Abstract
Bismuth has recently been shown to be able to maneuver between different oxidation states, enabling access to unique redox cycles that can be harnessed in the context of organic synthesis. Indeed, various catalytic Bi redox platforms have been discovered and revealed emerging opportunities in the field of main group redox catalysis. The goal of this perspective is to provide an overview of the synthetic methodologies that have been developed to date, which capitalize on the Bi redox cycling. Recent catalytic methods via low-valent Bi(II)/Bi(III), Bi(I)/Bi(III), and high-valent Bi(III)/Bi(V) redox couples are covered as well as their underlying mechanisms and key intermediates. In addition, we illustrate different design strategies stabilizing low-valent and high-valent bismuth species, and highlight the characteristic reactivity of bismuth complexes, compared to the lighter p-block and d-block elements. Although it is not redox catalysis in nature, we also discuss a recent example of non-Lewis acid, redox-neutral Bi(III) catalysis proceeding through catalytic organometallic steps. We close by discussing opportunities and future directions in this emerging field of catalysis. We hope that this Perspective will provide synthetic chemists with guiding principles for the future development of catalytic transformations employing bismuth.
Keywords: Main group catalysis, redox cycling, organobismuth, organic synthesis, heavy elements
1. Introduction
Catalysis based on transition metals (TMs) has undoubtedly had a profound impact in synthetic chemistry, ranging from laboratory routes to industrial-scale processes.1,2 One of the key underlying principles for the success of TM catalysis lies on their facile redox cycling between different oxidation states, which is coupled with bond formation or cleavage in its coordination sphere.3,4 For decades, such redox properties were believed to be the realm of TMs, and elements beyond the d-block were certainly not considered for such purposes. However, such dogmatic belief has been recently challenged, as many main group elements have been demonstrated to undergo elementary redox reactions that were once believed to be reserved for TMs. It is now common to see a growing number of articles in which s- and p-block elements undergo oxidative addition, reductive elimination, etc., leading to reactivity that resembles that of TMs and, in some instances, complements them.5−8 Despite the wealth of literature on the development of stoichiometric redox reactivity by main group elements, the development of efficient catalytic redox processes is challenging and is still in its infancy.
A major challenge is that unlocking catalytic redox cycle in main group elements requires a complete redesign of the approach when compared to TMs, because of the different electronic situation encountered for such elements (Figure 1). Generally, the facile access to various stable oxidation states in TM complexes is a consequence of the small energy gap between their frontier d-orbitals. On the other hand, the gap between the frontier s-/p-based orbitals in p-block elements is substantially larger.5 As a result, redox events at p-block elements generally occur unidirectionally in a stoichiometric fashion, facilitated by strong thermodynamic driving forces. Hence, the p-block products in the redox reactions are often thermodynamic sinks, and their recycling toward catalytic protocols becomes problematic. For example, oxidative addition of E–H or C–X (E = H, C, N, O; X = F, Cl, Br, I) bonds to low-valent group 13 or 14 compounds is often favorable,9−12 but reductive elimination from the high-valent compounds of these elements is challenging. In contrast, C–H, C–C, or C–X (X = F, Cl) bond formation via reductive elimination from high-valent group 16 or 17 compounds is well-established;13,14 yet, oxidative addition to low-valent compounds of these elements is rare. Because of such biased thermodynamics between low-valent and high-valent species, merging redox elementary steps into redox cycling remains a tremendous challenge for the aforementioned p-block elements.
The group 15 elements, or pnictogens, are located in a transitional position on the p-block from a reactivity standpoint. Hence, a bidirectional Pn(n)/Pn(n+2) redox couple at group 15 elements should be viable.15,16 Considering the foregoing pnictogen properties, the Radosevich group pioneered phosphorus redox catalysis via a P(III)/P(V) cycle.17 By virtue of strained tridentate and bidentate ligands around phosphorus, the group overcame the large energetic gap between the frontier orbitals, thereby enabling redox catalysis. In addition, by lowering the kinetic barrier of phosphine oxide reduction, numerous deoxygenative transformations via P(III)/P(V)=O cycling have been accomplished.16,18 While the heavier congeners, arsenic and antimony, have a few examples of redox catalysis,19,20 the heaviest stable element in group 15—bismuth—has had limited applications until recently, despite its low cost and low toxicity.21 Indeed, bismuth has presented its candidacy as a versatile element in redox catalysis, given its rich redox chemistry and labile bonding nature that can enable new organic transformations and small molecule activation through Bi redox cycling.
Traditionally, the catalytic utility of Bi(III) compounds in synthesis was limited to nonredox soft Lewis acid catalysis such as carbonyl and diene activation.22−24 Organobismuth(III) compounds also found synthetic applications as transmetalation reagents, because of the lability of Bi(III)–C bonds.25 In the presence of strong oxidants, these compounds undergo oxidation leading to organobismuth(V) compounds, which have been applied as oxidants in synthesis.26 Moreover, the labile nature of the Bi(V)–C bond has also been employed in several stoichiometric oxidative ligand coupling events toward C–C or C–X bond formation.27,28 In heterogeneous electrocatalysis or photocatalysis, Bi is frequently encountered as a dopant in materials in order to advance performance (e.g., selectivity and efficiency) in CO2 reduction, water splitting, and dinitrogen reduction.29 Despite these seminal reports indicating the rich redox chemistry of bismuth, the development of catalytic redox processes for organic synthesis have been largely overlooked. In this Perspective, we describe how the area of bismuth redox catalysis has flourished with illustration of recent developments in Bi(II)/Bi(III), Bi(I)/Bi(III), and Bi(III)/Bi(V) redox cycles. We aim to show design principles for different redox cycles toward new catalytic transformations. Albeit not via redox cycling, we also provide an example of redox-neutral Bi(III) catalysis via elementary organometallic steps. We believe that this Perspective provides a summary of the unique reactivity that makes bismuth a versatile element for redox catalysis, with properties that go beyond a mere TM alternative.
2. Low-Valent Catalysis
Unlike the lighter pnictogen elements, Bi has been shown to expand beyond the canonical Pn(III)/Pn(V) redox pair (see Section 3), enabling redox cycling at a rather unique low-valent platform, and hence, explore its reductive properties.30 In this section, we describe Bi(II)/Bi(III) and Bi(I)/Bi(III) redox cycles and the organic transformations derived thereof.
2.1. Low-Valent Bi(II)/Bi(III) Catalysis
The current examples of other group 15 elements in redox catalysis capitalize on the two-electron Pn(n)/Pn(n+2) redox pair for productive catalysis. In contrast, bismuth offers a unique opportunity for reactivity beyond such canonical redox pair through one-electron Bi(II)/Bi(III) redox processes.
As a seminal example of Bi(II) radical intermediates for synthetic use, Yamago reported organobismuthine-mediated radical polymerization reactions in 2007 (Figure 2).31 It was proposed that a bismuthine promotor 1 undergoes Bi–C bond homolysis to generate alkyl radical 2 and Bi(II) radical 3 through thermal decomposition or radical initiation by a substoichiometric amount of 2,2′-azobis(isobutyronitrile) (AIBN). The resulting Bi(II) species 3 promotes radical polymerization via the intermediacy of alkylbismuthine 4 that can further promote polymerization through homolytic Bi–C cleavage, enabling highly controlled living radical polymerization. In a follow-up report in 2009,32 arylthiobismuthine 5 is used as a co-catalyst for the synthesis of high-molecular-weight polystyrenes (Mn = 198 000 in the presence of 5; Mn = 107 000 in the absence of 5), given that thiyl radicals facilitate the homolytic substitution by alkyl radicals, as previously described by Barton.33 A cationic Bi(III) species 6 was reported by Lichtenberg and Okuda, which was also shown to drive the radical polymerization of styrene via Bi–C homolysis.34
While transient Bi(II) species have been proposed in polymerization reactions, in 2014, Iwamoto first reported a persistent Bi(II) radical that is in equilibrium with its dimer in solution.35 Later, Coles reported the first monomeric Bi(II) species 7 that is stable both in solution and in the solid state (Figure 3).36 Building on this seminal work on the isolable Bi(II) species, in 2018, Coles reported a Bi(II)/Bi(III) redox cycle in a catalytic dehydrocoupling reaction of TEMPO radical (11) and phenylsilane using Bi(II) complex 7 as a catalyst.37 This stable and isolable complex (7) supported by bis(amido)disiloxane ligand was previously shown to exhibit one-electron reducing character in a stoichiometric manner.38 Merging its reducing character and the reactive nature of Bi(III)–O bonds,39 the catalytic dehydrocoupling of TEMPO (11) and phenylsilane was achieved, albeit under high temperatures and long reaction times. In this reaction, 7 undergoes radical recombination with TEMPO (11) to generate Bi(III)–TEMPOxide intermediate 9. This resulting Bi(III) species is then proposed to undergo metathesis with the Si–H bond in phenylsilane to produce the dehydrocoupling product 12 and a Bi(III)–hydride species 10. The putative Bi(III)–hydride species 10 was previously shown to oxidatively generate H2,38b which would regenerate the Bi(II) catalyst 7 and close the Bi(II)/Bi(III) redox cycle.
An analogous Bi(II)/Bi(III) cycle was demonstrated by Lichtenberg using a diaryl(bismuth)thiolate precatalyst 8.40 It is proposed to generate a transient Bi(II) species upon UV irradiation via Bi–S bond homolysis and initiate the aforementioned Bi(II)/Bi(III) cycle toward dehydrocoupling. The comparatively faster reaction rates were achieved by in situ generation of the presumed catalytically active Bi(II) species.
In 2019, Lichtenberg also demonstrated a radical cycloisomerization of δ-iodo-olefins using a TM bismuthane 14, which would potentially proceed via a Bi(II)/Bi(III) cycle (Figure 4).41 It was shown that δ-iodo-olefins were readily cyclized to form 5-membered rings (19–22). In the reaction mixture, MnI(CO)5 and alkyldiphenylbismuth compound 15 were detected by IR and NMR spectroscopy, respectively. The Bi(III) compound 15 would potentially undergo Bi–C homolysis to generate Bi(II) species 16 and alkyl radical 18, and ultimately deliver 19 through a 5-exo-trig cyclization and radical trap by MnI(CO)5 or 17. In the presence of nitrone, the alkyl radical 18 can be trapped, supporting the proposed radical mechanism. Although complex 14 can behave both as a redox catalyst or a radical initiator, such radical chemistry is successfully triggered thermally because of the weak Bi–Mn bond.
2.2. Low-Valent Bi(I)/Bi(III) Catalysis
While the 6s2 nonbonding lone pair in Bi(III) compounds is relatively low in energy and rather inert,42 the 6p2 lone pair in Bi(I) compounds has been shown to be nucleophilic and readily oxidized. Since the seminal report by Tokitoh in 1997, organobismuth(I) compounds have largely been reported as oligomeric or dimeric structures.43 It was not until 2010, when Dostál reported the first monomeric, well-defined organobismuth(I) complex supported by an N,C,N-pincer ligand (Figure 5).44a Access to this isolable bismuthinidene 23 is feasible due to its stability rendered by the two N atoms in the imine arms, which coordinate to the low-lying Bi empty p-orbital.44b In addition, the 6p2 lone pair in the pz orbital perpendicular to the ligand plane, is stabilized further through delocalization across the π-bond of the ligand.
Building upon Dostál’s complex (23), in 2019, our group demonstrated that an analogous Bi(I) complex 24 is a competent catalyst in the context of transfer hydrogenation (Figure 6).45 Under the optimized reaction conditions, the complex 24 catalyzes the transfer hydrogenation from ammonia borane to either azoarenes or nitroarenes across the N–N or N–O π-bond, respectively, with good functional group tolerance. Azoarenes containing electron-donating or withdrawing groups can be reduced to the corresponding hydrazines in good yields (26–28), and an unsymmetrical azoarene in a push–pull situation was also well-tolerated (29). A bromide substituent on the ortho-position did not affect the reactivity (30). Analogously, nitroarenes were reduced into the corresponding N-hydroxylamines in good yields, which complements TM-catalyzed reduction reactions.46 Electron-donating group (31), halide substituents (32, 33), and unsaturated functionality (34) were well-tolerated. 2-Phenylnitrobenzene was selectively reduced to the N-hydroxylamine 35 without formation of carbazole, which is complementary to the Cadogan-type P(III) catalysis.47 In analogy with the intermediacy of the P(V)-dihydride in the previously described transfer hydrogenation via a P(III)/P(V) cycle, it is proposed that this catalytic cycle also involves the intermediacy of organobismuth(III) hydrides, such as 25, that can be detected by HRMS in the reaction mixture.17a However, the detailed mechanism including Bi–H bond formation/cleavage and N–H bond formation in the products is yet unknown.
Another example of the low-valent Bi(I)/Bi(III) redox platform can be found in the catalytic activation of nitrous oxide (N2O) (Figure 7A).48 Exposure of the Bi catalyst 24, used in the transfer hydrogenation, to N2O atmosphere (1 bar) resulted in the rapid liberation of N2 and the formation of a putative dimeric bismuth oxide species. The resulting oxo species can be detected by HRMS—also observed in related bismuth oxo or sulfido dimer precedents49—but it could not be isolated or crystallographically characterized because of its thermal instability. In an attempt to characterize the organobismuth(III) intermediates, two new Bi(I) complexes were synthesized which were supported by aldimines (36) and ketimines (37) bearing m-terphenyl substituents.50 Upon exposure to N2O, the complex 36 afforded a dimeric bismuth oxide 38 with a [Bi2O2] core as its anti-isomer which was identified by single-crystal X-ray crystallography (Figure 7B). By contrast, the ketimine-supported 37 gave a monomeric bismuth hydroxide (40), presumably as a result of deprotonation of one of the Me groups in the backbone. These observations suggest the involvement of a transient monomeric bismuth oxide (39) upon O-atom abstraction from N2O, which subsequently undergoes dimerization or deprotonation via the highly polarized Bi=O bond. Eventually, treating complexes 38 and 40 with pinacolborane (HBpin, 2 equiv) resulted in regeneration of 36 and 37, respectively, and formation of the mixture of HOBpin (41) and O(Bpin)2 (42).
With validated oxidation and reduction as potential catalytic elementary steps, catalytic reactions were performed using these Bi(I) catalysts 36, 37, and 24. Indeed, the catalysts drive N2O deoxygenation with TON of 54, 89, and 80, respectively. The most reactive 24 toward N2O enabled lowering the catalyst loading to 0.01 mol % and enhancing the TON to 6700 (TOF ≈ 52 min–1). This is an unprecedented catalytic activation of N2O by main group elements,51−53 which performs comparable or beyond TM deoxygenation catalysis.54−57 Albeit restricted to HBpin, this work reveals the potential of Bi catalysts for O-atom transfer (OAT), which can eventually be applied to organic substrates in catalytic oxidative transformations.
The catalytic utility of the N,C,N-chelated bismuthinidenes via a Bi(I)/Bi(III) redox cycle has been further expanded for hydrodefluorination (HDF) (Figure 8A).58 In this case, new bismuthinidenes 43 and 44 are employed, featuring oxazolines as pendant supporting arms, which is distinct from the previously reported Bi(I) complexes (i.e., 24, 36, 37). While HDF of hexafluorobenzene using the catalyst 24 (5 mol %) gave a trace amount of the product (53) in the presence of diethylsilane (2.4 equiv), the use of the newly designed catalysts 43 and 44 supported by a 2,6-bis(oxazolinyl)phenyl (Phebox) scaffold resulted in improvement to 40% and 74% yields, respectively. The structure of 43 and 44 that contains elongated Bi–C bonds and longer Bi–N bonds indicates that the advanced reactivity is achieved by virtue of more localized 6pz lone pair orbital in 43 and 44 compared to that in complex 24 (Bi–C: 2.193(6) Å for 43, 2.201(4) Å for 44, 2.146(18) Å for 24; Bi–N: 2.502(3) and 2.525(3) Å for 43, 2.5142(18) and 2.5359(19) Å for 44, 2.492(8) and 2.502(9) Å for 24). This method hydrodefluorinates a variety of polyfluoroarenes under mild conditions. Pentafluoropyridine and pentafluorobenzene with a strong electron-withdrawing group underwent HDF to give the products (49, 50) in quantitative yields at ambient temperature using 43. Di-HDF (43) and HDF (44) of highly fluorinated phosphine are viable at 60 °C. Partially fluorinated substrates can be defluorinated (54–56), using 44 as a catalyst in moderate yields. No directing effect was observed (55), providing orthogonal selectivity to TM catalysis.59 Polyfluoroarenes bearing an electron-donating group were found to be challenging, as similarly observed in HDFs using TM catalysts.60 Mechanistically, it is proposed that the Bi(I) catalyst 43 undergoes intermolecular oxidative addition (OA) of C(sp2)–F bond to give fluoroarylbismuthine 45 (Figure 8B). Subsequent ligand metathesis (LM) with diethylsilane (57) leads to the formation of hydridoarylbismuthine 47. It was found that a substoichiometric amount of 57 can be used as a 2 equiv hydride source, giving diethylfluorosilane (58) upon completion of the reactions. Lastly, reductive elimination (RE) of C–H bond gives the HDF product 49 and regenerates the Bi(I) catalyst 43 to close the catalytic redox loop.
Each elementary step of the cycle was studied and validated experimentally (Figure 8C). In the stoichiometric reaction with pentafluoropyridine (48), it was shown that Bi(I) complex 43 undergoes C–F OA to give a fluoroarylbismuthine 45 (Figure 8C, path a). It was observed that bismuthine 45 interconverts with other Bi species in the reaction mixture, but it could be characterized by multinuclear one-dimensional (1D) and two-dimensional (2D) nuclearmagnetic resonance (NMR) spectroscopy. Subsequent addition of diethylsilane to this mixture resulted in the formation of HDF product 49 and regeneration of Bi(I) complex 43. It was hypothesized that the labile Bi–F bond in 45 is attributed to the observed interconversion, and the isolation of cationic Bi(III) species 46 was attempted using LiOTf (Figure 8C, path b). Indeed, a stable cationic Bi(III) complex 46 was obtained, which was characterized by single-crystal X-ray crystallography. Upon treating 46 with LiAlH4 (0.5 equiv), the formation of hydridoarylbismuthine 47 was observed, which at elevated temperature, gave HDF product 49 and the Bi(I) complex 43. These three elementary steps, OA, LM, and RE, are akin to those in a well-defined P(III)/P(V) synthetic cycle for HDF previously demonstrated by Radosevich.61 While a catalytic OA/LM/RE cycle in a Pn(III)/Pn(V) manifold is considered yet challenging, the Bi(I)/Bi(III) redox cycle has enabled a catalytic cycle of such sequence by a main group complex, which is conventionally confined to TM catalysis.
It is also shown that this N,C,N-chelated Bi platform for a Bi(I)/Bi(III) redox cycle is sufficiently robust to be applied in an electrocatalytic hydrogen evolution reaction (HER). The bismuthinidene 24 was in situ generated from the parent Bi(III) by electrochemical reduction, which oxidatively evolves hydrogen in the presence of acetic acid.62
3. High-Valent Bi(III)/Bi(V) Catalysis
Seminal work that capitalize upon the redox properties of organobismuth(III) in catalysis was reported by Barton and Motherwell in 1981, in which a Ph3Bi catalyst (59) was engaged with N-bromosuccinimide (NBS) as a stoichiometric oxidant for the C–C scission in 1,2-diols (Figure 9).63 Similar to the previously reported oxidative reactions by Bi(V) reagents,26 the stochiometric oxidation of an α-glycol 61 by triphenylbismuth carbonate resulted in the formation of benzaldehyde (62) presumably via a Bi(V) intermediate 60. In this reaction, Ph3Bi was formed in quantitative yield (59), and thus, the catalytic reactions were readily achieved by slow addition of NBS for regeneration of Bi(V) species. Related oxidative reactivity has also been reported using Ph3Sb.20 However, in these examples, the organometallic compounds are mere oxidants and the Bi–C bonds remain intact throughout the catalysis.
Stoichiometric examples where the aryl ligands on organobismuth(V) complexes are engaged in reactivity can be found in the arylation of organic compounds through oxidative ligand coupling or formal reductive elimination.28,64 In order to advance these stoichiometric reactions to catalysis, the oxidation step (Bi(III) → Bi(V)) should be compatible with other elementary steps, and after ligand coupling, the commonly reactive aryl group must be reincorporated. In 2020, a demonstration of bismuth redox catalysis via canonical organometallic elementary steps in a Bi(III)/Bi(V) platform was described in the catalytic fluorination of aryl boronic esters (Figure 10A).65 A bismuth catalyst (63) supported by a dianionic bis-aryl sulfoximine ligand is used to enable Bi(III)/Bi(V) redox cycling, where the NCF3 coordinates to the Bi center as a weak ligand. With this ligand platform, Bi(III) catalyst 63 (10 mol %) drives fluorination of aryl boronic esters in the presence of 2,6-dichloro-1-fluoropyridinium tetrafluoroborate (67, 1 equiv) and sodium fluoride (5 equiv). Boronic esters with para-substitution are generally well-tolerated (70, 71). Those with meta- or ortho-substitution are amenable but relatively more challenging (72–75). Notably, the reaction did not proceed in the absence of the bismuth catalyst.66
The catalytic Bi(III)/Bi(V) redox cycle is proposed for this fluorination on the basis of stoichiometric reactivity and identification of key intermediates (Figure 10B). It is proposed that the Bi(III) catalyst 63 first undergoes transmetalation with an aryl boronic ester to generate phenylbismuthine 65. The complex 65 is then oxidized by fluoropyridinium reagent 67 to afford a high-valent Bi(V) species 66 that is stabilized by coordination of the pendant NCF3 moiety. The resulting Bi(V) species 66 undergoes reductive elimination of fluorobenzene 69 to regenerate Bi(III) species 64, enabling Bi(III)/Bi(V) redox cycling. These elementary steps, including transmetalation, oxidative addition, and reductive elimination, have all been confirmed to proceed stoichiometrically (Figure 10C). The transmetalation product 65 was crystallographically characterized, and the following oxidation step yielded the fluorination product 69 and regenerated the Bi(III) catalyst 63. The Bi(V) intermediate 66 was independently synthesized and shown to undergo reductive elimination of C–F bond to give 69. Indeed, they are reminiscent of those in classic TM-catalyzed cross-coupling reactions. However, this Bi(III)/Bi(V) redox catalysis does not merely mimic TM catalysis but enables a catalytic reaction that is considered to be challenging with TMs.67−70 This chemistry exhibits the potential of bismuth redox catalysis to unveil new reactions unique to bismuth.
In addition, it is interesting that the Bi(III)/Bi(V) manifold operates fluorination whereas the Bi(I)/Bi(III) manifold can drive defluorination of polyfluoroarenes. These two catalytic reactions explicitly exemplify that both catalytic oxidative and reductive chemistry can be achieved by employing relevant Bi redox couples.
The tethered bis-aryl ligand platform bearing a pendant neutral ligand was further exploited in an oxidative coupling of arylboronic acids with perfluoroalkyl sulfonate salts (Figure 11A).71 A related stoichiometric reaction was previously reported by Mukaiyama, in which a stoichiometric oxidative coupling between phenylbismuthine species and an excess of triflic acid in the presence of m-chloroperoxybenzoic acid (mCPBA) gives phenyl triflate (82) in 29% yield.28f This reductive elimination type reaction was advanced into Bi-catalyzed triflation and nonaflation of aryl boronic acids using the corresponding perfluoroalkyl sulfonate salts. In this catalysis, a modified Bi(III) catalyst 76 ligated with bis(m-trifluoromethyl-phenyl) sulfone is used. A more electron-deficient backbone is introduced to render the Bi(III) a better leaving group during C–O bond formation. Indeed, Bi(III) catalysts 76 and 77 (10 mol %) drive triflation and nonaflation of aryl boronic acids, respectively, in the presence of a perfluoroalkyl sulfonate salt (NaOTf or KONf, 1.1 equiv), an oxidant 67 (1.1 equiv), and Na3PO4 (2 equiv). This method accommodates aryl boronic acids with various types of substitution (82–91).
Analogous to the fluorination catalyzed by 63, the Bi(III) catalyst (76 or 77) is proposed to undergo transmetalation with aryl boronic acids (Figure 11B). Subsequent oxidation of the resulting phenylbismuthine 79 by oxidant 67 generates a putative high-valent Bi(V) species (80 or 81), which can be detected by HRMS in the corresponding stoichiometric reaction with NaOTf (Figure 11C). The Bi(V) intermediate (80 or 81) then undergoes reductive elimination of the corresponding aryl triflate 82 or aryl nonaflate 87 with the regeneration of Bi(III) species (78). DFT calculations suggest that ligand coupling occurs via a five-membered transition state in which two oxygens of triflate are involved. This transition state is shown to be favored over a three-membered transition state via which reductive elimination from transition metals typically occurs (ΔG5-mem,calc⧧ = 22.9 kcal/mol ΔG3-mem,calc⧧ = 24.9 kcal/mol).
Given that triflate and nonaflate are highly electronegative, poorly nucleophilic, and weakly coordinating,72 C–O bond formation is inherently challenging using TM complexes and a few chemistry is known for triflation steps by transition metals.73 As revealed in fluorination catalysis, this oxidative coupling reaction via a Bi(III)/Bi(V) cycle has also enabled new transformations that are relatively challenging in TM catalysis by taking advantage of the unique reactivity of bismuth complexes.
4. Redox-Neutral Bi(III) Catalysis
While non-redox Bi catalysis has been dominated by Lewis acid catalyzed transformations, redox-neutral Bi(III) catalysis in which a Bi(III) catalyst does not act as a mere Lewis acid has been recently demonstrated in a catalytic synthesis of (hetero)aryl sulfonyl fluorides from (hetero)aryl boronic acids using sulfur dioxide (SO2) and Selectfluor (96) (Figure 12).77 A Bi(III) catalyst 92 was found to be an optimal catalyst for this reaction, which was discovered through modification of the diarylsulfone ligands that were used in Bi(III)/Bi(V) redox catalysis. Analogous to the fluorination and the triflation catalysis (section 3), the catalyst 92 readily undergoes transmetallation with a (hetero)aryl boronic acid to generate (hetero)arylbismuthine 94. Although stoichiometric reactivities of both Bi(III)–C78 and Bi(V)–C79 toward SO2 or its surrogates have been known, it was found that Bi(III) intermediate 94 is an active species for the insertion of SO2 into the Bi–C bond which gives a diarylbismuth sulfinate 95. Upon oxidation by electrophilic fluorine source, complex 95 then affords the corresponding (hetero)aryl sulfonyl fluoride with regeneration of catalytically active Bi(III) species 93. In this catalytic reaction, it was demonstrated that the oxidant (96) does not oxidize the Bi center but the S(IV) to form the S–F bond. Thus, unlike aforementioned Bi(III)/Bi(V) redox catalysis, this reaction does not involve Bi(V) intermediates, achieving a redox-neutral Bi(III) catalysis with an insertion step. This method can tolerate various functional groups (97–102). In addition, by using a milder oxidant, N-fluorobenzenesulfonimide (NFSI), instead of 96, heteroaryl boronic acids can be tolerated (103–106), which has been challenging for Bi catalysis using strong oxidants. Compared to related two-step sequence TM-catalyzed reactions,80 Bi catalysis provides a convenient and competent synthetic method for sulfonyl fluorides with high functional group compatibility.
5. Conclusions and Future Outlook
In this Perspective, we provided an overview of recent efforts to leverage the Bi(II)/Bi(III), Bi(I)/Bi(III), and Bi(III)/Bi(V) redox couples for discovery of new reactivity and development of new main group catalysis. It is clear that the field of bismuth redox catalysis is still in its infancy, yet it has provided compelling evidence that new applications are awaiting in the area of organic synthesis, which could certainly reach beyond those catalyzed by transition metals. Looking ahead, we anticipate numerous opportunities for discovery in bismuth redox catalysis:
-
(1)
While other elementary steps (e.g., OA, LM, RE) comprising TM catalysis are analogously realized in Bi catalysis, combinations with migratory insertion-type elementary steps are relatively less explored.28i,49b,77,74 Incorporating this type of insertion step into a catalytic cycle would enable new transformations, using small molecules as building blocks.
-
(2)
Compared to organo–Bi(I), −Bi(III), and −Bi(V) compounds, open-shell −Bi(0), −Bi(II), and −Bi(IV) species are much less explored or unknown.75 These species might provide further opportunities to unveil new catalytic cycles and reaction pathways. For example, dual photoredox/bismuth catalysis or combined electrochemistry with bismuth, could be envisioned to precisely control oxidation or reduction of bismuth complexes and ultimately enable selective or challenging transformations.
-
(3)
Combined with diverse redox catalysis and the robust nature of Bi complexes, development of molecular Bi-based electrocatalysts is feasible, which could provide better understanding in heterogeneous electrocatalysis at the molecular level.62
-
(4)
Given the orthogonal or complementary reactivity of Bi catalysts, Bi redox cycles could be merged with TM catalysis to achieve challenging transformations via dual Bi/TM catalysis.
-
(5)
In more-complex settings, stereoselective redox transformations or bioconjugation reactions could be devised for organobismuth complexes, because of the facile introduction of chiral elements around its Bi center, and their benign nature and stability in aqueous solutions, respectively.
-
(6)
Finally, dinuclear or multinuclear complexes with mixed oxidation states would also allow one to discover new redox chemistry.76
We hope that this Perspective will draw more attention to the infant field of redox catalysis using bismuth, provide guiding principles for new advancements in catalysis and synthesis, and promote realization of the full potential of such a new area of expertise.
Acknowledgments
Financial support for this work was provided by Max-Planck-Gesellschaft, Max-Planck-Institut für Kohlenforschung and Fonds der Chemischen Industrie (FCI-VCI) and Deutsche Forschungsgemeinschaft. This project has received funding from European Union’s Horizon 2020 research and innovation programme under Agreement No. 850496 (ERC Starting Grant, J.C.). We thank Prof. Dr. A. Fürstner for insightful discussions and generous support.
Open access funded by Max Planck Society.
The authors declare no competing financial interest.
References
- Beller M.; Bolm C.; Eds. Transition Metals for Organic Synthesis, 2nd Edition; Wiley: Weinheim, Germany, 2004; pp 15–25. [Google Scholar]
- Diederich F.; Stang P. J.. Metal-Catalyzed Cross Coupling Reactions; Wiley-VCH: Weinheim, Germany, 1998. [Google Scholar]
- Crabtree R. H.The Organometallic Chemistry of the Transition Metals, 6th Edition; Wiley: Hoboken, NJ, 2014; pp 1–3, 163–181. [Google Scholar]
- Tamao K.; Miyaura N.. Introduction to Cross-Coupling Reactions. In Cross-Coupling Reactions: A Practical Guide; Miyaura N., Ed.; Topics in Current Chemistry219; Springer: Berlin, Heidelberg, 2002; pp 1–9. [Google Scholar]
- Power P. P. Main-group elements as transition metals. Nature 2010, 463, 171–177. 10.1038/nature08634. [DOI] [PubMed] [Google Scholar]
- Weetman C.; Inoue S. The Road Travelled: After Main-Group Elements as Transition Metals. ChemCatChem 2018, 10, 4213–4228. 10.1002/cctc.201800963. [DOI] [Google Scholar]
- Chu T.; Nikonov G. I. Oxidative Addition and Reductive Elimination at Main-Group Element Centers. Chem. Rev. 2018, 118, 3608–3680. 10.1021/acs.chemrev.7b00572. [DOI] [PubMed] [Google Scholar]
- Melen R. L. Frontiers in molecular p-block chemistry: From structure to reactivity. Science 2019, 363, 479–484. 10.1126/science.aau5105. [DOI] [PubMed] [Google Scholar]
- a Bonyhady S. J.; Collis D.; Frenking G.; Holzmann N.; Jones C.; Stasch A. Synthesis of a stable adduct of dialane(4) (Al2H4) via hydrogenation of a magnesium(I) dimer. Nat. Chem. 2010, 2, 865–869. 10.1038/nchem.762. [DOI] [PubMed] [Google Scholar]; b Chu T.; Korobkov I.; Nikonov G. I. Oxidative Addition of σ Bonds to an Al(I) Center. J. Am. Chem. Soc. 2014, 136, 9195–9202. 10.1021/ja5038337. [DOI] [PubMed] [Google Scholar]; c Protchenko A. V.; Birjkumar K. H.; Dange D.; Schwarz A. D.; Vidovic D.; Jones C.; Kaltsoyannis N.; Mountford P.; Aldridge S. A Stable Two-Coordinate Acyclic Silylene. J. Am. Chem. Soc. 2012, 134, 6500–6503. 10.1021/ja301042u. [DOI] [PubMed] [Google Scholar]
- a Arduengo A. J. III; Calabrese J. C.; Davidson F.; Rasika Dias H. V.; Goerlich J. R.; Krafczyk R.; Marshall W. J.; Tamm M.; Schmutzler R. C-H Insertion Reactions of Nucleophilic Carbenes. Helv. Chim. Acta 1999, 82, 2348–2364. . [DOI] [Google Scholar]; b Miller K. A.; Watson T. W.; Bender; Banaszak Holl M. M.; Kampf J. W. Intermolecular C-H Insertions and Cyclization Reactions Involving a Stable Germylene. J. Am. Chem. Soc. 2001, 123, 982–983. 10.1021/ja0026408. [DOI] [PubMed] [Google Scholar]; c Korotkikh N. I.; Rayenko G. F.; Shvaika O. P.; Pekhtereva T. M.; Cowley A. H.; Jones J. N.; Macdonald C. L. B. Synthesis of 1,2,4-Triazol-5-ylidenes and Their Interaction with Acetonitrile and Chalcogens. J. Org. Chem. 2003, 68, 5762–5765. 10.1021/jo034234n. [DOI] [PubMed] [Google Scholar]; d Vignolle J.; Asay M.; Miqueu K.; Bourissou D.; Bertrand G. Rearrangement of Biaryl Monoaminocarbenes via Concerted Asynchronous Insertion into Aromatic C-H Bonds. Org. Lett. 2008, 10, 4299–4302. 10.1021/ol801670z. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Moerdyk J. P.; Bielawski C. W. N,N′-Diamidocarbenes Facilitate Selective C-H Insertions and Transfer Hydrogenations. Chem. - Eur. J. 2013, 19, 14773–14776. 10.1002/chem.201302691. [DOI] [PubMed] [Google Scholar]; f Abdalla J. A. B.; Riddlestone I. M.; Tirfoin R.; Aldridge S. Cooperative Bond Activation and Catalytic Reduction of Carbon Dioxide at a Group 13 Metal Center. Angew. Chem., Int. Ed. 2015, 54, 5098–5102. 10.1002/anie.201500570. [DOI] [PubMed] [Google Scholar]
- a Lappert M. F.; Miles S. J.; Atwood J. L.; Zaworotko M. J.; Carty A. J. Oxidative addition of an alcohol to the alkylgermanium(II) compound Ge[CH(SiMe3)2]2; molecular structure of Ge[CH(SiMe3)2]2(H)OEt. J. Organomet. Chem. 1981, 212, C4–C6. 10.1016/S0022-328X(00)85535-7. [DOI] [Google Scholar]; b Frey G. D.; Lavallo V.; Donnadieu B.; Schoeller W. W.; Bertrand G. Facile Splitting of Hydrogen and Ammonia by Nucleophilic Activation at a Single Carbon Center. Science 2007, 316, 439–441. 10.1126/science.1141474. [DOI] [PubMed] [Google Scholar]; c Peng Y.; Guo J.-D.; Ellis B. D.; Zhu Z.; Fettinger J. C.; Nagase S.; Power P. P. Reaction of Hydrogen or Ammonia with Unsaturated Germanium or Tin Molecules under Ambient Conditions: Oxidative Addition versus Arene Elimination. J. Am. Chem. Soc. 2009, 131, 16272–16282. 10.1021/ja9068408. [DOI] [PubMed] [Google Scholar]; d Wang W.; Inoue S.; Yao S.; Driess M. Reactivity of N-Heterocyclic Germylene Toward Ammonia and Water. Organometallics 2011, 30, 6490–6494. 10.1021/om200970s. [DOI] [Google Scholar]; e Protchenko A. V.; Bates J. I.; Saleh L. M. A.; Blake M. P.; Schwarz A. D.; Kolychev E. L.; Thompson A. L.; Jones C.; Mountford P.; Aldridge S. Enabling and Probing Oxidative Addition and Reductive Elimination at a Group 14 Metal Center: Cleavage and Functionalization of E-H Bonds by a Bis(boryl)stannylene. J. Am. Chem. Soc. 2016, 138, 4555–4564. 10.1021/jacs.6b00710. [DOI] [PubMed] [Google Scholar]
- a Atwell W. H.; Weyenberg D. R. Divalent Silicon Intermediates. Angew. Chem., Int. Ed. Engl. 1969, 8, 469–477. 10.1002/anie.196904691. [DOI] [Google Scholar]; b Gynane M. J. S.; Lappert M. F.; Miles S. J.; Power P. P. Ready oxidative addition of an alkyl or aryl halide to a tin(II) alkyl or amide; evidence for a free-radical pathway. J. Chem. Soc., Chem. Commun. 1976, 256–257. 10.1039/c39760000256. [DOI] [Google Scholar]; c Eaborn C.; Hill M. S.; Hitchcock P. B.; Patel D.; Smith J. D.; Zhang S. Oxidative Addition to a Monomeric Stannylene To Give Four-Coordinate Tin Compounds Containing the Bulky Bidentate Ligand C(SiMe3)2SiMe2CH2CH2Me2Si(Me3Si)2C. Crystal Structures of CH2Me2Si(Me3Si)2CSnC(SiMe3)2SiMe2CH2, CH2Me2Si(Me3Si)2CSnMe(OCOCF3)C(SiMe3)2SiMe2CH2, and (CF3COO)2MeSnC(SiMe3)2SiMe2CH2CH2Me2Si-(Me3Si)2CSnMe(OCOCF3)2. Organometallics 2000, 19, 49–53. 10.1021/om990779p. [DOI] [Google Scholar]; d Xiong Y.; Yao S.; Driess M. Reactivity of a Zwitterionic Stable Silylene toward Halosilanes and Haloalkanes. Organometallics 2009, 28, 1927–1933. 10.1021/om801157u. [DOI] [Google Scholar]; e Jana A.; Samuel P. P.; Tavčar G.; Roesky H. W.; Schulzke C. Selective Aromatic C-F and C-H Bond Activation with Silylenes of Different Coordinate Silicon. J. Am. Chem. Soc. 2010, 132, 10164–10170. 10.1021/ja103988d. [DOI] [PubMed] [Google Scholar]; f Crimmin M. R.; Butler M. J.; White A. J. P. Oxidative addition of carbon-fluorine and carbon-oxygen bonds to Al(I). Chem. Commun. 2015, 51, 15994–15996. 10.1039/C5CC07140B. [DOI] [PMC free article] [PubMed] [Google Scholar]; g Chu T.; Boyko Y.; Korobkov I.; Nikonov G. I. Transition Metal-like Oxidative Addition of C-F and C-O Bonds to an Aluminum(I) Center. Organometallics 2015, 34, 5363–5365. 10.1021/acs.organomet.5b00793. [DOI] [Google Scholar]
- a Oae S. Ligand Coupling Reactions Through Hypervalent and Similar Valence-Shell Expanded Intermediates. Croat. Chem. Acta 1986, 59, 129–151. [Google Scholar]; b Sagae T.; Ogawa S.; Furukawa N. Stereochemical Proof for Front Side Deuteride Attack via σ-Sulfurane in the Reductive Desulfinylation of Sulfoxides with Lithium Aluminum Deuteride. Tetrahedron Lett. 1993, 34, 4043–4046. 10.1016/S0040-4039(00)60611-1. [DOI] [Google Scholar]; c Crivello J. V. Redox Initiated Cationic Polymerization: Reduction of Diaryliodonium Salts by 9-BBN. J. Polym. Sci., Part A: Polym. Chem. 2009, 47, 5639–5651. 10.1002/pola.23605. [DOI] [Google Scholar]
- a Oae S. Ligand coupling reactions of hypervalent species. Pure Appl. Chem. 1996, 68, 805–812. 10.1351/pac199668040805. [DOI] [Google Scholar]; b Stang P. J.; Zhdankin V. V. Organic Polyvalent Iodine Compounds. Chem. Rev. 1996, 96, 1123–1178. 10.1021/cr940424+. [DOI] [PubMed] [Google Scholar]; c Arnold A. M.; Ulmer A.; Gulder T. Advances in Iodine(III)-Mediated Halogenations: A Versatile Tool to Explore New Reactivities and Selectivities. Chem. - Eur. J. 2016, 22, 8728–8739. 10.1002/chem.201600449. [DOI] [PubMed] [Google Scholar]; d Kaiser D.; Klose I.; Oost R.; Neuhaus J.; Maulide N. Bond-Forming and -Breaking Reactions at Sulfur(IV): Sulfoxides, Sulfonium Salts, Sulfur Ylides, and Sulfinate Salts. Chem. Rev. 2019, 119, 8701–8780. 10.1021/acs.chemrev.9b00111. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Wei J.; Liang H.; Ni C.; Sheng R.; Hu J. Transition-Metal-Free Desulfinative Cross-Coupling of Heteroaryl Sulfinates with Grignard Reagents. Org. Lett. 2019, 21, 937–940. 10.1021/acs.orglett.8b03918. [DOI] [PubMed] [Google Scholar]; f Zhou M.; Tsien J.; Qin T. Sulfur(IV)-Mediated Unsymmetrical Heterocycle Cross-Couplings. Angew. Chem., Int. Ed. 2020, 59, 7372–7376. 10.1002/anie.201915425. [DOI] [PubMed] [Google Scholar]
- a For an overview of Bi redox catalysis, see: Janssen-Müller D.; Oestreich M.. Transition-Metal-Like Catalysis with a Main-Group Element: Bismuth-Catalyzed C-F Coupling of Aryl Boronic Esters. Angew. Chem., Int. Ed. 2020, 59, 8328−8330. [DOI] [PubMed] [Google Scholar]; b Planas O.; Cornella J. High-valent bismuth redox catalysis. Nachr. Chem. 2021, 69, 79–83. 10.1002/nadc.20214116551. [DOI] [Google Scholar]; c Planas O. Catálisis Redox con Bismuto. An. Quím. 2021, 117, 266–273. [Google Scholar]
- a For an overview of group 15 redox catalysis, see: Lipshultz J. M.; Li G.; Radosevich A. T.. Main Group Redox Catalysis of Organopnictogens: Vertical Periodic Trends and Emerging Opportunities in Group 15. J. Am. Chem. Soc. 2021, 143, 1699−1721. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Abbenseth J.; Goicoechea J. M. Recent developments in the chemistry of non-trigonal pnictogen pincer compounds: from bonding to catalysis. Chem. Sci. 2020, 11, 9728–9740. 10.1039/D0SC03819A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Dunn N. L.; Ha M. J.; Radosevich A. T. Main Group Redox Catalysis: Reversible PIII/PV Redox Cycling at a Phosphorus Platform. J. Am. Chem. Soc. 2012, 134, 11330–11333. 10.1021/ja302963p. [DOI] [PubMed] [Google Scholar]; b Reichl K. D.; Dunn N. L.; Fastuca N. J.; Radosevich A. T. Biphilic Organophosphorus Catalysis: Regioselective Reductive Transposition of Allylic Bromides via PIII/PV Redox Cycling. J. Am. Chem. Soc. 2015, 137, 5292–5295. 10.1021/jacs.5b01899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Guo H.; Fan Y. C.; Sun Z.; Wu Y.; Kwon O. Phosphine Organocatalysis. Chem. Rev. 2018, 118, 10049–10293. 10.1021/acs.chemrev.8b00081. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Xie C.; Smaligo A. J.; Song X.-R.; Kwon O. Phosphorus-Based Catalysis. ACS Cent. Sci. 2021, 7, 536–558. 10.1021/acscentsci.0c01493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Shi L.; Wang W.; Wang Y.; Huang Y. The first example of a catalytic Wittig-type reaction. Tri-n-butylarsine-catalyzed olefination in the presence of triphenyl phosphite. J. Org. Chem. 1989, 54, 2027–2028. 10.1021/jo00270a001. [DOI] [Google Scholar]; b Cao P.; Li C.-Y.; Kang Y.-B.; Xie Z.; Sun X.-L.; Tang Y. Ph3As-Catalyzed Wittig-Type Olefination of Aldehydes with Diazoacetate in the Presence of Na2S2O4. J. Org. Chem. 2007, 72, 6628–6630. 10.1021/jo0709899. [DOI] [PubMed] [Google Scholar]; c Wang P.; Liu C.-R.; Sun X.-L.; Chen S.-S.; Li J.-F.; Xie Z.; Tang Y. A newly-designed PE-supported arsine for efficient and practical catalytic Wittig olefination. Chem. Commun. 2012, 48, 290–292. 10.1039/C1CC16747B. [DOI] [PubMed] [Google Scholar]; d Inaba R.; Kawashima I.; Fujii T.; Yumura T.; Imoto H.; Naka K. Systematic Study on the Catalytic Arsa-Wittig Reaction. Chem. - Eur. J. 2020, 26, 13400–13407. 10.1002/chem.202002792. [DOI] [PubMed] [Google Scholar]
- a Akiba K.-Y.; Ohnari H.; Ohkata K. Oxidation of α-Hydroxyketones with Triphenylantimony Dibromide and Its Catalytic Cycle. Chem. Lett. 1985, 14, 1577–1580. 10.1246/cl.1985.1577. [DOI] [Google Scholar]; b Yasuike S.; Kishi Y.; Kawara S.-I.; Kurita J. Catalytic Action of Triarylstibanes: Oxidation of Benzoins into Benzyls Using Triarylstibanes Under an Aerobic Condition. Chem. Pharm. Bull. 2005, 53, 425–427. 10.1248/cpb.53.425. [DOI] [PubMed] [Google Scholar]
- Mohan R. Green bismuth. Nat. Chem. 2010, 2, 336. 10.1038/nchem.609. [DOI] [PubMed] [Google Scholar]
- Leonard N. M.; Wieland L. C.; Mohan R. S. Applications of bismuth(III) compounds in organic synthesis. Tetrahedron 2002, 58, 8373–8397. 10.1016/S0040-4020(02)01000-1. [DOI] [Google Scholar]
- Ollevier T.; Eds. Bismuth-Mediated Organic Reactions; Springer, 2012; pp 69–142, 179–198. [Google Scholar]
- Lichtenberg C. Molecular bismuth(III) monocations: structure, bonding, reactivity, and catalysis. Chem. Commun. 2021, 57, 4483–4495. 10.1039/D1CC01284C. [DOI] [PubMed] [Google Scholar]
- a Gillett-Kunnath M. M.; MacLellan J. G.; Forsyth C. M.; Andrews P. C.; Deacon G. B.; Ruhlandt-Senge K. BiPh3—A convenient synthon for heavy alkaline-earth metal amides. Chem. Commun. 2008, 4490–4492. 10.1039/b806948d. [DOI] [PubMed] [Google Scholar]; b Hébert M.; Petiot P.; Benoit E.; Dansereau J.; Ahmad T.; Le Roch A.; Ottenwaelder X.; Gagnon A. Synthesis of Highly Functionalized Triarylbismuthines by Functional Group Manipulation and Use in Palladium- and Copper-Catalyzed Arylation Reactions. J. Org. Chem. 2016, 81, 5401–5416. 10.1021/acs.joc.6b00767. [DOI] [PubMed] [Google Scholar]
- a Barton D. H. R.; Lester D. J.; Motherwell W. B.; Papoula M. T. B. Oxidation of organic substrates by pentavalent organobismuth reagents. J. Chem. Soc., Chem. Commun. 1979, 705–707. 10.1039/c39790000705. [DOI] [Google Scholar]; b Barton D. H. R.; Kitchin J. P.; Lester D. J.; Motherwell W. B.; Papoula M. T. B. Functional group oxidation by pentavalent organobismuth reagents. Tetrahedron 1981, 37, 73–79. 10.1016/0040-4020(81)85042-9. [DOI] [Google Scholar]
- a Barton D. H. R.; Lester D. J.; Motherwell W. B.; Papoula M. T. B. Obsevations on the cleavage of the bismuth-carbon bond in Bi compounds: a new arylation reaction. J. Chem. Soc., Chem. Commun. 1980, 246–247. 10.1039/C39800000246. [DOI] [Google Scholar]; b Barton D. H. R.; Blazejewski J.-C.; Charpiot B.; Lester D. J.; Motherwell W. B.; Papoula M. T. B. Comparative arylation reactions with pentaphenylbismuth and with triphenylbismuth carbonate. J. Chem. Soc., Chem. Commun. 1980, 827–829. 10.1039/c39800000827. [DOI] [Google Scholar]; c Barton D. H. R.; Kitchin J. P.; Lester D. J.; Motherwell W. B.; Papoula M. T. B. Functional group oxidation by pentavalent organobismuth reagents. Tetrahedron 1981, 37, 73–79. 10.1016/0040-4020(81)85042-9. [DOI] [Google Scholar]; d Barton D. H. R.; Charpiot B.; Motherwell W. B. Regiospecific arlation by acid/base controlled reactions of tetraphenylbismuth esters. Tetrahedron Lett. 1982, 23, 3365–3368. 10.1016/S0040-4039(00)87616-9. [DOI] [Google Scholar]; e Barton D. H. R.; Bhatnagar N. Y.; Blazejewski J.-C.; Charpiot B.; Finet J.-P.; Lester D. J.; Motherwell W. B.; Papoula M. T. B.; Stanforth S. P. Pentavalent organobismuth reagents. Part 2. The phenylation of phenols. J. Chem. Soc., Perkin Trans. 1 1985, 2657–2665. 10.1039/p19850002657. [DOI] [Google Scholar]; f Barton D. H. R.; Blazejewski J.-C.; Charpiot B.; Finet J.-P.; Motherwell W. B.; Papoula M. T. B.; Stanforth S. P. Pentavalent organobismuth reagents. Part 3. Phenylation of enols and of enolate and other anions. J. Chem. Soc., Perkin Trans. 1 1985, 2667–2675. 10.1039/p19850002667. [DOI] [Google Scholar]; g Barton D. H. R.; Bhatnagar N. Y.; Finet J.-P.; Motherwell W. B. Pentavalent organobismuth reagents. Part vi. Comparative migratory aptitudes of aryl groups in the arylation of phenols and enols by pentavalent bismuth reagents. Tetrahedron 1986, 42, 3111–3122. 10.1016/S0040-4020(01)87378-6. [DOI] [Google Scholar]
- a Ohkata K.; Takemoto S.; Ohnishi M.; Akiba K.-Y. Synthesis and chemical behaviors of 12-substituted dibenz[c,f][1,5]azastibocine and dibenz[c,f][1,5]azabismocine derivatives: evidences of 10-Pn-4 type hypervalent interaction. Tetrahedron Lett. 1989, 30, 4841–4844. 10.1016/S0040-4039(01)80523-2. [DOI] [Google Scholar]; b Suzuki H.; Murafuji T.; Azuma N. Synthesis and reactions of some new heterocyclic bismuth-(III) and -(V) compounds. 5,10-Dihydrodibenzo[b,e]bismine and related systems. J. Chem. Soc., Perkin Trans. 1 1992, 1593–1600. 10.1039/p19920001593. [DOI] [Google Scholar]; c Fedorov A. Y.; Finet J.-P. Synthesis and reactivity of pentavalent biphenyl-2,2′-ylenebismuth derivatives. J. Chem. Soc., Perkin Trans. 1 2000, 3775–3778. 10.1039/b006307j. [DOI] [Google Scholar]; d Ikegai K.; Mukaiyama T. Synthesis of N-Aryl Pyridin-2-ones via Ligand Coupling Reactions Using Pentavalent Organobismuth Reagents. Chem. Lett. 2005, 34, 1496–1497. 10.1246/cl.2005.1496. [DOI] [Google Scholar]; e Sakurai N.; Mukaiyama T. A New Preparative Method of Aryl Sulfonate Esters by Using Cyclic Organobismuth Reagents. Heterocycles 2007, 74, 771–790. 10.3987/COM-07-S(W)63. [DOI] [Google Scholar]; f Sakurai N.; Mukaiyama T. New Preparative Method of Aryl Tosylates by Using Organobismuth Reagents. Chem. Lett. 2007, 36, 928–929. 10.1246/cl.2007.928. [DOI] [Google Scholar]; g Imachi S.; Mukaiyama T. Oxidative Coupling of Carbonyl Compounds by Using Pentavalent Biphenyl-2,2′-ylenebismuth Reagents. Chem. Lett. 2007, 36, 718–719. 10.1246/cl.2007.718. [DOI] [Google Scholar]; h Jurrat M.; Maggi L.; Lewis W.; Ball L. T. Modular bismacycles for the selective C-H arylation of phenols and naphthols. Nat. Chem. 2020, 12, 260–269. 10.1038/s41557-020-0425-4. [DOI] [PubMed] [Google Scholar]; i Zhao F.; Wu X.-F. The first bismuth self-mediated oxidative carbonylative coupling reaction via BiIII/BiV redox intermediates. J. Catal. 2021, 397, 201–204. 10.1016/j.jcat.2021.03.029. [DOI] [Google Scholar]
- a Medina-Ramos J.; DiMeglio J. L.; Rosenthal J. Efficient Reduction of CO2 to CO with High Current Density Using in Situ or ex Situ Prepared Bi-Based Materials. J. Am. Chem. Soc. 2014, 136, 8361–8367. 10.1021/ja501923g. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Sun S.; Wang W.; Li D.; Zhang L.; Jiang D. Solar Light Driven Pure Water Splitting on Quantum Sized BiVO4 without any Cocatalyst. ACS Catal. 2014, 4, 3498–3503. 10.1021/cs501076a. [DOI] [Google Scholar]; c Liang L.; Lei F.; Gao S.; Sun Y.; Jiao X.; Wu J.; Qamar S.; Xie Y. Single Unit Cell Bismuth Tungstate Layers Realizing Robust Solar CO2 Reduction to Methanol. Angew. Chem., Int. Ed. 2015, 54, 13971–13974. 10.1002/anie.201506966. [DOI] [PubMed] [Google Scholar]; d Hao Y.-C.; Guo Y.; Chen L.-W.; Shu M.; Wang X.-Y.; Bu T.-A.; Gao W.-Y.; Zhang N.; Su X.; Feng X.; Zhou J.-W.; Wang B.; Hu C.-W.; Yin A.-X.; Si R.; Zhang Y.-W.; Yan C.-H. Promoting nitrogen electroreduction to ammonia with bismuth nanocrystals and potassium cations in water. Nat. Catal. 2019, 2, 448–456. 10.1038/s41929-019-0241-7. [DOI] [Google Scholar]
- Lichtenberg C. Well-Defined, Mononuclear BiI and BiII Compounds: Towards Transition-Metal-Like Behavior. Angew. Chem., Int. Ed. 2016, 55, 484–486. 10.1002/anie.201509234. [DOI] [PubMed] [Google Scholar]
- Kayahara E.; Yamago S. Development of an Arylthiobismuthine Cocatalyst in Organobismuthine-Mediated Living Radical Polymerization: Applications for Synthesis of Ultrahigh Molecular Weight Polystyrenes and Polyacrylates. J. Am. Chem. Soc. 2009, 131, 2508–2513. 10.1021/ja8092899. [DOI] [PubMed] [Google Scholar]
- Yamago S.; Kayahara E.; Kotani M.; Ray B.; Kwak Y.; Goto A.; Fukuda T. Highly Controlled Living Radical Polymerization through Dual Activation of Organobismuthines. Angew. Chem., Int. Ed. 2007, 46, 1304–1306. 10.1002/anie.200604473. [DOI] [PubMed] [Google Scholar]
- Barton D. H. R.; Bridon D.; Zard S. Z. The invention of radical reactions. Part XVIII. Decarboxylative radical addition to arsenic, antimony, and bismuth phenylsulphides - a novel synthesis of nor-alcohols from carboxylic acids. Tetrahedron 1989, 45, 2615–2626. 10.1016/S0040-4020(01)80092-2. [DOI] [Google Scholar]
- Lichtenberg C.; Pan F.; Spaniol T. P.; Englert U.; Okuda J. The Bis(allyl)bismuth Cation: A Reagent for Direct Allyl Transfer by Lewis Acid Activation and Controlled Radical Polymerization. Angew. Chem., Int. Ed. 2012, 51, 13011–13015. 10.1002/anie.201206782. [DOI] [PubMed] [Google Scholar]
- Ishida S.; Hirakawa F.; Furukawa K.; Yoza K.; Iwamoto T. Persistent Antimony- and Bismuth-Centered Radicals in Solution. Angew. Chem., Int. Ed. 2014, 53, 11172–11176. 10.1002/anie.201405509. [DOI] [PubMed] [Google Scholar]
- Schwamm R. J.; Harmer J. R.; Lein M.; Fitchett C. M.; Granville S.; Coles M. P. Isolation and Characterization of a Bismuth(II) Radical. Angew. Chem., Int. Ed. 2015, 54, 10630–10633. 10.1002/anie.201504632. [DOI] [PubMed] [Google Scholar]
- Schwamm R. J.; Lein M.; Coles M. P.; Fitchett C. M. Catalytic oxidative coupling promoted by bismuth TEMPOxide complexes. Chem. Commun. 2018, 54, 916–919. 10.1039/C7CC08402A. [DOI] [PubMed] [Google Scholar]
- a Schwamm R. J.; Lein M.; Coles M. P.; Fitchett C. M. Bi-P Bond Homolysis as a Route to Reduced Bismuth Compounds and Reversible Activation of P4. Angew. Chem., Int. Ed. 2016, 55, 14798–14801. 10.1002/anie.201608615. [DOI] [PubMed] [Google Scholar]; b Schwamm R. J.; Lein M.; Coles M. P.; Fitchett C. M. Bismuth(III) Complex of the [S4]•− Radical Anion: Dimer Formation via Pancake Bonds. J. Am. Chem. Soc. 2017, 139, 16490–16493. 10.1021/jacs.7b10454. [DOI] [PubMed] [Google Scholar]
- a Hanna T. A.; Rieger A. L.; Rieger P. H.; Wang X. Evidence for an Unstable Bi(II) Radical from Bi-O Bond Homolysis. Implications in the Rate-Determining Step of the SOHIO Process. Inorg. Chem. 2002, 41, 3590–3592. 10.1021/ic0202864. [DOI] [PubMed] [Google Scholar]; b Casely I. J.; Ziller J. W.; Fang M.; Furche F.; Evans W. J. Facile Bismuth-Oxygen Bond Cleavage, C-H Activation, and Formation of a Monodentate Carbon-Bound Oxyaryl Dianion, (C6H2tBu2–3,5-O-4)2−. J. Am. Chem. Soc. 2011, 133, 5244–5247. 10.1021/ja201128d. [DOI] [PubMed] [Google Scholar]
- Ramler J.; Krummenacher I.; Lichtenberg C. Well-Defined, Molecular Bismuth Compounds: Catalysts in Photochemically Induced Radical Dehydrocoupling Reactions. Chem. - Eur. J. 2020, 26, 14551–14555. 10.1002/chem.202002219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramler J.; Krummenacher I.; Lichtenberg C. Bismuth Compounds in Radical Catalysis: Transition Metal Bismuthanes Facilitate Thermally Induced Cycloisomerizations. Angew. Chem., Int. Ed. 2019, 58, 12924–12929. 10.1002/anie.201904365. [DOI] [PubMed] [Google Scholar]
- Fischer R. C.; Power P. P. π-Bonding and the Lone Pair Effect in Multiple Bonds Involving Heavier Main Group Elements: Developments in the New Millennium. Chem. Rev. 2010, 110, 3877–3923. 10.1021/cr100133q. [DOI] [PubMed] [Google Scholar]
- a Tokitoh N.; Arai Y.; Okazaki R.; Nagase S. Synthesis and Characterization of a Stable Dibismuthene: Evidence for a Bi-Bi Double Bond. Science 1997, 277, 78–80. 10.1126/science.277.5322.78. [DOI] [Google Scholar]; b Breunig H. J.; Rösler R.; Lork E. The First Organobismuth Rings: (RBi)3 and (RBi)4, R = (Me3Si)2CH. Angew. Chem., Int. Ed. 1998, 37, 3175–3177. . [DOI] [PubMed] [Google Scholar]; c Twamley B.; Sofield C. D.; Olmstead M. M.; Power P. P. Homologous Series of Heavier Element Dipnictenes 2,6-Ar2H3C6E = EC6H3–2,6-Ar2 (E = P, As, Sb, Bi; Ar = Mes = C6H2–2,4,6-Me3; or Trip = C6H2–2,4,6-iPr3) Stabilized by m-Terphenyl Ligands. J. Am. Chem. Soc. 1999, 121, 3357–3367. 10.1021/ja983999n. [DOI] [Google Scholar]
- a Šimon P.; de Proft F.; Jambor R.; Růžička A.; Dostál L. Monomeric Organoantimony(I) and Organobismuth(I) Compounds Stabilized by an NCN Chelating Ligand: Syntheses and Structures. Angew. Chem., Int. Ed. 2010, 49, 5468–5471. 10.1002/anie.201002209. [DOI] [PubMed] [Google Scholar]; b Vránová I.; Alonso M.; Lo R.; Sedlák R.; Jambor R.; Růžička A.; Proft F. D.; Hobza P.; Dostál L. From Dibismuthenes to Three- and Two-Coordinated Bismuthinidenes by Fine Ligand Tuning: Evidence for Aromatic BiC3N Rings through a Combined Experimental and Theoretical Study. Chem. - Eur. J. 2015, 21, 16917–16928. 10.1002/chem.201502724. [DOI] [PubMed] [Google Scholar]
- Wang F.; Planas O.; Cornella J. Bi(I)-Catalyzed Transfer-Hydrogenation with Ammonia-Borane. J. Am. Chem. Soc. 2019, 141, 4235–4240. 10.1021/jacs.9b00594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Tafesh A. M.; Weiguny J. A Review of the Selective Catalytic Reduction of Aromatic Nitro Compounds into Aromatic Amines, Isocyanates, Carbamates, and Ureas Using CO. Chem. Rev. 1996, 96, 2035–2052. 10.1021/cr950083f. [DOI] [PubMed] [Google Scholar]; b Kadam H. K.; Tilve S. G. Advancement in methodologies for reduction of nitroarenes. RSC Adv. 2015, 5, 83391–83407. 10.1039/C5RA10076C. [DOI] [Google Scholar]; c Orlandi M.; Brenna D.; Harms R.; Jost S.; Benaglia M. Recent Developments in the Reduction of Aromatic and Aliphatic Nitro Compounds to Amines. Org. Process Res. Dev. 2018, 22, 430–445. 10.1021/acs.oprd.6b00205. [DOI] [Google Scholar]
- Nykaza T. V.; Ramirez A.; Harrison T. S.; Luzung M. R.; Radosevich A. T. Biphilic Organophosphorus-Catalyzed Intramolecular Csp2-H Amination: Evidence for a Nitrenoid in Catalytic Cadogan Cyclizations. J. Am. Chem. Soc. 2018, 140, 3103–3113. 10.1021/jacs.7b13803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pang Y.; Leutzsch M.; Nöthling N.; Cornella J. Catalytic Activation of N2O at a Low-Valent Bismuth Redox Platform. J. Am. Chem. Soc. 2020, 142, 19473–19479. 10.1021/jacs.0c10092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Strîmb G.; Pöllnitz A.; Rat C. I.; Silvestru C. A general route to monoorganopnicogen(III) (M = Sb, Bi) compounds with a pincer (N,C,N) group and oxo ligands. Dalton Trans. 2015, 44, 9927–9942. 10.1039/C5DT00603A. [DOI] [PubMed] [Google Scholar]; b Breunig H. J.; Königsmann L.; Lork E.; Nema M.; Philipp N.; Silvestru C.; Soran A.; Varga R. A.; Wagner R. Hypervalent organobismuth(III) carbonate, chalcogenides and halides with the pendant arm ligands 2-(Me2NCH2)C6H4 and 2,6-(Me2NCH2)2C6H3. Dalton Trans. 2008, 1831–1842. 10.1039/b717127g. [DOI] [PubMed] [Google Scholar]; c Chovancová M.; Jambor R.; Ružička A.; Jirásko R.; Císařová I.; Dostál L. Synthesis, Structure, and Reactivity of Intramolecularly Coordinated Organoantimony and Organobismuth Sulfides. Organometallics 2009, 28, 1934–1941. 10.1021/om801194h. [DOI] [Google Scholar]
- a Twamley B.; Sofield C. D.; Olmstead M. M.; Power P. P. Homologous Series of Heavier Element Dipnictenes 2,6-Ar2H3C6E = EC6H3–2,6-Ar2 (E = P, As, Sb, Bi; Ar = Mes = C6H2–2,4,6-Me3; or Trip = C6H2–2,4,6-iPr3) Stabilized by m-Terphenyl Ligands. J. Am. Chem. Soc. 1999, 121, 3357–3367. 10.1021/ja983999n. [DOI] [Google Scholar]; b Hardman N. J.; Twamley B.; Power P. P. (2,6-Mes2H3C6)2BiH, a Stable, Molecular Hydride of a Main Group Element of the Sixth Period, and Its Conversion to the Dibismuthene (2,6-Mes2H3C6)BiBi(2,6-Mes2H3C6). Angew. Chem., Int. Ed. 2000, 39, 2771–2773. . [DOI] [PubMed] [Google Scholar]; c Breunig H. J.; Haddad N.; Lork E.; Mehring M.; Mügge C.; Nolde C.; Rat C. I.; Schürmann M. Novel Sterically Congested Monoorganobismuth(III) Compounds: Synthesis, Structure, and Bismuth-Arene π Interaction in ArBiXY (X, Y = Br, I, OH, 2,6-Mes2-4-t-Bu-C6H2PHO2). Organometallics 2009, 28, 1202–1211. 10.1021/om800934c. [DOI] [Google Scholar]
- For stoichiometric OAT from N2O to Si, see:; a Yao S.; Xiong Y.; Brym M.; Driess M. An Isolable Silanoic Ester by Oxygenation of a Stable Silylene. J. Am. Chem. Soc. 2007, 129, 7268–7269. 10.1021/ja072425s. [DOI] [PubMed] [Google Scholar]; b Xiong Y.; Yao S.; Driess M. An Isolable NHC-Supported Silanone. J. Am. Chem. Soc. 2009, 131, 7562–7563. 10.1021/ja9031049. [DOI] [PubMed] [Google Scholar]; c Yao S.; Xiong Y.; Driess M. N-Heterocyclic Carbene (NHC)-Stabilized Silanechalcogenones: NHC→Si(R2)=E (E = O, S, Se, Te). Chem. - Eur. J. 2010, 16, 1281–1288. 10.1002/chem.200902467. [DOI] [PubMed] [Google Scholar]; d Xiong Y.; Yao S.; Müller R.; Kaupp M.; Driess M. Activation of Ammonia by a Si=O Double Bond and Formation of a Unique Pair of Sila-Hemiaminal and Silanoic Amide Tautomers. J. Am. Chem. Soc. 2010, 132, 6912–6913. 10.1021/ja1031024. [DOI] [PubMed] [Google Scholar]; e Rosas-Sánchez A.; Alvarado-Beltran I.; Baceiredo A.; Saffon-Merceron N.; Massou S.; Hashizume D.; Branchadell V.; Kato T. Cyclic (Amino)(Phosphonium Bora-Ylide)Silanone: A Remarkable Room-Temperature-Persistent Silanone. Angew. Chem., Int. Ed. 2017, 56, 15916–15920. 10.1002/anie.201710358. [DOI] [PubMed] [Google Scholar]; f Wendel D.; Reiter D.; Porzelt A.; Altmann P. J.; Inoue S.; Rieger B. Silicon and Oxygen’s Bond of Affection: An Acyclic Three-Coordinate Silanone and Its Transformation to an Iminosiloxysilylene. J. Am. Chem. Soc. 2017, 139, 17193–17198. 10.1021/jacs.7b10634. [DOI] [PubMed] [Google Scholar]
- For stoichiometric OAT from N2O to Al, see:; a Hicks J.; Heilmann A.; Vasko P.; Goicoechea J. M.; Aldridge S. Trapping and Reactivity of a Molecular Aluminium Oxide Ion. Angew. Chem., Int. Ed. 2019, 58, 17265–17268. 10.1002/anie.201910509. [DOI] [PubMed] [Google Scholar]; b Anker M. D.; Coles M. P. Aluminium-Mediated Carbon Dioxide Reduction by an Isolated Monoalumoxane Anion. Angew. Chem., Int. Ed. 2019, 58, 18261–18265. 10.1002/anie.201911550. [DOI] [PubMed] [Google Scholar]
- For stoichiometric OAT from N2O to Ga, see:; Kassymbek A.; Vyboishchikov S. F.; Gabidullin B. M.; Spasyuk D.; Pilkington M.; Nikonov G. I. Sequential Oxidation and C-H Bond Activation at a Gallium(I) Center. Angew. Chem., Int. Ed. 2019, 58, 18102–18107. 10.1002/anie.201913028. [DOI] [PubMed] [Google Scholar]
- For catalytic reduction of N2O with HBpin (TON: 1400), see:Chen X.; Wang H.; Du S.; Driess M.; Mo Z.. Deoxygenation of Nitrous Oxide and Nitro Compounds Using Bis(N-Heterocyclic Silylene)Amido Iron Complexes as Catalysts. Angew. Chem., Int. Ed. 2021, 10.1002/anie.202114598. [DOI] [PubMed] [Google Scholar]
- For catalytic reduction of N2O with CO (TON: 561), see:; Zeng R.; Feller M.; Diskin-Posner Y.; Shimon L. J. W.; Ben-David Y.; Milstein D. CO Oxidation by N2O Homogeneously Catalyzed by Ruthenium Hydride Pincer Complexes Indicating a New Mechanism. J. Am. Chem. Soc. 2018, 140, 7061–7064. 10.1021/jacs.8b03927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- For catalytic reduction of N2O with hydrogen (TON: 307) and hydrosilanes (TON: 76), see:; Zeng R.; Feller M.; Ben-David Y.; Milstein D. Hydrogenation and Hydrosilylation of Nitrous Oxide Homogeneously Catalyzed by a Metal Complex. J. Am. Chem. Soc. 2017, 139, 5720–5723. 10.1021/jacs.7b02124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- For catalytic reduction of N2O with phosphine (TON: 17.2), see:; Gianetti T. L.; Rodríguez-Lugo R. E.; Harmer J. R.; Trincado M.; Vogt M.; Santiso-Quinones G.; Grützmacher H. Zero-Valent Amino-Olefin Cobalt Complexes as Catalysts for Oxygen Atom Transfer Reactions from Nitrous Oxide. Angew. Chem., Int. Ed. 2016, 55, 15323–15328. 10.1002/anie.201609173. [DOI] [PubMed] [Google Scholar]
- Pang Y.; Leutzsch M.; Nöthling N.; Katzenburg F.; Cornella J. Catalytic Hydrodefluorination via Oxidative Addition, Ligand Metathesis, and Reductive Elimination at Bi(I)/Bi(III) Centers. J. Am. Chem. Soc. 2021, 143, 12487–12493. 10.1021/jacs.1c06735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen Z.; He C.-Y.; Yin Z.; Chen L.; He Y.; Zhang X. Palladium-Catalyzed Ortho-Selective C-F Activation of Polyfluoroarenes with Triethylsilane: A Facile Access to Partially Fluorinated Aromatics. Angew. Chem., Int. Ed. 2013, 52, 5813–5817. 10.1002/anie.201300400. [DOI] [PubMed] [Google Scholar]
- a Lv H.; Zhan J.-H.; Cai Y.-B.; Yu Y.; Wang B.; Zhang J.-L. π-π Interaction Assisted Hydrodefluorination of Perfluoroarenes by Gold Hydride: A Case of Synergistic Effect on C-F Bond Activation. J. Am. Chem. Soc. 2012, 134, 16216–16227. 10.1021/ja305204y. [DOI] [PubMed] [Google Scholar]; b Lv H.; Cai Y.-B.; Zhang J.-L. Copper-Catalyzed Hydrodefluorination of Fluoroarenes by Copper Hydride Intermediates. Angew. Chem., Int. Ed. 2013, 52, 3203–3207. 10.1002/anie.201208364. [DOI] [PubMed] [Google Scholar]; c Nakai H.; Jeong K.; Matsumoto T.; Ogo S. Catalytic C-F Bond Hydrogenolysis of Fluoroaromatics by [(η5-C5Me5)RhI(2,2′-bipyridine)]. Organometallics 2014, 33, 4349–4352. 10.1021/om500647h. [DOI] [Google Scholar]
- Lim S.; Radosevich A. T. Round-Trip Oxidative Addition, Ligand Metathesis, and Reductive Elimination in a PIII/PV Synthetic Cycle. J. Am. Chem. Soc. 2020, 142, 16188–16193. 10.1021/jacs.0c07580. [DOI] [PubMed] [Google Scholar]
- Xiao W.-C.; Tao Y.-W.; Luo G.-G. Hydrogen formation using a synthetic heavier main-group bismuth-based electrocatalyst. Int. J. Hydrogen Energy 2020, 45, 8177–8185. 10.1016/j.ijhydene.2020.01.152. [DOI] [Google Scholar]
- Barton D. H. R.; Motherwell W. B.; Stobie A. A Catalytic Method for α-Glycol Cleavage. J. Chem. Soc., Chem. Commun. 1981, 1232–1233. 10.1039/C39810001232. [DOI] [Google Scholar]
- a Ooi T.; Goto R.; Maruoka K. Fluorotetraphenylbismuth: A New Reagent for Efficient Regioselective α-Phenylation of Carbonyl Compounds. J. Am. Chem. Soc. 2003, 125, 10494–10495. 10.1021/ja030150k. [DOI] [PubMed] [Google Scholar]; b Koech P. K.; Krische M. J. Phosphine Catalyzed α-Arylation of Enones and Enals Using Hypervalent Bismuth Reagents: Regiospecific Enolate Arylation via Nucleophilic Catalysis. J. Am. Chem. Soc. 2004, 126, 5350–5351. 10.1021/ja048987i. [DOI] [PubMed] [Google Scholar]; c Koech P. K.; Krische M. J. Enantioselective total and formal syntheses of paroxetine (PAXIL) via phosphine-catalyzed enone α-arylation using arylbismuth(V) reagents: a regiochemical complement to Heck arylation. Tetrahedron 2006, 62, 10594–10602. 10.1016/j.tet.2006.05.092. [DOI] [Google Scholar]
- Planas O.; Wang F.; Leutzsch M.; Cornella J. Fluorination of arylboronic esters enabled by bismuth redox catalysis. Science 2020, 367, 313–317. 10.1126/science.aaz2258. [DOI] [PubMed] [Google Scholar]
- a Clough J. M.; Diorazio L. J.; Widdowson D. A. A New Route to Fluoroaromatics from Boronic Acids and Caesium Fluoroxysulphate. Synlett 1990, 1990, 761–762. 10.1055/s-1990-21243. [DOI] [Google Scholar]; b Diorazio L. J.; Widdowson D. A.; Clough J. M. A new synthesis of aryl fluorides: the reaction of caesium fluoroxysulfate with arylboronic acids and derivatives. Tetrahedron 1992, 48, 8073–8088. 10.1016/S0040-4020(01)80478-6. [DOI] [Google Scholar]; c Cazorla C.; Métay E.; Andrioletti B.; Lemaire M. Metal-free electrophilic fluorination of alkyl trifluoroborates and boronic acids. Tetrahedron Lett. 2009, 50, 3936–3938. 10.1016/j.tetlet.2009.04.077. [DOI] [Google Scholar]
- For stoichiometric Cu, see:; a Fier P. S.; Luo J.; Hartwig J. F. Copper-Mediated Fluorination of Arylboronate Esters. Identification of a Copper(III) Fluoride Complex. J. Am. Chem. Soc. 2013, 135, 2552–2559. 10.1021/ja310909q. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Ye Y.; Sanford M. S. Mild Copper-Mediated Fluorination of Aryl Stannanes and Aryl Trifluoroborates. J. Am. Chem. Soc. 2013, 135, 4648–4651. 10.1021/ja400300g. [DOI] [PubMed] [Google Scholar]; c Tredwell M.; Preshlock S. M.; Taylor N. J.; Gruber S.; Huiban M.; Passchier J.; Mercier J.; Génicot C.; Gouverneur V. A General Copper-Mediated Nucleophilic 18F Fluorination of Arenes. Angew. Chem., Int. Ed. 2014, 53, 7751–7755. 10.1002/anie.201404436. [DOI] [PubMed] [Google Scholar]
- For stoichiometric Pd, see:; Furuya T.; Kaiser H. M.; Ritter T. Palladium-Mediated Fluorination of Arylboronic Acids. Angew. Chem., Int. Ed. 2008, 47, 5993–5996. 10.1002/anie.200802164. [DOI] [PubMed] [Google Scholar]
- For stoichiometric Ag, see:; Furuya T.; Ritter T. Fluorination of Boronic Acids Mediated by Silver(I) Triflate. Org. Lett. 2009, 11, 2860–2863. 10.1021/ol901113t. [DOI] [PubMed] [Google Scholar]
- For catalytic Pd, see:; Mazzotti A. R.; Campbell M. G.; Tang P.; Murphy J. M.; Ritter T. Palladium(III)-Catalyzed Fluorination of Arylboronic Acid Derivatives. J. Am. Chem. Soc. 2013, 135, 14012–14015. 10.1021/ja405919z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Planas O.; Peciukenas V.; Cornella J. Bismuth-Catalyzed Oxidative Coupling of Arylboronic Acids with Triflate and Nonaflate Salts. J. Am. Chem. Soc. 2020, 142, 11382–11387. 10.1021/jacs.0c05343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dhakal B.; Bohé L.; Crich D. Trifluoromethanesulfonate Anion as Nucleophile in Organic Chemistry. J. Org. Chem. 2017, 82, 9263–9269. 10.1021/acs.joc.7b01850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Camasso N. M.; Sanford M. S. Design, synthesis, and carbon-heteroatom coupling reactions of organometallic nickel(IV) complexes. Science 2015, 347, 1218–1220. 10.1126/science.aaa4526. [DOI] [PubMed] [Google Scholar]; b Canty A. J.; Ariafard A.; Camasso N. M.; Higgs A. T.; Yates B. F.; Sanford M. S. Computational study of C(sp3)-O bond formation at a PdIV centre. Dalton Trans. 2017, 46, 3742–3748. 10.1039/C7DT00096K. [DOI] [PubMed] [Google Scholar]; c Suero M. G.; Bayle E. D.; Collins B. S. L.; Gaunt M. J. Copper-Catalyzed Electrophilic Carbofunctionalization of Alkynes to Highly Functionalized Tetrasubstituted Alkenes. J. Am. Chem. Soc. 2013, 135, 5332–5335. 10.1021/ja401840j. [DOI] [PubMed] [Google Scholar]
- Magre M.; Cornella J. Redox-Neutral Organometallic Elementary Steps at Bismuth: Catalyt-ic Synthesis of Aryl Sulfonyl Fluorides. J. Am. Chem. Soc. 2021, 143, 21497–21502. 10.1021/jacs.1c11463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Smith B. C.; Waller C. B. The 1H Nuclear Magnetic Resonance Spectra of Some Organobismuth Compounds. J. Organomet. Chem. 1971, 32, C11–C12. 10.1016/S0022-328X(00)80147-3. [DOI] [Google Scholar]; b Marset X.; Torregrosa-Crespo J.; Martínez-Espinosa R. M.; Guillena G.; Ramón D. J. Multicomponent Synthesis of Sulfonamides from Triarylbismuthines, Nitro Compounds and Sodium Metabisulfite in Deep Eutectic Solvents. Green Chem. 2019, 21, 4127–4132. 10.1039/C9GC01541H. [DOI] [Google Scholar]; c Saavedra B.; Marset X.; Guillena G.; Ramón D. J. Multicomponent Synthesis of Sulfones and Sulfides from Tri-arylbismuthines and Sodium Metabisulfite in Deep Eutectic Solvents. Eur. J. Org. Chem. 2020, 3462–3467. 10.1002/ejoc.202000425. [DOI] [Google Scholar]
- a Sharutin V. V.; Ermoshkin A. E. Phenylation of Sulfur Dioxide by Pentaphenylbismuth. Izvsstiya Akademii Nauk SSSR, Seriya Khimicheskaya 1987, 11, 2598–2599. [Google Scholar]; b Zhao F.; Wu X.-F. Sulfonylation of Bismuth Reagents with Sulfinates or SO2 through BiIII/BiV Intermediates. Organometallics 2021, 40, 2400–2404. 10.1021/acs.organomet.1c00339. [DOI] [Google Scholar]; c Barton D. H. R.; Blazejewski J.-C.; Charpiot B.; Motherwell W. B. Tretraphenylbismuth Monotrifluoroacetate: a New Reagent for Regi-oselective Aryl Ether Formation. J. Chem. Soc., Chem. Commun. 1981, 503–504. 10.1039/c39810000503. [DOI] [Google Scholar]
- a Davies A. T.; Curto J. M.; Bagley S. W.; Willis M. C. One-pot palladium-catalyzed synthesis of sulfonyl fluorides from aryl bromides. Chem. Sci. 2017, 8, 1233–1237. 10.1039/C6SC03924C. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Lo P. K. T.; Chen Y.; Willis M. C. Nickel(II)-Catalyzed Synthesis of Sulfinates from Aryl and Heteroaryl Boronic Acids and the Sulfur Dioxide Surrogate DABSO. ACS Catal. 2019, 9, 10668–10673. 10.1021/acscatal.9b04363. [DOI] [Google Scholar]
- a Yin S.-F.; Maruyama J.; Yamashita T.; Shimada S. Efficient Fixation of Carbon Dioxide by Hypervalent Organobismuth Oxide, Hydroxide, and Alkoxide. Angew. Chem., Int. Ed. 2008, 47, 6590–6593. 10.1002/anie.200802277. [DOI] [PubMed] [Google Scholar]; b Kindra D. R.; Casely I. J.; Fieser M. E.; Ziller J. W.; Furche F.; Evans W. J. Insertion of CO2 and COS into Bi-C Bonds: Reactivity of a Bismuth NCN Pincer Complex of an Oxyaryl Dianionic Ligand, [2,6-(Me2NCH2)2C6H3]Bi(C6H2tBu2O). J. Am. Chem. Soc. 2013, 135, 7777–7787. 10.1021/ja403133f. [DOI] [PubMed] [Google Scholar]; c Kindra D. R.; Casely I. J.; Ziller J. W.; Evans W. J. Nitric Oxide Insertion Reactivity with the Bismuth-Carbon Bond: Formation of the Oximate Anion, [ON=(C6H2tBu2O)]−, from the Oxyaryl Dianion, (C6H2tBu2O)2–. Chem. - Eur. J. 2014, 20, 15242–15247. 10.1002/chem.201404910. [DOI] [PubMed] [Google Scholar]; d Lu W.; Hu H.; Li Y.; Ganguly R.; Kinjo R. Isolation of 1,2,4,3-Triazaborol-3-yl-metal (Li, Mg, Al, Au, Zn, Sb, Bi) Derivatives and Reactivity toward CO and Isonitriles. J. Am. Chem. Soc. 2016, 138, 6650–6661. 10.1021/jacs.6b03432. [DOI] [PubMed] [Google Scholar]; e Ramler J.; Poater J.; Hirsch F.; Ritschel B.; Fischer I.; Bickelhaupt F. M.; Lichtenberg C. Carbon monoxide insertion at a heavy p-block element: unprecedented formation of a cationic bismuth carbamoyl. Chem. Sci. 2019, 10, 4169–4176. 10.1039/C9SC00278B. [DOI] [PMC free article] [PubMed] [Google Scholar]; f Zhao F.; Wu X.-F. The first bismuth self-mediated oxidative carbonylative coupling reaction via BiIII/BiV redox intermediates. J. Catal. 2021, 397, 201–204. 10.1016/j.jcat.2021.03.029. [DOI] [Google Scholar]
- Oberdorf K.; Hanft A.; Ramler J.; Krummenacher I.; Bickelhaupt F. M.; Poater J.; Lichtenberg C. Bismuth Amides Mediate Facile and Highly Selective Pn–Pn Radical-Coupling Reactions (Pn = N, P, As). Angew. Chem., Int. Ed. 2021, 60, 6441–6445. 10.1002/anie.202015514. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Magre M.; Kuziola J.; Nöthling N.; Cornella J. Dibismuthanes in catalysis: from synthesis and characterization to redox behavior towards oxidative cleavage of 1,2-diols. Org. Biomol. Chem. 2021, 19, 4922–4929. 10.1039/D1OB00367D. [DOI] [PMC free article] [PubMed] [Google Scholar]