Open Access Article
Ryan A. Pohorenec
a and
Shiqing Xu
*ab
aDepartment of Chemistry, College of Arts and Sciences, Texas A&M University, College Station, TX 77843, USA. E-mail: shiqing.xu@tamu.edu
bDepartment of Pharmaceutical Sciences, Irma Lerma Rangel College of Pharmacy, Texas A&M University, College Station, Texas 77843, USA
First published on 18th December 2025
Functionalization of the C(sp2)–H bond is a challenging yet highly useful synthetic task due to its relative inertness and commonality among functional groups. Arene thianthrenation is an emerging C(sp2)–H bond activation strategy featuring high chemo- and regioselectivity in the modification of structurally diverse substrates. The resultant aryl thianthrenium (Ar–TT+) salts exhibit divergent single- and two-electron reactivity modes, which have enabled new C(sp2)–bond formation to the main group elements through mechanistically distinct routes. This review complements the existing literature on the two-electron behavior of Ar–TT+ salts by highlighting the enhanced control of radical reactivity achieved in the site-selective thianthrenation of arenes and subsequent C(sp2)–S bond homolysis to generate aryl radicals. Discussion of the structural and electronic properties of Ar–TT+ salts promoting single-electron reactivity under mild conditions, functional group tolerance, enhanced platform cross-compatibility, and increased potential for sustainability through catalysis is a thematic focus throughout. Identification of these prior methodological advancements culminates in a prospectus on the opportunities for future reaction development.
Thianthrenation as a C–H activation strategy features high chemo- and regioselectivity in the modification of structurally diverse substrates under mild reaction conditions.12 Activation of Csp3–H, Csp2–H, and Csp–H bonds by this approach has enabled new carbon bond formation to the non-metallic main group elements through multiple mechanistically distinct pathways. Further, this strategy enables reactivity complementary to other “onium”-based C–H functionalization platforms13 with superior selectivity and scope of compatible substrates.
Earlier accounts from the pioneering Ritter14 and Wang15 laboratories provided overviews of aryl sulfonium salt reactivity, and Ritter has recently expanded their work to highlight the recent growth in the synthetic utilization of Cspn-substituted thianthrenium salts.12 Shu has similarly outlined the synthesis and reactivity of Cspn-substituted thianthrenium reagents.16 Alkene Csp2-substituted thianthrenium salts display a wide range of distinct and synthetically valuable reactivity modes,17–30 as recently reviewed by Wickens.31 Chen reviewed the access to aryl radical intermediates from a variety of sulfonium salts.32 The field has continued to advance rapidly, with substantial progress made in understanding and exploiting the radical chemistry of thianthrenium salts. In this review, we focus specifically on the reactivity of aryl thianthrenium salts as aryl radical precursors, with an emphasis on the enhanced control of carbon-centered radical reactivity enabled by stepwise site-selective thianthrenation and subsequent functionalization. By providing a mechanistically focused overview of recent advances, we aim to clarify reactivity patterns and inspire continued innovation in this rapidly expanding area of research.
Synthetic access to and utilization of various subclasses of sulfonium salts have been recently reviewed.46,47 It is useful to distinguish cyclic sulfoniums, in which the sulfur atom contributes to a ring system, from their acyclic counterparts, as well as to refer to the most labile and reactive bond between sulfur and its substituents (e.g. “Cspn-substituted”), as reactivity patterns are distinct among these groups. While many sulfonium salts currently utilized synthetically are carbon-substituted, heteroatom-substituted sulfonium reagents have also been prepared.48 Among the common cyclic sulfonium salts are those constructed on the thianthrene (TT), dibenzothiophene (DBT),49 and phenoxathiin backbone. Notable transformations of Csp2- and Csp-substituted sulfoniums also involve cycloaddition reactions.46 Various cross-couplings of organosulfoniums have been reported, including Suzuki–Miyaura, Mizoroki–Heck,50 and Sonogashira reactions.51 Computational (DFT) study of the lowest unoccupied molecular orbital (LUMO) of aryl thianthrenium salts revealed antibonding character of the exocyclic Csp2–S bond, promoting selective reactivity at this site.52 Accordingly, the BDE of the exocyclic Csp2–S bond in Ph–TT+ has been calculated by DFT as 64.2 kcal mol−1.53 Endocyclic Csp2–S cleavage occurring through a ring-opening attack by a strong nucleophile, such as a carbanion equivalent, has been observed.54,55
The redox profile of thianthrene includes two reversible single-electron oxidations with Eox(TT+/TT) = +1.26 V and Eox(TT2+/TT+) = +1.77 V in MeCN vs. SCE.56 The singly-oxidized radical cation (TT+) was found to be air and water stable after 13 days in a solution of 96% H2SO4, as demonstrated by Shine and Petite's ultraviolet-visible (UV-vis) spectroscopic analysis. Early experimental evidence for the formation of thianthrene radical cation resulted from the first accurate EPR studies in 1962, undertaken by independent groups of Shine and Petite,57 Kinoshita,58 and Lucken.59 These experiments demonstrated five-line spectrum with isotropic hyperfine splitting of magnitude 1.32 G, consistent with spin-nuclear interactions with four equivalent hydrogen atoms. Characterization by single-crystal X-ray diffraction of the thianthrenium tetrachloroaluminate salt was achieved in 1994 by Bock and colleagues.60
Despite the early foundations set forth by these studies, thianthrenium salts were synthetically underutilized until only recently. A PubMed search on 24 November 2025 for “thianthrene cation” or “thianthrenium” or “thianthrenation” yielded 202 total results, with marked growth in annual reports since the year 2019 (Fig. 1). A structure-based search of the Reaxys database for aryl thianthrenium compounds on 16 August 2025 yielded results for 973 substances and 1321 reactions, as reported within 1973 documents divided among 215 primary or review articles and 1765 patents. The recently renewed interest in thianthrenation as a C–H functionalization strategy has largely been driven by the development of synthetic approaches to thianthrenium salt precursors, accompanied by mechanistic insights, recognition of favorable redox activity, compatibility with all major platforms for energy input, and demonstrated versatility in new-bond construction under mild conditions.
Following the preliminary works of the middle 20th century, a relative paucity in work occurred until the Dawn of the 21st century. Nine years after their unambiguous spectral characterization of thianthrene radical cation salts, the Shine group prepared aryl thianthrenium compounds by reaction of the explosive thianthrenium perchlorate salt with methylbenzene, methoxybenzene, and anisole.61 This work was extended to disubstituted benzenes three years later in 1974.62 Two decades later, [3 + 2] cycloadditions between the thianthrenium radical cation and olefins63,64 as well as additions to alkynes were observed.65,66 Cycloadducts of thianthrenium perchlorate salts and alkenes were later shown to decompose to oxazolines in the presence of acetonitrile and water,67 with a subsequent report demonstrating the capability for electrochemical reactivity and potential for regeneration of thianthrene upon formation of decomposition products.68 A paucity in new annual reports on the synthetic use of thianthrenium salts ensued, until 2019 (Fig. 1), presumably due to hazards associated with the perchlorate salts and difficulty in their purification. The preparation of the less hazardous thianthrenium tetrafluoroborate salt was described in 1988.69
The breakthrough came in 2019 when the Ritter group reported a new synthetic strategy that enabled continued development of the thianthrenation platform for Csp2–H activation and functionalization.70 When treated with trifluoroacetic anhydride (TFAA), (tetrafluoro)thianthrene (TFT or TT), and tetrafluoroboric acid (HBF4), (tetrafluoro)thianthrene S-oxide (TFTSO or TTSO), various arenes were readily converted to the corresponding thianthrenium tetrafluoroborate salts via Csp2–H thianthrenation. In the presence of the monosubstituted arenes such as ethylbenzene, high para-selectivity (>99%) was observed for Csp2–H thianthrenation. A similar result was obtained in the direct electrolysis of TFT in the presence of arene. Numerous synthetic applications were demonstrated in this report, including Suzuki, Negishi, Mizoroki–Heck, and Sonogashira couplings, in addition to carbonylation, halogenation, phosphonylation, and borylation. The initial mechanistic proposal for the formation of aryl thianthrenium salts involved the formation of a TFT+–/TT+–TFA adduct that, or TT, generated the radical cation (TFT˙+ or TT˙+), which subsequently reacts with arene to produce a dicationic Wheland intermediate prior to a selectivity-determining irreversible deprotonation event.70
Additional experimental and computational evidence later suggested that the dication (TT2+) is the reactive species toward arene, and it is generated by heterolysis of a TT+–TFA adduct (Scheme 1). Single electron transfer (SET) between TT2+ and arene affords a radical ion pair between TT˙+ and arene (Ar˙), which by comproportionation with a second equivalent of TFT subsequently undergoes radical recombination to form a dicationic Wheland intermediate. Reversibility of this radical recombination step, promoted by the relative stability of TT˙+ to other sulfur radicals, is permissive of thermodynamically controlled selection for the lowest energy regioisomer of the Wheland intermediate.71 In the case of monosubstituted arenes, steric considerations support a relative ordering in ΔG of ortho- > meta- ∼ para-, while the electronics of radical addition support ΔG meta- > ortho- ∼ para-. Taken together, an explanation for the generally observed exceptional para-selectivity in thianthrenation emerges based on arene stereoelectronics. Notably, an alternative pathway for generation of the Wheland intermediate via a concerted SEAr mechanism was considered operant in the absence of Brønsted acid, which proceeds through a higher energy transition state (ΔΔG‡ = 8.9 kcal mol−1) than formation of TT2+. In addition to the conditions used, dependence on the electronic nature of the arene influences the mechanistic trajectory, with stronger nucleophiles computationally favoring direct substitution on TT+–TFA.71
![]() | ||
| Scheme 1 Mechanistic basis for regioselectivity in thianthrenation. DFT-calculated Gibbs free energies are reported in kcal mol−1 with respect to TTO, where Ar = toluene and R = CH3.71 | ||
In 2020, Wang reported a strategy for the efficient synthesis of aryl thianthrenium triflate salts. Thianthrene S-oxide (TTSO) was reacted with triflic anhydride (Tf2O). High para-selectivity was observed in the thianthrenation of toluene as a model monosubstituted arene. A preliminary mechanistic hypothesis supported by computational investigation suggested that the dication TT2+ is the reactive species toward arene, which is generated by heterolysis of a TT+–OTf adduct. Experimentally observed regioselectivity was suggested to result from the electronics of the radical addition in combination with electrostatic interactions among accumulated charge on the arene and thianthrenium in formation of the Wheland intermediate, which together favor an exo-conformation leading to para-product formation.72 Given the mechanistic studies undertaken by the Ritter laboratory,71 a direct substitution pathway between arene and TT+–OTf adduct may be additionally considered for the formation of the Wheland intermediate.
Recently, distinct sets of thianthrenation conditions have been developed based on the electron density and acid–base sensitivity of the substrate to be functionalized, adapted for arenes that strongly activated (electron-rich) to those that are weakly activated (electron-deficient). Reported limitations include the undesirable reactivity of certain functional groups with the strong acids (e.g. TfOH, TFAA) added to enhance the efficiency of thianthrenation among weakly activated arenes, as well as a general incompatibility between deactivated arenes and any of the described sets of standard conditions.73 While few reports demonstrating the isolation of Ar–TT+ salts generated by the direct electrochemical conversion of arenes and thianthrene are available in the literature,74,75 this approach confers opportunity for the one-pot activation and functionalization of Csp2–H bonds as well as the catalytic reuse of substoichiometric quantities of thianthrene. Given the advantageous step- and resource economy of this method, we anticipate the increased future prevalence of this method of Ar–TT+ salt utilization. In addition to Csp2–H thianthrenation, alternative strategies have been developed to access a broader range of aryl-thianthrenium (Ar–TT+) salts.
These include Cu-mediated thianthrenation of (hetero)aryl boronic acids76 and photoinduced decarboxylative thianthrenation of (hetero)aryl carboxylic acids,77 which enable the synthesis of electron-deficient Ar–TT+ salts that are typically inaccessible via direct C–H thianthrenation.
Considering the remarkable selectivity observed in arene thianthrenation, its demonstrated versatile synthetic utility in Csp2–H activation, recent reviews have discussed the subject.12,14–16,31,32,46,51 Herein, we focus on the recent synthetic utilization of aryl thianthrenium (Ar–TT+) salts as aryl radical (Ar˙) precursors. Several routes to aryl radicals78–80 have been studied under thermal, photochemical, and electrochemical conditions, including aryl halides, aryldiazonoium salts,81 diaryliodonium salts, and aryl pyridinium salts.13 Compared to aryl halides (Ered(PhX/PhX˙–) = –2.97, −2.80, −2.44, and −2.24 V vs. SCE for X = F, Cl, Br, and I, respectively),82 Ar–TT+ salts are more reducible (Ered(Ph–TT+/Ph–TT˙) = –1.55 V vs. SCE in MeCN),52 enabling milder conditions and enhanced functional group tolerance among transformations of the latter. Aryl diazonium salts (Epc(PhN2+/PhN2˙) = –0.06 V vs. SCE)83 are highly reactive and therefore require special handling and storage precautions.84 Diaryliodonoium salts (Epc(PhIPh+/PhIPh˙) = –0.96 V vs. SCE)85 tend more reactive than Ar–TT+ salts, though their preparation usually proceeds with lower regioselectivity or by less direct routes than thianthrenation.86,87
Within the single-electron regime, deuteration of a variety of mono-, di-, and tri-substituted aryl thianthrenium salts in the presence of a CD3OD, base (Cs2CO3), and light (λ = 380–385 nm, P = 50 W) proceeded with moderate efficiency (up to 80% yield) and high regioselectivity, and high deuterium incorporation (up to 99%).97 Late stage functionalization of bioactive compounds of the agrochemicals pyriproxyfen and boscalid was reported therein. Mechanistic insights from experiments with the radical scavengers (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO), 2,6-di-tert-butyl-4-methylphenol (BHT), and1,1-diphenylethylene (DPE) suggested the formation of radical intermediates by observation of a significant reduction in yield when added to standard conditions (Scheme 2). While the mechanistic details remain unclear, observations on the extent of deuterium incorporation with varying solvent isotopomers suggested that the deuterium atom is sourced primarily from the CD3-group of CD3OD. This result is consistent with cleavage of the weakest D–X bond in methanol (C–H BDE = 96.1 ± 0.3 kcal mol−1 and the O–H BDE = 104.6 ± 0.7 kcal mol−1).98
![]() | ||
| Scheme 2 Hydrogen isotope exchange. TEMPO = (2,2,6,6-tetramethylpiperidin-1-yl)oxyl, BHT = 2,6-di-tert-butyl-4-methylphenol, DPE = 1,1-diphenylethylene. | ||
An organic photoredox-catalyzed deuteration of Ar–TT+ salts was reported by Yu in 2024.99 Under aerobic conditions and visible light irradiation (λ = 427 nm, P = 40 W; PS = 2,3,4,5,6-penta(carbazol-9-yl)benzonitrile, 5CzBN) in an aqueous medium, substituted biphenyl derivatives and (hetero)arenes were transformed.
Several bioactive and pharmaceutical compounds were functionalized with high deuterium incorporation (≥90%) from an inexpensive and widely available deuterium source (CDCl3) in a demonstration of applications in LSF. Select examples of extension of this approach to the borylation, cyanation, and selenocyanation of Ar–TT+ salts were described. Robust mechanistic study of the system was undertaken, which supports an energy transfer (EnT) pathway for the generation of aryl radicals; the first observation of this mode of activation of Ar–TT+ salts. Photoexcitation of the photosensitizer (PS) reflected increasingly less intense fluorescence with increasing concentration of the chosen model substrate, biphenylthianthrenium tetrafluoroborate, suggesting the capacity for Ar–TT+ salts to quench PS*. The triplet state (T1) of this substrate, resulting from an EnT interaction with PC*, has a calculated (DFT) energy of ET1 = 60.9 kcal mol−1, and reaction yields increased as ET1(PS*) approached and superceded this value (Scheme 3). The observation that reactivity does not depend as strongly on the relative magnitudes of Ered(PC˙+/PC*) and Ered(Ar–TT+) as it does on those of ET1(PS*) and ET1([Ar–TT+]*) is suggestive of an EnT mechanism. Additional support for an EnT mechanism was the observation of an absorption feature attributable to TT+ by transient absorption spectroscopy (TAS), where TT is the expected byproduct of reductive SET. Absent significant overlap between the PS* emission spectrum and substrate absorption spectrum, Förester resonance energy transfer (FRET) was deemed less likely than a Dexter-type EnT mechanism.100 Addition of the radical scavengers TEMPO, BHT, and DPE led to the detection of the corresponding adducts by HRMS analysis. Excluding a radical chain mechanism through on–off studies, and a mechanistic proposal was generated. An advantageous feature of this system is mitigation of the need for addition of exogenous redox reagents to restore the resting PS.
Sanford and Scott later reported the visible-light promoted, Cu-mediated [11C]-cyanation reaction of aryl thianthrenium salts under aerobic conditions at ambient temperature.109 A variety of electron-rich and electron-poor substrates were functionalized, including N-, O-, and S-heterocycles (Scheme 6). Functionalization of the pharmaceuticals febuxostat ethyl ester (xanthine oxidase inhibitor) and clopidogrel (P2Y12 inhibitor) was reported. Competing fragmentation of the endocyclic C–S bond during radical generation was detected by radio-HPLC analysis for some substrates. In consideration of applications of [11C]-radiolabeled compounds in pharmaceutical development,110 a low-cost assembly of an automation module was designed and implemented in the functionalization of febuxostat ethyl ester.
Patureau reported the direct photochemical (λ = 285 nm, P = 144 W) (hetero)arylation of aryl thianthrenium salts soon after.113 Electron-rich arenes were furanylated or conjugated to other N- or S-heterocyclic substrates under an inert atmosphere (N2) in DMSO (Scheme 8). Introduction of radical scavengers TEMPO and 1,4-dinitrobenzene led to a significant reduction (>35%) in yield when 5-(3-fluoro-4-methoxyphenyl)-thianthrenium tetrafluoroborate was irradiated in the presence of furan. Light-induced homolysis of the exocyclic C–S bond and subsequent addition of Ar to unsaturated substrate is proposed to precede SET from TT+ to the radical adduct, furnishing a cationic adduct that is reduced by BF4−. Reductions in yield were observed when the reaction was carried out under air and in organic solvents other than DMSO (including MeCN, THF, and DMF). Advantages of this system include the regeneration of neutral thianthrene, operational simplicity, and lack of sacrificial redox agents.
Sun and Yu achieved the site-selective, organic photoredox-catalyzed (λ = 457 nm, P = 10 W; PC = 4-CzIPN) arylation of 2-aryl-2H-indazoles at room temperature.114 Several electron-rich, mono- and di-substituted aryl thianthrenium salts were compatible with the transformation (yields 53–87%). Electron-donating substitutions on the N-phenyl group were well-tolerated (yields 55–83%), as were similar modifications to the indazole backbone (yields 71–83%) using p-tolylthainthrenium tetrafluoroborate salt coupling partners (Scheme 9). A prodrug of the COX1/2 inhibitor flurbiprofen and the agrochemical pyriproxyphen (71%) were thianthrenated and conjugated to 2-phenyl-2H-indazole. Treatment of the model substrates, 2-phenyl-2H-indazole and p-tolylthainthrenium tetrafluoroborate with radical scavengers under optimized reaction conditions abolished yield, and the corresponding TEMPO and BHT adducts were detected by high-resolution mass spectrometry (HRMS). Further, the reaction did not proceed in the dark or in the absence of PC, suggesting formation of radical intermediates. In consideration of redox potentials of the reactants and PC, a mechanistic proposal was generated in which aryl radicals are generated through SET with PC˙− prior to addition to indazole and subsequent DMAP-mediated HAT. Scalability was demonstrated by using model substrates in an attempted gram-scale synthesis (68% yield), as was the sustainable use of natural sunlight as an irradiation source (61% yield). Primary advantages of this system include functional group tolerance and operational simplicity.
The C2-arylation of azoles was achieved by metallaphotoredox-catalyzed (λ = 420 nm, P = 10 W, M = CuI) coupling to Ar–TT+ salts in 2025.115 Benzoxazoles were conjugated to electron-rich arenes in moderate yields at ambient temperature using LiOtBu as base under an inert (N2) atmosphere. From the absence of an observed kinetic isotope effect (KIE) upon C2-deuteration, it was suggested that C2–H bond cleavage is not the rate-limiting step. Radical trap experiments with TEMPO or DPE led to the detection of adducts with the arene substrate by HRMS along with a marked reduction in reaction yield, suggesting the intermediacy of aryl radicals. Base-mediated ligation of azole substrate to [Cu], followed by SET and coordination to the arene with subsequent reductive elimination of product, was proposed as the operant mechanism. A one-pot synthesis was demonstrated by the thianthrenation of ethylbenzene followed by addition of benzoxazole, [Cu], and base under irradiation in 54% yield.
In 2022, visible-light photoredox catalyzed (λ = 429 nm, P = 10 W; PC = DABCO) arylation of 6-azauracil derivatives (Scheme 10) and quinoxalin-2(1H)-ones was described by Yu.116 A wide scope of arenes were conjugated to the N-heterocyclic scaffolds. Bioactive compounds, including a derivative of the norepinephrine transporter (NET) inhibitor atomextine, flurbiprofen methyl ester, several nucleoside derivatives (38–71%), pyriproxyfen (64%), and boscalid (71%) were arylated in fair to good yields. Model substrates, 2,4-dibenzyl-1,2,4-triazine-3,5(2H,4H)-dione and p-tolyl-thianthrenium tetrafluoroborate, were selected for attempts at gram-scale synthesis (63%) using natural sunlight irradiation as energy source (79%). Treatment of the model system with the radical scavengers TEMPO and DPE led to observation of trapped adducts by HRMS and a significant reduction in reaction yield, suggesting formation of radical intermediates. UV-Vis absorption spectra recorded for independent and combined solutions of p-tolyl-thianthrenium tetrafluoroborate and 1,4-diazabicyclo[2.2.2]octane (DABCO) provided evidence for formation of an in situ generated EDA complex by observation of an absorption feature that is red-shifted compared to the stand-alone solutions. Taken with control experiments demonstrating a lack of reactivity in the absence of DABCO or light, a mechanism was proposed. Aryl radicals generated through SET within an EDA complex undergo radical addition to N-heterocyclic substrates. The resultant radical is oxidized by PC˙+ to generate a formal cationic species prior to HAT with DABCO to furnish product. This strategy offers the advantage of functionalizing nucleobase derivatives and bioactive compounds under mild conditions.
In 2025, Wu reported the cyclization of 2-isocyanobiphenyls initiated by the radical addition of aryl radicals derived from Ar–TT+ salts to afford the corresponding 6-aryl-phenanthridines (Scheme 11).117 Under visible light irradiation (λ = 450–455 nm, P = 30 W) and inert (N2) atmosphere at ambient temperature in the presence of DBU, the N-heterocyclic products were formed in moderate to fair yields within 12 h. Addition of either 1,4-dioxane or a diarylphosphine oxide to the reaction mixture formed the 6-alkylated or 6-phosphorylated product via HAT with Ar˙ prior to radical addition to the isocyanide. Thianthrenation of the bioactive compounds pyriproxyfen (46%) and boscalid (44%) for use in the reaction was demonstrated as a LSF strategy. Analysis of the system by UV-Vis spectroscopy indicated the strongest bathochromic shift in the absorption profile upon combination of a 2-isocyanobiphenyl substrate, Ar–TT+ salt, and DBU, suggesting the formation of a trimolecular EDA complex whose stoichiometry was further probed by Job's method. Electronic perturbations within the Ar–TT+ salt were also detected via 1H NMR spectroscopy upon addition of DBU and 4′-bromo-2-isocyanobiphenyl. Radical trap experiments with TEMPO, BHT, or DPE led to HRMS observation of the trapped adducts for the proposed alkyl, aryl, and phosphoryl radical intermediates. Radical cyclization under mild reaction conditions is advantageous to the efficient increase in molecular complexity through establishment of two new bonds in a single step.
An organic photoredox-catalyzed (λ = blue LED, P = 24 W, PC = eosin Y) hydroarylation of azine-substituted enamides was reported in 2021 by Wang.119 Electron-rich (hetero)arenes were investigated, as were substrates bearing substituted azines (Scheme 12). This strategy was extended to the functionalization of azine-substituted alkenes (52–63%) and the LSF of flurbiprofen (49%) and pyriproxyfen. A radical trap experiment with TEMPO was consistent with formation of aryl radicals, and incorporation of deuterium was observed when CD3OD was used solvent. Absence of PC or N,N-diisopropylethylamine (DIPEA) produced indeterminate yield. On this basis, a mechanism was proposed in which aryl radicals are generated via SET between Ar–TT+ and PC˙−, with subsequent radical addition to enamine. The resultant α-amino radical can either undergo HAT with DIPEȧ+ or sequential reduction by PC˙− and protonation by solvent to furnish product.
Difunctionalization of electron-deficient alkenes by irradiation (λ = white LED, P = 40 W) of methyl acrylate in the presence of aryl thianthrenium salts, N-phenyl-benzo[b]phenothiazine (PTH) as photocatalyst, and tert-butylammonium bromide (TBAB) at low temperature (−20 °C) was described by Ritter in 2022.120 Highly regio-selective alkene functionalization was observed, with aryl radical addition biased toward the least substituted olefinic carbon to produce a radical intermediate stabilized by the electron-withdrawing ester. Site-selective thianthrenation of flurbiprofen (66%), diclofenac amide (COX1/2 inhibitor), indomethacin, nimesulide (COX2 inhibitor), meclofenamic acid (COX1/2 inhibitor, 70%), and benzbromarone (URAT1 inhibitor, 46%) enabled their use as precursors for this reaction (Scheme 13). Further reactivity of the newly furnished Csp3–Br bond was demonstrated by treatment with various N-, O-, S-, and P-nucleophiles. No evidence for EDA complex formation between Ar–TT+ and PTH was observed in the UV-vis absorption spectra. Support for quenching of PC* by Ar–TT+ was obtained through Stern–Volmer analysis. Reduction of Br− to Br2 by PTH˙+ was observed indirectly by addition and halogenation of 1,3,5-trimethoxybenzene. Based on these results, a mechanistic proposal involving the formation of aryl radicals was generated. Photoexcited PTH* transfers an electron to Ar–TT+, producing Ar˙ and TT. Aryl radicals add to alkene substrates, followed by radical abstraction of bromine formed by reaction with PTH˙+.
A Cu-catalyzed three-component coupling of Ar–TT+ salts, sodium azide, and alkenes under visible light irradiation (λ = 453 nm, P = 40 W) generated 1,2-difunctionalized products under mild conditions and inert atmosphere (N2).121 A diverse substrate scope was evaluated, including N- and S-heterocyclic arenes, multi-substituted alkenes, and bioactive compounds (Scheme 14). Among the latter are included the pharmaceuticals diclofenac amide, flubiprofen, and meclofenamic acid, as well as agrochemicals boscalid (69%), cloquintocet-mexyl, and pyriproxyfen (60%). Stern–Volmer analysis suggested the role of [CuI(rac-BINAP)(N3)] (BINAP = racemic 2,2′-bis(diphenylphosphino)-1,1′-binaphthalene) as a photocatalyst on the basis of observed quenching by Ar–TT+ salts. Experimental evidence for the reactivity of the dimer [Cu2(rac-BINAP)2(μ-N3)2] with Ar˙ was obtained by formation of azidoarylation product when the complex was treated with benzoyl peroxide at 80 °C. Light-induced excitation of the PC and subsequent SET to Ar–TT+ was proposed to generate aryl radicals, which add to acrylonitrile to furnish an intermediate is α-cyano radical. Radical abstraction of an azide group from [CuII(rac-BINAP)(N3)2] ensues, forming product and regenerating the active catalyst. Attempts at enantioselective synthesis using (R,R)-BINAP ligand provided comparable yields with low enantiomeric excess (≤8% ee), purportedly due to the spatial separation of ligand from the site of enantio-determining radical abstraction. Nonetheless, high regioselectivity in the aryl radical addition to alkene was observed, presumably due to radical stabilization conferred by the electron withdrawing α-cyano group. Installation of the azido group enables further reactivity.
Difunctionalization of ethylene through a radical polar crossover pathway enabled by EnT-mediated activation of Ar–TT+ salts was reported by Ritter in 2025.53 Visible-light irradiation (λ = 390 nm, P = 40 W; PS = thioxanthone, TXO) of arenes under an atmosphere of ethylene gas led to the facile formation of 1,2-difunctionalized intermediates which, without isolation, are further transformed by the addition of various (C-, N-, P-, O-, S-, Se-, F-, Cl-, Br-, I-) nucleophiles (Scheme 15). Stern–Volmer analysis demonstrated the quenching ability of Ar–TT+ salts on PS*. Correlation of the reaction yield with ET(PS), and not the redox potential of the PC that would be expected in a SET-mediated process, was suggestive of an EnT mechanism. Additional support for EnT was generated through observation of a TAS absorption feature attributable to TT˙+. Certain properties of TXO favor EnT over SET, such as highly efficient intersystem crossing from S1 to T1 (ϕISC = 0.99)122 and a long triplet lifetime (τ = 95 μs in MeCN).123 An overall quantum yield of Φ = 0.12 was observed from a measured EnT efficiency of 0.33 and quantum yield of homolytic cleavage following EnT of ΦEnT = 0.36 when phenylthianthrenium triflate was used as substrate, demonstrating the efficient photophysical properties of this system. Exclusion of a radical chain mechanism was supported by this finding. Computational (DFT) study enabling consideration of the nature and spatial arrangement of the PS* HOMO relative to the substrate LUMO provided evidence for a hypothesized Dexter-type EnT mechanism. Further, the geometry of this interaction was suggested to impart chemoselectivity in EnT to the thianthrenyl moiety over other functional groups present in the arene.
Radical cascade cyclization reactions increase structural complexity by promoting multiple bond scission and formation events in a single reaction step. Initiation of an inter-intramolecular radical cyclization cascade via irradiation (λ = 390 nm, P = 30 W) of Ar–TT+ salts in the presence of N-methyl-N-aryl-methacrylamide led to formation of benzyl oxindoles under mild conditions and an inert atmosphere (Ar).124 Electron-rich, -neutral, and -poor arenes were studied, as were N-phenyl-substituted acrylamides (Scheme 16). Addition of TEMPO inhibited product formation suggesting formation of radical intermediates, and UV-vis absorption spectroscopy supported the formation of an EDA complex between Ar–TT+ and K2CO3 by observation of a spectral red-shift when a combined solution is prepared. Addition of aryl radicals, formed through EDA-complex-mediated SET, to acrylamides yields radical intermediates that undergo intramolecular cyclization. Cyclized radical intermediates can donate an electron to a second equivalent of Ar–TT+, thus initiating another cascade and generating a cationic product precursor converted to neutral form via deprotonation. The measured quantum yield of 6.1 supports the SET activation of the second Ar–TT+ equivalent by a radical intermediate.
Organic photoredox-catalyzed (λ = 427 nm, P = 30 W; PC = eosin Y) arene allylation by addition of aryl radicals to electron-deficient alkenes was reported in 2024.125 Electron-rich (hetero)arenes and gem-disubstituted alkenes were transformed in moderate to good yields (Scheme 17). The reaction did not proceed in the absence of light or PC, nor in the presence radical scavengers TEMPO and BHT, suggesting the formation of radical intermediates. Homolytic cleavage of the aryl thianthrenium salt promoted by SET with PC* generates aryl radicals that undergo addition to the alkene coupling partner at the sterically least hindered carbon, producing an intermediate radical adduct stabilized by the electron-withdrawing α-methyl ester. Decarboxylative loss of CO2 and tBuO˙ furnishes product. For some substrates, the decarboxylation step proceeds with moderate to high E/Z-stereoselectivity.
A Cu-catalyzed three-component reaction of Ar–TT+ salts, alkenes, and trimethylsilyl cyanide (TMSCN) under visible light irradiation (λ = 455 nm, P = 15 W) and inert atmosphere (N2) at low temperature (5 °C) generated 2,3-diarylpropionitrile products.126 Various electron-rich styrene derivatives and Ar–TT+ salts were coupled, and the LSF of several bioactive compounds was demonstrated (Scheme 18). Site-selective thianthrenation of ibuprofen (COX1/2 inhibitor), oxaprozin (COX1/2 inhibitor), and gemfibrozil enabled difunctionalization of p-phenyl styrene, or p-methoxystyrene in the case of flurbiprofen (83%) and pyriproxyfen (60%). Gram-scale coupling of model substrates p-methoxystyrene and biphenylthianthrenium tetrafluoroborate afforded product in 60% yield with efficient recovery of TT (89%). Independent and combined solutions of biphenylthianthrenium tetrafluoroborate, dibutamine (DBA), and TMSCN were analyzed by UV-vis absorption spectroscopy, which suggested formation of a trimolecular EDA complex by the observation of a spectral red shift and the appearance of an absorption feature at λ = 375–450 nm that is attributable to a charge-transfer band present only when all three analytes were combined. Trapping of radical intermediates upon addition of TEMPO was detected by HRMS, and the alkene difunctionalization product was formed in trace quantities. A control experiment in the absence of light annihilated reactivity to further support a radical mechanism. Activation of CuOAc precatalyst by ligation of rac-BINAP and reaction with TMSCN yields the resting catalyst, [CuI(rac-BINAP)(CN)]. After photoexcitation, [CuI(rac-BINAP)(CN)]* undergoes SET with [TMSCN-DBU]˙+, formed from SET within the EDA complex, to produce [CuII(rac-BINAP)(CN)2]. Alkyl radical intermediates formed by the addition of thianthrenium salt-derived aryl radicals to styrene substrates add to the [CuII] complex, generating a [CuIII] complex from which product is reductively eliminated.
Another three-component, metallaphotoredox-catalyzed (λ = 390 nm, P = 15 W, PC = PTH, [M] = Cu(MeCN)4PF6, L = 2,2′-bipyridine (bpy), B = KH2PO4) system involving TMSCN and 1,3-enynes provided access to the tetrasubstituted allenyl nitriles in 2025.127 The arene scope largely consisted of electron-rich substrates, although a few examples of disubstituted arenes bearing an ester substituent were included.
Mechanistic insight from radical trap experiments, Stern–Volmer analysis, and the UV-vis characterization of combined and independent solutions of reaction components enabled a proposal for the product formation pathway. Aryl radicals generated from the interaction of Ar–TT+ salts and PC* undergo radical addition to the terminal alkene site of the 1,3-enyne, and the progargyl radical resonance form of the adduct may ligate a [CuII(bpy)(MeCN)2] species to then reductively eliminate product. This system features the advantageous use of visible-light irradiation, inexpensive and earth-abundant metal catalyst, and ambient temperature.
Aryl radical addition to nonnatural amino acids (NAAs) derived from dehydroalanine (Dha) was reported by Noël in 2024 and represents an LSF strategy enabled by thianthrenation.128 Common to the NAAs studied is a modified N-terminus, incorporating a phthalimide moiety, and an α,β-unsaturated side chain and C-terminus. Under mild photocatalytic conditions (λ = 456 nm, PC = eosin Y) and inert atmosphere (N2), functionalization of peptides and bioactive compounds was achieved in batch and flow apparatus. Pharmaceutical compounds benzbromarone (62%), bifonazole, flurbiprofen, meclofenamate (64%), and the agrochemicals boscalid and pyriproxyfen (67%) were functionalized. Aryl thianthrenium salts bearing N- and O-heterocyclic or electron-donating substituents were well-tolerated (Scheme 19). The reaction was most efficient when run in a cosolvent system of MeCN/hexafluoroisopropanol (HFIP) (7
:
1 v/v) and with DIPEA as terminal reductant. Extension of this approach to the merger of Dha incorporation into proteins34 with the reactivity of Ar–TT+ salts represents a potential application in bioconjugate chemistry and a new route toward the selective modification of complex biomolecules under mild conditions.
![]() | ||
| Scheme 19 Arylation of nonnatural, dehydroalanine-derived amino acids. aBatch reaction, bContinuous flow reactor. AA = amino acid. | ||
Difunctionalization of the ring-strained tricyclic hydrocarbon [1.1.1]propellane by a metallaphotoredox catalysis (λ = 450–460 nm, P = 3 W, PC = Ir(ppy)3, [M] = Cu(acac)2, L = 2,2′-bis(dicyclohexylphosphino)-1,1′-biphenyl, B = DBU) approach was described by Zhang in 2025.129 Phenylacetlyenes with electron-donating aryl substituents and electron-rich Ar–TT+ salts were transformed in fair yields (45–60%) at ambient temperature under inert (Ar) atmosphere. On the basis of mechanistic studies involving radical trap and on–off experiments, it was hypothesized that formation of aryl radicals derived from Ar–TT+ salts following interaction with PC* enabled a strain-relieving radical addition to [1.1.1]propellane, where the resultant radical adduct is intercepted by a [Cu(II)(alkynide)] species to form a [Cu(III)] intermediate from which the 1-alkynyl-3-arylated product is reductively eliminated. Extension of this work to other ring-strained systems high in sp3 character may increase access to single-electron reactivity among these substrates under mild conditions.
A three-component coupling of substituted styrenes, (hetero)aromatic aldehydes, and Ar–TT+ salts using a visible-light (λ = 460–465 nm, P = 65 W) induced, NHC-catalyzed approach led to formation of the corresponding 1,2-arylacylated products (Scheme 20).130 The reaction was most efficient using sterically accessible and aldehydes with electron-withdrawing aryl substituents, electron-rich Ar–TT+ salts, and electron-rich olefins. Evidence for the generation of aryl radical intermediates was obtained from the addition of TEMPO to standard reaction conditions, as radical-trapped adducts were detected by HRMS and the desired product was not detected. Addition of base promotes the formation of an NHC(C2)-aldehyde adduct known as a Breslow intermediate, which provides access to enol-based precursors of persistent ketyl radicals after SET.131 Here, the radical combination of this ketyl radical and the radical intermediate formed upon addition of Ar˙ to the olefinic substrate furnishes a new Csp2–Csp3 bond and regenerates the active NHC organocatalyst. The use of a thiazole-based NHC organocatalyst enhances the scalability and sustainability of the transformation, while the formation of multiple bonds is beneficial from a step-efficiency standpoint. Exploration of generality of this transformation among other structural classes of aldehydes and olefins represents an opportunity for future investigations.
Amination of aryl thianthrenium salts under photochemical conditions (λ = 450 nm, P = 30 W; PC = [Ir(dtbbpy)(ppy)2]PF6, dtbbpy = 4,4′-di-tert-butyl-2,2′-dipyridine) using ammonia as a nitrogen source was reported by Jiang in 2023.133 Electron-rich Ar–TT+ salts, including those bearing N- and O-heterocyclic or aryl halide substituents, were transformed in up to 86% yield (Scheme 22). The bioactive compounds bifonazole, estradiol, fenbufen (COX1/2 inhibitor), flurbiprofen, and N-acetylmexiletene (antiarrhythmic), in addition to the agrochemical pyriproxyfen, were site-selectively functionalized. When p-tolylthianthrenium tetrafluoroborate was treated with TEMPO under standard conditions, trace amount of p-toluidine product was formed and the TEMPO adduct 2,2,6,6-tetramethyl-1-(p-tolyloxy)piperidine was detected by HRMS analysis, suggesting the formation of aryl radical intermediates. Stern–Volmer analysis demonstrated the ability of [Cu(CH3CN)4]PF6 to quench PC*. Taken together, a mechanistic proposal was generated in which the reduced form of the PC, [IrII], engages in SET with Ar–TT+ to furnish Ar˙ and the resting PC, [IrIII]. Upon photoexcitation, PC* oxidizes [CuI(L)2(NH3)]PF6 (L = 2,9-dichloro-1,10-phenanthroline) via SET and the resultant [CuII] complex intercepts Ar˙ and subsequently reductively eliminates aminated product from a transient [CuIII] complex. Although the precatalyst can theoretically be regenerated by reductive elimination, a reduction in yield from 83% to 67% was observed when the quantity of [Cu] was reduced from 1.0 equivalent to 0.6 equivalents with respect to p-tolylthianthrenium tetrafluoroborate, suggesting catalyst deactivation.
In extension of their prior work, the Ritter group reported visible-light-promoted (λ = white or blue LED), Ni-catalyzed aminations of Ar–TT+ salts in 2024.134 Under ambient conditions and inert atmosphere (Ar), several electron-rich arenes and N-nucleophiles were efficiently coupled by earth-abundant and inexpensive catalysts (Scheme 23). When used in supra-stoichiometric (≥2 equiv.) quantities relative to Ar–TT+ salts, substrate amine can serve as an endogenous base and reductant. Alternatively, exogenous DABCO or 2-tert-butyl-1,1,3,3-tetramethylguanidine (BTMG) was used to serve these functions. Late-stage functionalization of the bioactive compounds amoxapine (SNRI, serotonin-norepinephrine reuptake inhibitor), desloratadine (H1-receptor antagonist), fenbufen (54%), flurbiprofen, and the structurally complex (−)-strychnine (GlyR and AchR antagonist) was accomplished through reactivity of piperidine or piperazine moieties. Based on evidence of electron paramagnetic resonance spectroscopic evidence of Ni(I) formation under light irradiation and control experiments in the absence of light, a mechanistic proposal was generated. Coordination of N-nucleophiles to NiII-sources, such as NiCl2 or NiBr2, may precede the formation of amine radicals via light-induced ligand-to-metal charge transfer (LMCT) and subsequent SET from the resultant [NiI] complex to Ar–TT+. Oxidative ligation of Ar˙ followed by a subsequent reductive elimination formed product. Notably, the system is tolerant to added H2O (5.0 equiv.), producing a modest reduction in yield of ∼10% for the model substrates piperazine and 5-(4-((N,4-dimethylphenyl)sulfonamido)phenyl)-thianthrenium tetrafluoroborate.
:
1 stoichiometry. It was hypothesized that formation of Ar˙ from the EDA complex enables radical addition to phosphonate esters prior to sequential oxidation of the phosphoranyl radical intermediate by DMP˙+ and hydrolysis of the phosphonium ether to furnish product. Other tertiary amines screened in DMSO solvent, including DABCO, N,N,N′,N′-tetraethylenediamine (TMEDA), and 1,3,5-trimethyl-1,3,5-triazacyclohexane (TMTAC) provided comparable yields (within 6%) to DMP.
Granados reported the phosphonation of Ar–TT+ salts by a visible-light-promoted (λ = 390 nm, P = 30 W) EDA complex strategy with a P-donor.136 At ambient temperature under an inert atmosphere (Ar), electron-rich and electron-poor arenes were converted to aryl phosphonate esters in an acetonitrile solution of triethylphosphite and KHCO3 (Scheme 25). Alternative solvents (acetone and DCM), bases (DABCO, K2CO3, and K3PO4), irradiation wavelengths (456 nm), and phosphines (trimethyl-, triisopropyl-, and triphenyl phosphite) demonstrated compatibility for this transformation, with yields generally within 10% of the optimized conditions. Irradiation of the 5-(3-cyano-4-methoxyphenyl)thianthrenium or 5-(4-fluorophenyl)thianthrenium tetrafluoroborate salts with natural or simulated sunlight, respectively, provided the phosphonate products in 82% and 79% yield, demonstrating an opportunity for green energy input. Late-stage functionalization of fenbufen (74%), flurbiprofen (88%), gemfibrozil (65%), and clofibrate was described. Mechanistic insight from UV-vis spectroscopy suggested EDA complex formation, and a radical trap experiment with TEMPO indicated the presence of radical intermediates. The result of measurement of the quantum yield (Φ = 118) suggested a radical chain mechanism. Fragmentation of the EDA complex to generate aryl radicals with subsequent addition to phosphites generates a radical intermediate that is oxidized to the cation via SET with an additional equivalent of Ar–TT+. Formation of the phosphonium intermediate was supported by 1H-31P HMBC NMR and HRMS. An Arbuzov-type dealkylation step may then furnish product.
Ou and Su reported phosphination of Ar–TT+ salts under visible light irradiation (λ = 420 nm, P = 15 W) at room temperature by leveraging the capacity for an EDA complex with K2CO3 to generate aryl radicals.137 Treatment of electron-rich arenes with diaryl phosphines afforded asymmetrically substituted triaryl phosphines in moderate yields (Scheme 26).
Observation of a red-shifted UV-vis absorption spectrum upon combination of Ar–TT+, PR3, and K2CO3 provided experimental evidence for the formation of an EDA complex. A chemical shift was observed for thianthrenium 1H nuclei only in a combined solution of these three components, further supporting this hypothesis. Computational (DFT) studies identified an optimized ground state structure with intermolecular interactions between Ar–TT+ and phosphine as well as phosphine and K2CO3. On this basis, it was suggested that deprotonation of phosphine promotes SET within the EDA complex, forming aryl and phosphoryl radicals that recombine to form the product.
Patureau reported the direct photochemical (λ = 254 nm, P = 144 W) synthesis of phenols using (4-oxo)-TEMPO as an oxygen atom source in 2022.140 Substituted biaryls, including haloarenes, were functionalized in addition to O-/S-/N-heteroarenes (Scheme 29). Late-stage functionalization of flurbiprofen (56%) and bifonazole (62%) was reported.
Addition of the radical scavengers BHT, DPE, and 1,4-dinitrobenzene to the model system using biphenyl thianthrenium tetrafluoroborate as substrate led to a >40% reduction in phenol yield, suggesting the formation of an aryl radical intermediate. With thianthrenated fluorenone as substrate, a TEMPO adduct was isolated and characterized by single-crystal X-ray diffraction, supporting the intermediacy of arene-TEMPO adducts formed by radical recombination after light-induced homolysis of Ar–TT+ salts. Subsequently, fragmentation of the N–O bond within the adduct may furnish an oxygen-centered phenol radical capable of abstracting hydrogen to generate product.
A three-component coupling of amines, carbon dioxide, Ar–TT+ salts to produce O-aryl carbamates under metallaphotoredox conditions (λmax = 450 nm, P = 30 W; PC = [Ir(dtbbpy)(ppy)2]PF6) was reported by Qi in 2023.141 The reaction was most efficient using secondary amines, electron-rich arenes, DABCO as base, suprastoichiometric quantities of [Cu(MeCN)4]PF6, and BF3·OEt additive (Scheme 30). Radical trap experiments combining biphenyl thianthrenium tetrafluoroborate and TEMPO or DPE under optimized conditions led to observation of radical-trapped adducts and a ≥44% reduction in O-aryl carbamate product yield, suggesting formation of aryl radicals. Stern–Volmer analysis was consistent with the reduction of PC* by DABCO, supporting the following mechanistic proposal. Base-mediated deprotonation of amines facilitates carbamato complex formation with [CuI] prior to its oxidation by DABCO˙+ to [CuII]. Oxidative ligation of Ar˙ generates a [CuIII] complex from which product may be reductively eliminated. Increased product yield and suppression of hydrodefunctionalized product formation was observed upon increasing the molar equivalency of [CuI] from 5 mol% to 40 mol%, which further improved with the use of a suprastoichiometric quantity (3 equiv.) of [CuI] under the optimized conditions. This suggests a low TON for [CuI] prior to its deactivation or off-cycling.
A photoredox-catalyzed (λmax = 460–470 nm, P = 30 W; PC = Ir(ppy)3) three-component coupling of Ar–TT+ salts, [SO2], and silyl enol ethers was reported by Wu in 2022.143 At ambient temperatures under inert atmosphere (N2), electron-rich arenes and silyl enol ethers were coupled via a sulfone bridge to generate the corresponding (α,α-difluoro-)β-keto-arylsulfones in moderate to good yields (Scheme 32). Thianthrenated forms of the bioactive compounds estrone methyl ether, indomethacin methyl ester, and pyriproxyfen were successfully functionalized using this strategy. Recycling of thianthrene in 90% recovery was demonstrated for the model substrates 1-phenyl-1-trimethylsiloxyethylene and p-tolylthianthrenium triflate. Quenching of PC* by the model Ar–TT+ salt was observed in a Stern–Volmer analysis. Radical adducts were observed by HRMS when TEMPO or BHT were added to the model system, suggesting the formation of aryl radical intermediates. On this basis, it was hypothesized that aryl radicals formed upon SET with PC* undergo addition to [SO2] equivalents to generate an intermediate S-centered radical species, which may in turn add to silyl enol ether. The resultant C-centered radical is then oxidized to the corresponding cation, regenerating the resting PC, prior to desilylation to furnish product. Avoidance of sacrificial redox agents is established by two synthetically useful SET events between photocatalyst and these intermediates, although the redox involvement of DABCO in this system cannot be ruled out.
Arenesulfonohydrazides were synthesized via three-component coupling of Ar–TT+ salts, [SO2], and hydrazines under photoredox conditions (λν = blue LED, P = 30 W; PC = Ir(ppy)3) by Wu in 2022.144 Electron-rich arenes were functionalized (Scheme 33), including a demonstration of the LSF of pyriproxyfen and the protected carbohydrate salicin pentaacetate using morpholine as the hydrazine substrate. Mechanistic insight gained by addition of the radical scavenger TEMPO to the model system suggested the intermediacy of aryl radicals, corroborating the results of control experiments conducted in the absence of light or PC.
Quenching of PC* by the model Ar–TT+ salt, p-methoxyphenyl thianthrenium triflate, was indicated by Stern–Volmer analysis. On this basis, a mechanistic hypothesis was proposed. Aryl radicals engage [SO2] to yield S-centered radical adducts that couple with a hydrazine radical, formed in the reduction of [IrIV] to [IrIII], to generate product.
Cao reported a three-component fluorosulfonylation of Ar–TT+ salts under galvanostatic electrochemical conditions (I = 8 mA, ϕ = 1.5–2.0 V, cylindrical carbon felt electrodes [10 mm × 10 mm × 0.3 mm], electrolyte = n-Bu4NClO4) in 2023.145 Combination of electron-rich (hetero)aryl thianthrenium salts, DABSO·SO2, and potassium bifluoride (KHF2) in an undivided cell at room-temperature under inert atmosphere (N2) afforded the corresponding fluorosulfonylated arenes in moderate to good yields (46–86%) (Scheme 34). Application to the LSF of gemfibrozil methyl ester (42%) and the steroid epiandrosterone was demonstrated. A one-pot method in which the model Ar–TT+ salt, p-tert-butylphenyl thianthrenium triflate, was sequentially thianthrenated and fluorosulfonylated, generating product in 69% yield. A gram-scale experiment of the model system (4 mmol Ar–TT+) gave 81% yield, comparable to the 0.2 mmol-scale yield of 81%. Further, recovery of TTSO (85%) by exposure of the model system to air and extension of the reaction time was shown and did not affect product yield. Cyclic voltammetry studies supported the redox activity of the Ar–TT+ salt, with an irreversible reduction observed at Epc = −0.95 V vs. Ag/Ag+. A radical trap experiment using TEMPO led to detection of S–O–N and Csp2–O–N radical-trapped adducts by HRMS and trace product formation, suggesting the intermediacy of S- and C-centered radicals. When the model system was probed for activity in a divided cell, no product formation was observed by GC-MS analysis, suggesting synthetically productive interactions of species generated at both the cathode and anode. A mechanistic proposal was generated in which aryl radicals are formed by the cathodic reduction of the Ar–TT+ salt, which undergoes addition to [SO2] to produce a S-centered radical intermediate. This intermediate may then either combine with anodically generated DABCO˙+ or itself undergo anodic oxidation, followed by nucleophilic attack of F− to generate product in both cases.
Yang reported a visible-light-promoted (λmax = 455 nm, P = 20 W), Cu-catalyzed, four-component coupling of Ar–TT+ salts, alkenes, trimethylsilyl azide (TMSN3), and [SO2] in 2024.146 Difunctionalization of aryl-substituted alkenes generated the corresponding β-azido-arylsulfone products, which contain a reactive functional group capable of further transformation (Scheme 35). For instance, conversion of the β-azido moiety into a triazole by a Cu-catalyzed click reaction or reduction to primary amine was demonstrated in select cases; the latter was utilized in the synthesis of an apremilast (PDE4 inhibitor) analogue. High stereoselectivity and good regioselectivity was observed in the azidosulfonylation of some 1,3-butadiene substrates. Enhanced UV-vis absorption was observed upon addition of rac-BINAP to a solution of CuCl, suggesting the role of a [CuI(rac-BINAP)] complex as a photoactive species. Radical trapping experiments with TEMPO and BHT led to observation of S–O–N and Csp2–O–N radical-trapped adducts by HRMS, along with a significant reduction in product yield. Photoexcitation of CuI(rac-BINAP)(N3) enables SET to Ar–TT+, generating aryl radicals which may undergo addition to [SO2] and, subsequently, addition to alkene substrates. Outer-sphere radical abstraction of azide from CuII(rac-BINAP)(N3)2 then furnishes product and regenerates the catalyst. A similar system was reported by Wu in 2023 using Ir(ppy)3 as PC.147
Thioetherification of Ar–TT+ salts mediated by an EDA complex under aerobic conditions and visible light irradiation (λ = 427 nm, P = 30 W) was reported by Molander in 2022.148 Coupling of (N-hetero)aryl thiols and electron-rich Ar–TT+ salts was demonstrated (Scheme 36) in addition to the LSF of the bioactive compounds fenbufen, flurbiprofen, gemfibrozil, salicin, and xanthone. Longer irradiation wavelengths (λν = 455 nm/40 W or 525 nm/44 W) increased the formation of undesired thiolate dimerization products. Experimental evidence for the formation of an EDA complex was obtained by UV-vis spectroscopic analysis of reaction components in isolation and combination, and a 1
:
1 stoichiometry between thiolate and Ar–TT+ salt was suggested by Job's method. Formation of aryl radicals was indicated by a radical trap experiment with TEMPO and detection of the corresponding adduct by HRMS.
![]() | ||
| Scheme 36 Thioetherification of aryl thianthrenium salts promoted by a S-donor EDA complex. aHSAr2 (1.1 equiv.), K2CO3 (1.1 equiv.); bK2CO3 (2.0 equiv.). | ||
Determination of the quantum yield (Φ = 89) supported a radical chain mechanism. It was then hypothesized that base-mediated deprotonation of thiols promotes ground-state EDA complex stabilized by π–π interactions, which upon irradiation undergoes homolytic fragmentation to form Ar˙, TT, and disulfides. Aryl radicals may subsequently undergo addition to another equivalent of thiolate to form a radical anion adduct capable of reducing Ar–TT+ salts and thus initiating another cycle. Mild conditions and operational simplicity are advantageous features of this system.
A metallaphotoredox-catalyzed (λmax = 450 nm, P = 34 W; PC = [Ir(dtbbpy)(ppy)2]PF6; [M] = CuBr) thioetherification of Ar–TT+ salts was developed by Zha and Ji in 2024.149 Coupling of 1-thiosugars with (hetero)aryl thianthrenium salts led to the stereoretentive formation of aryl thioglycosides. Several pharmaceutical and agrochemical compounds were thianthrenated and glycosylated, including a gram-scale reaction estrone methyl ether (Scheme 37). Addition of TEMPO led to trace amounts of desired product formation and detection of an aryl radical-trapped adduct by HRMS and 1H NMR analysis. Stern–Volmer experiments demonstrated the most efficient quenching of PC* occurred in the presence of thiosugar and base, suggesting the role of thiolate as the kinetically preferred reducing agent among the redox-active reaction components. Through an on–off experiment in which product formation was inhibited in the absence of light, a self-propagating radical chain mechanism was excluded. It was then proposed that base-mediated ligation of thiosugar to [CuI] forms a complex that is oxidized to [CuII-SR] via SET with thiol radicals generated in the reduction of [IrIII]*. Aryl radicals oxidatively ligate this complex to generate a [Ar-CuIII-SR] complex before reductive elimination of product.
In 2023, Yang reported a visible-light-promoted (λ = 455 nm, P = 34 W) dithiocarbamoylation of Ar–TT+ salts mediated by EDA complex formation in an aqueous medium.150 This radical three-component coupling of Ar–TT+ salts, carbon disulfide, and secondary aryl- or alkyl amines (Scheme 38) was facilitated by surfactant-based (dodecyl trimethyl ammonium chloride; DTAC) micelle formation. In the absence of exogenous surfactant to the model substrates 5-(3-formyl-4-methoxyphenyl)thianthrenium tetrafluoroborate and pyrrolidine, S-aryl dithiocarbamoylate yield was reduced from 82% to 65%, improved to 70% in DMSO solvent. A gram-scale reaction of the model system generated product in 63% yield with 96% recovery of TT, and the of natural sunlight as the irradiation source offered 60% isolated product yield. Thianthrenation of fenofibrate and bifonazole enabled conjugation to diethylamine or N-methylaniline through a dithiocarbamates linker in moderate to good yield, demonstrating an application of the system to LSF. Radical trap experiments with TEMPO and DPE were indicative of aryl radical formation as the corresponding Csp2–O–N or Csp2–Csp2 adducts observed by HRMS with concomitant reduction in product formation. Experimental evidence for the 1
:
1 formation of an EDA complex consisting of a dithiocarbamate anion and Ar–TT+ salt was obtained through Job's method and a bathochromic shift in the UV-vis absorption spectrum. Thus, SET within the EDA complex followed by recombination of the dithiocarbamate and aryl radicals was proposed.
Zhang reported the visible-light-promoted (λ = 420 nm, P = 10 W) dithiocarbamoylation of Ar–TT+ salts using an EDA complex strategy.151 Three-component coupling Ar–TT+ salts, carbon disulfide, and secondary aryl- or alkyl amines in DMSO/H2O (1
:
1 v/v) yielded the corresponding S-aryl dithiocarbamates (Scheme 39). A gram-scale synthesis using the model substrates p-tolylthianthrenium salt and N-methylaniline under flow conditions generated product in 66% yield with 90% recovery of TT. Constant irradiation is necessary for product formation, as evidenced by an on–off experiment, a result that would be inconsistent with a chain radical mechanism. Analysis by UV-vis spectroscopy demonstrated the strongest bathochromic shift in a combined solution of the three substrates with base, indicating the contribution of each to the formation of an EDA complex. Job's method suggested a 1
:
1 stoichiometry of dithiocarbamate anion to Ar–TT+ salt within the EDA complex. From these experiments, a mechanistic proposal analogous to Yang's was formulated.
A three-component coupling of Ar–TT+ salts, amines, and carbon disulfide for the synthesis of S-aryl-dithiocarbamates under aerobic conditions and white-light irradiation (color temperature = 4000 K, P = 2 × 40 W) was reported by Chen and Yi in 2024 (Scheme 40).152 Structurally diverse amines and a selection of electron-deficient arenes were functionalized. Thianthrenation of bioactive compounds facilitated their conjugation to amines through a dithiocarbamate linker. In demonstrating green applications of this system, natural sunlight was used as irradiation source and water was utilized as the solvent. One-pot, two-step syntheses were carried out, further enhancing the efficiency of this approach. Design of a continuous flow photoreactor enabled a decagram-scale synthesis with 94% recovery of TT. Spectroscopic analysis revealed the strongest bathochromic shift in the UV-vis absorption profile of a combined solution of the three components with base, suggesting the contribution of each component to the formation of an EDA complex. Absorbance measurements on solutions of 5-(4-phenoxyphenyl)thianthrenium tetrafluoroborate and varying concentrations of morpholine-N-dithiocarbamate anion under were suggestive of a 1
:
1 stoichiometry in the EDA complex, as the strongest absorption was observed at a mole fraction of χ = 0.5. Operation of a radical chain mechanism was excluded through an on–off experiment in which product formation was not observed during dark cycles. The proposed mechanism is analogous to those previously discussed for multicomponent dithio-carbamoylation reactions proceeding through an EDA complex.
Wang reported the synthesis of sulfilimines from N-Bz-protected aryl sulfenamides and Ar–TT+ salts under visible-light irradiation (λmax = 452 nm, P = 48 W) in the presence of CuCl and DBU at ambient temperature under an inert (N2) atmosphere (Scheme 41).153 Electron-rich arenes as well as the pharmaceuticals fenbufen (56%), fenofibrate, gemfibrozil (59%), niflumic acid (COX2 inhibitor), and mexiletine (SCN5A inhibitor, 72%), were thianthrenated and utilized in the transformation. The selected model substrates (80% yield under standard conditions) were transformed in57% yield under air, in 35% yield with phosphate-buffered saline (PBS) as solvent, and in 40% yield with aqueous sodium dodecyl sulfate (SDS; 2 wt%) as solvent, suggesting some tolerance of the system to air and water. Also for the model substrates, the reaction proceeded in 37% yield in the absence of [Cu], suggesting a [Cu]-independent pathway to product formation. Characterization of independent and combined solutions of the reaction components was suggestive of EDA complex formation between deprotonated sulfenamide and the Ar–TT+ salt. It was proposed that, upon irradiation, the EDA complex can decompose into two substrate radicals which recombine to form product. Alternatively, [Cu] may intercept these radicals and facilitate their cross-coupling. Formation of the EDA complex prior preceding either pathway was supported by the absence of detectable product formation in the absence of DBU. Toward greater sustainability and scalability, experiments conducted using natural sunlight as the irradiation source (66% yield) or at gram-scale (62% yield) were undertaken, and the nearly quantitative recovery of TT was possible.
Arylation of phosphorothioates in the visible-light (λ = 455 nm, P = 20 W) promoted, [Cu]-catalyzed ([Cu] = Cu(OAc)2, B = K2HPO4) synthesis of S-aryl phosphorothioates at ambient temperatures under inert (N2) atmosphere in acetone was reported in 2024.154 While predominantly electron-rich Ar–TT+ salts were used, electron-withdrawing substituents (–CHO, –CO2Me, –CN) in the meta-position were also tolerated. Bioactive compounds were thianthrenated and used as substrates, including diclofenac amide, fenofibrate, gemfibrozil, and nimesulide. A gram-scale trial with highly efficient recovery of TT, as well as an experiment using natural sunlight irradiation provided favorable results for scalability and sustainability.
Through a three-component synthesis of aryl thioether-substituted imidazoles, Ar–TT+ salts were thioetherified under visible-light irradiation (λ = 455 nm, P = 25 W) and aerobic atmosphere by Yang in 2024.155 Aryl isothiocyanates, isonitriles, and Ar–TT+ salts were conjugated in the presence of strong base (LiOtBu) to furnish these trisubstituted N-hetereocycles under mild conditions (Scheme 42). Reaction of 1-fluoro-4-isothiocyanatobenzene, ethyl 2-cyanoacetate, and p-tolylthainthrenium tetrafluoroborate at the gram-scale afforded product in 62% yield with 94% recovery of TT.
Extending the reaction time to 3 days, these substrates reacted under natural sunlight irradiation in 72% yield. Radical trap experiments with TEMPO and DPE were suggestive of aryl and thiol radical formation by the detection of the corresponding adducts by HRMS and a significant reduction in the desired product yield. An on–off experiment provided negative evidence for a chain radical mechanism. A UV-vis analysis indicated EDA complex formation by observation of a bathochromic shift when all three substrates are combined in the presence of base. Deprotonation at the α-isocyano position and subsequent [3 + 2] cycloaddition with isothiocyanates establishes the imidazole backbone, and this anionic intermediate is proposed to form an EDA complex with Ar–TT+ salts. Single electron transfer within the complex generates aryl and sulfur-centered radical fragments that may recombine to yield product.
Ritter and Cornella further expanded the scope of their arene halogenation efforts in 2023 through the development of a thermally-promoted, Ni-catalyzed method for chlorination, bromination, and iodination.157 Treatment with Ar–TT+ salts with NiCl2·6H2O, Zn0, and either NaI, NaBr, or TBACl in DMA solvent under argon atmosphere at 25 °C yielded the corresponding haloarenes (Scheme 45). Application to LSF of bioactive compounds was demonstrated. Changing the [Ni] source to NiI(COD)(OPh*) (COD = 1,5-cyclooctadiene, OPh* = O(tBu)3C6H2) without addition of Zn0 led to production of N-(4-iodophenyl)-N,4-dimethylbenzenesulfonamide in 76% yield, compared to 95% under standard conditions. The activity of Ni(COD)2 at a 10 mol% loading was attributed to the oxidation of Ni0 to [NiI] by Ar–TT+ salts, as its use in stoichiometric quantity did not furnish product, an effect ascribed to the aggregation of inactive Ni0 at higher concentrations. Addition of [Cp2Fe](BArF4) (BArF4 = tetrakis[3,5-bis(trifluoromethyl)phenyl]borate) as oxidant with stoichiometric Ni(COD)2 loading promoted further reactivity (75% yield). Direct observation of [NiI] in the presence of [Cp2Fe](BArF4) through EPR spectroscopy further supported the role of NiI as the active species. Control experiments in the absence of [Ni] or [Zn] indicated their essential role in the transformation. Although the proposed intermediates were not isolated, a mechanistic hypothesis was proposed on this basis. Reduction of Ar–TT+ salts by [L–NiI–X] generates aryl radicals which oxidatively ligate [L–NiII–X]+ to afford a [Ar–NiIII–LX]+ complex. From this complex, product may be reductively eliminated and the resting catalyst regenerated through anion metathesis.
The origin of regioselectivity in thermally-promoted arene thianthrenation is the reversible formation of a Wheland intermediate followed by an irreversible, selectivity-determining deprotonation. Reversibility in the formation of the Wheland intermediate thermodynamically favors Csp2–S bond formation at the least sterically-hindered, electronically preferred para-position; positive charge accumulation within this intermediate biases the efficiency of this strategy toward electron-donating substituents at the ortho- and para-positions. Preferential reactivity subsequently occurs at the exocyclic Csp2–S site as the LUMO demonstrates antibonding character at this site and a dearomative enthalpic penalty is incurred by endocyclic reactivity.
Thianthrene is redox active, twice reversibly oxidized to the mono- and dication, and important implications of this were described in systems where these species functioned as a redox mediator or to sustain photoredox cycles. Further, the reversibility of these redox events enables the recycling of thianthrene and its catalytic reuse in one-pot, two-step, thianthrenation–functionalization reactions, increasing both the efficiency and sustainability of the platform. Formation of EDA complexes between Ar–TT+ salts and electron-donors enables the use of longer irradiation wavelengths, often within the visible region, and thus milder conditions and greater functional group tolerance of these systems. Late-stage functionalization of structurally complex substrates was achieved for a variety of transformations, demonstrating the chemoselectivity advantage of thianthrenation and the mild conditions required to promote radical reactivity of the corresponding Ar–TT+ salts. Achievement of LSF under mild conditions, aerobic atmosphere, aqueous media, and in batch and flow apparati, are promising features for scalability across academia and industry.
As a powerful Csp2–H functionalization platform, thianthrenation is methodologically limited in the directness of its approach, as is often the case for canonical methods, in that a minimum of two steps are required; activation (thianthrenation) and functionalization. However, further development of one-pot methods may mitigate this inefficiency. Sequential thianthrenation and functionalization of arenes and olefins under electrochemical conditions18,158 theoretically would enable the catalytic recycling of thianthrene. Development of redox neutral systems, enabled through regeneration of neutral TT, offers an opportunity for increased synthetic efficiency, energy efficiency, and sustainability through waste reduction. Although metallaphotoredox strategies featuring chiral ligand supports on the transition metal catalysts have been used in Csp2–Csp3 bond-forming systems of Ar–TT+ salts, very few examples of asymmetric catalysis have been reported to date. Nevertheless, this area presents substantial opportunities for further investigation. Further development of multicomponent and radical cascade reactions is desirable for the synthetic utility of multiple-bond formation events in a single reaction which may be otherwise challenging. Thianthrenation is therefore a promising Csp2–H activation method that merits additional research.
| This journal is © the Partner Organisations 2026 |