Next Article in Journal
Carotenoid Content and Profiles of Pumpkin Products and By-Products
Previous Article in Journal
The Electrochemical Behavior of Unmodified and Pd-NPs Modified AB5 Hydrogen Storage Alloy in Selected Protic and Aprotic Ionic Liquids (ILs): Towards ILs-Based Electrolytes for Ni-MH Batteries
Previous Article in Special Issue
Radical-Induced Cascade Annulation/Hydrocarbonylation for Construction of 2-Aryl-4H-chromen-4-ones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electroreductively Induced Radicals for Organic Synthesis

Key Laboratory of Pesticides & Chemical Biology Ministry of Education, College of Chemistry, Central China Normal University, 152 Luoyu Road, Wuhan 430079, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(2), 857; https://doi.org/10.3390/molecules28020857
Submission received: 29 November 2022 / Revised: 30 December 2022 / Accepted: 9 January 2023 / Published: 15 January 2023
(This article belongs to the Special Issue New Insights in Organic Radicals)

Abstract

:
Organic electrochemistry has attracted tremendous interest within the novel sustainable methodologies that have not only reduced the undesired byproducts, but also utilized cleaner and renewable energy sources. Particularly, oxidative electrochemistry has gained major attention. On the contrary, reductive electrolysis remains an underexplored research direction. In this context, we discuss advances in transition-metal-free cathodically generated radicals for selective organic transformations since 2016. We highlight the electroreductive reaction of alkyl radicals, aryl radicals, acyl radicals, silyl radicals, fluorosulfonyl radicals and trifluoromethoxyl radicals.

1. Introduction

Radicals have played vital roles in drug discovery, agrochemicals, material science, and fine chemical manufacturing [1,2,3,4,5,6]. The current intense developments in radical chemistry particularly rely on the utilization of photoredox catalysis for the generation of radical species [7,8,9]. However, photoredox catalysts can absorb only a minor part of the solar spectrum (300–450 nm) [10]. As a complementary alternative, organic electrosynthesis has recently witnessed a remarkable renaissance for green and sustainable chemistry [11,12,13,14,15,16,17]. Meanwhile, electrochemically driven radical transformations have been attracting increasing interest in recent years [18,19,20,21,22]. The emergence of the first electro-induced radical processes can be traced back to the Kolbe electrochemical decarboxylation, which delivered dimerization of alkyl radicals by anodic oxidation [23]. In the 1960s, the first development of a cathodically generated radicals based method for synthesis of adiponitrile has been identified as the creation of an economically attractive approach for a scalable commodity chemical [24]. Compared with extensive studies on oxidative electrochemistry [25,26,27,28,29], reductive radical chemistry remains underdeveloped [30,31,32], due to several intrinsic challenges [33]. Firstly, the radical precursors usually have high reduction potentials (absolute value). Therefore, careful reaction design is often necessary to avoid the competitive side-reactions (e.g., protonation or substrate decomposed). In addition, the counter-anodic oxidation for cathodic reduction commonly employs the oxidation of a sacrificial anode wherein the anode is corroded, which has resulted in high current densities and irrelevant potential control. Despite these challenges, considerable advances have been made in cathodic electrosynthesis [34,35]. This review aims to highlight the key features of recent developments (2016–2022) in transition-metal-free electroreductive synthesis by cathodically generated radicals.
Radicals can be generated by cathodic reduction from a variety of radical precursors, such as alkyl or aryl halides, carbonyl compounds, and alkenes, among others (Figure 1A). The resulted radicals can be divided into three main categories of chemical transformations: (A) addition to a π-system by delivery of a new C-centered radical, which could be further reduced to carbanion, which then reacts with an electrophile; (B) a second cathodic reduction offers carbanion and then reacts with an electrophile directly; and (C) radical-radical cross-coupling occurent in persistent radical systems [36]. The reactivity and selectivity of cathodically generated radicals depends on their reduction potentials and relative rates (Figure 1B) [37,38].

2. Electrochemical Reduction

2.1. Electroreductive Reaction of Alkyl Radicals

Alkyl radicals are important synthetic intermediates that have widely contributed to the development of synthetic radical chemistry in recent decades [39]. The approaches for generating alkyl radicals are the reduction of alkyl halides, epoxides, cyclopropanes and ketones as well as iminium salts. The generated alkyl radicals could then proceed via two processes, in which (i) the alkyl radicals are further reduced to alkyl anions (a radical-polar crossover pathway) and then are trapped with the electrophile (E+) to deliver the functionalized alkylation products; and (ii) the alkenes act as radical acceptors to afford the new C-centered radicals, which are further reduced at the cathode to generate the corresponding carbanions and then react with another electrophile (E+) to construct the design products (Figure 2).

2.1.1. Electroreductive Reaction of Alkyl Halides

Alkyl halides are the most simple and abundant reactants in organic synthesis and can be used as excellent alkyl radical precursors. They have frequently been engaged in novel organic reactions by their versatile reactivities and readily available properties. The high reduction potentials (absolute value) of alkyl halides (E = −1.5 to −2.5 V vs. SCE) are conceivably difficult to reduce to alkyl radicals [40].
Instead, in 2020, Lin group reported the electroreductive carbo-functionalization of alkenes with alkyl bromides via the radical-polarity crossover mechanism (Figure 3) [41]. Thus, value-added di-functionalization of alkenes was formed by the addition of two distinct electrophiles in a highly chemo- and regio-selective fashion. The new electroreductive strategies would be amenable to carboformylation, hydroalkylation, and carbocarboxylation of alkenes with a broad substrate scope and good functional group compatibility. The reaction started with the cathodic reduction of the alkyl bromide leading to an alkyl radical, which then was rapidly added to an alkene to give a new C-centered radical 6. In the next step, the nascent C-centered radical 6 will be readily reduced to the carbanion intermediate 7. Finally, the more inertly reductive electrophiles such as DMF (CHO donor), MeCN (H+ donor), and CO2 were suitable for the second C-C bond formation (Figure 3C).
Deuterated compounds have been widely applied in the realm of chemistry, mechanistic studies, and pharmaceutical science [42,43,44,45,46]. However, corresponding reports on unactivated alkyl halides have not been well elucidated [47,48]. Recently, the Qiu group showed electrochemically dehalogenative deuteration of un-activated alkyl halides (X = Cl, Br, I) or pseudo halides (X = OMs) using D2O as the economical deuterium source (Figure 4) [49]. This strategy is a good complement to the electrochemical deuteration of aryl halides, which has recently been successfully developed [50,51]. Here, a variety of natural products and pharmaceuticals were fully tolerated in the direct external catalysts-free reaction system. The proposed catalytic scenario consists of initial formation of the alkyl radical, which resulted from the cathodically reduced alkyl halides. Subsequent second electroreduction furnishes the alkyl carbanion intermediate 11 which then reacts with D2O and forms the final deuterated product. The N, N-diisopropylethylamine (DIPEA) combined with iodine anions (I) serves as sacrificial reductants oxidized to iminium ion or iodine radical, respectively (Figure 4).
Although a few examples have been known for electroreductive or electro-photocatalytic reductive borylation of aryl halides [52,53]. The un-activated alkyl halides, especially bromides or chlorides, have rarely been explored for the electrochemical borylation. In 2021, Lu and co-workers first reported electroreductive transition-metal-free borylation of un-activated alkyl halides (X = I, Br, Cl) (Figure 5) [54]. The reaction proved compatible with primary, secondary, and tertiary alkyl halides to provide the desired boronic esters in a highly efficient manner. The mild electrochemical reaction conditions offered high chemoselectivities and various synthetically meaningful natural products and drug derivatives (Figure 5B). Based on detailed experimental and computational mechanistic investigations, a plausible mechanism was proposed for the electroreductive borylation of alkyl halides. It was thus suggested to initiate by cathodic reduction of B2cat2 the generation of B2cat2 radical anion species 25. DFT studies showed that B2cat2 radical anion 25 both mediated the formation of the alkyl radical intermediate and trapped the alkyl radical, resulting in the formation of the final product 27 (Figure 5C).
The development of efficient C(sp3)-C(sp3) bonds construction has stimulated continued interest [55]. A wide variety of the couplings between an electrophilic organohalide and a nucleophilic organometallic agent, and a transition metal-catalyzed cross-coupling of two different carbon electrophiles have been reported [56,57,58]. These strategies, based on the prefunctionalized starting materials, underwent the undesired homocoupling side reactions. More recently, Lin described the electroreductive cross-electrophile coupling (XEC) [59,60] of the differential active alkyl halides driven by their disparate electronic and steric properties (Figure 6) [61]. The transition-metal-free methods provided high selectivity for the generation of a C-centered radical from the cathodic reduction of a more substituted alkyl halide, owing to its lower negative reduction potential (absolute value) compared to the less substituted alkyl halide (−2.13 V vs. −2.69 V, measured vs. Ag wire). Subsequent reduction of the resulting radical species 35 generated a carbanion 36. This anionic nucleophile 36 was then subjected to react with various less substituted alkyl halide, thus completing the C(sp3)-C(sp3) bonds construction. The second reduction of radical intermediate 35 to the carbanion was difficult due to the facile proto-debromination and elimination side steps. It was critical to introduce an anion-stabilizing substituent to a more substituted alkyl halide, in order to reduce the potential for the second reduction event. Various functionalities substituents, such as boryl, aryl, vinyl, alkynyl and silyl, were compatible with the electroreductive conditions, providing the functional complexity and synthetic value of the cross-coupling products.

2.1.2. Electroreductive Reaction of Epoxides and Cyclopropanes

In 2022, Lu and co-workers introduced an electroreductive methodology for regioselective hydrogenation of epoxides (Figure 7A) [62]. According to the density functional theory (DFT) calculations, anti-Markovnikov selective hydrogenation occurred preferentially with aryl-substituted epoxides, which originated from the thermodynamic stabilities of the in situ generated benzyl radicals 40a. On the other hand, alkyl-substituted epoxides favored Markovnikov selective hydrogenation, which can be attributed to the kinetic preference for the formation intermediate 40b. It was shown that the key to the success of this strategy was the implementation of utilizing TPPA ((pyrrolidino)phosphoramide), which had been applied as cosolvent and tripped the Mg anode as well as activated the epoxide. The proposed catalytic scenario consists of initially reductive epoxide to corresponding radical anion 39. Subsequent rapid irreversible C−O bond cleavage of 39 furnished the C-centered radical intermediate 40a or 40b, which can further undergo electro-proton transfer to achieve the final products.
Concurrently, Yu reported the first electrochemically reductive dicarboxylation of small rings with CO2 (Figure 7B) [63]. Various valuable dicarboxylic acids were achieved efficiently by electroreduction of cyclopropanes and cyclobutanes with CO2. Moreover, different from electro-oxidative ring-opening difunctionalization, which are limited to the anodic oxidative radical cations’ intermediates [64,65,66]. The author first realized an electro-reductive ring-opening strategy, one which involved the key radical anions and carbanions intermediates. Under electroreduction, corresponding radical anion 42 was preferentially generated from single electron transfer (SET) reduction of the substituted small rings on the cathode, and then underwent nucleophilic attack on CO2 to give the carboxylated carbon radical 43, which can be further SET reduced to yield C-centered anion 44. Therefore, attacking another CO2 would generate the diacid products.
Subsequently, the Qiu group and the Zhang group independently developed the electroreduction of aryl epoxides with CO2 to afford valuable hydroxy acids with high selectivity and efficiency (Figure 7C) [67,68]. Moreover, three, four, five and six-membered aryl cyclic esters were suitable substrates, providing the corresponding β-hydroxy acids, γ-hydroxy acids and δ-hydroxy acids as well as ε-hydroxy acids, respectively. The proposed mechanism for these electroreductive carboxylations is depicted in Figure 7C. The formation of epoxide radical anion 46 was involved in two possible pathways in the initial step, which either underwent direct reduction on the cathode or was reduced by the generated CO2 radical anion. Then, intermediate 46 underwent C-O bond cleavage to give the C-centered radicals 47. Subsequently, a further SET reduction of this C-centered radical 47 gave C-nucleophile species, and then the following attack of CO2 resulted in the formation of intermediate 48. Finally work-up with aqueous HCl yielded the hydroxy acids product.

2.1.3. Electroreductive Reaction of Ketones, Alkenes and Alkynes

As exemplified by the prior case study, the alkenes as the radical acceptors to be added with the cathodically generated radicals could be both a strategic and tactical assent for organic synthesis. Recent efforts by Baran and co-workers demonstrated that ketones and simple unactivated olefins were also viable substrates for access tertiary alcohols (Figure 8) [69]. The authors found that the sacrificial anode (Zn), electrolyte (nBu4NBr) and current density all played critical roles to facilitate a high-yielding olefin-ketone coupling. The ketone 59 was initially reduced on the Sn cathode, whereupon reaction of the resulting ketyl radical anion 60 with alkenes gave rise to a C-centered radical 61. The radical 61 can be reduced into the corresponding C-centered nucleophilic species 62 which can get a proton from α-position of ketone. Finally, the desired tertiary alcohols were achieved by workup (Figure 8C).
The hydrogenations of C-C multiple bonds as well as C-O double bonds are fundamental in organic synthesis. Traditionally, substrates are reduced with risky reagents (H2) [70,71], silicon hydrides [72,73], or hydrogen donors (formats, alcohols) [74,75,76]. However, safety hazards and undesired byproducts have compromised the principles of green chemistry. In 2019, Li and Cheng disclosed the electrochemical hydrogenation of alkenes, alkynes and ketones with gaseous ammonia as the proton source (Figure 9A) [77]. The use of ammonia was the key to ensure the lack of need of a sacrificial anode. The mild reaction conditions featured a broad tolerance of valuable functional groups, such as esters, amides and nitrile groups (Figure 9B). The selective hydrogenation of ketones, either to the corresponding alcohols or the diphenyl methane, relied on the applied voltage. The same group further expanded the scope of reductive electrosynthesis to the deuteration of α,β-unsaturated carbonyl compounds (Figure 9C) [78]. In this reaction, D2O severed as deuteration source as well as a sacrificial reductant avoiding cathodic corrosion. In addition, the detailed cyclic voltammetry investigations highlighted an initial reduction of the substrate 77 at −2.1 V vs. SCE, while the reduction potential of D2O (−2.5 V vs. SCE) was found to be higher (absolute value). These findings indicated that the substrate 77 had been reduced preferentially. The analogous electroreductive mechanism was depicted in Figure 9E. First, the alkene substrates 77 processed through SET cathodic reduction to form the anionic radical species 78, which then coupled with a proton (deuton) transferring from ammonia or D2O to yield a radical 79 intermediate. The second cathodic reduction of resulting radical 79 generated the C-centered anion 80. Finally, proton (deuton) transfer delivered the desired products 81a or 81b, respectively. Concurrently, N2 was generated from ammonia at the anode to fulfil the hydrogenation task in the first example. Notably, the 18O2 was captured with PPh3 to deliver the 18O-labelled triphenylphosphine oxide, which suggested the oxidation of D2O at the anode in the electrochemical deuteration reaction.
In 2019, another hydrogenation work was reported by the group of Xia using NH4Cl and methanol as hydrogen donors (Figure 10) [79]. Compared with previous hydrogenation of C=O double bonds in ketones [77], Xia reported the different selectivity hydrogenation of C=C double bonds in α,β-unsaturated ketones. The transformation proceeded through the aforementioned hydrogenation mechanism. Importantly, the authors reported that the solvent DMSO as the sacrificial reductant oxidized to methyl sulfone.

2.1.4. Electroreductive Reaction of Cyanoheteroarenes

Reductions of cyanoheteroarenes have been well known to generate persistent radicals via photoinduced single-electron transformations [80,81,82]. A radical-radical coupling approach was proposed between this persistent and a transient radical. In 2020, Rovis and co-workers developed an electroreductive method for synthesis of the hindered primary amines (Figure 11A) [83]. Benchtop-stable iminium salts and cyanoheteroarenes were used as the radical precursors under mild reductive conditions. A proton-coupled electron-transfer (PCET) mechanism for the generation of radical 103 was supported by cyclic voltammetry studies and DFT calculations, as well as NMR spectroscopy. By the comparison of reduction potential of iminium salt 100 (−0.76 V vs. SCE) with cyanoheteroarene 102 (−1.3 V vs. SCE), the iminium salt 100 was preferentially reduced, furnishing α-amino radical 101. The formation of 103 radical via the PCET pathway could undergo radical-radical coupling to forge the intermediate 104. Finally, loss of H+ and CN would deliver the design hindered amine 105. Building on this precedent, Xia group reported that the ketyl radicals, generated through reduction of aldehydes and ketones, can undergo facile radical-radical cross-coupling with the heteroaryl radical anion 106 providing secondary and tertiary alcohols (Figure 11C) [84]. The generated crucial aliphatic ketyl radicals were proved by electron paramagnetic resonance (EPR) studies. Concurrently, Jiang disclosed a mechanistically related transformation using thioester-activated alkenes as the radical precursors (Figure 12) [85]. This method was successfully applied to the synthesis of β-pyridine thioester. It should be mentioned that very recently Cheng group reported an electroreductively enantioselective Pd-catalyzed allylic 4-pyridinylation [86]. This method successfully employed the chiral η3 allyl Pd complex to capture the persistent radical via radical rebound.

2.2. Electroreductive Reaction of Aryl Radicals

Aryl halides served as essential reagents in organic chemistry and were common aryl radical precursors because of the inherent reactivity of aryl halogen (Ar-X) bonds. However, the high reduction potentials (absolute value) of aryl halides (E = −1.88 to −3.2 V vs. SCE), particularly the uncontrolled side reactions, compromised chemoselectivity, which remained a challenging area in electrochemistry. The general mechanism for the transformations of aryl radicals can proceed via three processes: (1) cathodic reduction to deliver the radicals anion, which falls in two categories: direct cathodic reduction or indirect cathodic reduction; (2) halogen heterolysis cleavage to form aryl radicals, and (3) the generated aryl radicals then proceed to the diversity of functionalized products (Figure 13).

2.2.1. Direct Electroreductive Reaction of Aryl Radicals

The direct electroreductive system enables substrates to undergo electron transfer directly at the cathode surfaces. The direct approach results in the accumulation of highly reactive radical and radical anion intermediates in the electric double layer. Thus, these high-energy species can lead to further undesirable side reactions. Chi and co-workers reported a direct reductive progress to access hydro-dehalogenation of organic halides, which was significant for detoxification of environmentally hazardous organic halides (Figure 14A) [87]. The use of trialkylamines as suitable reductants and hydrogen atom donors to provide an electrochemical reductive system for replacing halogen atoms with hydrogen atoms. Later, the Zhang group proposed a procedure with copper nanowire arrays (Cu NWAs) as the recycled cathode to deliver the deutero-dehalogenation products (Figure 14B) [88]. The EPR measurements rationalized the existent aryl and hydrogen radicals. Moreover, the methods could be successfully applied to the paired electrolysis involving two desirable half-reactions. In 2021, Xia group disclosed the electroreductive methodology which was successfully extended to reactions other than hydro-dehalogenation, such as the reduction of Ts-protected amines and aromatic cyanides [89].
Due to the unique property of deuterated compounds, the development of user-friendly deuteron-dehalogenation has been reported by Lei and co-workers (Figure 14C) [51]. The commercially available electrodes, common battery and more user-friendly undivided cell highlight the potential practicability of this deuteration method. Initiation with cathodic reduction of aryl halides afforded the aryl radical (•Ar) 117 and bromine anion (Br). D2O was reduced simultaneously to deliver the deuterium radical (•D), which can be trapped by the aryl radical to form the desired product 118 via radical-radical cross-coupling. The oxidation of bromine anion (Br) at the anode followed by reaction with NPh3 delivered N(Ph)2PhBr.
The methodology for generating aryl radicals was not limited to the electroreduction of aryl halides, and electron-deficient arenes could be reduced at the cathode to generate the corresponding radical anions. Lei group developed the reductive radical cross-coupling of electro-deficient arenes and aryldiazonium (Figure 15) [90]. Various electro-deficient arenes and aryl diazonium tetrafluoroborate were employed to generate the corresponding arylation products. Moreover, anilines can be applied to synthesis diazonium reagents in situ, which can be further applied to the electrochemical arylation reaction. The highly effective nature of this reductive transformation is owed to the relatively low reduction potentials (absolute value) of quinoxaline (E = −1.06 V vs. Ag/Ag+) and aryl diazonium salt (E = −0.62 V vs. Ag/Ag+). It was also found that the EPR signals of quinoxaline radical anion and aryl radical were observed.

2.2.2. Indirect Electroreductive Reaction of Aryl Radicals

In contrast, with indirect electroreductive reaction, a redox mediator, more easily reduced than the substrates, acts as the electron transfer shuttle from the heterogeneous electrode surface to the homogeneous dissolved substrates [91,92]. The indirect approach offers several advantages. In many cases, the introduction of a mediator results in improved reaction efficacy and better chemoselectivity by decreasing the overpotential required for substrate reduction. In 2016, Wan and coworkers showed that indirect cathodic reduction of arylhalides and pyrroles produced arylation of pyrroles using perylene disimide (PDI) as a mediator (Figure 16A) [93]. Mechanistically, aryl radical anion (Ar•−) 125 was produced through electro gain at the reduced PDI radical anion (PDI•−) species. The resulting (Ar•−) 125 underwent cleavage of C-X bond to deliver an aryl radical (•Ar) 126. The pyrrole partner then reacted with aryl radical 126 to yield the arylation of pyrroles products. It has to be considered, though, that the radical anion (PDI•−) could further reduced to dianion (PDI2−). The radical anion (PDI•−) is proposed to be the more active catalyst species.
The electroreductive borylation of aryl iodides via aryl radical pathway was reported by the Mo group [54], which used pinacol diboron (B2pin2) as the borylating agent without the transition metal catalysts, reductants or sacrificial anodes (Figure 16B). The authors explored a series of mechanistic studies, such as EPR experiments and the cyclic voltammetry studies (CV), to reveal the electroreductive generation of aryl radicals.
In a subsequent report, Lin and Lambert jointly demonstrated electro-photocatalytic reductive transformation of aryl halides by using a dicyanoanthracene (DCA) as the electro-photocatalyst (Figure 17A) [53]. The transformation provided access to arylboronate, arylstannane and biaryl products by employing different radical trapping agents. The electro-photocatalysis relied on the electrochemical reduction of DCA to the corresponding radical anion (DCA•−) and followed by visible light photoexcitation to generate the stronger reductant [DCA•−]* radical anion (E = −3.2 V vs. SCE). The highly reducing photoexcited radical anion [DCA•−]* enabled facile reduction of aryl chlorides, which displayed high reduction potentials (absolute value) (E = −2.94 V vs. SCE) and strong bond dissociation energies (>97 kal/mol). Concurrently, a similar strategy was published by Wickens and co-workers who employed a naphthalene-based monoimide (NpMI) as electro-photocatalyst (Figure 17B) [94]. The resulting aryl radicals engaged with various radical traps to furnish phosphorylation and heteroarylation of aryl chloride. In this method, photoexcited reductant [NpMI•−]* has capabilities of reduction potential beyond those of Na metal.
The versatile electroreductive generation of aryl radicals were further applied to the intramolecular cyclisation of aryl halides. Recently, Brown group has showcased how the organic mediator-catalyzed electroreductive cyclisation reaction enabled the synthetically valuable five, six-membered cyclic or tricyclic ethers in flow (Figure 18) [95]. The flow reductive electrolysis was performed in efficient mass transport and narrow electrode area with the absence of a sacrificial anode. Sensitive functional groups cyano and 1-(allyloxy)-2-heterocyclic halides were thereby well tolerated. It was suggested to initiate by cathodic reduction of the phenanthrene mediator in order to deliver the strongly reducing phenanthrene radical anion. Then, electroreduction of aryl halides with the phenanthrene radical anion resulted in the formation of 148 intermediate, which undergoes the cleavage of halogen anion (X) to form aryl radical intermediate 149. The C-centered radical was trapped intramolecularly by a terminal alkene to form the five- or six-membered C-centered radical 150. The latter C-centered was reduced by the second phenanthrene radical anion and protonation to give the final product 151. The author unraveled the formation of hydrogenolysis byproduct by cathodic reduction 149 and then protonation in the absence of the phenanthrene mediator.
Very recently, Qiu group reported a mechanistically related carboxylation with CO2 as the C1 feedstock (Figure 19) [96]. Introducing the naphthalene as an efficient organic mediator led to unprecedented carboxylation of organic halides. A set of substituted aryl halides and heteroaryl halides as well as structurally complex drugs could be smoothly converted to the desired carboxylation products (Figure 19B). Based on detailed cyclic voltammetry (CV) studies, Qiu proposed a plausible reaction pathway involving the reductive mediator strategy. The relatively lower reduction potential (absolute value) of naphthalene (E = −3.09 V vs. Ag/Ag+) compared with aryl bromide (E = −3.53 V vs. Ag/Ag+) made it an available mediator to catalyze a wide array of organic halides. The cathodic reduction of naphthalene to form a naphthalene radical anion followed reduction with aryl halides forming intermediate 159. The radical anion species 159 was dissociated to the active aryl radical (•Ar) 160, which underwent cathodic reduction to give aryl anion (Ar) 161. Finally, nucleophilic attacking to CO2 afforded the design carboxylation product 162 (Figure 19C).

2.3. Electroreductive Reaction of Acyl Radicals

The electroreductive reactions were further extended to radical-radical cross-coupl- ing with benzoyl chloride and benzenesulfinic acid. Lei and co-workers recently demonstrated the synthesis of thioester derivatives via cathodic reductive acyl radicals and thiyl radicals in a user-friendly undivided cell (Figure 20) [97]. A wide range of substrates including heteroaromatic acyl chlorides, such as 2-furoyl chloride, 2-thiophenecarbonyl chloride and 2-chloronicotinyl chloride, were well tolerated. It is worthy of note that various alkyl acyl chlorides and alkyl sulfinic acid showed good compatibility in this electroreductive cross-coupling reaction (Figure 20B). Based on detailed mechanistic studies and the previous photochemical reductive reports [98], a plausible mechanism was proposed (Figure 20C). The reaction started with the cathodic reduction of Tf2O to generate [Tf2O]•− radical anion and showed Tf2O mediate the formation of the acyl radical 172. Therefore, the acyl radical 172 with benzenesulfinic acid furnished intermediate 173, which was followed by cathodic reduction to an O-centered radical 174. Trapping of 174 with acyl radical 172 resulted in the formation of the intermediate 175, which underwent a second cathodic reduction to generate the thiyl radical 176. Finally, radical-radical cross-coupling between acyl radical 172 and thiyl radical 176 led to the formation of the final thioester 177.

2.4. Electroreductive Reaction of Silyl Radicals

In 2020, the group of Lin et al. reported an effective method to construct C-Si bonds for the synthetically valuable vicinal disilane by electroreduction of alkenes with commercially available chlorosilanes as the Si source (Figure 21) [99]. The protocol allowed for the construction of vicinal C-Si bonds of styrene, heterocycles and enynes. Moreover, a variety of chlorosilanes were efficiently transformed into the corresponding organosilanes (Figure 21B). The reaction started with the cathodic reduction of chlorosilanes to generate silyl radicals (•SiR3) in the presence of the anodic generation of magnesium salts (Mg2+). The producing •SiR3 sequential addition to the alkene provided the typically short-lived C-centered radical 187, which can be further cathodically reduced to give 188 carbanion. Another electrophilic species, chlorosilanes, can readily undergo substitution to form designed disilylation products 189. Notably, more challenging silacycles also proved viable substrates to deliver the corresponding five- and six-membered silacycles 190. Additionally, a wide range of hydrosilylation products 191 were achieved by employing weakly acidic acetonitrile as the solvent.

2.5. Electroreductive Reaction of Fluorosulfonyl Radical

The photocatalyzed fluorosulfonylation has been achieved through sulfuryl chlorofluoride (FSO2Cl) as an effective fluorosulfonyl radical precursor [100,101]. Based on those results, the same research group also discovered an electrochemical synthesis of β-keto sulfonyl fluorides through reductive FSO2Cl with alkynes (Figure 22) [102]. The unstable FSO2 radical (•FSO2) could be generated via cathodic reduction of FSO2Cl. The alkene radical 196 underwent oxidation by O2 in the air to give the peroxy radical 197, which then provided radical intermediate 198 following the Russell mechanism. Therefore, the resulting intermediate 198 can be the secondary cathodically reduced to deliver β-keto sulfonyl fluorides. Interestingly, α-chloro β-keto sulfonyl fluoride also can be achieved under THF as the solvent. The same synthetic strategy can also be applied to the fluorosulfonylation of vinyl triflates [103]. Note that this protocol avoided the use of a sacrificial anode and metal catalysis.

2.6. Electroreductive Reaction of OCF3 Radicals (•OCF3)

The tremendous progress have been made in electrochemical oxidative trifluoromethylation (CF3) of alkenes by CF3SO2Na as radical precursor [104]. Unfortunately, the electrochemical trifluoromethoxylation (CF3O) of (hetero)aromatics are inherently inefficient. Recently, the Qing group has gained significant momentum in this area by involving trifluoromethyl 2-pyridyl sulfone and oxygen as a trifluoromethoxylation source (Figure 23) [105]. As shown in Figure 23B, the mild electrochemical reaction conditions proved compatible with various synthetically meaningful functional groups, such as chloro, sulfonate, cyano on the (hetero)aromatics as well as bio-relevant molecules. The mechanism involved the formation of the •CF3, which was reduced from 2-pyridyl sulfone (206) with assistance of the hydrogen bond of HFIP solvent at the cathode. Subsequently, •CF3 combined with O2 to afford •OOCF3, which was secondarily reduced for the formation of the key •OCF3 intermediate. Next, the •OCF3 attacked the (hetero)aromatics to generate a transient C-centered radical species 207, which was then oxidized at the anode and followed by rearomatization to provide the desired trifluoromethoxylated product 209. Here, the additive Et3N served as the sacrificial reductant oxidized to the corresponding iminium cation.

3. Conclusions and Outlook

Electrochemical-reduction-induced radicals have been useful intermediates in organic chemistry. This Review has cited a large number of useful organic transformations, including alkenes functionalization, borylation, hydrogenation, silylation, arylation, dehalogenation and so on. Despite these advances, there are still many opportunities for the development of new and robust electrochemically reductive methods. The electroreductive reaction involves only one cathodic half-reaction, with the anode oxidation usually designed to be sacrificed. In order to achieve more efficient and selective molecular syntheses, electrochemical processes can be designed as a combination of two simultaneous desirable half-reactions, oxidation and reduction. Therefore, paired electrolysis can be considered as the gold standard for two desirable progressions on both electrodes [106,107]. On the other hand, there should be more advances made in asymmetric electrochemical radical reductive transformation [86,108]. We hope that this Review will inspire further developments in this exciting area of electrochemistry.

Author Contributions

Conceptualization, C.Z. and H.X. (Huaming Xiang); validation, C.Z., H.X. (Hao Xu) and M.Q.; writing—original draft preparation, C.Z.; writing—review and editing, C.Z., H.X. (Huaming Xiang), J.H., W.Q., H.X. (Hao Xu), W.D., Y.O., Y.W.; visualization, C.Z.; supervision, C.Z.; project administration, C.Z.; funding acquisition, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China and grant number was 22201089.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are available in the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jasperse, C.P.; Curran, D.P.; Fevig, T.L. Radical Reactions in Natural Product Synthesis. Chem. Rev. 1991, 91, 1237–1286. [Google Scholar] [CrossRef]
  2. Curran, D.P. The Design and Application of Free Radical Chain Reactions in Organic Synthesis. Part 1. Synthesis 1988, 417–439. [Google Scholar] [CrossRef]
  3. Curran, D.P. The Design and Application of Free Radical Chain Reactions in Organic Synthesis. Part 2. Synthesis 1988, 489–513. [Google Scholar] [CrossRef]
  4. Yan, M.; Lo, J.C.; Edwards, J.T.; Baran, P.S. Radicals: Reactive Intermediates with Translational Potential. J. Am. Chem. Soc. 2016, 138, 12692–12714. [Google Scholar] [CrossRef] [PubMed]
  5. Smith, J.M.; Harwood, S.J.; Baran, P.S. Radical Retrosynthesis. Acc. Chem. Res. 2018, 51, 1807–1817. [Google Scholar] [CrossRef] [PubMed]
  6. Parsaee, F.; Senarathna, M.C.; Kannangara, P.B.; Alexander, S.N.; Arche, P.D.E.; Welin, E.R. Radical Philicity and Its Role in Selective Organic Transformations. Nat. Rev. Chem. 2021, 5, 486–499. [Google Scholar] [CrossRef]
  7. Goddard, J.-P.; Ollivier, C.; Fensterbank, L. Photoredox Catalysis for the Generation of Carbon Centered Radicals. Acc. Chem. Res. 2016, 49, 1924–1936. [Google Scholar] [CrossRef]
  8. Chan, A.Y.; Perry, I.B.; Bissonnette, N.B.; Buksh, B.F.; Edwards, G.A.; Frye, L.I.; Garry, O.L.; Lavagnino, M.N.; Li, B.X.; Liang, Y.; et al. Metallaphotoredox: The Merger of Photoredox and Transition Metal Catalysis. Chem. Rev. 2022, 122, 1485–1542. [Google Scholar] [CrossRef] [PubMed]
  9. Yu, X.-Y.; Chen, J.-R.; Xiao, W.-J. Visible Light-Driven Radical-Mediated C–C Bond Cleavage/Functionalization in Organic Synthesis. Chem. Rev. 2021, 121, 506–561. [Google Scholar] [CrossRef]
  10. Schultz, D.M.; Yoon, T.P. Solar Synthesis: Prospects in Visible Light Photocatalysis. Science 2014, 343, 1239176. [Google Scholar] [CrossRef]
  11. Gandeepan, P.; Finger, L.H.; Meyer, T.H.; Ackermann, L. 3d Metallaelectrocatalysis for Resource Economical Syntheses. Chem. Soc. Rev. 2020, 49, 4254–4272. [Google Scholar] [CrossRef] [PubMed]
  12. Meyer, T.H.; Choi, I.; Tian, C.; Ackermann, L. Powering the Future: How Can Electrochemistry Make a Difference in Organic Synthesis? Chem 2020, 6, 2484–2496. [Google Scholar] [CrossRef]
  13. Zhu, C.; Ang, N.W.J.; Meyer, T.H.; Qiu, Y.; Ackermann, L. Organic Electrochemistry: Molecular Syntheses with Potential. ACS Cent. Sci. 2021, 7, 415–431. [Google Scholar] [CrossRef] [PubMed]
  14. Wiebe, A.; Gieshoff, T.; Möhle, S.; Rodrigo, E.; Zirbes, M.; Waldvogel, S.R. Electrifying Organic Synthesis. Angew. Chem. Int. Ed. 2018, 57, 5594–5619. [Google Scholar] [CrossRef]
  15. Siu, J.C.; Fu, N.; Lin, S. Catalyzing Electrosynthesis: A Homogeneous Electrocatalytic Approach to Reaction Discovery. Acc. Chem. Res. 2020, 53, 547–560. [Google Scholar] [CrossRef]
  16. Liu, J.; Lu, L.; Wood, D.; Lin, S. New Redox Strategies in Organic Synthesis by Means of Electrochemistry and Photochemistry. ACS Cent. Sci. 2020, 6, 1317–1340. [Google Scholar] [CrossRef] [PubMed]
  17. He, J.-Y.; Qian, W.-F.; Wang, Y.-Z.; Yao, C.; Wang, N.; Liu, H.; Zhong, B.; Zhu, C.; Xu, H. Sustainable Electrochemical Dehydrogenative C(sp3)–H Mono/di-alkylations. Green Chem. 2022, 24, 2483–2491. [Google Scholar] [CrossRef]
  18. Pham, P.H.; Petersen, H.A.; Katsirubas, J.L.; Luca, O.R. Recent Synthetic Methods Involving Carbon Radicals Generated by Electrochemical Catalysis. Org. Biomol. Chem. 2022, 20, 5907–5932. [Google Scholar] [CrossRef]
  19. Chen, N.; Xu, H.-C. Electrochemical Generation of Nitrogen-Centered Radicals for Organic Synthesis. Green Synth. Catal. 2021, 2, 165–178. [Google Scholar] [CrossRef]
  20. Yuan, Y.; Yang, J.; Lei, A. Recent Advances in Electrochemical Oxidative Cross-Coupling with Hydrogen Evolution involving Radicals. Chem. Soc. Rev. 2021, 50, 10058–10086. [Google Scholar] [CrossRef]
  21. Sauer, G.S.; Lin, S. An Electrocatalytic Approach to the Radical Difunctionalization of Alkenes. ACS Catal. 2018, 8, 5175–5187. [Google Scholar] [CrossRef]
  22. Xiong, P.; Xu, H.-C. Chemistry with Electrochemically Generated N-Centered Radicals. Acc. Chem. Res. 2019, 52, 3339–3350. [Google Scholar] [CrossRef] [PubMed]
  23. Kolbe, H. Zersetzung der Valeriansäure durch den elektrischen Strom. Justus Liebigs Ann. Chem. 1848, 64, 339–341. [Google Scholar]
  24. Baizer, M.M. Electrolytic Reductive Coupling: I. Acrylonitrile. J. Electrochem. 1964, 111, 215. [Google Scholar] [CrossRef]
  25. Ackermann, L. Metalla-Electrocatalyzed C–H Activation by Earth-Abundant 3d Metals and Beyond. Acc. Chem. Res. 2020, 53, 84–104. [Google Scholar] [CrossRef]
  26. Tang, S.; Liu, Y.; Lei, A. Electrochemical Oxidative Cross-Coupling with Hydrogen Evolution: A Green and Sustainable Way for Bond Formation. Chem 2018, 4, 27–45. [Google Scholar] [CrossRef] [Green Version]
  27. Kingston, C.; Palkowitz, M.D.; Takahira, Y.; Vantourout, J.C.; Peters, B.K.; Kawamata, Y.; Baran, P.S. A Survival Guide for the “Electro-curious”. Acc. Chem. Res. 2020, 53, 72–83. [Google Scholar] [CrossRef]
  28. Rosen, B.R.; Werner, E.W.; O’Brien, A.G.; Baran, P.S. Total Synthesis of Dixiamycin B by Electrochemical Oxidation. J. Am. Chem. Soc. 2014, 136, 5571–5574. [Google Scholar] [CrossRef]
  29. Shi, S.-H.; Liang, Y.; Jiao, N. Electrochemical Oxidation Induced Selective C–C Bond Cleavage. Chem. Rev. 2021, 121, 485–505. [Google Scholar] [CrossRef]
  30. Park, S.H.; Ju, M.; Ressler, A.J.; Shim, J.; Kim, H.; Lin, S. Reductive Electrosynthesis: A New Dawn. Aldrichimica Acta 2021, 54, 17–27. [Google Scholar]
  31. Yan, M.; Kawamata, Y.; Baran, P.S. Synthetic Organic Electrochemical Methods Since 2000: On the Verge of a Renaissance. Chem. Rev. 2017, 117, 13230–13319. [Google Scholar] [CrossRef] [PubMed]
  32. Sperry, J.B.; Wright, D.L. The Application of Cathodic Reductions and Anodic Oxidations in the Synthesis of Complex Molecules. Chem. Soc. Rev. 2006, 35, 605–621. [Google Scholar] [CrossRef]
  33. Wirtanen, T.; Prenzel, T.; Tessonnier, J.-P.; Waldvogel, S.R. Cathodic Corrosion of Metal Electrodes—How to Prevent It in Electroorganic Synthesis. Chem. Rev. 2021, 121, 10241–10270. [Google Scholar] [CrossRef] [PubMed]
  34. Huang, B.; Sun, Z.; Sun, G. Recent Progress in Cathodic Reduction-Enabled Organic Electrosynthesis: Trends, Challenges, and Opportunities. eScience 2022, 2, 243–277. [Google Scholar] [CrossRef]
  35. Zhang, S.-K.; Samanta, R.C.; Del Vecchio, A.; Ackermann, L. Evolution of High-Valent Nickela-Electrocatalyzed C−H Activation: From Cross(-Electrophile)-Couplings to Electrooxidative C−H Transformations. Chem. Eur. J. 2020, 26, 10936–10947. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, S.; Tang, S.; Lei, A. Tuning Radical Reactivity for Selective Radical/radical Cross-Coupling. Sci. Bull. 2018, 63, 1006–1009. [Google Scholar] [CrossRef] [Green Version]
  37. Andrieux, C.P.; Gallardo, I.; Saveant, J.M. Outer-Sphere Electron-Transfer Reduction of Alkyl Halides. A Source of Alkyl Radicals or of Carbanions? Reduction of Alkyl Radicals. J. Am. Chem. Soc. 1989, 111, 1620–1626. [Google Scholar] [CrossRef]
  38. Benderskii, V.A.; Krivenko, A.G. The Electrochemistry of Short-Lived Intermediate Species. Russ. Chem. Rev. 1990, 59, 1–22. [Google Scholar] [CrossRef]
  39. Schäfer, H.J. Electrochemical Generation of Radicals. In Radicals in Organic Synthesis; Renaud, P., Sibi, M.P., Eds.; Wiley: New York, NY, USA, 2001; pp. 250–297. [Google Scholar]
  40. Roth, H.G.; Romero, N.A.; Nicewicz, D.A. Experimental and Calculated Electrochemical Potentials of Common Organic Molecules for Applications to Single-Electron Redox Chemistry. Synlett 2016, 27, 714–723. [Google Scholar]
  41. Zhang, W.; Lin, S. Electroreductive Carbofunctionalization of Alkenes with Alkyl Bromides via a Radical-Polar Crossover Mechanism. J. Am. Chem. Soc. 2020, 142, 20661–20670. [Google Scholar] [CrossRef]
  42. Wiberg, K.B. The Deuterium Isotope Effect. Chem. Rev. 1955, 55, 713–743. [Google Scholar] [CrossRef]
  43. Kushner, D.J.; Baker, A.; Dunstall, T.G. Pharmacological Uses and Perspectives of Heavy Water and Deuterated Compounds. Can. J. Physiol. Pharmacol. 1999, 77, 79–88. [Google Scholar] [CrossRef] [PubMed]
  44. Gómez-Gallego, M.; Sierra, M.A. Kinetic Isotope Effects in the Study of Organometallic Reaction Mechanisms. Chem. Rev. 2011, 111, 4857–4963. [Google Scholar] [CrossRef] [PubMed]
  45. Pirali, T.; Serafini, M.; Cargnin, S.; Genazzani, A.A. Applications of Deuterium in Medicinal Chemistry. J. Med. Chem. 2019, 62, 5276–5297. [Google Scholar] [CrossRef] [PubMed]
  46. Kopf, S.; Bourriquen, F.; Li, W.; Neumann, H.; Junge, K.; Beller, M. Recent Developments for the Deuterium and Tritium Labeling of Organic Molecules. Chem. Rev. 2022, 122, 6634–6718. [Google Scholar] [CrossRef] [PubMed]
  47. Li, W.; Rabeah, J.; Bourriquen, F.; Yang, D.; Kreyenschulte, C.; Rockstroh, N.; Lund, H.; Bartling, S.; Surkus, A.-E.; Junge, K.; et al. Scalable and Selective Deuteration of (Hetero)arenes. Nat. Chem. 2022, 14, 334–341. [Google Scholar] [CrossRef]
  48. Norcott, P.L. Current Electrochemical Approaches to Selective Deuteration. Chem. Comm. 2022, 58, 2944–2953. [Google Scholar] [CrossRef]
  49. Li, P.; Guo, C.; Wang, S.; Ma, D.; Feng, T.; Wang, Y.; Qiu, Y. Facile and General Electrochemical Deuteration of Unactivated Alkyl Halides. Nat. Commun. 2022, 13, 3774. [Google Scholar] [CrossRef]
  50. Lu, L.; Li, H.; Zheng, Y.; Bu, F.; Lei, A. Facile and Economical Electrochemical Dehalogenative Deuteration of (Hetero)Aryl Halides. CCS Chem. 2021, 3, 2669–2675. [Google Scholar] [CrossRef]
  51. Mitsudo, K.; Okada, T.; Shimohara, S.; Mandai, H.; Suga, S. Electro-reductive Halogen-Deuterium Exchange and Methylation of Aryl Halides in Acetonitrile. Electrochemistry 2013, 81, 362–364. [Google Scholar] [CrossRef] [Green Version]
  52. Kim, H.; Kim, H.; Lambert, T.H.; Lin, S. Reductive Electrophotocatalysis: Merging Electricity and Light To Achieve Extreme Reduction Potentials. J. Am. Chem. Soc. 2020, 142, 2087–2092. [Google Scholar] [CrossRef] [PubMed]
  53. Hong, J.; Liu, Q.; Li, F.; Bai, G.; Liu, G.; Li, M.; Nayal, O.S.; Fu, X.; Mo, F. Electrochemical Radical Borylation of Aryl Iodides. Chin. J. Chem. 2019, 37, 347–351. [Google Scholar] [CrossRef]
  54. Wang, B.; Peng, P.; Ma, W.; Liu, Z.; Huang, C.; Cao, Y.; Hu, P.; Qi, X.; Lu, Q. Electrochemical Borylation of Alkyl Halides: Fast, Scalable Access to Alkyl Boronic Esters. J. Am. Chem. Soc. 2021, 143, 12985–12991. [Google Scholar] [CrossRef] [PubMed]
  55. Choi, J.; Fu, G.C. Transition Metal-Ctalyzed Alkyl-Alkyl Bond Formation: Another Dimension in Cross-Coupling Chemistry. Science 2017, 356, eaaf7230. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, X.; Dai, Y.; Gong, H. Nickel-Catalyzed Reductive Couplings. Top. Curr. Chem. 2016, 374, 43. [Google Scholar] [CrossRef] [PubMed]
  57. Lucas, E.L.; Jarvo, E.R. Stereospecific and Stereoconvergent Cross-Couplings Between Alkyl electrophiles. Nat. Rev. Chem. 2017, 1, 0065. [Google Scholar] [CrossRef]
  58. Jana, R.; Pathak, T.P.; Sigman, M.S. Advances in Transition Metal (Pd, Ni, Fe)-Catalyzed Cross-Coupling Reactions Using Alkyl-organometallics as Reaction Partners. Chem. Rev. 2011, 111, 1417–1492. [Google Scholar] [CrossRef] [Green Version]
  59. Everson, D.A.; Weix, D.J. Cross-Electrophile Coupling: Principles of Reactivity and Selectivity. J. Org. Chem. 2014, 79, 4793–4798. [Google Scholar] [CrossRef]
  60. Weix, D.J. Methods and Mechanisms for Cross-Electrophile Coupling of Csp2 Halides with Alkyl Electrophiles. Acc. Chem. Res. 2015, 48, 1767–1775. [Google Scholar] [CrossRef] [Green Version]
  61. Zhang, W.; Lu, L.; Zhang, W.; Wang, Y.; Ware, S.D.; Mondragon, J.; Rein, J.; Strotman, N.; Lehnherr, D.; See, K.A.; et al. Electrochemically Driven Cross-Electrophile Coupling of Alkyl Halides. Nature 2022, 604, 292–297. [Google Scholar] [CrossRef] [PubMed]
  62. Huang, C.; Ma, W.; Zheng, X.; Xu, M.; Qi, X.; Lu, Q. Epoxide Electroreduction. J. Am. Chem. Soc. 2022, 144, 1389–1395. [Google Scholar] [CrossRef]
  63. Liao, L.-L.; Wang, Z.-H.; Cao, K.-G.; Sun, G.-Q.; Zhang, W.; Ran, C.-K.; Li, Y.; Chen, L.; Cao, G.-M.; Yu, D.-G. Electrochemical Ring-Opening Dicarboxylation of Strained Carbon–Carbon Single Bonds with CO2: Facile Synthesis of Diacids and Derivatization into Polyesters. J. Am. Chem. Soc. 2022, 144, 2062–2068. [Google Scholar] [CrossRef] [PubMed]
  64. Kolb, S.; Petzold, M.; Brandt, F.; Jones, P.G.; Jacob, C.R.; Werz, D.B. Electrocatalytic Activation of Donor–Acceptor Cyclopropanes and Cyclobutanes: An Alternative C(sp3)−C(sp3) Cleavage Mode. Angew. Chem., Int. Ed. 2021, 60, 15928–15934. [Google Scholar] [CrossRef] [PubMed]
  65. Peng, P.; Yan, X.; Zhang, K.; Liu, Z.; Zeng, L.; Chen, Y.; Zhang, H.; Lei, A. Electrochemical C−C Bond Cleavage of Cyclopropanes Towards the Synthesis of 1,3-Difunctionalized Molecules. Nat. Commun. 2021, 12, 3075. [Google Scholar] [CrossRef]
  66. Kolb, S.; Ahlburg, N.L.; Werz, D.B. Friedel–Crafts-Type Reactions with Electrochemically Generated Electrophiles from Donor–Acceptor Cyclopropanes and -Butanes. Org. Lett. 2021, 23, 5549–5553. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, Y.; Tang, S.; Yang, G.; Wang, S.; Ma, D.; Qiu, Y. Electrocarboxylation of Aryl Epoxides with CO2 for the Facile and Selective Synthesis of beta-Hydroxy Acids. Angew. Chem. Int. Ed. 2022, 61, e202207746. [Google Scholar]
  68. Zhang, K.; Ren, B.-H.; Liu, X.-F.; Wang, L.-L.; Zhang, M.; Ren, W.-M.; Lu, X.-B.; Zhang, W.-Z. Direct and Selective Electrocarboxylation of Styrene Oxides with CO2 for Accessing β-Hydroxy Acids. Angew. Chem., Int. Ed. 2022, 61, e202207660. [Google Scholar]
  69. Hu, P.; Peters, B.K.; Malapit, C.A.; Vantourout, J.C.; Wang, P.; Li, J.; Mele, L.; Echeverria, P.-G.; Minteer, S.D.; Baran, P.S. Electroreductive Olefin–Ketone Coupling. J. Am. Chem. Soc. 2020, 142, 20979–20986. [Google Scholar] [CrossRef]
  70. Mori, A.; Miyakawa, Y.; Ohashi, E.; Haga, T.; Maegawa, T.; Sajiki, H. Pd/C-Catalyzed Chemoselective Hydrogenation in the Presence of Diphenylsulfide. Org. Lett. 2006, 8, 3279–3281. [Google Scholar] [CrossRef]
  71. Mendes-Burak, J.; Ghaffari, B.; Copéret, C. Selective Hydrogenation of α,β-Unsaturated Carbonyl Compounds on Silica-Supported Copper Nanoparticles. Chem. Comm. 2019, 55, 179–181. [Google Scholar] [CrossRef]
  72. Mori, A.; Fujita, A.; Nishihara, Y.; Hiyama, T. Copper(I) Salt Mediated 1,4-Reduction of α,β-Unsaturated Ketones Using Hydrosilanes. Chem. Comm. 1997, 2159–2160. [Google Scholar] [CrossRef] [Green Version]
  73. Sugiura, M.; Sato, N.; Kotani, S.; Nakajima, M. Lewis Base-Catalyzed Conjugate Reduction and Reductive Aldol Reaction of α,β-Unsaturated Ketones Using Trichlorosilane. Chem. Comm. 2008, 4309–4311. [Google Scholar] [CrossRef] [PubMed]
  74. Broggi, J.; Jurčík, V.; Songis, O.; Poater, A.; Cavallo, L.; Slawin, A.M.Z.; Cazin, C.S.J. The Isolation of [Pd{OC(O)H}(H)(NHC)(PR3)] (NHC = N-Heterocyclic Carbene) and Its Role in Alkene and Alkyne Reductions Using Formic Acid. J. Am. Chem. Soc. 2013, 135, 4588–4591. [Google Scholar]
  75. Baán, Z.; Finta, Z.; Keglevich, G.; Hermecz, I. Unexpected Chemoselectivity in the Rhodium-Catalyzed Transfer Hydrogenation of α,β-Unsaturated Ketones in Ionic Liquids. Green Chem. 2009, 11, 1937–1940. [Google Scholar] [CrossRef]
  76. Wang, Y.; Huang, Z.; Leng, X.; Zhu, H.; Liu, G.; Huang, Z. Transfer Hydrogenation of Alkenes Using Ethanol Catalyzed by a NCP Pincer Iridium Complex: Scope and Mechanism. J. Am. Chem. Soc. 2018, 140, 4417–4429. [Google Scholar] [PubMed]
  77. Li, J.; He, L.; Liu, X.; Cheng, X.; Li, G. Electrochemical Hydrogenation with Gaseous Ammonia. Angew. Chem. Int. Ed. 2019, 58, 1759–1763. [Google Scholar] [CrossRef]
  78. Liu, X.; Liu, R.; Qiu, J.; Cheng, X.; Li, G. Chemical-Reductant-Free Electrochemical Deuteration Reaction using Deuterium Oxide. Angew. Chem. Int. Ed. 2020, 59, 13962–13967. [Google Scholar] [CrossRef]
  79. Huang, B.; Li, Y.; Yang, C.; Xia, W. Electrochemical 1,4-Reduction of α,β-Unsaturated Ketones with Methanol and Ammonium Chloride as Hydrogen Sources. Chem. Comm. 2019, 55, 6731–6734. [Google Scholar] [CrossRef]
  80. Vittimberga, B.M.; Minisci, F.; Morrocchi, S. Photoinitiated Electron Transfer-Substitution Reaction of Diphenyl Ketyl with Protonated 4-Cyanopyridine. J. Am. Chem. Soc. 1975, 97, 4397–4398. [Google Scholar] [CrossRef]
  81. McNally, A.; Prier, C.K.; MacMillan, D.W.C. Discovery of an α-Amino C–H Arylation Reaction Using the Strategy of Accelerated Serendipity. Science 2011, 334, 1114–1117. [Google Scholar] [CrossRef]
  82. Pirnot, M.T.; Rankic, D.A.; Martin, D.B.C.; MacMillan, D.W.C. Photoredox Activation for the Direct β-Arylation of Ketones and Aldehydes. Science 2013, 339, 1593–1596. [Google Scholar] [CrossRef]
  83. Lehnherr, D.; Lam, Y.-h.; Nicastri, M.C.; Liu, J.; Newman, J.A.; Regalado, E.L.; DiRocco, D.A.; Rovis, T. Electrochemical Synthesis of Hindered Primary and Secondary Amines via Proton-Coupled Electron Transfer. J. Am. Chem. Soc. 2020, 142, 468–478. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, X.; Yang, C.; Gao, H.; Wang, L.; Guo, L.; Xia, W. Reductive Arylation of Aliphatic and Aromatic Aldehydes with Cyanoarenes by Electrolysis for the Synthesis of Alcohols. Org. Lett. 2021, 23, 3472–3476. [Google Scholar] [CrossRef] [PubMed]
  85. Xu, H.; Liu, J.; Nie, F.; Zhao, X.; Jiang, Z. Metal-Free Hydropyridylation of Thioester-Activated Alkenes via Electroreductive Radical Coupling. J. Org. Chem. 2021, 86, 16204–16212. [Google Scholar] [CrossRef]
  86. Ding, W.; Li, M.; Fan, J.; Cheng, X. Palladium-Catalyzed Asymmetric Allylic 4-Pyridinylation via Electroreductive Substitution Reaction. Nat. Commun. 2022, 13, 5642. [Google Scholar] [CrossRef]
  87. Ke, J.; Wang, H.; Zhou, L.; Mou, C.; Zhang, J.; Pan, L.; Chi, Y.R. Hydrodehalogenation of Aryl Halides through Direct Electrolysis. Chem. Eur. J. 2019, 25, 6911–6914. [Google Scholar] [CrossRef]
  88. Liu, C.; Han, S.; Li, M.; Chong, X.; Zhang, B. Electrocatalytic Deuteration of Halides with D2O as the Deuterium Source over a Copper Nanowire Arrays Cathode. Angew. Chem. Int. Ed. 2020, 59, 18527–18531. [Google Scholar] [CrossRef]
  89. Huang, B.; Guo, L.; Xia, W. A Facile and Versatile Electro-Reductive System for Hydrodefunctionalization Under Ambient Conditions. Green Chem. 2021, 23, 2095–2103. [Google Scholar] [CrossRef]
  90. Wang, P.; Yang, Z.; Wang, Z.; Xu, C.; Huang, L.; Wang, S.; Zhang, H.; Lei, A. Electrochemical Arylation of Electron-Deficient Arenes through Reductive Activation. Angew. Chem. Int. Ed. 2019, 58, 15747–15751. [Google Scholar] [CrossRef]
  91. Francke, R.; Little, R.D. Redox Catalysis in Organic Electrosynthesis: Basic Principles and Recent Developments. Chem. Soc. Rev. 2014, 43, 2492–2521. [Google Scholar] [CrossRef]
  92. Tamirat, A.G.; Guan, X.; Liu, J.; Luo, J.; Xia, Y. Redox Mediators as Charge Agents for Changing Electrochemical Reactions. Chem. Soc. Rev. 2020, 49, 7454–7478. [Google Scholar] [CrossRef]
  93. Sun, G.; Ren, S.; Zhu, X.; Huang, M.; Wan, Y. Direct Arylation of Pyrroles via Indirect Electroreductive C–H Functionalization Using Perylene Bisimide as an Electron-Transfer Mediator. Org. Lett. 2016, 18, 544–547. [Google Scholar] [CrossRef] [PubMed]
  94. Cowper, N.G.W.; Chernowsky, C.P.; Williams, O.P.; Wickens, Z.K. Potent Reductants via Electron-Primed Photoredox Catalysis: Unlocking Aryl Chlorides for Radical Coupling. J. Am. Chem. Soc. 2020, 142, 2093–2099. [Google Scholar] [CrossRef] [PubMed]
  95. Folgueiras-Amador, A.A.; Teuten, A.E.; Salam-Perez, M.; Pearce, J.E.; Denuault, G.; Pletcher, D.; Parsons, P.J.; Harrowven, D.C.; Brown, R.C.D. Cathodic Radical Cyclisation of Aryl Halides Using a Strongly-Reducing Catalytic Mediator in Flow. Angew. Chem. Int. Ed. 2022, 61, e202203694. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, Y.; Zhao, Z.; Pan, D.; Wang, S.; Jia, K.; Ma, D.; Yang, G.; Xue, X.S.; Qiu, Y. Metal-Free Electrochemical Carboxylation of Organic Halides in the Presence of Catalytic Amounts of an Organomediator. Angew. Chem. Int. Ed. 2022, 134, e202210201. [Google Scholar]
  97. Xu, J.; Lu, F.L.; Sun, L.H.; Huang, M.N.; Jiang, J.W.; Wang, K.; Ouyang, D.D.; Lu, L.J.; Lei, A. Electrochemical Reductive Cross-coupling of Acyl Chlorides and Sulfinic Acids Towards the Synthesis of Thioesters. Green Chem. 2022, 24, 7350–7354. [Google Scholar] [CrossRef]
  98. Bogonda, G.; Patil, D.V.; Kim, H.Y.; Oh, K. Visible-Light-Promoted Thiyl Radical Generation from Sodium Sulfinates: A Radical–Radical Coupling to Thioesters. Org. Lett. 2019, 21, 3774–3779. [Google Scholar] [CrossRef]
  99. Lu, L.; Siu, J.C.; Lai, Y.; Lin, S. An Electroreductive Approach to Radical Silylation via the Activation of Strong Si–Cl Bond. J. Am. Chem. Soc. 2020, 142, 21272–21278. [Google Scholar] [CrossRef]
  100. Xu, R.; Xu, T.; Yang, M.; Cao, T.; Liao, S. A Rapid Access to Aliphatic Sulfonyl Fluorides. Nat. Commun. 2019, 10, 3752. [Google Scholar] [CrossRef] [Green Version]
  101. Nie, X.; Xu, T.; Song, J.; Devaraj, A.; Zhang, B.; Chen, Y.; Liao, S. Radical Fluorosulfonylation: Accessing Alkenyl Sulfonyl Fluorides from Alkenes. Angew. Chem. Int. Ed. 2021, 60, 3956–3960. [Google Scholar] [CrossRef]
  102. Chen, D.; Nie, X.; Feng, Q.; Zhang, Y.; Wang, Y.; Wang, Q.; Huang, L.; Huang, S.; Liao, S. Electrochemical Oxo-Fluorosulfonylation of Alkynes under Air: Facile Access to β-Keto Sulfonyl Fluorides. Angew. Chem. Int. Ed. 2021, 60, 27271–27276. [Google Scholar] [CrossRef]
  103. Feng, Q.; Fu, Y.; Zheng, Y.; Liao, S.; Huang, S. Electrochemical Synthesis of β-Keto Sulfonyl Fluorides via Radical Fluorosulfonylation of Vinyl Triflates. Org. Lett. 2022, 24, 3702–3706. [Google Scholar] [CrossRef]
  104. Zhang, L.; Zhang, G.; Wang, P.; Li, Y.; Lei, A. Electrochemical Oxidation with Lewis-Acid Catalysis Leads to Trifluoromethylative Difunctionalization of Alkenes Using CF3SO2Na. Org. Lett. 2018, 20, 7396–7399. [Google Scholar] [CrossRef]
  105. Ouyang, Y.; Xu, X.-H.; Qing, F.-L. Electrochemical Trifluoromethoxylation of (Hetero)aromatics with a Trifluoromethyl Source and Oxygen. Angew. Chem. Int. Ed. 2022, 61, e202114048. [Google Scholar] [CrossRef] [PubMed]
  106. Liu, D.; Liu, Z.-R.; Wang, Z.-H.; Ma, C.; Herbert, S.; Schirok, H.; Mei, T.-S. Paired Electrolysis-Enabled Nickel-Catalyzed Enantioselective Reductive Cross-Coupling Between α-Chloroesters and Aryl Bromides. Nat. Commun. 2022, 13, 7318. [Google Scholar] [CrossRef] [PubMed]
  107. Li, Z.; Sun, W.; Wang, X.; Li, L.; Zhang, Y.; Li, C. Electrochemically Enabled, Nickel-Catalyzed Dehydroxylative Cross-Coupling of Alcohols with Aryl Halides. J. Am. Chem. Soc. 2021, 143, 3536–3543. [Google Scholar] [CrossRef] [PubMed]
  108. Lin, Q.; Luo, S. Tailoring Radicals by Asymmetric Electrochemical Catalysis. Org. Chem. Front. 2020, 7, 2997–3000. [Google Scholar] [CrossRef]
Figure 1. Cathodically generated radicals and the general mechanism of electroreductive transformations.
Figure 1. Cathodically generated radicals and the general mechanism of electroreductive transformations.
Molecules 28 00857 g001
Figure 2. General mechanism of electroreductive transformations via alkyl radicals.
Figure 2. General mechanism of electroreductive transformations via alkyl radicals.
Molecules 28 00857 g002
Figure 3. Electroreductive reaction of alkyl halides with alkenes.
Figure 3. Electroreductive reaction of alkyl halides with alkenes.
Molecules 28 00857 g003
Figure 4. Electrochemical deuteration of unactivated alkyl halides by Qiu group.
Figure 4. Electrochemical deuteration of unactivated alkyl halides by Qiu group.
Molecules 28 00857 g004
Figure 5. Electrochemical borylation of alkyl halides for alkyl boronic esters.
Figure 5. Electrochemical borylation of alkyl halides for alkyl boronic esters.
Molecules 28 00857 g005
Figure 6. Electrochemically driven cross-electrophile coupling of alkyl halides.
Figure 6. Electrochemically driven cross-electrophile coupling of alkyl halides.
Molecules 28 00857 g006
Figure 7. Electroreductionof epoxides or cyclopropanes.
Figure 7. Electroreductionof epoxides or cyclopropanes.
Molecules 28 00857 g007
Figure 8. Electroreductive olefin−ketone coupling for the mild synthesis of tertiary alcohols.
Figure 8. Electroreductive olefin−ketone coupling for the mild synthesis of tertiary alcohols.
Molecules 28 00857 g008
Figure 9. Electrochemical hydrogenation with gaseous ammonia or deuteration with D2O of alkenes, alkynes and ketones.
Figure 9. Electrochemical hydrogenation with gaseous ammonia or deuteration with D2O of alkenes, alkynes and ketones.
Molecules 28 00857 g009
Figure 10. Electrochemical 1,4-reduction of α,β-unsaturated ketones with MeOH and NH4Cl as hydrogen sources.
Figure 10. Electrochemical 1,4-reduction of α,β-unsaturated ketones with MeOH and NH4Cl as hydrogen sources.
Molecules 28 00857 g010
Figure 11. Electroreductive of cyanoheteroarenes with iminium salts or ketones.
Figure 11. Electroreductive of cyanoheteroarenes with iminium salts or ketones.
Molecules 28 00857 g011
Figure 12. Electrochemical hydropyridylation of α,β-unsaturated carbonyl compounds.
Figure 12. Electrochemical hydropyridylation of α,β-unsaturated carbonyl compounds.
Molecules 28 00857 g012
Figure 13. General mechanism for electroreductive transformation via aryl radical.
Figure 13. General mechanism for electroreductive transformation via aryl radical.
Molecules 28 00857 g013
Figure 14. Direct electroreductive hydrodehalogenation and deuterodehalogenation of aryl halides.
Figure 14. Direct electroreductive hydrodehalogenation and deuterodehalogenation of aryl halides.
Molecules 28 00857 g014
Figure 15. Direct electroreductive arylation of electro-deficient arenes.
Figure 15. Direct electroreductive arylation of electro-deficient arenes.
Molecules 28 00857 g015
Figure 16. Electroreductive arylation and borylation of aryl halides.
Figure 16. Electroreductive arylation and borylation of aryl halides.
Molecules 28 00857 g016
Figure 17. Electrophotochemical reductive transformation of aryl halides, DCA = dicyanoanthracene.
Figure 17. Electrophotochemical reductive transformation of aryl halides, DCA = dicyanoanthracene.
Molecules 28 00857 g017
Figure 18. Phenanthrene mediated electroreductive intramolecular cyclisation of aryl halides by Brown group.
Figure 18. Phenanthrene mediated electroreductive intramolecular cyclisation of aryl halides by Brown group.
Molecules 28 00857 g018
Figure 19. Naphthalene mediated electroreductive carboxylation of aryl halides by Qiu group.
Figure 19. Naphthalene mediated electroreductive carboxylation of aryl halides by Qiu group.
Molecules 28 00857 g019
Figure 20. Electroreductive cross-coupling between acyl radicals and thiyl radicals by Lei group.
Figure 20. Electroreductive cross-coupling between acyl radicals and thiyl radicals by Lei group.
Molecules 28 00857 g020
Figure 21. Electroreductive silylation of alkenes by Lin group.
Figure 21. Electroreductive silylation of alkenes by Lin group.
Molecules 28 00857 g021
Figure 22. Electroreductive fluorosulfonylation of alkynes or vinyl triflates for the synthesis of β-keto sulfonyl fluorides by Liao group.
Figure 22. Electroreductive fluorosulfonylation of alkynes or vinyl triflates for the synthesis of β-keto sulfonyl fluorides by Liao group.
Molecules 28 00857 g022
Figure 23. Electroreductive trifluoromethoxylation of (hetero)aromatics with a trifluoromethyl source and oxygen by Qing group.
Figure 23. Electroreductive trifluoromethoxylation of (hetero)aromatics with a trifluoromethyl source and oxygen by Qing group.
Molecules 28 00857 g023
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xiang, H.; He, J.; Qian, W.; Qiu, M.; Xu, H.; Duan, W.; Ouyang, Y.; Wang, Y.; Zhu, C. Electroreductively Induced Radicals for Organic Synthesis. Molecules 2023, 28, 857. https://doi.org/10.3390/molecules28020857

AMA Style

Xiang H, He J, Qian W, Qiu M, Xu H, Duan W, Ouyang Y, Wang Y, Zhu C. Electroreductively Induced Radicals for Organic Synthesis. Molecules. 2023; 28(2):857. https://doi.org/10.3390/molecules28020857

Chicago/Turabian Style

Xiang, Huaming, Jinyu He, Weifeng Qian, Mingqiang Qiu, Hao Xu, Wenxi Duan, Yanyan Ouyang, Yanzhao Wang, and Cuiju Zhu. 2023. "Electroreductively Induced Radicals for Organic Synthesis" Molecules 28, no. 2: 857. https://doi.org/10.3390/molecules28020857

Article Metrics

Back to TopTop