities. eScience 2022, 2, 243–277. 10.1016/j.esci.2022.04.006. [DOI] [Google Scholar]
  846. Yan M.; Kawamata Y.; Baran P. S. Synthetic Organic Electrochemical Methods Since 2000: On the Verge of a Renaissance. Chem. Rev. 2017, 117, 13230–13319. 10.1021/acs.chemrev.7b00397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  847. Tang S.; Liu Y.; Lei A. Electrochemical Oxidative Cross-coupling with Hydrogen Evolution: A Green and Sustainable Way for Bond Formation. Chem. 2018, 4, 27–45. 10.1016/j.chempr.2017.10.001. [DOI] [Google Scholar]
  848. Mazzucato M.; Isse A. A.; Durante C. Dissociative electron transfer mechanism and application in the electrocatalytic activation of organic halides. Curr. Opin. Electrochem. 2023, 39, 101254. 10.1016/j.coelec.2023.101254. [DOI] [Google Scholar]
  849. Novaes L. F. T.; Liu J.; Shen Y.; Lu L.; Meinhardt J. M.; Lin S. Electrocatalysis as an enabling technology for organic synthesis. Chem. Soc. Rev. 2021, 50, 7941–8002. 10.1039/D1CS00223F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  850. Zhu C.; Ang N. W. J.; Meyer T. H.; Qiu Y.; Ackermann L. Organic Electrochemistry: Molecular Syntheses with Potential. ACS Cent. Sci. 2021, 7, 415–431. 10.1021/acscentsci.0c01532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  851. Yuan Y.; Lei A. Electrochemical Oxidative Cross-Coupling with Hydrogen Evolution Reactions. Acc. Chem. Res. 2019, 52, 3309–3324. 10.1021/acs.accounts.9b00512. [DOI] [PubMed] [Google Scholar]
  852. Horn E. J.; Rosen B. R.; Baran P. S. Synthetic Organic Electrochemistry: An Enabling and Innately Sustainable Method. ACS Cent. Sci. 2016, 2, 302–308. 10.1021/acscentsci.6b00091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  853. Möhle S.; Zirbes M.; Rodrigo E.; Gieshoff T.; Wiebe A.; Waldvogel S. R. Modern Electrochemical Aspects for the Synthesis of Value-Added Organic Products. Angew. Chem., Int. Ed. 2018, 57, 6018–6041. 10.1002/anie.201712732. [DOI] [PMC free article] [PubMed] [Google Scholar]
  854. Waldvogel S. R.; Lips S.; Selt M.; Riehl B.; Kampf C. J. Electrochemical Arylation Reaction. Chem. Rev. 2018, 118, 6706–6765. 10.1021/acs.chemrev.8b00233. [DOI] [PubMed] [Google Scholar]
  855. Yang L.; Zhuang Q.; Wu M.; Long H.; Lin C.; Lin M.; Ke F. Electrochemical-induced hydroxylation of aryl halides in the presence of Et3N in water. Org. Biomol. Chem. 2021, 19, 6417–6421. 10.1039/D1OB00931A. [DOI] [PubMed] [Google Scholar]
  856. Ke J.; Wang H.; Zhou L.; Mou C.; Zhang J.; Pan L.; Chi Y. R. Hydrodehalogenation of Aryl Halides through Direct Electrolysis. Chem.—Eur. J. 2019, 25, 6911–6914. 10.1002/chem.201901082. [DOI] [PubMed] [Google Scholar]
  857. Huang B.; Guo L.; Xia W. A facile and versatile electro-reductive system for hydrodefunctionalization under ambient conditions. Green Chem. 2021, 23, 2095–2103. 10.1039/D1GC00317H. [DOI] [Google Scholar]
  858. Lu L.; Li H.; Zheng Y.; Bu F.; Lei A. Facile and Economical Electrochemical Dehalogenative Deuteration of (Hetero)Aryl Halides. CCS Chem. 2021, 3, 2669–2675. 10.31635/ccschem.020.202000512. [DOI] [Google Scholar]
  859. Liu C.; Han S.; Li M.; Chong X.; Zhang B. Electrocatalytic Deuteration of Halides with D2O as the Deuterium Source over a Copper Nanowire Arrays Cathode. Angew. Chem., Int. Ed. 2020, 59, 18527–18531. 10.1002/anie.202009155. [DOI] [PubMed] [Google Scholar]
  860. Hong J.; Liu Q.; Li F.; Bai G.; Liu G.; Li M.; Nayal O. S.; Fu X.; Mo F. Electrochemical Radical Borylation of Aryl Iodides. Chin. J. Chem. 2019, 37, 347–351. 10.1002/cjoc.201900001. [DOI] [Google Scholar]
  861. Wang S.; Yang C.; Sun S.; Wang J. Catalyst-free phosphorylation of aryl halides with trialkyl phosphites through electrochemical reduction. Chem. Commun. 2019, 55, 14035–14038. 10.1039/C9CC07069A. [DOI] [PubMed] [Google Scholar]
  862. Sengmany S.; Ollivier A.; Le Gall E.; Léonel E. A mild electroassisted synthesis of (hetero)arylphosphonates. Org. Biomol. Chem. 2018, 16, 4495–4500. 10.1039/C8OB00500A. [DOI] [PubMed] [Google Scholar]
  863. Bai Y.; Liu N.; Wang S.; Wang S.; Ning S.; Shi L.; Cui L.; Zhang Z.; Xiang J. Nickel-Catalyzed Electrochemical Phosphorylation of Aryl Bromides. Org. Lett. 2019, 21, 6835–6838. 10.1021/acs.orglett.9b02475. [DOI] [PubMed] [Google Scholar]
  864. Kurono N.; Honda E.; Komatsu F.; Orito K.; Tokuda M. Regioselective Radical Cyclization by Electrochemical Reduction Using an Arene Mediator. Environmentally Benign Method. Chem. Lett. 2003, 32, 720–721. 10.1246/cl.2003.720. [DOI] [Google Scholar]
  865. Kurono N.; Honda E.; Komatsu F.; Orito K.; Tokuda M. Regioselective synthesis of substituted 1-indanols, 2,3-dihydrobenzofurans and 2,3-dihydroindoles by electrochemical radical cyclization using an arene mediator. Tetrahedron 2004, 60, 1791–1801. 10.1016/j.tet.2003.12.038. [DOI] [Google Scholar]
  866. Mitsudo K.; Nakagawa Y.; Mizukawa J.-i.; Tanaka H.; Akaba R.; Okada T.; Suga S. Electro-reductive cyclization of aryl halides promoted by fluorene derivatives. Electrochim. Acta 2012, 82, 444–449. 10.1016/j.electacta.2012.03.130. [DOI] [Google Scholar]
  867. Mitsudo K.; Okada T.; Shimohara S.; Mandai H.; Suga S. Electro-reductive Halogen-Deuterium Exchange and Methylation of Aryl Halides in Acetonitrile. Electrochemistry 2013, 81, 362–364. 10.5796/electrochemistry.81.362. [DOI] [Google Scholar]
  868. Sun G.; Ren S.; Zhu X.; Huang M.; Wan Y. Direct Arylation of Pyrroles via Indirect Electroreductive C-H Functionalization Using Perylene Bisimide as an Electron-Transfer Mediator. Org. Lett. 2016, 18, 544–547. 10.1021/acs.orglett.5b03581. [DOI] [PubMed] [Google Scholar]
  869. Folgueiras-Amador A. A.; Teuten A. E.; Salam-Perez M.; Pearce J. E.; Denuault G.; Pletcher D.; Parsons P. J.; Harrowven D. C.; Brown R. C. D. Cathodic Radical Cyclisation of Aryl Halides Using a Strongly-Reducing Catalytic Mediator in Flow. Angew. Chem., Int. Ed. 2022, 61, e202203694 10.1002/anie.202203694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  870. Yu W.; Wang S.; He M.; Jiang Z.; Yu Y.; Lan J.; Luo J.; Wang P.; Qi X.; Wang T.; Lei A. Electroreduction Enables Regioselective 1,2-Diarylation of Alkenes with Two Electrophiles. Angew. Chem., Int. Ed. 2023, 62, e202219166 10.1002/anie.202219166. [DOI] [PubMed] [Google Scholar]
  871. Senboku H.; Michinishi J.-y.; Hara S. Facile Synthesis of 2,3-Dihydrobenzofuran-3-ylacetic Acids by Novel Electrochemical Sequential Aryl Radical Cyclization-Carboxylation of 2-Allyloxy-bromobenzenes Using Methyl 4-tert-Butylbenzoate as an Electron-Transfer Mediator. Synlett 2011, 2011, 1567–1572. 10.1055/s-0030-1260794. [DOI] [Google Scholar]
  872. Katayama A.; Senboku H.; Hara S. Aryl radical cyclization with alkyne followed by tandem carboxylation in methyl 4-tert-butylbenzoate-mediated electrochemical reduction of 2-(2-propynyloxy)bromobenzenes in the presence of carbon dioxide. Tetrahedron 2016, 72, 4626–4636. 10.1016/j.tet.2016.06.032. [DOI] [Google Scholar]
  873. Wang Y.; Zhao Z.; Pan D.; Wang S.; Jia K.; Ma D.; Yang G.; Xue X.-S.; Qiu Y. Metal-Free Electrochemical Carboxylation of Organic Halides in the Presence of Catalytic Amounts of an Organomediator. Angew. Chem., Int. Ed. 2022, 61, e202210201 10.1002/anie.202210201. [DOI] [PubMed] [Google Scholar]
  874. Huang H.; Steiniger K. A.; Lambert T. H. Electrophotocatalysis: Combining Light and Electricity to Catalyze Reactions. J. Am. Chem. Soc. 2022, 144, 12567–12583. 10.1021/jacs.2c01914. [DOI] [PubMed] [Google Scholar]
  875. Yu Y.; Guo P.; Zhong J.-S.; Yuan Y.; Ye K.-Y. Merging photochemistry with electrochemistry in organic synthesis. Org. Chem. Front. 2020, 7, 131–135. 10.1039/C9QO01193E. [DOI] [Google Scholar]
  876. Capaldo L.; Quadri L. L.; Ravelli D. Merging Photocatalysis with Electrochemistry: The Dawn of a new Alliance in Organic Synthesis. Angew. Chem., Int. Ed. 2019, 58, 17508–17510. 10.1002/anie.201910348. [DOI] [PubMed] [Google Scholar]
  877. Chernowsky C. P.; Chmiel A. F.; Wickens Z. K. Electrochemical Activation of Diverse Conventional Photoredox Catalysts Induces Potent Photoreductant Activity. Angew. Chem., Int. Ed. 2021, 60, 21418–21425. 10.1002/anie.202107169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  878. Barham J. P.; König B. Synthetic Photoelectrochemistry. Angew. Chem., Int. Ed. 2020, 59, 11732–11747. 10.1002/anie.201913767. [DOI] [PMC free article] [PubMed] [Google Scholar]
  879. Hardwick T.; Ahmed N. C-H Functionalization via Electrophotocatalysis and Photoelectrochemistry: Complementary Synthetic Approach. ACS Sustainable Chem. Eng. 2021, 9, 4324–4340. 10.1021/acssuschemeng.0c08434. [DOI] [Google Scholar]
  880. Wu S.; Kaur J.; Karl T. A.; Tian X.; Barham J. P. Synthetic Molecular Photoelectrochemistry: New Frontiers in Synthetic Applications, Mechanistic Insights and Scalability. Angew. Chem., Int. Ed. 2022, 61, e202107811 10.1002/anie.202107811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  881. Liu J.; Lu L.; Wood D.; Lin S. New Redox Strategies in Organic Synthesis by Means of Electrochemistry and Photochemistry. ACS Cent. Sci. 2020, 6, 1317–1340. 10.1021/acscentsci.0c00549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  882. Capaldo L.; Ravelli D. The Dark Side of Photocatalysis: One Thousand Ways to Close the Cycle. Eur. J. Org. Chem. 2020, 2020, 2783–2806. 10.1002/ejoc.202000144. [DOI] [Google Scholar]
  883. Caby S.; Bouchet L. M.; Argüello J. E.; Rossi R. A.; Bardagi J. I. Excitation of Radical Anions of Naphthalene Diimides in Consecutive- and Electro-Photocatalysis. ChemCatChem. 2021, 13, 3001–3009. 10.1002/cctc.202100359. [DOI] [Google Scholar]
  884. Medici F.; Chiroli V.; Raimondi L.; Benaglia M. Latest updates in ElectroPhotoChemical reactions. Tetrahedron Chem. 2024, 9, 100061. 10.1016/j.tchem.2023.100061. [DOI] [Google Scholar]
  885. Qian L.; Shi M. Contemporary photoelectrochemical strategies and reactions in organic synthesis. Chem. Commun. 2023, 59, 3487–3506. 10.1039/D3CC00437F. [DOI] [PubMed] [Google Scholar]
  886. Ghosh I.; Ghosh T.; Bardagi J. I.; König B. Reduction of aryl halides by consecutive visible light-induced electron transfer processes. Science 2014, 346, 725–728. 10.1126/science.1258232. [DOI] [PubMed] [Google Scholar]
  887. Marchini M.; Gualandi A.; Mengozzi L.; Franchi P.; Lucarini M.; Cozzi P. G.; Balzani V.; Ceroni P. Mechanistic insights into two-photon-driven photocatalysis in organic synthesis. Phys. Chem. Chem. Phys. 2018, 20, 8071–8076. 10.1039/C7CP08011E. [DOI] [PubMed] [Google Scholar]
  888. Schmalzbauer M.; Ghosh I.; Konig B. Utilising excited state organic anions for photoredox catalysis: activation of (hetero) aryl chlorides by visible light-absorbing 9-anthrolate anions. Faraday Discuss. 2019, 215, 364–378. 10.1039/C8FD00176F. [DOI] [PubMed] [Google Scholar]
  889. Ghosh I.; König B. Chromoselective Photocatalysis: Controlled Bond Activation through Light-Color Regulation of Redox Potentials. Angew. Chem., Int. Ed. 2016, 55, 7676–7679. 10.1002/anie.201602349. [DOI] [PubMed] [Google Scholar]
  890. Gong H. X.; Cao Z.; Li M. H.; Liao S. H.; Lin M. J. Photoexcited perylene diimide radical anions for the reduction of aryl halides: a bay-substituent effect. Org. Chem. Front. 2018, 5, 2296–2302. 10.1039/C8QO00445E. [DOI] [Google Scholar]
  891. Zeman C. J. IV; Kim S.; Zhang F.; Schanze K. S. Direct Observation of the Reduction of Aryl Halides by a Photoexcited Perylene Diimide Radical Anion. J. Am. Chem. Soc. 2020, 142, 2204–2207. 10.1021/jacs.9b13027. [DOI] [PubMed] [Google Scholar]
  892. Horsewill S. J.; Hierlmeier G.; Farasat Z.; Barham J. P.; Scott D. J. Shining Fresh Light on Complex Photoredox Mechanisms through Isolation of Intermediate Radical Anions. ACS Catal. 2023, 13, 9392–9403. 10.1021/acscatal.3c02515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  893. Yang Y.-Y.; Zhang P.; Hadjichristidis N. Two-Photon Excitation Photoredox Catalysis Enabled Atom Transfer Radical Polymerization. J. Am. Chem. Soc. 2023, 145, 12737–12744. 10.1021/jacs.3c02832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  894. Kim D.; Teets T. S. Strategies for accessing photosensitizers with extreme redox potentials. Chem. Phys. Rev. 2022, 3, 021302. 10.1063/5.0084554. [DOI] [Google Scholar]
  895. Lee Y.-M.; Nam W.; Fukuzumi S. Redox catalysis via photoinduced electron transfer. Chem. Sci. 2023, 14, 4205–4218. 10.1039/D2SC07101K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  896. Lepori M.; Schmid S.; Barham J. P. Photoredox catalysis harvesting multiple photon or electrochemical energies. Beilstein J. Org. Chem. 2023, 19, 1055–1145. 10.3762/bjoc.19.81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  897. Kim H.; Kim H.; Lambert T. H.; Lin S. Reductive Electrophotocatalysis: Merging Electricity and Light to Achieve Extreme Reduction Potentials. J. Am. Chem. Soc. 2020, 142, 2087–2092. 10.1021/jacs.9b10678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  898. Cowper N. G. W.; Chernowsky C. P.; Williams O. P.; Wickens Z. K. Potent Reductants via Electron-Primed Photoredox Catalysis: Unlocking Aryl Chlorides for Radical Coupling. J. Am. Chem. Soc. 2020, 142, 2093–2099. 10.1021/jacs.9b12328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  899. Chen Y.-J.; Lei T.; Hu H.-L.; Wu H.-L.; Zhou S.; Li X.-B.; Chen B.; Tung C.-H.; Wu L.-Z. Tandem photoelectrochemical and photoredox catalysis for efficient and selective aryl halides functionalization by solar energy. Matter 2021, 4, 2354–2366. 10.1016/j.matt.2021.05.004. [DOI] [Google Scholar]
  900. Zeng L.; Qin J.-H.; Lv G.-F.; Hu M.; Sun Q.; Ouyang X.-H.; He D.-L.; Li J.-H. Electrophotocatalytic Reductive 1,2-Diarylation of Alkenes with Aryl Halides and Cyanoaromatics. Chin. J. Chem. 2023, 41, 1921–1930. 10.1002/cjoc.202300174. [DOI] [Google Scholar]
  901. Kang W.-J.; Zhang Y.; Li B.; Guo H. Electrophotocatalytic hydrogenation of imines and reductive functionalization of aryl halides. Nat. Commun. 2024, 15, 655. 10.1038/s41467-024-45015-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  902. Anastas P. T.; Warner J. C. In Green Chemistry: Theory and Practice; Oxford University Press, New York, 1998. [Google Scholar]
  903. Horváth I. T.; Anastas P. T. Introduction: Green Chemistry. Chem. Rev. 2007, 107, 2167–2168. 10.1021/cr0783784. [DOI] [PubMed] [Google Scholar]
  904. Anastas P. T.; Williamson T. C. In Green Chemistry: Designing Chemistry for the Environment; American Chemical Series Books, Washington, DC, 1996; pp 1–20. [Google Scholar]
  905. Anastas P. T.; Eghbali N. Green Chemistry: Principles and Practice. Chem. Soc. Rev. 2010, 39, 301–312. 10.1039/B918763B. [DOI] [PubMed] [Google Scholar]
  906. Roschangar F.; Sheldon R. A.; Senanayake C. H. Overcoming barriers to green chemistry in the pharmaceutical industry - the Green Aspiration LevelTM concept. Green Chem. 2015, 17, 752–768. 10.1039/C4GC01563K. [DOI] [Google Scholar]
  907. Sheldon R. A. Metrics of Green Chemstry and Sustainability: Past, Present, and Future. ACS Sustain. Green Chem. 2018, 6, 32–48. 10.1021/acssuschemeng.7b03505. [DOI] [Google Scholar]
  908. Trost B. M. The atom economy: a search for synthetic efficiency. Science 1991, 254, 1471–1477. 10.1126/science.1962206. [DOI] [PubMed] [Google Scholar]
  909. Trost B. M.; A Challenge for Organic Synthesis: Homogeneous Catalysis Leads the Way. Angew. Chem., Int. Ed. 1995, 34, 259–281. 10.1002/anie.199502591. [DOI] [Google Scholar]
  910. Sheldon R. A. Organic synthesis; past, present and future. Chem. Ind. 1992, 903–906. [Google Scholar]
  911. Sheldon R. A. The E factor: fifteen years on. Green Chem. 2007, 9, 1273–1283. 10.1039/b713736m. [DOI] [Google Scholar]
  912. Merkushev E. B. Advances in the Synthesis of Iodoaromatic Compounds. Synthesis 1988, 1988, 923–937. 10.1055/s-1988-27758. [DOI] [Google Scholar]
  913. Urch C. J. In Comprehensive Organic Functional Group Transformations Vol. 2; Ley S. V., Ed.; Pergamon Press, Oxford, 1995, Chapter 2.12. [Google Scholar]
  914. Hanson J. R. Advances in the direct iodination of aromatic compounds. J. Chem. Res. 2006, 2006, 277–280. 10.3184/030823406777410981. [DOI] [Google Scholar]
  915. Stavber S.; Jereb M.; Zupan M. Electrophilic Iodination of Organic Compounds Using Elemental Iodine or Iodides. Synthesis 2008, 2008, 1487–1513. 10.1055/s-2008-1067037. [DOI] [Google Scholar]
  916. Sy W.-H.; Lodge B. A. Iodination of alkylbenzenes with iodine and silver nitrite. Tetrahedron Lett. 1989, 30, 3769–3772. 10.1016/S0040-4039(01)80650-X. [DOI] [Google Scholar]
  917. Mulholland G. K.; Zheng Q.-H. A direct iodination method with iodine and silver triflate for the synthesis of SPECT and PET imaging agent precursors. Synth. Commun. 2001, 31, 3059–3068. 10.1081/SCC-100105877. [DOI] [Google Scholar]
  918. Barluenga J.; Campos P. J.; Gonzalez J. M.; Asensio G. A simple and general route to aryl iodides from arenes. J. Chem. Soc., Perkin Trans. 1 1984, 2623–2624. 10.1039/P19840002623. [DOI] [Google Scholar]
  919. Orito K.; Hatakeyama T.; Takeo M.; Suginome H. Iodination of Alkyl Aryl Ethers by Mercury(II) Oxide-Iodine Reagent in Dichloromethane. Synthesis 1995, 1995, 1273–1277. 10.1055/s-1995-4089. [DOI] [Google Scholar]
  920. Krassowska-Rwiebocka B.; Lulinski P.; Skulski L. Aromatic Iodination of Activated Arenes and Heterocycles with Lead Tetraacetate as the Oxidant. Synthesis 1995, 1995, 926–928. 10.1055/s-1995-4044. [DOI] [Google Scholar]
  921. Lulinski P.; Skulski L. The Direct Iodination of Arenes with Chromium(VI) Oxide as the Oxidant. Bull. Chem. Soc. Jpn. 1997, 70, 1665–1669. 10.1246/bcsj.70.1665. [DOI] [Google Scholar]
  922. Sugiyama T. A. Mild and Convenient Procedure for Conversion of Aromatic Compounds into Their Iodides Using Ammonium Hexanitratocerate(IV). Bull. Chem. Soc. Jpn. 1981, 54, 2847–2848. 10.1246/bcsj.54.2847. [DOI] [Google Scholar]
  923. Tilve R. D.; Alexander V. M.; Khadilkar B. M. Iodination of activated arenes using silfen: an improved protocol. Tetrahedron Lett. 2002, 43, 9457–9459. 10.1016/S0040-4039(02)02262-1. [DOI] [Google Scholar]
  924. Firouzabadi H; Iranpoor N; Shiri M Direct and regioselective iodination and bromination of benzene, naphthalene and other activated aromatic compounds using iodine and bromine or their sodium salts in the presence of the Fe(NO3)3·1.5N2O4/charcoal system. Tetrahedron Lett. 2003, 44, 8781–8785. 10.1016/j.tetlet.2003.09.189. [DOI] [Google Scholar]
  925. Baird W. C. Jr; Surridge J. H. Halogenation with copper(II) halides. Synthesis of aryl iodides. J. Org. Chem. 1970, 35, 3436–3442. 10.1021/jo00835a055. [DOI] [Google Scholar]
  926. Datta R. L.; Chatterjee N. R. HALOGENATION. XV. DIRECT IODINATION OF HYDROCARBONS BY MEANS OF IODINE AND NITRIC ACID. J. Am. Chem. Soc. 1917, 39, 435–441. 10.1021/ja02248a012. [DOI] [Google Scholar]
  927. Yusubov M. S.; Filimonov V. D.; Jin H.-W.; Chi K.-W. A Novel lodination of Aromatic Rings Using Iodine/Metallic Nitrate. Bull. Korean Chem. Soc. 1998, 19, 400–401. [Google Scholar]
  928. Yang S. G.; Kim Y. H. A practical iodination of aromatic compounds using tetrabutylammonium peroxydisulfate and iodine. Tetrahedron Lett. 1999, 40, 6051–6054. 10.1016/S0040-4039(99)01236-8. [DOI] [Google Scholar]
  929. Tajik H.; Esmaeili A. A.; Mohammadpoor-Baltork I.; Ershadi A.; Tajmehri H. Iodination of Aromatic Compounds with Iodine and n-Butyltriphenylphosphonium Peroxodisulfate. Synth. Commun. 2003, 33, 1319–1323. 10.1081/SCC-120018691. [DOI] [Google Scholar]
  930. Jereb M.; Zupan M.; Stavber S. Effective and selective iodofunctionalisation of organic molecules in water using the iodine-hydrogen peroxide tandem. Chem. Commun. 2004, 2614–2615. 10.1039/B409919B. [DOI] [PubMed] [Google Scholar]
  931. Lulinski P.; Kryska A.; Sosnowski M.; Skulski L. Eco-friendly Oxidative Iodination of Various Arenes with a Urea-Hydrogen Peroxide Adduct (UHP) as the Oxidant. Synthesis 2004, 441–445. 10.1055/s-2004-815955. [DOI] [Google Scholar]
  932. Pavlinac J.; Zupan M.; Stavber S. ‘Green’ Iodination of Dimethoxy- and Trimethoxy-Substituted Aromatic Compounds Using an Iodine-Hydrogen Peroxide Combination in Water. Synthesis 2006, 2006, 2603–2607. 10.1055/s-2006-942470. [DOI] [Google Scholar]
  933. Zupan M.; Iskra J.; Stavber S. Room Temperature Regioselective Iodination of Aromatic Ethers Mediated by SelectfluorTM Reagent F-TEDA-BF4. Tetrahedron Lett. 1997, 38, 6305–6306. 10.1016/S0040-4039(97)01414-7. [DOI] [Google Scholar]
  934. Pavlinac J.; Zupan M.; Stavber S. Effect of Water on the Functionalization of Substituted Anisoles with Iodine in the Presence of F-TEDA-BF4 or Hydrogen Peroxide. J. Org. Chem. 2006, 71, 1027–1032. 10.1021/jo052021n. [DOI] [PubMed] [Google Scholar]
  935. Brazdil L. C.; Cutler C. J. Selective Production of Diiodobenzene and Iodobenzene from Benzene. J. Org. Chem. 1996, 61, 9621–9622. 10.1021/jo961493m. [DOI] [Google Scholar]
  936. Togo H.; Abe S.; Nogami G.; Yokoyama M. Synthetic Use of Poly[4-(diacetoxyiodo)styrene] for Organic Reactions. Bull. Chem. Soc. Jpn. 1999, 72, 2351–2356. 10.1246/bcsj.72.2351. [DOI] [Google Scholar]
  937. Sosnowski M.; Skulski L. A Comparison of Microwave-Accelerated and Conventionally Heated Iodination Reactions of Some Arenes and Heteroarenes, Using ortho-Periodic Acid as the Oxidant. Molecules 2005, 10, 401–406. 10.3390/10020401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  938. Kraszkiewicz L.; Sosnowski M.; Skulski L. Easy, inexpensive and effective oxidative iodination of deactivated arenes in sulfuric acid. Tetrahedron 2004, 60, 9113–9119. 10.1016/j.tet.2004.07.095. [DOI] [Google Scholar]
  939. Kraszkiewicz L.; Sosnowski M.; Skulski L. Oxidative Iodination of Deactivated Arenes in Concentrated Sulfuric Acid with I2/NaIO4 and KI/NaIO4 Iodinating Systems. Synthesis 2006, 1195–1199. 10.1055/s-2006-926374. [DOI] [Google Scholar]
  940. Pasha M. A.; Myint Y. Y. Acid-Catalysed Selective Monoiodination of Electron-Rich Arenes by Alkali Metal Iodides. Synth. Commun. 2004, 34, 2829–2833. 10.1081/SCC-200026240. [DOI] [Google Scholar]
  941. Iskra J.; Stavber S.; Zupan M. Nonmetal-Catalyzed Iodination of Arenes with Iodide and Hydrogen Peroxide. Synthesis 2004, 2004, 1869–1873. 10.1055/s-2004-829136. [DOI] [Google Scholar]
  942. Emmanuvel L.; Shukla R. K.; Sudalai A.; Gurunath S.; Sivaram S. NaIO4/KI/NaCl: a new reagent system for iodination of activated aromatics through in situ generation of iodine monochloride. Tetrahedron Lett. 2006, 47, 4793–4796. 10.1016/j.tetlet.2006.05.062. [DOI] [Google Scholar]
  943. Radner F. Lower nitrogen oxide species as catalysts in a convenient procedure for the iodination of aromatic compounds. J. Org. Chem. 1988, 53, 3548–3553. 10.1021/jo00250a024. [DOI] [Google Scholar]
  944. Sathiyapriya R.; Karunakaran R. J. A Convenient Procedure for the Iodination of Arenes. J. Chem. Res. 2006, 2006, 575–576. 10.3184/030823406778521383. [DOI] [Google Scholar]
  945. Sathiyapriya R.; Karunakaran R. J. Novel, Water-Based Procedure for the Mono Iodination of Aromatic Amines and Phenols. Synth. Commun. 2006, 36, 1915–1917. 10.1080/00397910600602750. [DOI] [Google Scholar]
  946. Adimurthy S.; Ramachandraiah G.; Ghosh P. K.; Bedekar A. V. A new, environment friendly protocol for iodination of electron-rich aromatic compounds. Tetrahedron Lett. 2003, 44, 5099–5101. 10.1016/S0040-4039(03)01144-4. [DOI] [Google Scholar]
  947. Edgar K. J.; Falling S. N. An efficient and selective method for the preparation of iodophenols. J. Org. Chem. 1990, 55, 5287–5291. 10.1021/jo00305a026. [DOI] [Google Scholar]
  948. Krishna Mohan K. V. V.; Narender N.; Kulkarni S. J. Simple and regioselective oxyiodination of aromatic compounds with ammonium iodide and Oxone®. Tetrahedron Lett. 2004, 45, 8015–8018. 10.1016/j.tetlet.2004.09.010. [DOI] [Google Scholar]
  949. Song S.; Sun X.; Li X.; Yuan Y.; Jiao N. Efficient and Practical Oxidative Bromination and Iodination of Arenes and Heteroarenes with DMSO and Hydrogen Halide: A Mild Protocol for Late-Stage Functionalization. Org. Lett. 2015, 17, 2886–2889. 10.1021/acs.orglett.5b00932. [DOI] [PubMed] [Google Scholar]
  950. Benson W. R.; McBee E. T.; Rand L. N-IODOSUCCINIMIDE. Org. Synth. 1962, 42, 73. 10.15227/orgsyn.042.0073. [DOI] [Google Scholar]
  951. Olah G. A.; Wang Q.; Sandford G.; Prakash G. K. S. Iodination of deactivated aromatics with N-iodosuccinimide in trifluoromethanesulfonic acid (NIS-CF3SO3H) via in situ generated superelectrophilic iodine(I) trifluoromethanesulfonate. J. Org. Chem. 1993, 58, 3194–3195. 10.1021/jo00063a052. [DOI] [Google Scholar]
  952. Prakash G. K. S.; Mathew T.; Hoole D.; Esteves P. M.; Wang Q.; Rasul G.; Olah G. A. N-Halosuccinimide/BF3-H2O, Efficient Electrophilic Halogenating Systems for Aromatics. J. Am. Chem. Soc. 2004, 126, 15770–15776. 10.1021/ja0465247. [DOI] [PubMed] [Google Scholar]
  953. Castanet A.-S.; Colobert F.; Broutin P.-E. Mild and regioselective iodination of electron-rich aromatics with N-iodosuccinimide and catalytic trifluoroacetic acid. Tetrahedron Lett. 2002, 43, 5047–5048. 10.1016/S0040-4039(02)01010-9. [DOI] [Google Scholar]
  954. Carreño M. C.; Ruano J. L. G.; Sanz G.; Toledo M. A.; Urbano A. Mild and Regiospecific Nuclear Iodination of Methoxybenzenes and Naphthalenes with N-Iodosuccinimide in Aeetonitrile. Tetrahedron Lett. 1996, 37, 4081–4084. 10.1016/0040-4039(96)00738-1. [DOI] [Google Scholar]
  955. Orazi O. O.; Corral R. A.; Bertorello H. E. N-Iodohydantoins. II. Iodinations with 1,3-Diiodo-5,5-Dimethylhydantoin. J. Org. Chem. 1965, 30, 1101–1104. 10.1021/jo01015a036. [DOI] [PubMed] [Google Scholar]
  956. Jakab G.; Hosseini A.; Hausmann H.; Schreiner P. R. Mild and Selective Organocatalytic Iodination of Activated Aromatic Compounds. Synthesis 2013, 45, 1635–1640. 10.1055/s-0033-1338468. [DOI] [Google Scholar]
  957. Iida K.; Ishida S.; Watanabe T.; Arai T. Disulfide-Catalyzed Iodination of Electron-Rich Aromatic Compounds. J. Org. Chem. 2019, 84, 7411–7417. 10.1021/acs.joc.9b00769. [DOI] [PubMed] [Google Scholar]
  958. Andrews L. J.; Keefer R. M. The Reaction of Alkylbenzenes with Iodine Monochloride in Carbon Tetrachloride and in Trifluoroacetic Acid. J. Am. Chem. Soc. 1957, 79, 1412–1416. 10.1021/ja01563a038. [DOI] [Google Scholar]
  959. Barluenga J. Transferring iodine: more than a simple functional group exchange in organic synthesis. Pure Appl. Chem. 1999, 71, 431–436. 10.1351/pac199971030431. [DOI] [Google Scholar]
  960. Barluenga J.; Gonzalez J. M.; Garcia-Martin M. A.; Campos P. J.; Asensio G. Acid-mediated reaction of bis(pyridine)iodonium(I) tetrafluoroborate with aromatic compounds. A selective and general iodination method. J. Org. Chem. 1993, 58, 2058–2060. 10.1021/jo00060a020. [DOI] [Google Scholar]
  961. Dolenc D. N-Iodosaccharin - a New Reagent for Iodination of Alkenes and Activated Aromatics. Synlett 2000, 544–546. [Google Scholar]
  962. Suzuki H.; Nonoyama N. Nitrogen dioxide - sodium iodide as an efficient reagent for the one-pot conversion of aryl amines to aryl iodides under nonaqueous conditions. Tetrahedron Lett. 1998, 39, 4533–4536. 10.1016/S0040-4039(98)00824-7. [DOI] [Google Scholar]
  963. Krasnokutskaya E. A.; Semenischeva N. I.; Filimonov V. D.; Knochel P. A New, One-Step, Effective Protocol for the Iodination of Aromatic and Heterocyclic Compounds via Aprotic Diazotization of Amines. Synthesis 2007, 2007, 81–84. 10.1055/s-2006-958936. [DOI] [Google Scholar]
  964. Zarei A.; Hajipour A.; Khazdooz L. A One-Pot Method for the Iodination of Aryl Amines via Stable Aryl Diazonium Silica Sulfates Under Solvent-Free Conditions. Synthesis 2009, 2009, 941–944. 10.1055/s-0028-1087981. [DOI] [Google Scholar]
  965. Trusova M. E.; Krasnokutskaya E. A.; Postnikov P. S.; Choi Y.; Chi K.-W.; Filimonov V. D. A Green Procedure for the Diazotization-Iodination of Aromatic Amines under Aqueous, Strong-Acid-Free Conditions. Synthesis 2011, 2011, 2154–2158. 10.1055/s-0030-1260046. [DOI] [Google Scholar]
  966. Uemura S.; Tanaka S.; Okano M.; Hamana M. Decarboxylative ipso halogenation of mercury(II) pyridinecarboxylates. Facile formation of 3-iodo- and 3-bromopyridines. J. Org. Chem. 1983, 48, 3297–3301. 10.1021/jo00167a027. [DOI] [Google Scholar]
  967. Barton D. H. R.; Serebryakov E. P. A Convenient Procedure for the Decarboxylation of Acids. Proc. Chem. Soc. 1962, 309. [Google Scholar]
  968. Barton D. H. R.; Lacher B.; Zard S. Z. The invention of radical reactions: Part XVI. Radical decarboxylative bromination and iodination of aromatic acids. Tetrahedron 1987, 43, 4321–4328. 10.1016/S0040-4020(01)90307-2. [DOI] [Google Scholar]
  969. Singh R.; Just G. An Efficient Synthesis of Aromatic Iodides from Aromatic Carboxylic Acids. Synth. Commun. 1988, 18, 1327–1330. 10.1080/00397918808078799. [DOI] [Google Scholar]
  970. Hamamoto H.; Hattori S.; Takemaru K.; Miki Y. Hypervalent iodine(III)-LiX combination in fluoroalcohol solvent for aromatic halogenation of electron-rich arenecarboxylic acids. Synlett. 2011, 2011, 1563–1566. 10.1055/s-0030-1260791. [DOI] [Google Scholar]
  971. Li Z.; Wang K.; Liu Z.-Q. Transition-metal-free Hunsdiecker reaction of electron-rich arenecarboxylic acids and aryl aldehydes in water. Synlett. 2014, 25, 2508–2512. 10.1055/s-0034-1378583. [DOI] [Google Scholar]
  972. Perry G. J. P.; Quibell J. M.; Panigrahi A.; Larrosa I. Transition-Metal-Free Decarboxylative Iodination: New Routes for Decarboxylative Oxidative Cross-Couplings. J. Am. Chem. Soc. 2017, 139, 11527–11536. 10.1021/jacs.7b05155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  973. Bennett G. M.; Vernon I. H. Reversible replacement of aromatic halogen atoms. J. Chem. Soc. 1938, 1783–1786. 10.1039/jr9380001783. [DOI] [Google Scholar]
  974. Takagi K.; Hayama N.; Inokawa S. The in Situ-generated Nickel(0)-catalyzed Reaction of Aryl Halides with Potassium Iodide and Zinc Powder. Bull. Chem. Soc. Jpn. 1980, 53, 3691–3695. 10.1246/bcsj.53.3691. [DOI] [Google Scholar]
  975. Yang S. H.; Li C. S.; Cheng C. H. Halide exchange reactions between aryl halides and alkali halides catalyzed by nickel metal. J. Org. Chem. 1987, 52, 691–694. 10.1021/jo00380a041. [DOI] [Google Scholar]
  976. Klapars A.; Buchwald S. L. Copper-Catalyzed Halogen Exchange in Aryl Halides: An Aromatic Finkelstein Reaction. J. Am. Chem. Soc. 2002, 124, 14844–14845. 10.1021/ja028865v. [DOI] [PubMed] [Google Scholar]
  977. Li L.; Liu W.; Zeng H.; Mu X.; Cosa G.; Mi Z.; Li C.-J. Photo-induced Metal-Catalyst-Free Aromatic Finkelstein Reaction. J. Am. Chem. Soc. 2015, 137, 8328–8331. 10.1021/jacs.5b03220. [DOI] [PubMed] [Google Scholar]
  978. Kabalka G. W.; Sastry K. A. R.; Sastry U.; Somayaji V. SYNTHESIS OF ARYL HALIDES via ORGANOBORANE CHEMISTRY. Org. Prep. Proced. Int. 1982, 14, 359–362. 10.1080/00304948209354932. [DOI] [Google Scholar]
  979. Thiebes C.; Prakash G. K. S.; Petasis N. A.; Olah G. A. Mild Preparation of Haloarenes by Ipso-Substitution of Arylboronic Acids with N-Halosuccinimides. Synlett 1998, 1998, 141–142. 10.1055/s-1998-1614. [DOI] [Google Scholar]
  980. Kabalka G. W.; Mereddy A. R. A facile synthesis of aryl iodides via potassium aryltrifluoroborates. Tetrahedron Lett. 2004, 45, 343–345. 10.1016/j.tetlet.2003.10.149. [DOI] [Google Scholar]
  981. Liu W.; Yang X.; Gao Y.; Li C.-J. Simple and Efficient Generation of Aryl Radicals from Aryl Triflates: Synthesis of Aryl Boronates and Aryl Iodides at Room Temperature. J. Am. Chem. Soc. 2017, 139, 8621–8627. 10.1021/jacs.7b03538. [DOI] [PubMed] [Google Scholar]
  982. International Labour Organization . Iodine. International Chemical Safety Cards (ICSC) 0167. https://chemicalsafety.ilo.org/dyn/icsc/showcard.display?p_card_id=0167&p_lang=en&p_version=2.
  983. Tadross P. M.; Stoltz B. M. A Comprehensive History of Arynes in Natural Product Total Synthesis. Chem. Rev. 2012, 112, 3550–3577. 10.1021/cr200478h. [DOI] [PubMed] [Google Scholar]
  984. Himeshima Y.; Sonoda T.; Kobayashi H. Fluoride-Induced 1,2-Elimination of o-trimethylsilylphenyl Triflate to Benzyne Under Mild Conditions. Chem. Lett. 1983, 12, 1211–1214. 10.1246/cl.1983.1211. [DOI] [Google Scholar]
  985. Matsumoto T.; Hosoya T.; Katsuki M.; Suzuki K. New Efficient Protocol for Aryne Generation. Selective Synthesis of Differentially Protected 1,4,5-Naphthalenetriols. Tetrahedron Lett. 1991, 32, 6735–6736. 10.1016/S0040-4039(00)93589-5. [DOI] [Google Scholar]
  986. Sapountzis I.; Lin W.; Fischer M.; Knochel P. Preparation of Polyfunctional Arynes via 2-Magnesiated Diaryl Sulfonates. Angew. Chem., Int. Ed. 2004, 43, 4364–4366. 10.1002/anie.200460417. [DOI] [PubMed] [Google Scholar]
  987. Kitamura T.; Aoki Y.; Isshiki S.; Wasai K.; Fujiwara Y. Efficient generation and trapping of acylbenzynes from hypervalent iodine compounds. Tetrahedron Lett. 2006, 47, 1709–1712. 10.1016/j.tetlet.2006.01.042. [DOI] [Google Scholar]
  988. Miao Y.-H.; Hu Y.-H.; Yang J.; Liu T.; Sun J.; Wang X.-J. Natural source, bioactivity and synthesis of benzofuran derivatives. RSC Adv. 2019, 9, 27510–27540. 10.1039/C9RA04917G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  989. Radadiya A.; Shah A. Bioactive benzofuran derivatives: An insight on lead developments, radioligands and advances of the last decade. Eur. J. Med. Chem. 2015, 97, 356–376. 10.1016/j.ejmech.2015.01.021. [DOI] [PubMed] [Google Scholar]
  990. Khanam; Shamsuzzaman H. Bioactive Benzofuran derivatives: A review. Eur. J. Med. Chem. 2015, 97, 483–504. 10.1016/j.ejmech.2014.11.039. [DOI] [PubMed] [Google Scholar]
  991. Keay B.; Dibble P. In Comprehensive Heterocyclic Chemistry II; Katritzky A., Rees C., Scriven E., Eds.; Pergamon: Oxford, UK, 1996; p 395. [Google Scholar]
  992. Kadieva M. G.; Oganesyan É. T. Methods for The Synthesis of Benzofuran Derivatives (Review). Chem. Heterocycl. Compd. 1997, 33, 1245–1258. 10.1007/BF02320323. [DOI] [Google Scholar]
  993. Chiummiento L.; D’Orsi R.; Funicello M.; Lupattelli P. Last Decade of Unconventional Methodologies for the Synthesis of Substituted Benzofurans. Molecules 2020, 25, 2327. 10.3390/molecules25102327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  994. Fürstner A.; Davies P. W. Heterocycles by PtCl2-Catalyzed Intramolecular Carboalkoxylation or Carboamination of Alkynes. J. Am. Chem. Soc. 2005, 127, 15024–15025. 10.1021/ja055659p. [DOI] [PubMed] [Google Scholar]
  995. Trost B. M.; McClory A. Rhodium-Catalyzed Cycloisomerization: Formation of Indoles, Benzofurans, and Enol Lactones. Angew. Chem., Int. Ed. 2007, 46, 2074–2077. 10.1002/anie.200604183. [DOI] [PubMed] [Google Scholar]
  996. Liang Y.; Tang S.; Zhang X.-D.; Mao L.-Q.; Xie Y.-X.; Li J.-H. Novel and Selective Palladium-Catalyzed Annulations of 2-Alkynylphenols to Form 2-Substituted 3-Halobenzo[b]furans. Org. Lett. 2006, 8, 3017–3020. 10.1021/ol060908f. [DOI] [PubMed] [Google Scholar]
  997. Damera K.; Ke B.; Wang K.; Dai C.; Wang L.; Wang B. A novel base-promoted cyclization: synthesis of substituted benzo[b]furans. RSC Adv. 2012, 2, 9403–9405. 10.1039/c2ra21302h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  998. Nakamura M.; Ilies L.; Otsubo S.; Nakamura E. 3-Zinciobenzofuran and 3-Zincioindole: Versatile Tools for the Construction of Conjugated Structures Containing Multiple Benzoheterole Units. Angew. Chem., Int. Ed. 2006, 45, 944–947. 10.1002/anie.200502920. [DOI] [PubMed] [Google Scholar]
  999. Nakamura M.; Ilies L.; Otsubo S.; Nakamura E. 2,3-Disubstituted Benzofuran and Indole by Copper-Mediated C-C Bond Extension Reaction of 3-Zinciobenzoheterole. Org. Lett. 2006, 8, 2803–2805. 10.1021/ol060896y. [DOI] [PubMed] [Google Scholar]
  1000. Boyer A.; Isono N.; Lackner S.; Lautens M. Domino rhodium(I)-catalysed reactions for the efficient synthesis of substituted benzofurans and indoles. Tetrahedron 2010, 66, 6468–6482. 10.1016/j.tet.2010.05.106. [DOI] [Google Scholar]
  1001. Morozov O. S.; Lunchev A. V.; Bush A. A.; Tukov A. A.; Asachenko A. F.; Khrustalev V. N.; Zalesskiy S. S.; Ananikov V. P.; Nechaev M. S. Expanded-Ring N-Heterocyclic Carbenes Efficiently Stabilize Gold(I) Cations, Leading to High Activity in π-Acid-Catalyzed Cyclizations. Chem.—Eur. J. 2014, 20, 6162–6170. 10.1002/chem.201303760. [DOI] [PubMed] [Google Scholar]
  1002. Castro C. E.; Gaughan E. J.; Owsley D. C. Indoles, Benzofurans, Phthalides, and Tolanes via Copper(I) Acetylides. J. Org. Chem. 1966, 31, 4071–4078. 10.1021/jo01350a045. [DOI] [Google Scholar]
  1003. Bates C. G.; Saejueng P.; Murphy J. M.; Venkataraman D. Synthesis of 2-Arylbenzo[b]furans via Copper(I)-Catalyzed Coupling of o-Iodophenols and Aryl Acetylenes. Org. Lett. 2002, 4, 4727–4729. 10.1021/ol0272040. [DOI] [PubMed] [Google Scholar]
  1004. Kabalka G. W.; Wang L.; Pagni R. M. Sonogashira Coupling and Cyclization Reactions on Alumina: a Route to Aryl Alkynes, 2-Substituted-Benzo[b]furans and 2-Substituted-Indoles. Tetrahedron 2001, 57, 8017–8028. 10.1016/S0040-4020(01)00774-8. [DOI] [Google Scholar]
  1005. Carril M.; Correa A.; Bolm C. Iron-Catalyzed Sonogashira Reactions. Angew. Chem., Int. Ed. 2008, 47, 4862–4865. 10.1002/anie.200801539. [DOI] [PubMed] [Google Scholar]
  1006. Zhou R.; Wang W.; Jiang Z.-J.; Wang K.; Zheng X.-L.; Fu H.-Y.; Chen H.; Li R.-X. One-pot synthesis of 2-substituted benzo[b]furans via Pd-tetraphosphine catalyzed coupling of 2-halophenols with alkynes. Chem. Commun. 2014, 50, 6023–6026. 10.1039/C4CC00815D. [DOI] [PubMed] [Google Scholar]
  1007. Saha D.; Dey R.; Ranu B. C. A Simple and Efficient One-Pot Synthesis of Substituted Benzo[b]furans by Sonogashira Coupling-5-endo-dig Cyclization Catalyzed by Palladium Nanoparticles in Water Under Ligand- and Copper-Free Aerobic Conditions. Eur. J. Org. Chem. 2010, 2010, 6067–6071. 10.1002/ejoc.201000980. [DOI] [Google Scholar]
  1008. Bosiak M. J. A Convenient Synthesis of 2-Arylbenzo[b]furans from Aryl Halides and 2-Halophenols by Catalytic One-Pot Cascade Method. ACS Catal. 2016, 6, 2429–2434. 10.1021/acscatal.6b00190. [DOI] [Google Scholar]
  1009. Wang J.-R.; Manabe K. Hydroxyterphenylphoshine-Palladium Catalyst for Benzo[b]furan Synthesis from 2-Chlorophenols. Bifunctional Ligand Strategy for Cross-Coupling of Chloroarenes. J. Org. Chem. 2010, 75, 5340–5342. 10.1021/jo1007948. [DOI] [PubMed] [Google Scholar]
  1010. Batail N.; Bendjeriou A.; Lomberget T.; Barret R.; Dufaud V.; Djakovitch L. First Heterogeneous Ligand- and Salt-Free Larock Indole Synthesis. Adv. Synth. Catal. 2009, 351, 2055–2062. 10.1002/adsc.200900386. [DOI] [Google Scholar]
  1011. Singh C.; Prakasham A. P.; Ghosh P. Palladium Acyclic Diaminocarbene (ADC) Triflate Complexes as Effective Precatalysts for the Hiyama Alkynylation/Cyclization Reaction Yielding Benzofuran Compounds: Probing the Influence of the Triflate Co-Ligandin the One-Pot Tandem Reaction. ChemistrySelect 2019, 4, 329–336. 10.1002/slct.201803292. [DOI] [Google Scholar]
  1012. Yuan H.; Bi K.-J.; Li B.; Yue R.-C.; Ye J.; Shen Y.-H.; Shan L.; Jin H.-Z.; Sun Q.-Y.; Zhang W.-D. Construction of 2-Substituted-3- Functionalized Benzofurans via Intramolecular Heck Coupling: Application to Enantioselective Total Synthesis of Daphnodorin B. Org. Lett. 2013, 15, 4742–4745. 10.1021/ol4021095. [DOI] [PubMed] [Google Scholar]
  1013. Tabolin A. A.; Ioffe S. L. Rearrangement of N-Oxyenamines and Related Reactions. Chem. Rev. 2014, 114, 5426–5476. 10.1021/cr400196x. [DOI] [PubMed] [Google Scholar]
  1014. Contiero F.; Jones K. M.; Matts E. A.; Porzelle A.; Tomkinson N. C. O. Direct Preparation of Benzofurans from O-Arylhydroxylamines. Synlett 2009, 2009, 3003–3006. 10.1055/s-0029-1218273. [DOI] [Google Scholar]
  1015. Maimone T. J.; Buchwald S. L. Pd-Catalyzed O-Arylation of Ethyl Acetohydroximate: Synthesis of O-Arylhydroxylamines and Substituted Benzofurans. J. Am. Chem. Soc. 2010, 132, 9990–9991. 10.1021/ja1044874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  1016. Boronic Acids: Preparation and Applications in Organic Synthesis, Medicine and Materials, 2nd ed.; Hall D. G., Ed.; Wiley-VCH: Weinheim, Germany, 2011. [Google Scholar]
  1017. Miyaura N.; Suzuki A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95, 2457–2483. 10.1021/cr00039a007. [DOI] [Google Scholar]
  1018. Chan D. M. T.; Monaco K. L.; Wang R.-P.; Winters M. P. New N- and O-arylations with phenylboronic acids and cupric acetate. Tetrahedron Lett. 1998, 39, 2933–2936. 10.1016/S0040-4039(98)00503-6. [DOI] [Google Scholar]
  1019. Evans D. A.; Katz J. L.; West T. R. Synthesis of diaryl ethers through the copper-promoted arylation of phenols with arylboronic acids. An expedient synthesis of thyroxine. Tetrahedron Lett. 1998, 39, 2937–2940. 10.1016/S0040-4039(98)00502-4. [DOI] [Google Scholar]
  1020. Lam P. Y. S.; Clark C. G.; Saubern S.; Adams J.; Winters M. P.; Chan D. M. T.; Combs A. New aryl/heteroaryl C-N bond cross-coupling reactions via arylboronic acid/cupric acetate arylation. Tetrahedron Lett. 1998, 39, 2941–2944. 10.1016/S0040-4039(98)00504-8. [DOI] [Google Scholar]
  1021. Brown H. C.; Cole T. E. Organoboranes. 31. A simple preparation of boronic esters from organolithium reagents and selected trialkoxyboranes. Organometallics 1983, 2, 1316–1319. 10.1021/om50004a009. [DOI] [Google Scholar]
  1022. Brown H. C.; Srebnik M.; Cole T. E. Organoboranes. 48. Improved procedures for the preparation of boronic and borinic esters. Organometallics 1986, 5, 2300–2303. 10.1021/om00142a020. [DOI] [Google Scholar]
  1023. Baron O.; Knochel P. Preparation and Selective Reactions of Mixed Bimetallic Aromatic and Heteroaromatic Boron-Magnesium Reagents. Angew. Chem., Int. Ed. 2005, 44, 3133–3135. 10.1002/anie.200462928. [DOI] [PubMed] [Google Scholar]
  1024. Ishiyama T.; Miyaura N. Metal-catalyzed reactions of diborons for synthesis of organoboron compounds. Chem. Rec. 2004, 3, 271–280. 10.1002/tcr.10068. [DOI] [PubMed] [Google Scholar]
  1025. Merino P.; Tejero T. Expanding the Limits of Organoboron Chemistry: Synthesis of Functionalized Arylboronates. Angew. Chem., Int. Ed. 2010, 49, 7164–7165. 10.1002/anie.201003191. [DOI] [PubMed] [Google Scholar]
  1026. Pilarski L. T.; Szabó K. J. Palladium-Catalyzed Direct Synthesis of Organoboronic Acids. Angew. Chem., Int. Ed. 2011, 50, 8230–8232. 10.1002/anie.201102384. [DOI] [PubMed] [Google Scholar]
  1027. Chow W. K.; Yuen O. Y.; Choy P. Y.; So C. M.; Lau C. P.; Wong W. T.; Kwong F. Y. A decade advancement of transition metal-catalyzed borylation of aryl halides and sulfonates. RSC Adv. 2013, 3, 12518–12539. 10.1039/c3ra22905j. [DOI] [Google Scholar]
  1028. Ishiyama T.; Miyaura N. Transition metal-catalyzed borylation of alkanes and arenes via C-H activation. J. Organomet. Chem. 2003, 680, 3–11. 10.1016/S0022-328X(03)00176-1. [DOI] [Google Scholar]
  1029. Mkhalid I. A. I.; Barnard J. H.; Marder T. B.; Murphy J. M.; Hartwig J. F. C-H Activation for the Construction of C-B Bonds. Chem. Rev. 2010, 110, 890–931. 10.1021/cr900206p. [DOI] [PubMed] [Google Scholar]
  1030. Hartwig J. F. Regioselectivity of the borylation of alkanes and arenes. Chem. Soc. Rev. 2011, 40, 1992–2002. 10.1039/c0cs00156b. [DOI] [PubMed] [Google Scholar]
  1031. Hartwig J. F. Borylation and Silylation of C-H Bonds: A Platform for Diverse C-H Bond Functionalizations. Acc. Chem. Res. 2012, 45, 864–873. 10.1021/ar200206a. [DOI] [PubMed] [Google Scholar]
  1032. Lee K.-s.; Zhugralin A. R.; Hoveyda A. H. Efficient C-B Bond Formation Promoted by N-Heterocyclic Carbenes: Synthesis of Tertiary and Quaternary B-Substituted Carbons through Metal-Free Catalytic Boron Conjugate Additions to Cyclic and Acyclic α,β-Unsaturated Carbonyls. J. Am. Chem. Soc. 2009, 131, 7253–7255. 10.1021/ja902889s. [DOI] [PMC free article] [PubMed] [Google Scholar]
  1033. Lillo V.; Prieto A.; Bonet A.; Díaz-Requejo M. M.; Ramírez J.; Pérez P. J.; Fernández E. Asymmetric β-Boration of α,β-Unsaturated Esters with Chiral (NHC)Cu Catalysts. Organometallics 2009, 28, 659–662. 10.1021/om800946k. [DOI] [Google Scholar]
  1034. Tian Y.-M.; Guo X.-N.; Braunschweig H.; Radius U.; Marder T. B. Photoinduced Borylation for the Synthesis of Organoboron Compounds. Chem. Rev. 2021, 121, 3561–3597. 10.1021/acs.chemrev.0c01236. [DOI] [PubMed] [Google Scholar]
  1035. Jiang M.; Yang H.; Fu H. Visible-Light Photoredox Borylation of Aryl Halides and Subsequent Aerobic Oxidative Hydroxylation. Org. Lett. 2016, 18, 5248–5251. 10.1021/acs.orglett.6b02553. [DOI] [PubMed] [Google Scholar]
  1036. Tyman J. H. P.Synthetic and Natural Phenols; Elsevier, New York, 1996. [Google Scholar]
  1037. Rappoport Z.The Chemistry of Phenols; Wiley-VCH, Weinheim, 2003. [Google Scholar]
  1038. Scott K. A.; Cox P. B.; Njardarson J. T. Phenols in Pharmaceuticals: Analysis of a Recurring Motif. J. Med. Chem. 2022, 65, 7044–7072. 10.1021/acs.jmedchem.2c00223. [DOI] [PubMed] [Google Scholar]
  1039. George T.; Mabon R.; Sweeney G.; Sweeney J. B.; Tavassoli A. Alcohols, ethers and phenols. J. Chem. Soc. Perkin Trans. 1 2000, 2529–2574. 10.1039/a808133f. [DOI] [Google Scholar]
  1040. Hoarau C.; Pettus T. R. R. Strategies for the Preparation of Differentially Protected ortho-Prenylated Phenols. Synlett 2003, 34, 127–137. 10.1002/chin.200310256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  1041. Anderson K. W.; Ikawa T.; Tundel R. E.; Buchwald S. L. The Selective Reaction of Aryl Halides with KOH: Synthesis of Phenols, Aromatic Ethers, and Benzofurans. J. Am. Chem. Soc. 2006, 128, 10694–10695. 10.1021/ja0639719. [DOI] [PubMed] [Google Scholar]
  1042. Chen G.; Chan A. S. C.; Kwong F. Y. Palladium-catalyzed C-O bond formation: direct synthesis of phenols and aryl/alkyl ethers from activated aryl halides. Tetrahedron Lett. 2007, 48, 473–476. 10.1016/j.tetlet.2006.11.036. [DOI] [Google Scholar]
  1043. Schulz T.; Torborg C.; Schäffner B.; Huang J.; Zapf A.; Kadyrov R.; Börner A.; Beller M. Practical Imidazole-Based Phosphine Ligands for Selective Palladium-Catalyzed Hydroxylation of Aryl Halides. Angew. Chem., Int. Ed. 2009, 48, 918–921. 10.1002/anie.200804898. [DOI] [PubMed] [Google Scholar]
  1044. Sergeev A. G.; Schulz T.; Torborg C.; Spannenberg A.; Neumann H.; Beller M. Palladium-catalyzed hydroxylation of aryl halides under ambient conditions. Angew. Chem., Int. Ed. 2009, 48, 7595–7599. 10.1002/anie.200902148. [DOI] [PubMed] [Google Scholar]
  1045. Tlili A.; Xia N.; Monnier F.; Taillefer M. A Very Simple Copper-Catalyzed Synthesis of Phenols Employing Hydroxide Salts. Angew. Chem., Int. Ed. 2009, 48, 8725–8728. 10.1002/anie.200903639. [DOI] [PubMed] [Google Scholar]
  1046. Zhao D.; Wu N.; Zhang S.; Xi P.; Su X.; Lan J.; You J. Synthesis of Phenol, Aromatic Ether, and Benzofuran Derivatives by Copper-Catalyzed Hydroxylation of Aryl Halides. Angew. Chem., Int. Ed. 2009, 48, 8729–8732. 10.1002/anie.200903923. [DOI] [PubMed] [Google Scholar]
  1047. Maurer S.; Jiang Y. W.; Ma D. W.; Liu W.; Zhang X. J. An Efficient Synthesis of Phenol via CuI/8-Hydroxyquinoline-Catalyzed Hydroxylation of Aryl Halides and Potassium Hydroxide. Synlett 2010, 2010, 976–978. 10.1055/s-0029-1219548. [DOI] [Google Scholar]
  1048. Yu C. W.; Chen G. S.; Huang C. W.; Chern J. W. Efficient Microwave-Assisted Pd-Catalyzed Hydroxylation of Aryl Chlorides in the Presence of Carbonate. Org. Lett. 2012, 14, 3688–3691. 10.1021/ol301523q. [DOI] [PubMed] [Google Scholar]
  1049. Wang K.; Tong M.; Yang Y.; Zhang B.; Liu H.; Li H.; Zhang F. Visible light-catalytic hydroxylation of aryl halides with water to phenols by carbon nitride and nickel complex cooperative catalysis. Green Chem. 2020, 22, 7417–7423. 10.1039/D0GC02367A. [DOI] [Google Scholar]

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

RESOURCES