Next Article in Journal
RNA-Seq Analysis of MCF-7 Breast Cancer Cells Treated with Methyl Gallate Isolated from the Rhizomes of Nymphaea Odorata L. Shows Upregulation of Apoptosis, Autophagy, and Unfolded Protein Canonical Pathways
Previous Article in Journal
Photophysical Properties of 1,3-Diphenylisobenzofuran as a Sensitizer and Its Reaction with O2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in α-Arylation of Carbonyl Compounds: Diaryliodonium Salts as Arylating Agents

1
School of Chemistry and Chemical Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
2
State Key Laboratory of Applied Organic Chemistry, College of Chemistry and Chemical Engineering, Lanzhou University, Lanzhou 730000, China
3
Animal Husbandry and Veterinary Bureau of Anding District, Dingxi 743021, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(14), 3019; https://doi.org/10.3390/molecules30143019
Submission received: 11 June 2025 / Revised: 12 July 2025 / Accepted: 16 July 2025 / Published: 18 July 2025

Abstract

Diaryliodonium salts are an important part of hypervalent iodine chemistry, owing to their highly electrophilic character, non-toxicity, and air and moisture stability, have been identified as an important arylating agent. It has been widely applied in the synthesis of natural products, drugs, and bioactive molecules bearing active α-arylation carbonyl units. Within the domain of α-arylation of carbonyl compounds using diaryliodonium salts, there is a notable absence in the literature of a comprehensive compilation dedicated to exclusive arylation processes involving these compounds. In this review, we focus on the overview of the recent advancements in utilizing diaryliodonium salts for α-arylation reactions, encompassing both racemic and asymmetric approaches to various carbonyl compounds including ketones, esters, enolates, and amides. Furthermore, we discuss the unique advantages and inherent limitations of diaryliodonium salts as arylating agents, as well as the underexplored application potentials that warrant further investigation in this rapidly evolving field.

1. Introduction

The α-aryl carbonyl compounds are prevalent in numerous natural products, bioactive molecules, and drugs of significance. Such scaffolds also serve as key building blocks in the total synthesis of biologically active molecules (Figure 1) [1,2,3,4,5,6,7,8,9,10]. Over the past few decades, numerous synthetic chemists, both domestically and internationally, have dedicated significant effort to the study of α-arylation of carbonyl groups. This research has given rise to a multitude of exceptional chemical synthesis methods and strategies [11,12,13,14,15,16] such as transition metal-catalyzed cross-coupling, transition metal-free α-arylation, metal-catalyzed photoredox reaction [17,18] and metal-free photoredox α-arylation [19]. According to the type of arylating agents employed, these agents can be classified into two major categories: (i) arylating nucleophilic, aryl metal, or boron species; (ii) arylating electrophiles, such as aryl halides, aryl triflates, diaryliodonium salts, etc. Among them, the diaryliodonium salt was first synthesized in 1894 by Meyer and Hartmann [20]. The structure features a trivalent iodine atom bonded to two aryl groups and one counter-anion. In the solid state, this compound adopts a three-coordinate, T-shaped geometry. Within this structure, the iodine atom forms a hypervalent, linear 3-center 4-electron bond with two of its substituents (Figure 2a).
This bonding imparts electrophilic character to the iodine center, resulting in excellent reactivity (Figure 2b). Consequently, a large variety of synthetic routes to diaryliodonium salts have been developed [21,22,23,24,25,26,27,28,29,30]. With the rapid development of the related synthetic methods, diaryliodonium salts have been widely recognized and utilized in the field of organic synthesis chemistry, such as metal-catalyzed cross-coupling, transition-metal-catalyzed C-H bond activation/aromatization, α-position arylation of carbonyl compounds, asymmetric tandem cyclization and arylation of heteroatoms [31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49], and photoredox catalysis involving diaryliodonium salts [50,51,52]. Meanwhile, these protocols have been successfully applied to the rapid synthesis of natural products, pharmaceutical candidates, treatment of cancer, oral topical analgesic (S)-Ketoprofen (1-e), oral non-steroidal anti-inflammatory drug (S)-Naproxen (1-f) and Ibuprofen (1-g), as well as clinical drugs ketamine (1-h) and tiletamine (1-i).
Over the past few decades, the α-arylation of carbonyl compounds is a significant area of research within this domain, with diaryliodonium salts emerging as efficient, mild, selective, and robust arylating agents to surmount this difficult problem. In 2022, Zhou and Yu [53] published a review article that covers the enantioselective α-arylation of carbonyl enolates and related compounds by iodonium salts. In 2025, Simlandy’s group [54] conducted a comprehensive review of metal-catalyzed enantioselective aryl transfer methodologies utilizing diaryliodonium salts. Despite significant achievements, to the best of our knowledge, there lacks a comprehensive review article to summarize the recent advances in the α-arylation of α-carbonyl compounds.
Within the domain of α-arylation of carbonyl compounds using diaryliodonium salts, there is a notable absence in the literature of a comprehensive compilation dedicated to exclusive arylation processes involving these compounds. In this review article, we summarize recent advances in the α-arylation of carbonyl compounds (such as ketones, esters, enolates, and amides, etc.) (Figure 2c).

2. The Racemic α-Arylation of Carbonyl Compounds

2.1. α-Arylation of Keto Carbonyl Compounds

In 1960, Yudis’ research group [55] reported the α-arylation of carbonyl compounds using diphenyliodonium salts as arylating agents. The reaction was conducted in t-BuOH using t-BuOK as base, wherein 1,3-dicarbonyls (3) (such as dimedone, dibenzoylmethane, and tribenzoylmethane) reacted with diphenylsulfonium salt (2-a) to afford α-phenylated ketones (4-a4-b) and side products. Systematic screening of solvents and counter-anions identified tert-butanol and the iodide anion as optimal components.
In subsequent studies, Huang’s group [56] and Beringer’s group [57] employed the same reaction system to achieve the α-phenylation of indandione derivatives (4-d4-f) and monoketones (4-g4-k), respectively. The phenylated indandione derivatives product was obtained in moderate yield, whereas the phenylated monoketone product was afforded in moderate-to-good yield (Scheme 1).
Based on the observed reaction pathway, the group proposed a radical-pair mechanism: electron transfer from a carbanion (R) to the iodonium ion (Ar–I+–Ar) generates transient radical pairs (R· and Ar·), which predominantly recombine via radical substitution on diphenyliodine or direct coupling to yield arylative products. Competing pathways involving radical dissociation generate side products (Ar-R, Ar-Ar, R-R) through stochastic recombination (Scheme 2). Furthermore, the byproduct observed in Beringer’s group’s study could also be explained by this reaction mechanism.
The synthesis of these arylation products demonstrates the utilization of diaryliodonium salts as arylating agents in organic synthesis. These experiments preliminarily investigated the reaction conditions, applicability, and mechanism of aryl iodonium salts as arylating reagents, laying the foundation for subsequent research on arylation.
In 1997, Ryan’s research group [58] reported the transition metal-mediated α-arylation of cyclic ketones using diaryliodonium triflates (Scheme 3). Employing CuCN as a catalyst, the reaction between diaryliodonium triflate (2-b) and cyclic ketones (5) selectively produced α-phenylated derivatives (6). Notably, 5-membered cyclic ketones exclusively underwent diarylation (6-a, 6-b), while larger rings (6–8-membered) favored monoarylation (6-c, 6-d, 6-e.). This research group has developed efficient methods for synthesizing a wide range of stable diaryliodonium triflates and successfully applied them to the arylation of cyclic ketones. This work significantly expands both the synthetic methodologies for preparing diaryliodonium salts and their applicability as arylating agents.
In 2015, Liu and co-workers [59] reported a t-BuOK-promoted method for the α-arylation of indolone derivatives (7) with diaryliodonium salts, achieving monoarylated indolone derivatives (8) in moderate-to-good yields (24–60%) (Scheme 4). This procedure demonstrated remarkable compatibility with a variety of functional groups and displayed a wide substrate applicability, showing adaptability to varying electronic and steric conditions. Moreover, transformation of monoarylated indolinones with diverse electrophiles has been demonstrated, enabling access to bioactive compounds featuring 2-arylated indolinone cores.
In 2015, Zhang’s research group [60] reported Cu-mediated one-pot Michael addition/α-arylation strategy to construct α-aryl-β-substituted cyclic ketone scaffolds (10) using cyclohex-2-en-1-one (9) as the substrate and diaryliodonium salts as arylating agents (Scheme 5). The transformation exhibits advantages including commercially accessible reagents, broad substrate tolerance, moderate-to-excellent yields, and high diastereoselectivity.
Based on the experimental results and previous studies, a possible mechanism for the current one-pot reaction was proposed (Scheme 6). Firstly, a catalytically active Cu(I) species—generated either in situ via formation of a copper enolate intermediate or derived from Cu(II) through reduction/disproportionation—reacts with the diaryliodonium salt to afford a key Cu(III)-aryl intermediate. This electrophilic Cu(III) species subsequently engages with the Michael addition-derived enolate intermediate through either path a or path b, followed by phenyl group transfer and C–C(Ar) bond formation via reductive elimination. This sequence delivers the product while regenerating the Cu(I) catalyst.
In 2017, Wen and co-workers [61] reported a palladium-catalyzed dual C–C arylation reaction using cyclic diaryliodonium salts (2-c) and methylene-containing 1,3-dicarbonyl compounds (11) as substrates (Scheme 7). This method enabled the concise and efficient construction of structurally diverse fluorene derivatives (12) bearing all-carbon quaternary carbon centers. Moreover, the protocol was extended to synthesize novel pharmacologically active spirofluorene derivatives incorporating barbituric acid or indanone moieties, demonstrating distinctive synthetic value.
In 2019, Zhang’s research group [62] extended the frontiers of metal-free ketone functionalization by pioneering the transition metal-free α-arylation of α-nitro ketones (13) mediated by diaryliodonium salts. The protocol delivered structurally diverse α-aryl-α-nitro ketones (14) in synthetically viable yields (40–76%), while demonstrating exceptional functional group compatibility and extensive substrate generality (Scheme 8). Additionally, this group achieved a concise and efficient synthesis of the clinical drug Tiletamine (1-i) via this method through a three-step transformation with a 32% overall yield. In summary, the synthetic strategy developed by this research group not only features mild reaction conditions and broad substrate scope, but enables the production of synthetically valuable products.
In 2023, our group [63] further achieved enantioselective divergent total syntheses of several meroterpenoids. (Scheme 9) The key transformations involved a one-pot Michael addition/α-arylation of an enone (16), utilizing diaryliodonium salts (2-d) as the arylating reagent to install the α-aryl group, yielding the desired arylation products (17 and 18). The authors have further expanded the use of diaryliodonium salts as arylation reagents in the total synthesis of meroterpenoid natural products.

2.2. α-Arylation of Ester Carbonyl Compounds

In 1963, Forgione and co-workers [64] pioneered the α-phenylation of ester carbonyl compounds (19) using diphenyliodonium chloride as the arylating reagent. This transformation was achieved in a t-BuOK/t-BuOH system under reflux conditions, delivering α-arylated products (20-a20-e) in modest yields (28–70%). They discovered that spatial steric hindrance significantly impacted the reaction. Furthermore, the benzhydrylation products, synthesized in an early exploratory phase without optimization, yielded suboptimally but are amenable to improvement. In 1987, Chen et al. [65] utilized the same reaction system to achieve the α-arylation of isopropylidene malonate and its 5-substituted derivates. The reaction afforded diarylated or monoarylated products (20-f20-v) in good-to-excellent yields (Scheme 10).
A key feature of the above approach is that the α-arylated diesters can be hydrolyzed to α-aryl carboxylic acids or their esters. Therefore, this method provides an indirect yet efficient route for the preparation of α-aryl carboxylic acids or esters. Furthermore, the α-arylation of isopropylidene malonate and its 5-substituted derivatives proceeded in satisfactory yields regardless of whether the aryl ring bore electron-withdrawing groups (Cl, NO2) or electron-donating groups (such as CH3, OCH3).
In 1999, Chang and colleagues [66] reported the α-arylation of β-ketoester substrates (20) employing diaryliodonium salts as arylating agents under basic promotion (Scheme 11). The experimental results demonstrated that the corresponding aryl derivatives (21) were obtained in moderate-to-good yields under mild conditions. However, when Ar1 and Ar2 are distinct groups, the arylation reaction exhibits poor regioselectivity. This limitation significantly constrained the synthetic utility of unsymmetrical diaryliodonium salts. This challenge was subsequently addressed in later research.
In 2014, Shibata and co-workers [67] synthesized a series of sterically hindered, unsymmetrical diaryliodonium salts, including pentafluorophenyl and triisopropylphenyl moieties (2-f). Employing these electrophilic aryl transfer reagents in conjunction with K3PO4 as the base, the group achieved α-pentafluorophenylation of β-ketoesters and β-ketoamides (22), affording α-pentafluorophenylated 1,3-dicarbonyl compounds featuring all-carbon quaternary stereocenters (23) (Scheme 12).
Subsequently, the group [68] developed a one-pot synthesis of unsymmetrical SF5-aryl-λ3-iodanes (2-g), employing them for SF5-arylation of 1,3-dicarbonyl compounds (24) under mild conditions (NaH/DMF) to yield SF5-functionalized derivatives (25-a–25-l). This methodology was extended to the arylation of diverse nucleophiles (e.g., phenols, anilines, carboxylates) using base or Cu catalysts. In 2017, the same group [69] synthesized SO2CF3-aryl/pyridyl iodonium salts (2-h), enabling trifluoromethanesulfonylation of C-/heteroatom-nucleophiles and trifluoromethanesulfonylpyridylation under some conditions, affording α-aryl/pyridyl trifluoromethanesulfonyl ketones in high yields (Scheme 13). Collectively, these approaches address regioselectivity challenges in unsymmetrical diaryliodonium chemistry and provide efficient routes to SF5- and SO2CF3-containing aryl/heteroaryl scaffolds—valuable precursors for bioactive compounds, catalysts, and chiral molecules.
In 2015, Manetsch and co-workers [70] demonstrated metal-free α-arylation of ethyl acetoacetate (26) using diaryliodonium triflates under ambient conditions. Whereas electron-donating/withdrawing substituents (excluding 4-methoxy groups) delivered moderate yields of α-arylated adducts, sterically encumbered systems (e.g., mesityl iodonium salts) proved substantially less competent. This methodology enabled the key arylation step in the synthesis of the antimalarial lead compound ELQ-300 (28-a), further demonstrating the promising utility of diaryliodonium salts as arylating agents (Scheme 14).
Subsequently, researchers expanded the substrate scope from dicarbonyl compounds to monocarbonyl compounds bearing ester groups with electron-withdrawing groups at the α-position. In 2015, Olofsson’s group [71] developed an efficient, direct, and metal-free approach to achieve the α-arylation of α-nitro esters. This transformation requires stoichiometric amounts of nitro esters and diaryliodonium salts, affording α-substituted nitro esters (30-a30-e) in moderate-to-good yields with broad functional group tolerance. In 2016, Han and Wang et al. [72] developed a transition metal-free direct arylation of 2-substituted cyanoacetates with diaryliodonium salts. The synthetic methodology was successfully extended to substrates bearing diverse functional groups, enabling the preparation of various 2-substituted α-arylacetonitriles (30-f30-j) in good yields. These compounds serve as valuable chemical intermediates for medicinal research. Significantly, the effective application of this method afforded streamlined access to glutethimide (33), a therapeutically relevant compound exhibiting neuroparalytic, immunomodulatory, and anticancer activities—thereby demonstrating its strategic utility in complex molecule synthesis. (Scheme 15).
In 2023, Kikushima et al. [73] reported catalyst-free decarboxylative arylation reaction employing α, α-difluoro-β-ketoesters (34) as substrates and diaryliodonium salts (2-i) as arylating reagents. This reaction can be applied to the synthesis of diverse α, α-difluoromethyl ketone derivatives (35) (Scheme 16). The synthetic significance of this reaction is twofold: it not only resides in the capacity of the α,α-difluoromethyl ketone group within the synthesized products to be converted into esters, amides, and difluoromethyl motifs prevalent in biologically active compounds, but concurrently establishes an environmentally benign methodology for constructing fluorinated pharmaceuticals and functional molecules.

2.3. α-Arylation of Silyl Enol Ethers

Due to silyl enol ethers being more nucleophilic than ketones, the arylation of silyl enol ethers derivatives is a strategic C–C bond-forming process that has found widespread application in organic synthesis. Over the past few decades, the diaryliodonium salts have facilitated major breakthroughs in enolate α-arylation.
Koser’s group [74] investigated the α-phenylation of silyl enol ethers (36) using fluorinated diphenyliodonium salts (DIF). This reaction afforded α-phenyl ketones (37) or α, α-diphenyl ketones (38) in yields ranging from 20% to 88% (Scheme 17). During this study, the group examined the influence of kinetic and thermodynamic factors on regioselectivity. Their findings demonstrated that regioselectivity in α-phenylation can be controlled through appropriate condition selection.
When this group subjected 3,3-dimethyl-2-(silyloxy)-1-butene (37-f) to DIF, the reaction not only generated the α-phenylated product (major) but also afforded a dehydrogenated dimeric byproduct of the ketone (minor) (Scheme 18). Based on experimental evidence suggesting a radical pathway, the research group proposed a plausible mechanism for this reaction.
In 1999, Rawal and co-workers [75] developed a controlled synthetic approach for carbocyclic-fused indoles (Scheme 19). This strategy commenced with the regioselective α-arylation of a silyl enol ether (38) using an o-nitrobenzene-substituted phenyliodonium salt (NPIF, 2-j). Subsequently, the resulting nitro compound underwent TiCl3-mediated reduction. The generated aniline intermediate then spontaneously condensed with the adjacent ketone functionality. This efficient two-step sequence ultimately afforded a series of indole derivatives (39).
In 2021, Taillefer and co-workers [76] reported a ligand-free and base-free Cu-catalyzed protocol for the α-arylation of arylketone-derived enolsilanes (Scheme 20). This methodology employed ketone-derived silyl enol ethers (41) as substrates paired with symmetric diaryliodonium salts or unsymmetric mesityl(phenyl)iodonium salts as arylating agents, achieving intermolecular C–C coupling under mild conditions, affording a series of aromatic ketone derivatives. (42) Critically, the ligand-free and base-free system exhibited broad functional group tolerance, accommodating sensitive moieties such as triflates and iodides (Scheme 20a).
In this case, the authors proposed the following plausible reaction mechanism (Scheme 20b): First, a Cu(III) species (Int 1) is generated via oxidative addition of Cu(I) into the Ar–I bond of the diaryliodonium salt. Subsequently, the key intermediate (Int 1) reacts with the silyl ether to afford a mixed Cu(III) aryl–alkyl complex (Int 2). This intermediate undergoes reductive elimination to regenerate the Cu(I) catalytic species, while simultaneously releasing the α-aryl ketone product after desilylation, mediated by the triflate anion.

2.4. α-Arylation of Amides

In 2020, Mohanan’s research group [77] reported a metal-free, mild, and efficient α-arylation reaction between diaryliodonium salts and α-nitro-fluoroacetamides (43) (Scheme 21). This reaction successfully utilized a broad range of diaryliodonium salts bearing sensitive functional groups and α-nitro-fluoroacetamides to synthesize α-arylated α-nitro-fluoroacetamide derivatives (44-a44-k) in moderate-to-high yields. These compounds exhibit unique potential value due to their benzyl-fluorinated quaternary carbon stereocenters. Subsequently, the authors extended this methodology to α-cyano-α-fluoroacetamides with diaryliodonium salt under briefly optimized conditions, affording a series of α-aryl-α-cyano-α-fluoroacetamide derivatives (44-l44-p). This approach exhibited excellent functional group tolerance and broad substrate scope, further highlighting the synthetic utility of the method.
Concurrently, Taillefer’s group [78] reported a metal-free methodology for synthesizing α-arylated α-fluoroacetamides (Scheme 22). This approach leveraged unsymmetric diaryliodonium salts to arylate α-aceto-α-fluoroacetamide (45), affording tetrasubstituted α-aceto-α-fluoroacetate derivatives (46) in good yields. Notably, electron-deficient diaryliodonium salts triggered a spontaneous arylation/deacylation cascade, directly yielding α-arylated-α-fluoroacetamides (47). The protocol’s versatility was further demonstrated through a sequential base-mediated deacylation of arylated intermediates (46), enabling efficient access to diverse α-aryl-fluoroacetamides (47). This work significantly broadened the synthetic utility of unsymmetric diaryliodonium salts as modular arylating agents.
The authors proposed a plausible mechanism (Scheme 23): Deprotonation under basic conditions generates an enolate intermediate, which subsequently coordinates with the diaryliodonium salt via C–I or O–I bond formation to afford tricoordinated iodonium intermediates Int 3 and Int 4. These intermediates rapidly equilibrate, followed by characteristics [1,2]-ligand coupling or [2,3]-rearrangement to yield the arylated product (46). Finally, the electron-deficient aryl group enhances the electrophilicity of the fluorinated quaternary carbon center, facilitating base-mediated diacylation to deliver the deacylated product (47).

3. The Asymmetric α-Arylation of Carbonyl Compounds

The catalytic asymmetric synthesis of α-aryl carbonyl compounds has long posed a significant challenge in asymmetric catalysis. The core difficulty stems from the frequent requirement to construct a quaternary carbon stereocenter or the propensity for facile racemization at the newly formed α-aryl α-carbonyl-substituted secondary carbon stereocenter. Despite these inherent challenges, the past two decades have seen the development of diverse innovative strategies for the catalytic asymmetric α-arylation of racemic carbonyl compounds. Among these approaches, the catalytic enantioselective α-arylation employing diaryliodonium salts as arylating agents stands out as a particularly efficient and robust method. This strategy provides a streamlined route to enantioenriched α-aryl carbonyl compounds—structurally important motifs with substantial synthetic value. In the following, we will introduce the related progress on the basis of the diaryliodonium salts as arylating agents employed.
In 1999, Ochiai et al. [79] pioneered the synthesis of chiral diaryliodonium salts (2-l) via a BF3·Et2O-catalyzed tin-λ3-iodane exchange. Concurrently, the group developed an asymmetric α-phenylation of cyclic β-ketoesters (48) employing a chiral binaphthyl(phenyl)iodonium salt as the aryl source. This transformation afforded the corresponding products (49) in 34–53% yield with enantioselectivities ranging from 34% to 53% ee. This work highlighted the feasibility of asymmetric α-arylation using chiral diaryliodonium salts. Although reports on the asymmetric synthesis of chiral diaryliodonium salts remain scarce, this strategy provides a novel approach for asymmetric α-phenylation and significantly expands the synthetic utility of diaryliodonium salts (Scheme 24).
In 2005, Olofsson’s group [80] reported a direct asymmetric α-arylation of cyclohexanones. This method involves the asymmetric enolization of 4-substituted cyclohexanones (50) using Simpkins’ base, followed by coupling with diaryliodonium salts to afford α-aryl ketones (51) in moderate-to-good yields with high enantioselectivity. Notably, this methodology was further applied to the short and efficient total synthesis of (-)-epibatidine (53) and the formal synthesis of (+)-epibatidine (55) (Scheme 25).
In 2011, MacMillan and co-workers [81] disclosed an enantioselective α-arylation of unactivated aldehydes (56) employing diaryliodonium salts as arylating agents. This transformation was achieved through a dual catalytic system comprising chiral amine catalyst TCA (Cat. 1, 10 mol%) and CuBr (10 mol%). This method achieved excellent yields (67–95%) and outstanding enantioselectivity (91–94% ee), delivering a series of α-chiral tertiary carbon-containing esters (57). These mild catalytic conditions provided a novel strategy for the enantioselective construction and retention of stereocenters at the enolizable α-formyl benzylic position. Furthermore, this group successfully applied this methodology as a key step to the rapid synthesis of (S)-ketoprofen (1-e), an oral topical analgesic (Scheme 26).
Building on this work, the group further explored the reaction mechanism (Scheme 27). They suggested that this tandem catalytic pathway starts with the condensation of amine catalyst (Cat. 1) with aldehyde (56) substrate to form activated chiral enamine Int 5, followed by the oxidative addition of Cu(I) to the C-I bond of diaryliodonium triflate system to form electron-deficient aryl Cu(III) species, which is then followed by a highly electrophilic Cu(III) aryl system that will be rapidly coordinated with the Re face of the chiral enamine to form the π-Cu complex (Int 6). After which, rapid bond isomerization occurs to form the η1-imino Int 7, which is eliminated by reduction to release the optically enriched α-site aryl imine, rebuild the Cu(I)Br catalyst, and hydrolyze the imine to give the desired α-site aryl aldehyde product (57), whilst allowing the organocatalyst to be released and participate in the next cycle.
In a nearly concurrent publication, MacMillan’s group [82] reported a copper-catalyzed asymmetric α-arylation reaction involving diaryliodonium salts. This reaction employs a preformed Cu complex (Cat. 2, derived from (S, S)-Ph-Box and CuOTf, 10 mol%), to achieve arylation of N-acyl oxazolidinones (60) derived from lactones or acyl oxazolidinones with diaryliodonium hexafluorophosphates, delivering arylative products (61). This reaction provided the arylative products in good-to-excellent yields (65–96%) and high enantioselectivity (87–95% ee) under mild conditions, offering a novel strategy for constructing stereocenters at enolizable α-carbonyl benzylic positions. To demonstrate its utility, the group applied this approach to the rapid enantioselective synthesis of the NSAID (S)-naproxen (1-f). Concurrently, Gaunt et al. [83] reported the arylation of N-acyl oxazolidinones (60) using a chiral Cu(II)-bis(oxazoline) (Cat. 3) catalytic system. Similarly effective under mild conditions, their method afforded excellent yields and enantioselectivity, delivering valuable building blocks (61-g–61-m) and enabling the synthesis of important NSAIDs and analogs. Notably, they achieved gram-scale synthesis of ibuprofen (1-g) with 99% yield and 93% ee (Scheme 28).
They proposed the following mechanism (Scheme 29): The ligand-bound Cu(I) complex undergoes oxidative insertion into the diaryliodonium salt, generating a highly electrophilic chiral Cu (III) species (Int 8). This intermediate is captured by the silyl enol ether nucleophile (60), followed by reductive elimination and hydrolysis of the silyl group (Int 9) to yield the desired α-aryl carbonyl product (61), while the Cu(I) catalyst is regenerated to participate in the next catalytic cycle.
In 2013, Feng and co-workers [84] demonstrated that chiral N, N′-dioxide-derived metal complexes also serve as effective catalysts for the enantioselective α-arylation of carbonyl enolates with electrophilic diaryliodonium salts (Scheme 30). Under the presence of NaHCO3, the combination of 10 mol% chiral N, N′-dioxide (L1), and 10 mol% Sc (OTf)3 complexes as a chiral Lewis acid efficiently enabled the enantioselective α-arylation of N-unprotected 3-substituted oxindoles (63) with diaryliodonium triflates. This reaction afforded oxindole derivatives bearing quaternary carbon centers (64) in good-to-excellent yields (up to 99%) and high enantioselectivity (up to 99% ee). Notably, this methodology was successfully applied to the enantioselective synthesis of an antiproliferative agent (1-d) for cancer treatment.
Three years later, this research group [85] achieved highly enantioselective α-arylation/vinylation/alkynylation of cyclic β-keto amides/esters (65) with diaryliodonium triflates, using a chiral N, N′-dioxide (L2)/Ni (OTf)2 complex as the catalyst. Under optimized conditions (10 mol% catalyst loading), a series of α-aryl/α-vinyl/α-alkynyl-substituted β-keto amides/esters (66) were synthesized with yields and enantiomeric excess (ee) values reaching up to 99% (Scheme 31).
In 2021, Orlandi’s group [86] discovered a novel catalytic system based on Cu(II) and a chiral bisphosphine dioxide ligand (L3), successfully achieving the asymmetric α-arylation of acyclic ketones at the carbonyl α-position (Scheme 32). Under mild catalytic conditions with 12 mol% of the chiral bisphosphine dioxide ligand and 4 mol% Cu(OTf)2·Tol, the reaction utilized silyl enol ethers of acyclic ketones (67) as substrates and diaryliodonium salts as arylating agents, delivering a series of α-chiral tertiary carbon-containing carbonyl compounds (68) in excellent yields with high enantioselectivity. The experimental results showed that the silyl enol ethers bearing electron-donating substituents in meta- or para-positions were found to give lower selectivity, likely due to the geometrical bias provided by the ortho-OMe group. Additionally, the steric hindrance next to the reaction site was detrimental, such as iPr or Bn groups (68-h), onto the nucleophilic α-carbon.
Up to date, the catalytic asymmetric α-arylation of carbonyl compounds using diaryliodonium salts have remarkably developed in the past two decades, this field of research is still in its early stages and daunting challenge. The research groups’ results showed that the substrates bearing electron-donating substituents in meta- or para-positions were found to give lower enantioselectivity, likely due to the geometrical bias provided by the ortho-OMe group (Table 1, Entry 3). Additionally, the steric hindrance next to the reaction site was detrimental, such as iPr, Bn groups, etc. (Table 1, Entry 2,4), and diaryliodonium salts featuring the ortho-positions were found to give lower reactivity (Table 1, Entry 4).
Based on the above asymmetric α-arylation reaction protocols. Scheme 33 summarizes the several possible mechanisms for α-arylation of carbonyl compounds reported so far. For example, (a) α-Arylation by cross-coupling, (b) α-arylation by rearrangement, and (c) C-linked intermediate and [1,2] rearrangement.
Currently, diaryliodonium salts are generally believed to react by the reductive elimination pathway, delivering the equatorial aryl moiety to the axially installed nucleophile (Scheme 33a). Furthermore, the involvement of electrophilic addition and aryl moiety rearrangement was also proposed. Olofsson and co-workers, through computational study, confirmed that the addition of diaryliodonium salts to the enolate results in almost isoenergetic oxygen–iodine- and carbon–iodine-bonded isomers and that reaction could follow associative or dissociative ligand-exchange pathways. The [2,3]-rearrangement pathway was favored over the [1,2] process (Scheme 33b) [87]. Nevertheless, Feng and co-workers [84] proposed that by introducing chiral Lewis acids with strong oxygen affinity, C-linked enantiomeric intermediates could be formed, which could then undergo [1,2] rearrangement to achieve asymmetric arylation (Scheme 33c). The proposal of this mechanism has important implications for the development of arylation reactions involving non-oxidative metal participation or organocatalysis.

4. Summary and Outlook

In summary, the great advantages of diaryliodonium salts, for example, with their low toxicity compared with heavy-metal reagents, mild reaction conditions, fast accessibility of a large variety of reagents, and easy handling, have established themselves as versatile arylation reagents in organic synthesis over recent decades. Their exceptional reactivity has driven significant progress in α-arylation strategies for carbonyl compounds, enabling transformations under mild conditions with remarkable functional group tolerance and regioselectivity. Furthermore, advancements in chiral ligand design, organocatalysis, and asymmetric catalytic systems have unlocked unprecedented potential for enantioselective α-arylation reactions using these iodonium salts. In spite of the remarkable achievements to date, this field of research is still in its early stages and brimming with potential opportunities for further synthetic exploration. For example: (a) The reaction mechanism for the arylation of diaryliodonium salts is worthy of further exploration [87]. (b) Chemoselectivity and regioselective migration in unsymmetric diaryliodonium salt reactions remains a critical challenge [66,88,89]. (c) The reaction involving diaryliodonium salts fails to meet the atom economy requirement, as only one aryl group undergoes productive coupling while the counterpart is converted into stoichiometric aryl iodide byproducts. This inherent stoichiometric inefficiency underscores the ongoing challenge of developing methodologies for dual-aryl utilization of these reagents [90,91,92,93,94]. (d) The current focus of synthetic applications lies in the preparation of bioactive compounds using diaryliodonium salts, and the total synthesis of natural products and functional materials using diaryliodonium salts as arylating agents is rarely reported. Thus, the further expansion of diaryliodonium salts as arylating agents in the preparation of the above-mentioned compounds is needed to be explored, especially in practical applications [63,95]. (e) The development of novel chiral ligands, organocatalysts, and chiral catalyst systems for the enantioselective α-arylation reactions of carbonyls and related compounds will persist as a pivotal research frontier. This is because the diverse access to optically active α-aryl carbonyl molecules will continue to be explored, and such exploration will be beneficial for the discovery and development of related drugs.

Author Contributions

Conceptualization, X.-W.C. and X.F.; methodology, J.-L.C. and H.Z.; software, X.-W.C. and J.-L.C.; writing—original draft preparation, X.-W.C., J.-L.C. and L.-H.Z.; writing—review and editing, X.-W.C. and X.F.; visualization, X.-W.C.; supervision, X.-W.C.; project administration, X.-W.C.; funding acquisition, X.-W.C. and X.C. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by Gansu Provincial Science and Technology Program under Grant No. 25YFWA016 and 24CXNJ013, Tianyou Youth Talent Lift Program of Lanzhou Jiaotong University (1520260514), and Young Scholars Science Foundation of Lanzhou Jiaotong University (1200061308).

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

We acknowledge financial support by Gansu Provincial Science and Technology Program, Tianyou Youth Talent Lift Program of Lanzhou Jiaotong University, and Young Scholars Science Foundation of Lanzhou Jiaotong University.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bellina, F.; Rossi, R. Transition metal-catalyzed direct arylation of substrates with activated sp3-hybridized C-H bonds and some of their synthetic equivalents with aryl halides and pseudohalides. Chem. Rev. 2010, 110, 1082–1146. [Google Scholar] [CrossRef] [PubMed]
  2. Aoki, S.; Watanabe, Y.; Sanagawa, M.; Setiawan, A.; Kotoku, N.; Kobayashi, M. Cortistatins A, B, C, and D, Anti-angiogenic steroidal alkaloids, from the marine sponge Corticium simplex. J. Am. Chem. Soc. 2006, 128, 3148–3149. [Google Scholar] [CrossRef] [PubMed]
  3. Doan, T.H.; Ho, V.D.; Le, T.B.H.; Le, T.A.; Pham, T.K.; Nguyen, T.H.; Raal, A. Two new abietane diterpenes huperphlegmarins A and B from Huperzia Phlegmaria. Nat. Prod. Res. 2019, 33, 2051–2059. [Google Scholar] [CrossRef] [PubMed]
  4. Ito, T.; Tanaka, T.; Iinuma, M.; Nakaya, K.; Takahashi, Y.; Sawa, R.; Murata, J.; Darnaedi, D. Three new resveratrol oligomers from the stem bark of Vatica pauciflora. J. Nat. Prod. 2004, 67, 932–937. [Google Scholar] [CrossRef] [PubMed]
  5. Mitchell, J.A.; Akarasereenont, P.; Thiemermann, C.; Flower, R.J.; Vane, J.R. Selectivity of nonsteroidal anti-inflammatory drugs as inhibitors of constitutive and inducible cyclooxygenase. Proc. Natl. Acad. Sci. USA 1993, 90, 11693–11697. [Google Scholar] [CrossRef] [PubMed]
  6. Katayev, D.; Kündig, E.P. Catalytic enantioselective synthesis of a 3-aryl-3-benzyloxindole (3-aryl-3-benzyl-1,3-dihydro-2H-indol-2-one) exhibiting antitumor activity. Helv. Chim. Acta 2012, 95, 2287–2295. [Google Scholar] [CrossRef]
  7. Khwaja, F.; Allen, J.; Lynch, J.; Andrews, P.; Djakiew, D. Ibuprofen inhibits survival of bladder cancer cells by induced expression of the p75NTR tumor suppressor protein. Cancer Res. 2004, 64, 6207–6213. [Google Scholar] [CrossRef] [PubMed]
  8. Palomer, A.; Pérez, J.J.; Navea, S.; Llorens, O.; Pascual, J.; García, L.; Mauleón, D. Modeling cyclooxygenase inhibition. Implication of active site hydration on the selectivity of ketoprofen analogues. J. Med. Chem. 2000, 43, 2280–2284. [Google Scholar] [CrossRef] [PubMed]
  9. Zanos, P.; Moaddel, R.; Morris, P.J.; Riggs, L.M.; Highland, J.N.; Georgiou, P.; Pereira, E.F.R.; Albuquerque, E.X.; Thomas, C.J.; Zarate, C.A.; et al. Ketamine and ketamine metabolite pharmacology: Insights into therapeutic mechanisms. Pharmacol. Rev. 2018, 70, 621–660. [Google Scholar] [CrossRef] [PubMed]
  10. Zhou, B.; Yang, S.; Zhou, X.; Chen, Q.; Tu, E.; Zhang, B.; Shi, L.; Zhou, X. Severe tremors induced by tiletamine e-cigarette and alcohol use: A case report. Front. Psychiatry 2025, 16, 1537822. [Google Scholar] [CrossRef] [PubMed]
  11. Han, B.; He, X.H.; Liu, Y.Q.; He, G.; Peng, C.; Li, J.L. Asymmetric organocatalysis: An enabling technology for medicinal chemistry. Chem. Soc. Rev. 2021, 50, 1522–1586. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, J.; Rueping, M. Metallaphotoredox catalysis for sp3C-H functionalizations through hydrogen atom transfer (HAT). Chem. Soc. Rev. 2023, 52, 4099–4120. [Google Scholar] [CrossRef] [PubMed]
  13. Chakraborty, P.; Mandal, R.; Garg, N.; Sundararaju, B. Recent advances in transition metal-catalyzed asymmetric electrocatalysis. Coord. Chem. Rev. 2021, 444, 214065–241088. [Google Scholar] [CrossRef]
  14. Biffis, A.; Centomo, P.; Del Zotto, A.; Zecca, M. Pd metal catalysts for cross-couplings and related reactions in the 21st century: A critical review. Chem. Rev. 2018, 118, 2249–2295. [Google Scholar] [CrossRef] [PubMed]
  15. Swathy, K.J.; Saranya, P.V.; Anilkumar, G. Recent advances in palladium-catalyzed α-arylation reactions. Appl. Organomet. Chem. 2024, 38, 7508. [Google Scholar] [CrossRef]
  16. Mamgain, R.; Sakthivel, K.; Singh, F.V. Recent advances in transition-metal-free arylation reactions involving hypervalent iodine salts. Beilstein J. Org. Chem. 2024, 20, 2891–2920. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, T.Q.; MacMillan, D.W.C. A metallaphotoredox strategy for the cross-electrophile coupling of α-chloro carbonyls with aryl halides. Angew. Chem. Int. Ed. 2019, 58, 14584–14588. [Google Scholar] [CrossRef] [PubMed]
  18. Pandey, G.; Tiwari, S.K.; Budakoti, A.; Sahani, P.K. Transition-metal-free photoredox intermolecular α-arylation of ketones. Org. Chem. Front. 2018, 5, 2610–2614. [Google Scholar] [CrossRef]
  19. Hossain, M.M.; Shaikh, A.C.; Moutet, J.; Gianetti, T.L. Photocatalytic α-arylation of cyclic ketones. Nat. Synth. 2022, 1, 147–157. [Google Scholar] [CrossRef]
  20. Hartmann, C.; Meyer, V. Ueber eine neue klasse jodhaltiger, stickstofffreier organischer basen. Ber. Dtsch. Chem. Ges. 1894, 27, 426–432. [Google Scholar] [CrossRef]
  21. Takenaga, N.; Kumar, R.; Dohi, T. Heteroaryliodonium(III) salts as highly reactive electrophiles. Front. Chem. 2020, 8, 599026. [Google Scholar] [CrossRef] [PubMed]
  22. Bielawski, M.; Olofsson, B. High-yielding one-pot synthesis of diaryliodonium triflates from arenes and iodine or aryl iodides. Chem. Commun. 2007, 24, 2521–2523. [Google Scholar] [CrossRef] [PubMed]
  23. 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. [Google Scholar] [CrossRef]
  24. Bielawski, M.; Krämer, K.; Olofsson, B. Efficent.one-pot synthesis of bis (4-tert butylphenyl) iodonium triflate. Org. Synth. 2009, 86, 308–314. [Google Scholar]
  25. 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. [Google Scholar] [CrossRef]
  26. 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. [Google Scholar] [CrossRef] [PubMed]
  27. Cardinale, J.; Ermert, J.; Coenen, H.H. Convenient preparation of (4-iodophenyl) aryliodonium salts. Tetrahedron 2012, 68, 4112–4116. [Google Scholar] [CrossRef]
  28. Chun, J.-H.; Pike, V.W. Regiospecific syntheses of functionalized diaryliodonium tosylates via [hydroxy(tosyloxy)iodo] arenes generated in situ from (Diacetoxyiodo) arenes. J. Org. Chem. 2012, 77, 1931–1938. [Google Scholar] [CrossRef] [PubMed]
  29. Ito, M.; Itani, I.; Toyoda, Y.; Morimoto, K.; Dohi, T.; Kita, Y. Synthesis of boron-substituted diaryliodonium salts and selective transformation into functionalized aryl boronates. Angew. Chem. Int. Ed. 2012, 51, 12555–12558. [Google Scholar] [CrossRef] [PubMed]
  30. Jalalian, N.; Olofsson, B. Design and asymmetric synthesis of chiral diaryliodonium salts. Tetrahedron 2010, 66, 5793–5800. [Google Scholar] [CrossRef]
  31. Yoshimura, A.; Zhdankin, V.V. Advances in synthetic applications of hypervalent iodine compounds. Chem. Rev. 2016, 116, 3328–3435. [Google Scholar] [CrossRef] [PubMed]
  32. Peng, J.; Chen, C.; Xi, C.J. β-Arylation of oxime ethers using diaryliodonium salts through activation of inert C(sp3)–H bonds using a palladium catalyst. Chem. Sci. 2016, 7, 1383–1387. [Google Scholar] [CrossRef] [PubMed]
  33. Peng, J.; Li, M.; Xi, C.J.; Chen, C. Direct vicinal disubstitution of diaryliodonium salts by pyridine N-oxides and N-amidates by a 1,3-radical rearrangement. Angew. Chem. Int. Ed. 2013, 52, 7574–7578. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, Y.; Li, M.; Chen, C. Copper (II)-catalyzed three-component cascade annulation of diaryliodoniums, nitriles, and alkynes: A regioselective synthesis of multiply substituted quinolines. Angew. Chem. Int. Ed. 2013, 52, 5323–5327. [Google Scholar] [CrossRef] [PubMed]
  35. Cao, C.K.; Sheng, J.; Chen, C. Cu-catalyzed cascade annulation of diaryliodonium salts and nitriles: Synthesis of nitrogen-containing heterocycles. Synlett 2017, 49, 5081–5092. [Google Scholar] [CrossRef]
  36. Merritt, E.A.; Olofsson, B. Diaryliodonium salts: A journey from obscurity to fame. Angew. Chem. Int. Ed. 2009, 48, 9052–9070. [Google Scholar] [CrossRef] [PubMed]
  37. Yusubov, M.S.; Maskaev, A.V.; Zhdankin, V.V. Iodonium salts in organic synthesis. Arkivoc 2011, 2011, 370–409. [Google Scholar] [CrossRef]
  38. Aradi, K.; Tóth, B.; Tolnai, G.; Novák, Z. Diaryliodonium salts in organic syntheses: A useful compound class for novel arylation strategies. Synlett 2016, 27, 1456–1485. [Google Scholar] [CrossRef]
  39. Fañanás-Mastral, M. Copper-catalyzed arylation with diaryliodonium salts. Synthesis 2017, 49, 1905–1930. [Google Scholar] [CrossRef]
  40. Stuart, D.R. Aryl transfer selectivity in metal-free reactions of unsymmetrical diaryliodonium salts. Chem. Eur. J. 2017, 23, 15852–15863. [Google Scholar] [CrossRef] [PubMed]
  41. Singhal, R.; Choudhary, S.P.; Malik, B.; Pilania, M. Cyclic diaryliodonium salts: Applications and overview. Org. Biomol. Chem. 2023, 21, 4358–4378. [Google Scholar] [CrossRef] [PubMed]
  42. Xiao, Z.; Xia, C. Progresses of diaryliodonium salts in organic reactions. Chin. J. Org. Chem. 2013, 33, 2119–2130. [Google Scholar] [CrossRef]
  43. Wang, D.; Li, Q.; Li, M.; Du, Z.; Fu, Y. Recent progress in arylation reactions with diaryliodonium salts. Curr. Org. Chem. 2021, 25, 1298–1320. [Google Scholar] [CrossRef]
  44. Silva, L.F. Hypervalent iodine–mediated ring contraction reactions. Molecules 2006, 21, 421–434. [Google Scholar] [CrossRef] [PubMed]
  45. Karandikar, S.S.; Bhattacharjee, A.; Metze, B.E.; Javaly, N.; Valente, E.J.; McCormick, T.M.; Stuart, D.R. Orbital analysis of bonding in diarylhalonium salts and relevance to periodic trends in structure and reactivity. Chem. Sci. 2022, 13, 653–6540. [Google Scholar] [CrossRef] [PubMed]
  46. Le Du, E.; Waser, J. Recent progress in alkynylation with hypervalent iodine reagents. Chem. Commun. 2023, 59, 1589–1604. [Google Scholar] [CrossRef] [PubMed]
  47. Yoshimura, A.; Zhdankin, V.V. Recent progress in synthetic applications of hypervalent iodine (III) reagents. Chem. Rev. 2024, 124, 11108–11186. [Google Scholar] [CrossRef] [PubMed]
  48. Dong, D.-Q.; Hao, S.-H.; Wang, Z.-L.; Chen, C. Hypervalent iodine: A powerful electrophile for asymmetric α-functionalization of carbonyl compounds. Org. Biomol. Chem. 2014, 12, 4278–4289. [Google Scholar] [CrossRef] [PubMed]
  49. Olofsson, B. Arylation with diaryliodonium salts. In Hypervalent Iodine Chemistry; Wirth, T., Ed.; Topics in Current Chemistry; Springer: Cham, Switzerland, 2015; Volume 373, pp. 135–166. [Google Scholar]
  50. Prier, C.K.; Rankic, D.A.; MacMillan, D.W.C. Visible light photoredox catalysis with transition metal complexes: Applications in organic synthesis. Chem. Rev. 2013, 113, 5322–5363. [Google Scholar] [CrossRef] [PubMed]
  51. Romero, N.A.; Nicewicz, D.A. Organic photoredox catalysis. Chem. Rev. 2016, 116, 10075–10166. [Google Scholar] [CrossRef] [PubMed]
  52. 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, 92–100. [Google Scholar] [CrossRef] [PubMed]
  53. Hao, Y.-J.; Hu, X.-S.; Zhou, Y.; Zhou, J.; Yu, J.-S. Catalytic enantioselective α-arylation of carbonyl enolates and related compounds. ACS Catal. 2020, 10, 955–993. [Google Scholar] [CrossRef]
  54. Singh, P.R.; Banerjee, A.; Simlandy, A.K. Advances in catalytic enantioselective transformations using diaryliodonium reagents. ACS Catal. 2025, 15, 3096–3115. [Google Scholar] [CrossRef]
  55. Marshall Beringer, F.; Forgione, P.S.; Yudis, M.D. Diaryliodonium salts—XII: The phenylation of dimedone, dibenzoylmethane and tribenzoylmethane. Tetrahedron 1960, 8, 49–63. [Google Scholar] [CrossRef]
  56. Beringer, F.M.; Galton, S.A.; Huang, S.J. Diaryliodonium salts. XVII. The phenylation of 1,3-indandiones. J. Am. Chem. Soc. 1962, 84, 2819–2823. [Google Scholar] [CrossRef]
  57. Beringer, F.M.; Daniel, W.J.; Galton, S.A.; Rubin, G. Phenylation of monoketones with diphenyliodonium chloride. J. Org. Chem. 1966, 31, 4315–4318. [Google Scholar] [CrossRef]
  58. Ryan, J.H.; Stang, P.J. Direct α-arylation of ketones: The reaction of cyclic ketone enolates with diphenyliodonium triflate. Tetrahedron Lett. 1997, 38, 5061–5064. [Google Scholar] [CrossRef]
  59. Liu, Z.-J.; Han, J.; Zhang, Y. tert-Butoxide-mediated arylation of 1-acetylindolin-3-ones with diaryliodonium salts. Synlett 2015, 26, 2593–2597. [Google Scholar]
  60. Pan, J.-L.; Chen, T.; Zhang, Z.-Q.; Li, Y.-F.; Zhang, X.-M.; Zhang, F.-M. A Cu-mediated one-pot Michael addition/α-arylation strategy using diaryliodonium salt: Direct and efficient approach to α-aryl-β substituted cyclic ketone scaffolds. Chem. Commun. 2016, 52, 2382–2385. [Google Scholar] [CrossRef] [PubMed]
  61. Peng, X.; Luo, H.; Wu, F.; Zhu, D.; Ganesan, A.; Huang, P.; Wen, S. Synthesis of fluorenes with an all-carbon quaternary center via palladium-catalyzed dual arylation using cyclic diaryliodonium triflates. Adv. Syn. Catal. 2017, 359, 1152–1156. [Google Scholar] [CrossRef]
  62. 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. [Google Scholar] [CrossRef] [PubMed]
  63. Chen, X.-W.; Hou, Z.-C.; Chen, C.; Zhang, L.-H.; Chen, M.-E.; Zhang, F.-M. Enantioselective total syntheses of six natural and two proposed meroterpenoids from Psoralea Corylifolia. Chem. Sci. 2023, 14, 5699–5704. [Google Scholar] [CrossRef] [PubMed]
  64. Beringer, F.M.; Forgione, P.S. Diaryliodonium salts. XVIII. The phenylation of esters in t-butyl alcohol1-3. J. Org. Chem. 1963, 28, 714–717. [Google Scholar] [CrossRef]
  65. Chen, Z.; Jin, Y.; Stang, P.J. Polyvalent iodine in synthesis. 1. An efficient route to isopropylidene arylmalonates (5-aryl-substituted Meldrum’s acid). J. Org. Chem. 1987, 52, 4115–4117. [Google Scholar] [CrossRef]
  66. Oh, C.H.; Kim, J.S.; Jung, H.H. Highly efficient arylation of malonates with diaryliodonium salts. J. Org. Chem. 1999, 64, 1338–1340. [Google Scholar] [CrossRef]
  67. 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. [Google Scholar] [CrossRef] [PubMed]
  68. 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. [Google Scholar] [CrossRef] [PubMed]
  69. 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. [Google Scholar] [CrossRef] [PubMed]
  70. 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)-quinolones. J. Org. Chem. 2015, 80, 2513–2520. [Google Scholar] [CrossRef] [PubMed]
  71. Dey, C.; Lindstedt, E.; Olofsson, B. Metal-free C-arylation of nitro compounds with diaryliodonium salts. Org. Lett. 2015, 17, 4554–4557. [Google Scholar] [CrossRef] [PubMed]
  72. Qian, X.; Han, J.; Wang, L. tert-Butoxide-mediated arylation of 2-substituted cyanoacetates with diaryliodonium salts. Adv. Syn. Catal. 2016, 358, 940–946. [Google Scholar] [CrossRef]
  73. 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. [Google Scholar] [CrossRef]
  74. 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. [Google Scholar] [CrossRef]
  75. 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. [Google Scholar] [CrossRef] [PubMed]
  76. Bouquin, M.; Jaroschik, F.; Taillefer, M. Versatile and base-free copper-catalyzed α-arylations of aromatic ketones using diaryliodonium salts. Tetrahedron Lett. 2021, 75, 153208–153212. [Google Scholar] [CrossRef]
  77. 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. [Google Scholar] [CrossRef] [PubMed]
  78. 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. [Google Scholar] [CrossRef] [PubMed]
  79. 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. [Google Scholar] [CrossRef]
  80. Aggarwal, V.K.; Olofsson, B. Enantioselective α-arylation of cyclohexanones with diaryliodonium salts: Application to the synthesis of (-)-epibatidine. Angew. Chem. Int. Ed. 2005, 44, 5516–5519. [Google Scholar] [CrossRef] [PubMed]
  81. Allen, A.E.; MacMillan, D.W.C. Enantioselective α-Arylation of aldehydes via the productive merger of iodonium salts and organocatalysis. J. Am. Chem. Soc. 2011, 133, 4260–4263. [Google Scholar] [CrossRef] [PubMed]
  82. Harvey, J.S.; Simonovich, S.P.; Jamison, C.R.; MacMillan, D.W. Enantioselective α-arylation of carbonyls via Cu(I)-bisoxazoline catalysis. J. Am. Chem. Soc. 2011, 133, 13782–13785. [Google Scholar] [CrossRef] [PubMed]
  83. Bigot, A.; Williamson, A.E.; Gaunt, M.J. Enantioselective α-arylation of N-acyloxazolidinones with copper(II)-bisoxazoline catalysts and diaryliodonium salts. J. Am. Chem. Soc. 2011, 133, 13778–13781. [Google Scholar] [CrossRef] [PubMed]
  84. Guo, J.; Dong, S.; Zhang, Y.; Kuang, Y.; Liu, X.; Lin, L.; Feng, X. Chiral scandium(III)-catalyzed enantioselective α-arylation of N-unprotected 3-substituted oxindoles with diaryliodonium salts. Angew. Chem. Int. Ed. 2013, 52, 10245–10249. [Google Scholar] [CrossRef] [PubMed]
  85. Guo, J.; Lin, L.; Liu, Y.; Li, X.; Liu, X.; Feng, X. Nickel(II)-catalyzed enantioselective α-vinylation of β-keto amides/esters with hypervalent iodine salts. Org. Lett. 2016, 18, 5540–5543. [Google Scholar] [CrossRef] [PubMed]
  86. Escudero-Casao, M.; Licini, G.; Orlandi, M. Enantioselective α-arylation of ketones via a novel Cu(I)-bis(phosphine) dioxide catalytic system. J. Am. Chem. Soc. 2021, 143, 3289–3294. [Google Scholar] [CrossRef] [PubMed]
  87. 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. [Google Scholar] [CrossRef] [PubMed]
  88. Wang, B.; Graskemper, J.W.; Qin, L.; DiMagno, S.G. Regiospecific reductive elimination from diaryliodonium salts. Angew. Chem. Int. Ed. 2010, 49, 4079–4083. [Google Scholar] [CrossRef] [PubMed]
  89. 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. [Google Scholar] [CrossRef] [PubMed]
  90. Trost, B.M. The atom economy-A search for synthetic efficiency. Science 1991, 254, 1471–1477. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, M.; Chen, S.; Jiang, X. Atom-economical applications of diaryliodonium salts. Chem. Asian J. 2018, 13, 2195–2207. [Google Scholar] [CrossRef] [PubMed]
  92. Modha, S.G.; Greaney, M.F. Atom-Economical Transformation of Diaryliodonium Salts: Tandem C−H and N−H Arylation of Indoles. J. Am. Chem. Soc. 2015, 137, 1416–1419. [Google Scholar] [CrossRef] [PubMed]
  93. 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. [Google Scholar] [CrossRef]
  94. Wu, T.; Greaney, M.F. Chemoselective cascade arylation. Chem 2022, 8, 609–611. [Google Scholar] [CrossRef]
  95. Silva, L.F., Jr.; Olofsson, B. Hypervalent iodine reagents in the total synthesis of natural products. Nat. Prod. Rep. 2011, 28, 1722–1754. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Selected natural products and bioactive molecules bearing an α-aryl carbonyl scaffold.
Figure 1. Selected natural products and bioactive molecules bearing an α-aryl carbonyl scaffold.
Molecules 30 03019 g001
Figure 2. (a) General and T-shaped structure. (b) Orbitals in the hypervalent ω bond. (c) α-arylation of carbonyl compounds using diaryliodonium salts.
Figure 2. (a) General and T-shaped structure. (b) Orbitals in the hypervalent ω bond. (c) α-arylation of carbonyl compounds using diaryliodonium salts.
Molecules 30 03019 g002
Scheme 1. α-Phenylation of ketones.
Scheme 1. α-Phenylation of ketones.
Molecules 30 03019 sch001
Scheme 2. Proposed radical-pair mechanism.
Scheme 2. Proposed radical-pair mechanism.
Molecules 30 03019 sch002
Scheme 3. Cu-mediated α-phenylation of cyclic ketones.
Scheme 3. Cu-mediated α-phenylation of cyclic ketones.
Molecules 30 03019 sch003
Scheme 4. α-Arylation of indolone derivatives.
Scheme 4. α-Arylation of indolone derivatives.
Molecules 30 03019 sch004
Scheme 5. Cu-mediated one-pot Michael addition/α-arylation of an enone.
Scheme 5. Cu-mediated one-pot Michael addition/α-arylation of an enone.
Molecules 30 03019 sch005
Scheme 6. The mechanism of Cu-mediated one-pot Michael addition/α-arylation of an enone.
Scheme 6. The mechanism of Cu-mediated one-pot Michael addition/α-arylation of an enone.
Molecules 30 03019 sch006
Scheme 7. Synthesis of fluorene and spirofluorene derivatives.
Scheme 7. Synthesis of fluorene and spirofluorene derivatives.
Molecules 30 03019 sch007
Scheme 8. α-Arylation of α-nitro ketones and synthesis of tiletamine.
Scheme 8. α-Arylation of α-nitro ketones and synthesis of tiletamine.
Molecules 30 03019 sch008
Scheme 9. Total synthesis of several meroterpenoid natural products.
Scheme 9. Total synthesis of several meroterpenoid natural products.
Molecules 30 03019 sch009
Scheme 10. α-Phenylation of ester carbonyl compounds and isopropyl malonate.
Scheme 10. α-Phenylation of ester carbonyl compounds and isopropyl malonate.
Molecules 30 03019 sch010
Scheme 11. α-Arylation of β-ketoester.
Scheme 11. α-Arylation of β-ketoester.
Molecules 30 03019 sch011
Scheme 12. α-Pentafluorophenylation of β-ketoesters and β-ketoamides.
Scheme 12. α-Pentafluorophenylation of β-ketoesters and β-ketoamides.
Molecules 30 03019 sch012
Scheme 13. α-Arylation with SF5-aryl-λ3-iodanes and SO2CF3-containing aryl/pyridyl groups.
Scheme 13. α-Arylation with SF5-aryl-λ3-iodanes and SO2CF3-containing aryl/pyridyl groups.
Molecules 30 03019 sch013
Scheme 14. α-Arylation of ethyl acetoacetate and synthesis of ELQ-300.
Scheme 14. α-Arylation of ethyl acetoacetate and synthesis of ELQ-300.
Molecules 30 03019 sch014
Scheme 15. α-Arylation of α-nitro esters and ethyl 2-cyanobutanoate derivatives.
Scheme 15. α-Arylation of α-nitro esters and ethyl 2-cyanobutanoate derivatives.
Molecules 30 03019 sch015
Scheme 16. Synthesis of fluorene and spirofluorene derivatives via cyclic diaryliodonium salts.
Scheme 16. Synthesis of fluorene and spirofluorene derivatives via cyclic diaryliodonium salts.
Molecules 30 03019 sch016
Scheme 17. α-Phenylation of silyl enol ethers.
Scheme 17. α-Phenylation of silyl enol ethers.
Molecules 30 03019 sch017
Scheme 18. The mechanisms of α-phenylation of silyl enol ethers.
Scheme 18. The mechanisms of α-phenylation of silyl enol ethers.
Molecules 30 03019 sch018
Scheme 19. Synthesis of carbocyclic-thickened indoles.
Scheme 19. Synthesis of carbocyclic-thickened indoles.
Molecules 30 03019 sch019
Scheme 20. α-Arylation of arylketone-derived enolsilanes and possible mechanism.
Scheme 20. α-Arylation of arylketone-derived enolsilanes and possible mechanism.
Molecules 30 03019 sch020
Scheme 21. α-Arylation of α-nitro-fluoroacetamides and α-cyano-α-fluoroacetamides.
Scheme 21. α-Arylation of α-nitro-fluoroacetamides and α-cyano-α-fluoroacetamides.
Molecules 30 03019 sch021
Scheme 22. α-Arylation of α-aceto-fluoroacetamide.
Scheme 22. α-Arylation of α-aceto-fluoroacetamide.
Molecules 30 03019 sch022
Scheme 23. Base-promoted reaction mechanism for the α-arylation of α-fluoroacetamide.
Scheme 23. Base-promoted reaction mechanism for the α-arylation of α-fluoroacetamide.
Molecules 30 03019 sch023
Scheme 24. Asymmetric α-phenylation involving chiral diaryliodonium salts.
Scheme 24. Asymmetric α-phenylation involving chiral diaryliodonium salts.
Molecules 30 03019 sch024
Scheme 25. Simpkins’ base-mediated asymmetric α-arylation.
Scheme 25. Simpkins’ base-mediated asymmetric α-arylation.
Molecules 30 03019 sch025
Scheme 26. Enantioselective α-arylation catalyzed by chiral amine catalyst TCA.
Scheme 26. Enantioselective α-arylation catalyzed by chiral amine catalyst TCA.
Molecules 30 03019 sch026
Scheme 27. Catalytic mechanism of the chiral amine catalyst TCA.
Scheme 27. Catalytic mechanism of the chiral amine catalyst TCA.
Molecules 30 03019 sch027
Scheme 28. Asymmetric α-arylation catalyzed by a chiral Cu(II)-bis(oxazoline) complex.
Scheme 28. Asymmetric α-arylation catalyzed by a chiral Cu(II)-bis(oxazoline) complex.
Molecules 30 03019 sch028
Scheme 29. Mechanism of cooperative catalysis reaction by ligand-bound Cu(I) complex.
Scheme 29. Mechanism of cooperative catalysis reaction by ligand-bound Cu(I) complex.
Molecules 30 03019 sch029
Scheme 30. Enantioselective α-Arylation of N-unprotected 3-substituted oxindoles.
Scheme 30. Enantioselective α-Arylation of N-unprotected 3-substituted oxindoles.
Molecules 30 03019 sch030
Scheme 31. Enantioselective α-arylation/vinylation/alkynylation of cyclic β-keto amides/esters.
Scheme 31. Enantioselective α-arylation/vinylation/alkynylation of cyclic β-keto amides/esters.
Molecules 30 03019 sch031
Scheme 32. Enantioselective α-arylation of acyclic ketones.
Scheme 32. Enantioselective α-arylation of acyclic ketones.
Molecules 30 03019 sch032
Scheme 33. Possible mechanism for α-arylation of carbonyl compounds.
Scheme 33. Possible mechanism for α-arylation of carbonyl compounds.
Molecules 30 03019 sch033
Table 1. Synthetic protocols and strategy for enantioselective α-arylation of carbonyl compounds.
Table 1. Synthetic protocols and strategy for enantioselective α-arylation of carbonyl compounds.
Molecules 30 03019 i001
EntrySubstrateCatalyst SystemsEnantioselectivityEfficiencyFunctional Group ToleranceSubstrate ScopeRef.
SubstrateDiaryliodonium Salts
1Molecules 30 03019 i002Combination of copper and organic catalysts
(Cat. 3·TCA and CuBr)
90–94% eeCat. 3·TCA (10–40 mol%)
CuBr (5–10 mol%)
NaHCO3, 23 °C
67–95% yield
CO2Et, NBoc, NHCbz, Alkene, etc.F, CF3, OMe, NO2, etc.R = Alkyl, iPr, Bn;
Ar1 = Aryl, Thiophene, Pyridine, Naphthalene group, etc.
[81]
2Molecules 30 03019 i003Cu(II)-bis(oxazoline)17–94% eeCat. 3 (5–10 mol%),
0–10 °C, 6–20 h,
40–99% yield
NMeCbz, Alkene, etc.CO2Et, F, CF3, OMe, etc.R = Alkyl, Aryl, Indole, iPr, Bn;
Ar1 = Aryl, Thiophene, Naphthalene group, etc.
[82]
390–94% eeCat. 4 (10–20 mol%),
−20–0 °C, 24 h
65–96% yield
NMeCbz, Alkene, etc.CO2Et, F, CF3, OMe, etc.R = Alkyl, Indole, iPr, Bn;
Ar1 = Aryl, Thiophene, Naphthalene group, etc.
[83]
4Molecules 30 03019 i004Cu(I)−
Bis(phosphine) Dioxide
45–95% eeL3 (12 mol%)
Cu(OTf)2·Tol (4 mol%)
rt, 16 h
20–90% yield
NCbz, F, CF3, OMe, NO2, etc.CO2Et, F, Cl, Br, CF3, OMe, NO2, etc.R = Alkyl, iPr, Bn, Tetralone;
Ar1 = Aryl, Pyridine group, etc.
[86]
5Molecules 30 03019 i005N,N′-dioxide-Sc(OTf)3 complex17–99% eeSc(OTf)3/L1 =1:1 10 mol%, 3 Å M.S.,
NaHCO3, 35 °C
35 °C, 24–120 h
47–99% yield
F, Cl, Br, CN, OMe, NO2, etc.F, Cl, Br, Ph, Me, etc.R = Alkyl, 2-naphthylmethyl;
Ar1 = Aryl, Pyridine group, etc.
[84]
6Molecules 30 03019 i006N,N′-dioxide (L2)/Ni(OTf)2 complex91–99% eeL-PisEPh/Ni(OTf)2 (1:1.2; 10 mol %) Na2CO3, 35 °C,
24–120 h
48–99% yield
F, Cl, Br, Ph, etc.F, Cl, Br, Ph, Me, etc.Ar1 = Aryl group[85]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, X.-W.; Chen, J.-L.; Zhang, L.-H.; Zhang, H.; Chen, X.; Fan, X. Advances in α-Arylation of Carbonyl Compounds: Diaryliodonium Salts as Arylating Agents. Molecules 2025, 30, 3019. https://doi.org/10.3390/molecules30143019

AMA Style

Chen X-W, Chen J-L, Zhang L-H, Zhang H, Chen X, Fan X. Advances in α-Arylation of Carbonyl Compounds: Diaryliodonium Salts as Arylating Agents. Molecules. 2025; 30(14):3019. https://doi.org/10.3390/molecules30143019

Chicago/Turabian Style

Chen, Xiao-Wei, Jia-Le Chen, Ling-Hui Zhang, Huhu Zhang, Xiaojun Chen, and Xiaohui Fan. 2025. "Advances in α-Arylation of Carbonyl Compounds: Diaryliodonium Salts as Arylating Agents" Molecules 30, no. 14: 3019. https://doi.org/10.3390/molecules30143019

APA Style

Chen, X.-W., Chen, J.-L., Zhang, L.-H., Zhang, H., Chen, X., & Fan, X. (2025). Advances in α-Arylation of Carbonyl Compounds: Diaryliodonium Salts as Arylating Agents. Molecules, 30(14), 3019. https://doi.org/10.3390/molecules30143019

Article Metrics

Back to TopTop