Skip to main content
ACS Central Science logoLink to ACS Central Science
. 2021 Mar 9;7(3):415–431. doi: 10.1021/acscentsci.0c01532

Organic Electrochemistry: Molecular Syntheses with Potential

Cuiju Zhu , Nate W J Ang , Tjark H Meyer †,, Youai Qiu , Lutz Ackermann †,‡,*
PMCID: PMC8006177  PMID: 33791425

Abstract

graphic file with name oc0c01532_0021.jpg

Efficient and selective molecular syntheses are paramount to inter alia biomolecular chemistry and material sciences as well as for practitioners in chemical, agrochemical, and pharmaceutical industries. Organic electrosynthesis has undergone a considerable renaissance and has thus in recent years emerged as an increasingly viable platform for the sustainable molecular assembly. In stark contrast to early strategies by innate reactivity, electrochemistry was recently merged with modern concepts of organic synthesis, such as transition-metal-catalyzed transformations for inter alia C–H functionalization and asymmetric catalysis. Herein, we highlight the unique potential of organic electrosynthesis for sustainable synthesis and catalysis, showcasing key aspects of exceptional selectivities, the synergism with photocatalysis, or dual electrocatalysis, and novel mechanisms in metallaelectrocatalysis until February of 2021.

Short abstract

Organic electrosynthesis has unique potential and showcases exceptional selectivities, synergism with photocatalysis, or dual electrocatalysis, and novel mechanisms in metallaelectrocatalysis.

1. Introduction

Organic synthesis arguably represents the key discipline for the bottom-up assembly and late-stage diversification of molecular compounds with transformative applications to inter alia medicinal chemistry, drug development, and material sciences as well as chemical and pharmaceutical industries.1 While scientists have in the past decade increasingly exploited the enabling platforms of photochemistry,2 artificial intelligence,3 mechanochemistry,4 or flow technology,5 molecular electrosynthesis has largely laid dormant until very recently.6 Particularly, organic electrochemistry has in recent years overcome some of its past limitations as a niche technique.7 Electroorganic synthesis can indeed be traced back to the 19th century with Faraday’s hydrolysis of acetic acid to hydrocarbons8 and Kolbe’s electrochemical decarboxylative dimerization (Figure 1).9 Hence, in the early twentieth century, Hickeling proposed that reactions could be conducted under potentiostatic control, rather than constant-current electrolysis.10 These findings were based on the emerging interest in polarographic methods at that time, such as voltammetry—the detection of current as a function of the potential at a solid working electrode—developed by Heyrovský among others.11 In the mid-twentieth century, electrochemistry has been identified as an economically attractive approach for scalable commodity chemicals, for instance, the Simons fluorination process,12 Monsanto adiponitrile processes,13 or later, the BASF Lysmeral synthesis via anodic benzylic oxidation,14 indicating the scalability of electroorganic synthesis, particularly in a paired electrosynthesis regime.14b,15 Yoshida introduced the concept of electroauxiliaries to selectively lower the electrochemical potential of substrates in the late twentieth century.16 The use of redox mediators for indirect electrolysis could be traced back to 1900, when inorganic redox mediators were applied to the synthesis of quinones,17 while the principles of indirect electrolysis were formalized by Steckhan in the 1980s.18 Additional key achievements on the direct exploitation of electric current were made by Little,19 Schäfer,20 Lund,21 Moeller,22 Amatore,23 Jutand,24 and Yoshida.25 On the basis of these pioneering contributions, electrosynthesis gained significant momentum for sustainable organic syntheses (Figure 1).26

Figure 1.

Figure 1

Selected milestones of representative electroorganic chemistry.

The resurgence of this strategy stems from, among other things, an alternative array of reaction mechanisms that are exclusively feasible by electrochemistry or synergistically by photoelectrochemistry.27 Here, electrons can be used as traceless redox equivalents to achieve exceptional selectivities, thus avoiding stoichiometric redox reagents and undesired byproduct generation.28 The development of commercial electrochemical equipment,29 has enabled most user-friendly access of electrosynthesis. Furthermore, the often exceedingly mild reaction conditions and the frequent use of protic solvents encompasses electrosynthesis as an environmentally benign strategy for molecular assembly.30 On a different note, electrochemistry has the ability to regulate reactivity and selectivity by the precise control of the applied potential.6a,31 Thus, this unique tunability translates into an unmatched chemoselectivity of electrochemistry compared with commonly used chemical redox reagents.32 Particularly, the concept of indirect electrolysis with redox mediators can improve the efficiency and chemoselectivity of electrosynthesis (Figure 2). Synergistic electrocatalysis has gained particular recent impetus within the renaissance of electroorganic syntheses, which has addressed several challenges encountered in modern organic syntheses.33 Specifically, electrophotochemistry combined the electrochemical and photochemical steps in tandem pathways to generate a highly reactive intermediate, thus providing new avenues for contemporary reaction design and molecular transformations.27,34 The merger of electrosynthesis with transition metal catalysis enabled novel resource-economic bond functionalizations, which unearthed a variety of new reaction mechanisms.35 Electrochemical reduction shows largely untapped potential for reductive organic syntheses through cathodic reduction with the aid of a sacrificial anode material.36 Enantioselective electrosynthesis, a key research arena which is highly relevant for pharmaceutical and crop-protecting industries, provides greener synthetic methods unattainable by traditional means (Figure 2).37

Figure 2.

Figure 2

Prospects of modern concepts in organic electrosynthesis and electrocatalysis.

The current Outlook highlights representative recent concepts for the electrocatalytic diversification of organic molecules beyond their innate reactivity, until 2021. Thus, we focus on the selectivity control by modern redox mediators, dual electrocatalysis, new mechanistic innovation, electrochemical reduction, and asymmetric electrocatalysis, while conventional organic electrosynthesis has been comprehensively summarized elsewhere.31,38

2. REDOX MEDIATORS: UNIQUE SELECTIVITY CONTROL IN ORGANIC ELECTROCHEMISTRY

Direct electrolysis enables molecules to undergo electron transfer directly at the electrode surface. In contrast, with indirect electrosynthesis, a redox mediator which is more easily oxidized or reduced than the substrate, acts as the electron-transfer-shuttle from the heterogeneous electrode surface to the homogeneous dissolved substrates.19a,39 The concept of indirect electrolysis offers several advantages. In many cases, the indirect approach results in improved reaction efficacy and better chemoselectivity by avoiding undesirable side reactions. The facile homogeneous redox process of the mediator eliminates kinetic inhibition of the heterogeneous electron transfer and structural modification of the mediator offer direct selectivity control (Scheme 1A).18b,19a,40 In this context, the beneficial effects of redox mediators were recently exemplified for the challenging selective anodic oxidation of activated C–H bonds, which proved to be useful for the transformation of natural products. The electrochemical direct allylic oxidation pioneered by Shono, can engage the electrogenerated alkyl radical with the subsequent fragmentation of the cyclobutane ring to enable the direct oxidation of α-pinene (1).41 A major advance was made by Masui, employing N-hydroxyphthalimide (NHPI) as an electron carrier to facilitate allylic C–H oxidations.42

Scheme 1. Case Studies in Redox Mediator in Organic Electrosynthesis.

Scheme 1

Cl4NHPI = tetrachloro-N-hydroxyphthalimide, BQ = p-benzoquinone, CCE = constant current electrolysis.

Taking inspiration from these findings, Baran discovered that the use of a modified tetrachloro-derivative of NHPI, namely tetrachloro-N-hydroxyphthalimide (Cl4NHPI), significantly improved the reaction outcome (Scheme 1B).43 Comparing the reversible redox potentials of NHPI (E1/2 = 0.78 V vs. Ag/AgCl) and Cl4NHPI (E1/2 = 0.87 V vs. Ag/AgCl), a slightly higher redox couple of Cl4NHPI was indicative of the stronger oxidative character of the more electron-deficient N-oxyl radical. A 100 gram-scale reaction highlighted the synthetic utility of the approach, including the selective functionalization of steroids, monoterpenoids, and triterpenoids.

The redox mediator strategy was likewise implemented in the oxygenation of annulated hydrocarbons with quinuclidine as the mediator (Scheme 1C).44 The indirect electrochemical process enabled the anodic oxidation to proceed at relatively low potentials compared with direct oxidation. Thus, chemoselective oxidation of unactivated C–H bonds proved to be amenable. The accessible electrooxidation allowed for the successful oxidation of sclareolide 4 on a 50 g-scale.

In contrast to traditional Shono oxidations by direct electron transfer to the electrode, Stahl developed the α-C–H oxygenation of cyclic carbamates 5 using bicyclic aminoxyl as a mediator (Scheme 1D).45 Here, the anodically generated oxoammonium species promoted the oxidation of the substrate 5 to form the substrate-derived iminium ion, which reacted with H2O in a two-electron oxidation to afford the desired product 6. Notably, the redox mediator benefited from an oxidizing potential being 1.0 V lower than the direct one-electron oxidation of the substrate. Hence, a good functional-group tolerance and broad substrate scope were shown to be viable.

Recently, the redox mediator strategy was successfully extended to metalla-electrocatalyzed C–H activation. With regard to studies by Jutand, on p-benzoquinone as a redox mediator for Fujiwara–Moritani type reactions,39c Ackermann showed the beneficial effect of redox mediators for iridaelectro-catalyzed C–H alkenylations of benzoic acids 7 (Scheme 1E).46 Here, a variety of sensitive functional groups, including cyano, ester, halide, and even labile iodo, were fully tolerated. The efficacy and chemoselectivity of the iridium electrocatalysis were considerably improved by the aid of the redox mediator. The robustness of the iridaelectro-catalyzed C–H activation47 was further exploited for the assembly of medicinally relevant steroids and peptides.

It is noteworthy that the concept of modern redox mediators has proven to be beneficial for various aspects of organic electrosynthesis and innovative electrocatalysis (vide infra).48 Also, novel electrode materials49 have recently proven to have a major impact on the selectivity of specific organic electrochemical transformations. For instance, on the basis of the pioneering studies on biaryl formations in electrosynthesis,50 recent impetus has been gained in electrooxidative coupling reactions to enable unprecedented substitution patterns and selectivity regimes, particularly by the aid of boron-doped diamond (BDD) electrodes, along with 1,1,1,3,3,3-hexafluoro-propan-2-ol (HFIP) as the solvent.26c,51

3. DUAL ELECTROCATALYSIS IN ORGANIC SYNTHESES

3.1. Synergistic Dual Catalysis for Electrochemistry

In contrast to indirect electrolysis, dual electrocatalysis enables two distinct catalytic transformations that go beyond electron transfer, including but not being limited to, group transfer reactions, hydrogen atom transfer, or two individual catalytic concepts, such as mediated electrochemistry and photoredox catalysis. It was shown to be an efficient strategy to improve catalytic performance, chemoselectivity, and overall catalytic efficacy.52 In a representative elegant example, Lin recently utilized a novel chiral bisoxazoline (sBOX) ligand to establish asymmetric alkene hydrocyanation reactions within a dual electrocatalytic manifold by employing Co(salen) 9 and Cu(sBOX) as the catalysts (Scheme 2A).53 The proposed catalytic scenario consists of initial formation of the catalytically competent [CoIII]–H species, formed from the anodically oxidized cobalt(III)salen complex and a hydrosilane. Subsequent hydrogen-atom transfer (HAT)54 between the formal [CoIII]–H intermediate and olefin 8 furnishes the new C–H bond, along with a carbon-centered radical. The radical species now enters a second electrocatalytic cycle that is responsible for the asymmetric cyanide transfer. Here, the thus formed radical species is proposed to undergo single-electron oxidative addition to a [CuII]–CN complex, generating a copper(III) adduct. Finally, reductive elimination from the chiral complex delivers the enantio-enriched nitrile 10 and a reduced copper(I) complex, which is easily reoxidized via anodic oxidation. The merger of two distinct electrocatalytic radical reactions, namely, cobalt-catalyzed hydrogen-atom transfer (HAT) and copper-catalyzed radical cyanation,55 was the key to put into practice the asymmetric hydrocyanation of alkenes. Moreover, the electrocatalysis protocol featured significantly improved yields, chemoselectivities, and enantioselectivities compared with transformations using chemical oxidants, as was exemplified by the failure of topical oxidants such as N-fluorobenzenesulfonimide (NFSI), tert-butyl hydroperoxide (TBHP), Cu(OAc)2, and PhI(OAc)2 (Scheme 2B).

Scheme 2. Synergism of Dual Electrocatalysis in Practice: (A) and (B) Enantioselective Dual Electrocatalysis; (C) C–H Oxygenation by Synergistic Catalysis.

Scheme 2

NFSI = N-fluorobenzenesulfonimide, TBHP = tert-butyl hydroperoxide.

Hypervalent iodine(III) reagents have been extensively studied as strong chemical oxidants for oxidative transformations.56 However, they commonly need to be synthesized and meticulously handled, and their use often results in stoichiometric waste products. Recently, Ackermann merged the catalytic electro-regeneration of hypervalent iodine(III) reagents57 with ruthenaelectro-catalyzed C–H oxygenations (Scheme 2C).58 The iodoarenes and ruthenium dual electrocatalytic strategy provided a unique avenue toward sustainable C–H oxygenations. Commonly used chemical oxidants, required to generate the hypervalent iodine(III) reagent, such as m-CPBA or ozone, indeed failed to deliver the desired product 13 with satisfactory yields. These results highlight that the hypervalent iodine reagent does not only operate as the electron-shuttle (vide supra) but can rather be classified as a transfer reagent of carboxylate anions, while likewise mediating the oxidative generation of high-valent ruthenium(IV) intermediates.59 Notably, a series of weakly-O-coordinating amides 11 was thereby selectively converted to the corresponding oxygenated products 13. The dual electrocatalysis proved also viable for selective C–H oxygenations of aromatic ketones or simple arenes in the absence of an external.

3.2. Electrophotochemistry

The use of a dual electrocatalysis approach by electrophotochemistry27,34,60 broadens the possibilities for elegant reaction design and expands the viable scope of photoredox catalysis. While important contributions for electrophotosynthesis have been made by inter alia Xu,61,62 Lambert,63 Lin,64,67 Wickens,65 and our group,66 among others,34c,34d selected examples shall be discussed in the following section. Seminal work by Lin for the oxidation of alcohols 14 relied on riboflavin tetraacetate (RFT) and thiourea as the cocatalyst. Previous photochemical flavin-catalyzed aerobic oxidation of alcohols was thus far limited to benzylic alcohols. The dual electrophotocatalytic system enabled the oxidation of more challenging unactivated aliphatic alcohols 14 under exceedingly mild reaction conditions (Scheme 3A).67 Likewise, a chemical oxidant-free C–H alkylation of heteroarenes 16 with organotrifluoroborate salts was elegantly carried out by means of electrophotochemical activation (Scheme 3B).61 Highly oxidizing excited state organic dye [Mes-Acr+]* (Ered = 2.06 V vs SCE in MeCN) was generated from irradiation of the organic dye ion [Mes-Acr+]. Then, a single-electron transfer (SET) delivers acridinyl radical (Mes-Acr) and an alkyl radical, respectively. Anodic electrooxidation of the acridinyl radical (Mes-Acr) subsequently regenerates the ground state cationic catalyst [Mes-Acr+]. On a different note, the C–H/N–H coupling of azoles 19 was realized by Lambert using a trisaminocyclopropenium (TAC) ion as the electrophotocatalyst (Scheme 3C).63c The electrophotocatalysis relied on the electrochemical oxidation of trisaminocyclopropenium ion (TAC+) to the corresponding radical dication (TAC2+), followed by visible light photoexcitation to generate the highly potent oxidizer TAC2+* (E = 3.33 V vs SCE). The highly oxidizing photoexcited TAC radical dication (TAC2+*) enabled facile oxidation of inert simple arenes 18.

Scheme 3. Recent Progress in Electrophotochemistry.

Scheme 3

RFT = riboflavin tetraacetate, TFA = trifluoroacetic acid, TAC = trisaminocyclopropenium, HFIP = 1,1,1,3,3,3-hexafluoro-propan-2-ol, CCE = constant current electrolysis.

4. UNRAVELLING NOVEL MECHANISTIC INSIGHTS: ORGANIC METALLA-ELECTROCATALYZED TRANSFORMATIONS

4.1. Novel Findings for Electrochemical Cross-Couplings

Despite considerable progress, cross-couplings by nickel catalysis often suffered from the need of air-sensitive nickel(0) catalysts, strong alkoxide bases, and high temperatures.68 Electrochemical nickel-catalyzed cross-couplings have set the stage for C–C and C–heteroatom formations under significantly milder reaction conditions.69 Recently, electrochemical nickel-catalyzed aminations of aryl halides and triflates 21 were demonstrated (Scheme 4).70 Detailed mechanistic studies of the electrocatalytic aryl aminations with the aid of cyclic voltammetry, kinetic studies, and DFT calculations unraveled the novel paired electrolysis working mode, responsible for the exceedingly mild reaction conditions (Scheme 5).71 Initially, cathodic reduction of the nickel(II) precursor 24 delivers a nickel(I) species 25. Then, oxidative addition of aryl halide 21 furnishes the nickel(III) intermediate 26, and a second cathodic reduction generates the nickel(II) species 27. Along with ligand exchange with the amine, intermediate 28 is anodically oxidized to the nickel(III) complex 29. Thereafter, the desired product 23 is formed via reductive elimination, while the catalytically competent complex 25 is regenerated.

Scheme 4. Nickelaelectro-Catalyzed Amination.

Scheme 4

CCE = constant current electrolysis.

Scheme 5. Mechanistic Rationale.

Scheme 5

4.2. New Mechanistic Insights: Metalla-Electrocatalyzed C–H Activation

Transition-metal-catalyzed C–H activation has surfaced as a particularly powerful tool for step-economical molecular syntheses, with major prospects for materials sciences and the pharmaceutical industry, among others.72 The merger of C–H activation and electrochemistry thus acts as an ideal platform for modern sustainable molecular syntheses by the aid of electrochemical anodic oxidation and cathodic reduction processes.73 The application of analytical techniques such as voltammetry, spectrophotometry or in operando techniques, such as React-IR, NMR spectroscopy, and electrospray ionization mass spectrometry (ESI-MS), as well as computation, have set the stage for novel insights into the catalyst’s modes of action. Thus, electrochemistry allowed for the identification of new paradigms in metallaelectrocatalysis by characterization of short-lived intermediates and detailed insights into often fundamental single electron transfer (SET) processes.

As a pertinent example, cobalt electrocatalysis allowed for a series of C–H transformations, such as C–H oxygenations,74 C–H aminations,75 C–H/N–H annulations with alkynes,76 isocyanides and allenes,77 as well as with carbon monoxide.78 The catalyst’s modus operandi was probed through in-depth mechanistic studies.79 The well-defined cobalt(III) cyclometalated complex 31 was utilized in a series of key stoichiometric reactions. In contrast to alkyne and allene annulations, the C–H oxygenation (Scheme 6) could only be induced under anodic oxidative conditions, thus providing unique mechanistic support for an oxidation-induced reductive elimination80 by a cobalt(III/IV/II) regime. The hypothesis was further supported by means of cyclic voltammetry (CV) of the isolated cyclometalated complex 31, a Hammett–Zuman plot, and DFT calculations (Figure 3).

Scheme 6. Cobaltaelectro-Catalyzed C–H Oxygenation.

Scheme 6

CCE = constant current electrolysis.

Figure 3.

Figure 3

Cyclic voltammograms of cobaltacycles 31 in MeOH (3.5 mM) at different scan rates. The voltammograms were recorded in 0.1 M n-Bu4NPF6 at 273 K. Reprinted with permission from ref (79). Copyright 2020 Wiley-VCH Verlag GmbH & Co. KGaA.

Electrochemical transformations of unactivated C–H bonds were not limited to cobalt catalysis. Nickel catalysts, in comparison to precious metals, are earth-abundant, cost-effective, and less toxic.81 In this context, 3d nickelaelectro-catalyzed C–H activation has recently been recognized as amenable strategy for the construction of organic molecules in a sustainable and user-friendly fashion.69 Hence, C–H amination of amides proved to be viable, representing the proof-of-concept for the feasibility of C–H transformations by nickelaelectrocatalysis.82 In addition, the versatile electrocatalysis proved to be effective for the challenging C–H alkoxylations with secondary alcohols 14.83 Particularly, chemical oxidants were not able to efficiently promote the envisioned C–H alkoxylation reaction with challenging secondary alcohols 14 (Scheme 7). Cyclic voltammograms of nickela(III)cycles 34 displayed an oxidation wave at a relatively low potential of 0.50 V vs Fc0/+ (Figure 4). Moreover, detailed mechanistic investigations provided strong support for an oxidation-induced reductive elimination via a nickel(III/IV) manifold (Scheme 8), in contrast to previously reported nickel(II/III/I)-catalytic cycles facilitated by chemical oxidants.84 The broadly applicable nickelaelectro-catalyzed C–H activation approach gained further momentum for the effective conversion of synthetically useful alkyl iodides under ambient condition.85

Scheme 7. Nickelaelectrocatalysis: Electricity vs Chemical Oxidants.

Scheme 7

CCE = constant current electrolysis.

Figure 4.

Figure 4

Cyclic voltammograms of nickelacycle 34 in DMA (0.1 mM) at 100 mV/s scan rate. The voltammograms were recorded in 0.1 M n-Bu4NBF4.

Scheme 8. Proposed Catalytic Cycle for Nickelaelectro(II/III/IV)-Catalyzed C–H Activation.

Scheme 8

Iron-catalyzed C–H arylations were realized with dichloroisobutane (DCIB) as the chemical oxidant.86 DCIB is a cost-intensive vicinal dihalide ($4086/mol);87 however, it is often an essential oxidant. In order to address these limitations, direct electrochemical C–H arylation by iron catalysis was devised under mild reaction temperature. Thus, the vicinal-dichloride DCIB could be replaced by electricity as the green terminal oxidant (Scheme 9A).87 The performance of the electrocatalysis was indeed significantly increased compared with the DCIB-mediated transformation. Detailed mechanistic studies by experiments and computation (Scheme 9B) featured an iron(II) complex as the active catalyst. Initially, a ligand-to-ligand hydrogen transfer (LLHT) delivers cyclometalated iron(II) intermediate 45.88 The following transmetalation generates the iron-complex 46 and set the stage for an oxidation-induced reductive elimination. Thus, subsequent anodic oxidation delivered the iron(III) complex 47, which then undergoes reductive elimination to generate the desired product 43. Finally, the active catalyst is regenerated through anodic oxidation. Furthermore, the metallaelectrocatalysis strategy was implemented in an electrochemical manganese-catalyzed C–H arylation of amides without zinc additives.

Scheme 9. Electrochemical C–H Arylation under Iron and Manganese Catalysis.

Scheme 9

dppe = 1,2-bis(diphenylphosphino)ethane, TAM = triazolyldimethylmethyl, CCE = constant current electrolysis.

While 3d metallaelectro-catalyzed C–H activations largely involve bidentate directing groups to encompass positional-selectivity, 4d and 5d transition metals have likewise benefited from key mechanistic insights by electrocatalysis. For instance, within the synthesis of bridgehead N-fused [5,6]-bicyclic heteroarenes 51 through ruthenaelectro-catalyzed dehydrogenative C–H/N–H annulation of imidazoles 48 with alkynes 49, novel mechanistic scenarios were uncovered (Scheme 10).89 Notably, two ruthenium(II) complexes were isolated and fully characterized. The formation of product 51 was observed when electricity was applied, thus providing support for an oxidation-induced reductive elimination within an unusual ruthenium(II/III/I) regime, which is in stark contrast to the previously reported mechanisms with chemical oxidants.90

Scheme 10. Ruthenaelectro(II/III/I)-Catalyzed Alkyne Annulations.

Scheme 10

CCE = constant current electrolysis.

The robust rhodaelectrocatalytic C–H activation91 allowed electrochemical flow techniques92 to establish challenging flow-rhodaelectro-catalyzed C–H/N–H alkyne annulations with imidates 52 (Scheme 11).93 Using cyclic voltammetry and further detailed mechanistic studies by experiment provided novel insights for the rhodaelectrocatalytic manifold. Sufficient formation of product 55 was solely observed when electricity was applied to oxidize the isolated rhoda(III)cycle 54, thus providing support for an oxidation-induced reductive elimination within a rhodium(III/IV/II) regime. The generality of this strategy was demonstrated in several innovative transformations. A multiple C–H domino electrooxidative alkyne annulation was developed for accessing aza-polycyclic aromatic hydrocarbons (aza-PAHs).94 Novel rhodium-cyclometalated complexes were fully characterized and identified as key intermediates, which demonstrated the order of the three subsequent C–H activation events. Very recently, Xu concurrently reported a mechanistically related phosphorylation with N-coordinating directing groups.95 Ackermann further developed a rhodium-catalyzed electrochemical C–C activation. The rhodaelectro-catalyzed C–C alkenylation confers advantages of chemo- and position-selectivities to access the hindered 1,2,3-substituted arenes, which were not accessible by a C–H scission strategy.96

Scheme 11. Rhodaelectro-Catalyzed C–H Activation.

Scheme 11

CPE = constant potential electrolysis.

5. ELECTROCHEMICAL REDUCTION FOR UNIQUE CHEMOSELECTIVITIES AND SMALL MOLECULE TRANSFORMATIONS

Electrochemical reduction offers molecular synthesis by means of cathodic reductive electrolysis to generate radical-anions.6c,30c,36 This strategy was successfully applied for the reduction of various functional groups such as aldehydes,97 ketones,98 esters, or amides.99 An early example of electrochemical Birch reduction of benzene was disclosed by Kashimura, who employed Mg electrodes, LiClO4 as a supporting electrolyte, and t-BuOH as a proton donor in dry THF to afford 1,4-cyclohexadienes 58 as the related product.100 However, the approach was limited to simple hydrocarbons and continuous sonication was required, which impedes its application to practical synthesis. To address these drawbacks, a scalable reductive electrosynthetic strategy was devised by Baran (Scheme 12A), introducing tris(pyrrolidino)phosphoramide (TPPA) as an overcharge protectant, and dimethylurea (DMU) was used as the proton source, as well as manganese as the sacrificial anode material of choice (Scheme 12B).101 The robustness of this electrochemical Birch reduction led to unprecedented levels of functional group tolerance as well as broadly reductive transformations including ketone 58 deoxygenations and reductive ring-opening of epoxides 60 and furans (Scheme 12C).

Scheme 12. Electrochemical Birch Reduction.

Scheme 12

TPPA = tris(pyrrolidino)phosphoramide, DMU = 1,3-dimethylurea, CCE = constant current electrolysis.

Carbon dioxide evolution is of major environmental concern as a significant source of greenhouse gas emissions.102 However, CO2 has also proven to be an easily available, yet non-toxic source of C1 synthon in modern organic synthesis. The utilization of CO2 as a building block in organic synthesis have been studied in detail,103 whether for carboxylations,104 carbonylations as CO surrogates,105 methylation of amines,106 synthesis of polycarbonates, or cyclic carbonates107 from epoxides. The kinetic and thermodynamic stability of CO2, however, translates into the requirement of superstoichiometric amounts of strong reducing agents, such as Mn or Et2Zn. Innovative electrochemical approaches for the environmentally benign reduction of CO2 for C–C bond formation reactions are highly desirable.24,108 There is precedent for electroreductive cross-coupling of organohalides with CO2.104d,109 Lu and Wang reported an enantioselective electrochemical carboxylation, utilizing chiral cobalt salen complexes (Scheme 13A). This example illustrated the possibility of an asymmetric electroreductive carboxylation of inexpensive and optically inactive alkyl chloride 62 for the first time.110 Subsequently, Mei displayed an elegant palladium-catalyzed electroreductive carboxylation of allyl esters 64 with CO2 in a highly regioselectivity manner (Scheme 13B).111 In contrast, Ackermann showcased earth-abundant cobalt-catalysis for electroreductive carboxylation of allylic chlorides 66 with a simplified undivided cell setup and nontoxic solvent to achieve the synthesis of styrylacetic acid derivatives 67 in good yield and regioselectivity (Scheme 13C).112 Despite efforts into designing and employing electrochemical means for the reduction of CO2 for organic syntheses, multifarious challenges, such as the high selectivity of CO2 transformations and the need for a sacrificial anode remained unsolved to model a multifacet cross-coupling reaction with CO2 as a synthon to enrich organic synthesis with sustainable methods.

Scheme 13. Case Studies in Electrochemical Reduction of CO2.

Scheme 13

DPPPh = 1,2-Bis(diphenylphosphino)benzene.

The technique of switching polarity direction of the electrodes, such as alternating current (AC), during the reaction could avoid passivation on the electrode surface, thereby increasing the lifetime of the electrodes. This underexplored approach could be the key in unlocking some of the problematic reaction, chemoselectivity issues and avoiding the use of sacrificial anodes.113 Recent developments by Reid highlighted the ability of inverting chemoselectivity by alternating current electrolysis for the selective oxidation of 4-methyl anisole (68). The in-depth mechanistic studies confers deeper understanding of AC vs non-alternating constant potential (CPE) electrolysis (Scheme 14A),114 here enabled with the aid of the redox mediator NHPI (vide supra). Luo and Nguyen likewise devised the successful usage of AC electrolysis for the trifluoromethylation of heteroarenes 71 (Scheme 14B).115 Within their study, alternating current significantly enhanced the overall efficacy of the electrosynthetic transformation by the direct conversion of otherwise unstable intermediates in a confined space.

Scheme 14. Alternating Current (AC) Electrolysis.

Scheme 14

6. ENANTIOSELECTIVE ELECTROCHEMICAL TRANSFORMATIONS

Enantioselective electrochemical synthesis was envisioned for the synthesis of enantiomerically enriched, chiral compounds by electrochemical synthetic methods.37 Organic electrosynthesis offers the possibility to perform reactions under exceedingly mild reaction conditions such as low temperatures, which are typically required to achieve highly enantioselective transformations. The existing methods of asymmetric electrochemical synthesis relied on the participation of an external source of chirality, such as chiral auxiliaries, chiral catalysts, chiral reagents, chiral electrodes,116 chiral electrolytes,117 or chiral solvents.118 Very recently, Lin devised the enantioselective cupraelectro-catalyzed cyanophosphinoylation of alkenes 8 (Scheme 15A).119 By merging two distinct oxidative events, challenging enantioselective transformations of alkenes 8 with diphenylphosphine oxide 73 were made possible.120 The stereoselectivities were well-controlled with the aid of chiral ligands for the envisioned heterodifunctionalization of alkenes.37d,121 The designed bis(oxazoline) (sBOX)122 was found to be the key to success for the enantio-determining C–CN formation. The same copper catalyst was thereafter found to be essential for the chiral induction within a dual electrocatalytic approach (vide supra).53 As to reductive transformations, Reisman established enantioselective electrochemical nickel-catalyzed cross-couplings of alkenyl bromides 75 and benzyl chlorides 62 by employing a chiral bis(oxazoline) ligand L1 (Scheme 15B).123 Likewise, Mei reported the nickel-catalyzed electroreductive enantioselective homocoupling of aryl bromides 77 to furnish axially chiral biaryl compounds 78 (Scheme 15C).124 The high levels of asymmetric induction relied on the newly designed chiral pyridine-oxazoline ligands L2. Compared with previously proposed nickel(II) intermediates, Mei suggested a nickel(0/II/I) pathway for this transformation. To avoid elements of prefunctionalization, Ackermann developed asymmetric palladaelectro-catalyzed C–H olefinations with high position-, diastereo-, and enantio-control under mild reaction conditions (Scheme 15D).125 A transient directing group strategy was utilized leading to the atroposelective organometallic C–H activation. Here, the authors were able to assemble axially chiral biaryl scaffolds 80 by organic electrocatalysis.

Scheme 15. Enantioselective Electrocatalytic Transformations.

Scheme 15

7. OUTLOOK

Recent years have witnessed a remarkable renaissance of organic electrosynthesis. The resurgence of molecular electrochemistry was spurred by key conceptual developments, which are considerably environmentally benign yet economically attractive transformations. Particularly, the merger of electrosynthesis with transition-metal catalysts within synergistic catalysis regimes has set the stage for novel functionalizations. These findings had partially involved redox mediators and were guided by model mechanistic insights. Thereby, innovative strategies for the full control of chemo-, position-, diastereo-, and even enantio-control were identified, also enabling asymmetric electrocatalysis. In spite of these indisputable advances, key challenges remain to be overcome to render organic electrosynthesis the central position it fully deserves (Figure 5). While reductive cross-couplings are efficient tools for C–C formation, electroreductive reactions have largely incited by the need for sacrificial electrodes.126 However, there were advances made to bypass this limitation by paired electrolysis or alternating current.113

Figure 5.

Figure 5

Selected perspectives of the synthetic electrochemistry.

Paired electrolysis refers to parallel electrochemical processes where two desirable half reactions are performed simultaneously. The combination of two half-reactions that consequently give the desired product(s), thereby maximizing the energy efficiency. Paired electrolysis can be considered as the gold standard for industrial settings due to the optimal overall usage of applied energy for two simultaneous desirable processes on both electrodes; this could be either through parallel, sequential divergent, or convergent processes.15c,127 The approach of paired electrolysis was recently exploited by Waldvogel and Morandi for reversible halide-shuttle reactions.128 The innovative concept was particularly powerful for electrochemical vicinal dihalogenations of various olefins, using simple dihalogenated solvents as the halide source.

The synergism of electrochemistry and photochemistry has sparked significant current interest with numerous electro-photocatalytic reactions being disclosed to achieve extreme redox potentials for otherwise difficult molecular transformations.27,34a,129 The quest for a well-designed electrophotocatalyst for an allied co-operation with electrocatalysis as well as inherent scalability remains a considerable hurdle.

Electrochemical conversion of small molecules has indeed thrived within recent years; however, there have been limited efforts to further utilize them in organic synthesis.130 Introducing CO2 as a C1 synthon provides a greener alternative for organic synthesis in conjunction with electrochemical methods. The global outlook for chemical synthesis has been reaching out far for more renewable methodologies and chemical sources, and we expect more advances to be made in this direction toward efficient conversion of CO2 and N2 to useful synthetic materials.

Recent developments in asymmetric electrochemical transformations have led to synthetic applications in the organic synthesis with good functional group tolerance and high enantioselectivity.37

The rapid development of electrosynthesis strongly depends on detailed mechanistic insights into electroorganic reactions. In this Outlook, we detailed how the oxidation-induced reductive elimination progress operates with inter alia cobalt, nickel, ruthenium, and rhodium cyclometalated intermediates by electrochemical analyses. Thus, this technique is anticipated to continue to unravel the mechanism of electrosynthesis by characterization of short-lived intermediates and provide further insights into fundamental single electron transfer (SET) processes.80a,131

In summary, electrosynthesis offers a green platform with the prospect for sustainable molecular synthesis for peptide chemistry,132 biochemistry,133 and material sciences.134 Furthermore, the ideal levels of resource economy of electrochemistry hold great potential for large-scale industrial manufacturing.

Acknowledgments

Generous support by the DFG (Gottfried-Wilhelm-Leibniz award) and the CSC (Ph.D. fellowship to C.Z.) is gratefully acknowledged.

Author Contributions

§ C.Z. and N.W. J.A. contributed equally.

The authors declare no competing financial interest.

References

  1. a Börgel J.; Ritter T. Late-Stage Functionalization. Chem. 2020, 6, 1877–1887. 10.1016/j.chempr.2020.07.007. [DOI] [Google Scholar]; b Wu F.; Zhang J.; Song F.; Wang S.; Guo H.; Wei Q.; Dai H.; Chen X.; Xia X.; Liu X.; Zhang L.; Yu J. Q.; Lei X. Chrysomycin A Derivatives for the Treatment of Multi-Drug-Resistant Tuberculosis. ACS Cent. Sci. 2020, 6, 928–938. 10.1021/acscentsci.0c00122. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Wang W.; Lorion M. M.; Shah J.; Kapdi A. R.; Ackermann L. Late-Stage Peptide Diversification by Position-Selective C–H Activation. Angew. Chem., Int. Ed. 2018, 57, 14700–14717. 10.1002/anie.201806250. [DOI] [PubMed] [Google Scholar]; d Cernak T.; Dykstra K. D.; Tyagarajan S.; Vachal P.; Krska S. W. The medicinal chemist’s toolbox for late stage functionalization of drug-like molecules. Chem. Soc. Rev. 2016, 45, 546–576. 10.1039/C5CS00628G. [DOI] [PubMed] [Google Scholar]; e Pouliot J.-R. m.; Grenier F. O.; Blaskovits J. T.; Beaupré S.; Leclerc M. Direct (hetero)arylation polymerization: simplicity for conjugated polymer synthesis. Chem. Rev. 2016, 116, 14225–14274. 10.1021/acs.chemrev.6b00498. [DOI] [PubMed] [Google Scholar]; f Noisier A. F.; Brimble M. A. C–H functionalization in the synthesis of amino acids and peptides. Chem. Rev. 2014, 114, 8775–8806. 10.1021/cr500200x. [DOI] [PubMed] [Google Scholar]; g Schipper D. J.; Fagnou K. Direct arylation as a synthetic tool for the synthesis of thiophene-based organic electronic materials. Chem. Mater. 2011, 23, 1594–1600. 10.1021/cm103483q. [DOI] [Google Scholar]
  2. a Guillemard L.; Wencel-Delord J. When metal-catalyzed C–H functionalization meets visible-light photocatalysis. Beilstein J. Org. Chem. 2020, 16, 1754–1804. 10.3762/bjoc.16.147. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Marzo L.; Pagire S. K.; Reiser O.; König B. Visible-Light Photocatalysis: Does It Make a Difference in Organic Synthesis?. Angew. Chem., Int. Ed. 2018, 57, 10034–10072. 10.1002/anie.201709766. [DOI] [PubMed] [Google Scholar]; c Skubi K. L.; Blum T. R.; Yoon T. P. Dual catalysis strategies in photochemical synthesis. Chem. Rev. 2016, 116, 10035–10074. 10.1021/acs.chemrev.6b00018. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Kärkäs M. D.; Porco J. A. Jr; Stephenson C. R. Photochemical approaches to complex chemotypes: applications in natural product synthesis. Chem. Rev. 2016, 116, 9683–9747. 10.1021/acs.chemrev.5b00760. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Brimioulle R.; Lenhart D.; Maturi M. M.; Bach T. Enantioselective Catalysis of Photochemical Reactions. Angew. Chem., Int. Ed. 2015, 54, 3872–3890. 10.1002/anie.201411409. [DOI] [PubMed] [Google Scholar]; f Prier C. K.; Rankic D. A.; MacMillan D. W. Visible light photoredox catalysis with transition metal complexes: applications in organic synthesis. Chem. Rev. 2013, 113, 5322–5363. 10.1021/cr300503r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. a de Almeida A. F.; Moreira R.; Rodrigues T. Synthetic organic chemistry driven by artificial intelligence. Nat. Rev. Chem. 2019, 3, 589–604. 10.1038/s41570-019-0124-0. [DOI] [Google Scholar]; b Zhavoronkov A. Artificial Intelligence for Drug Discovery, Biomarker Development, and Generation of Novel Chemistry. Mol. Pharmaceutics 2018, 15, 4311–4313. 10.1021/acs.molpharmaceut.8b00930. [DOI] [PubMed] [Google Scholar]
  4. a de Oliveira P. F. M.; Torresi R. M.; Emmerling F.; Camargo P. H. C. Challenges and opportunities in the bottom-up mechanochemical synthesis of noble metal nanoparticles. J. Mater. Chem. A 2020, 8, 16114–16141. 10.1039/D0TA05183G. [DOI] [Google Scholar]; b Huskić I.; Lennox C. B.; Friščić T. Accelerated ageing reactions: towards simpler, solvent-free, low energy chemistry. Green Chem. 2020, 22, 5881–5901. 10.1039/D0GC02264K. [DOI] [Google Scholar]; c Bolm C.; Hernández J. G. Mechanochemistry of Gaseous Reactants. Angew. Chem., Int. Ed. 2019, 58, 3285–3299. 10.1002/anie.201810902. [DOI] [PubMed] [Google Scholar]
  5. a Santoro S.; Ferlin F.; Ackermann L.; Vaccaro L. C–H functionalization reactions under flow conditions. Chem. Soc. Rev. 2019, 48, 2767–2782. 10.1039/C8CS00211H. [DOI] [PubMed] [Google Scholar]; b Folgueiras-Amador A. A.; Wirth T. Perspectives in flow electrochemistry. J. Flow Chem. 2017, 7, 94–95. 10.1556/1846.2017.00020. [DOI] [Google Scholar]; c Plutschack M. B.; Pieber B. u.; Gilmore K.; Seeberger P. H. The hitchhiker’s guide to flow chemistry. Chem. Rev. 2017, 117, 11796–11893. 10.1021/acs.chemrev.7b00183. [DOI] [PubMed] [Google Scholar]; d Pletcher D.; Green R. A.; Brown R. C. Flow electrolysis cells for the synthetic organic chemistry laboratory. Chem. Rev. 2018, 118, 4573–4591. 10.1021/acs.chemrev.7b00360. [DOI] [PubMed] [Google Scholar]; e Gemoets H. P.; Su Y.; Shang M.; Hessel V.; Luque R.; Noel T. Liquid phase oxidation chemistry in continuous-flow microreactors. Chem. Soc. Rev. 2016, 45, 83–117. 10.1039/C5CS00447K. [DOI] [PubMed] [Google Scholar]; f Pastre J. C.; Browne D. L.; Ley S. V. Flow chemistry syntheses of natural products. Chem. Soc. Rev. 2013, 42, 8849–8869. 10.1039/c3cs60246j. [DOI] [PubMed] [Google Scholar]; g Noël T.; Buchwald S. L. Cross-coupling in flow. Chem. Soc. Rev. 2011, 40, 5010–5029. 10.1039/c1cs15075h. [DOI] [PubMed] [Google Scholar]; h Saaby S.; Knudsen K. R.; Ladlow M.; Ley S. V. The use of a continuous flow-reactor employing a mixed hydrogen–liquid flow stream for the efficient reduction of imines to amines. Chem. Commun. 2005, 2909–2911. 10.1039/b504854k. [DOI] [PubMed] [Google Scholar]
  6. a Little R. D. A Perspective on Organic Electrochemistry. J. Org. Chem. 2020, 85, 13375–13390. 10.1021/acs.joc.0c01408. [DOI] [PubMed] [Google Scholar]; b Horn E. J.; Rosen B. R.; Baran P. S. Synthetic Organic Electrochemistry: An Enabling and Innately Sustainable Method. ACS Cent. Sci. 2016, 2, 302–308. 10.1021/acscentsci.6b00091. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Hammerich O.; Speiser B.. Organic electrochemistry: revised and expanded, 5th ed.; CRC Press, 2015. [Google Scholar]
  7. a Pollok D.; Waldvogel S. R. Electro-organic synthesis – a 21st century technique. Chem. Sci. 2020, 11, 12386–12400. 10.1039/D0SC01848A. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Kawamata Y.; Baran P. S. Electrosynthesis: Sustainability Is Not Enough. Joule 2020, 4, 701–704. 10.1016/j.joule.2020.02.002. [DOI] [Google Scholar]; c Meyer T. H.; Choi I.; Tian C.; Ackermann L. Powering the Future: How Can Electrochemistry Make a Difference in Organic Synthesis?. Chem. 2020, 6, 2484–2496. 10.1016/j.chempr.2020.08.025. [DOI] [Google Scholar]
  8. Faraday M. Siebente Reihe von Experimental-Untersuchungen über Elektricität. Ann. Phys. 1834, 109, 481–520. 10.1002/andp.18341093102. [DOI] [Google Scholar]
  9. a Leech M. C.; Lam K. Electrosynthesis Using Carboxylic Acid Derivatives: New Tricks for Old Reactions. Acc. Chem. Res. 2020, 53, 121–134. 10.1021/acs.accounts.9b00586. [DOI] [PubMed] [Google Scholar]; b Schäfer H. J. Recent contributions of Kolbe electrolysis to organic synthesis. Top. Curr. Chem. 1990, 152, 91–151. 10.1007/BFb0034365. [DOI] [Google Scholar]; c Becking L.; Schäfer H. J. Pyrrolidines by intramolecular addition of Kolbe radicals generated from β-allylaminoalkanoates. Tetrahedron Lett. 1988, 29, 2797–2800. 10.1016/0040-4039(88)85212-2. [DOI] [Google Scholar]; d Kolbe H. Beobachtungen über die oxydirende Wirkung des Sauerstoffs, wenn derselbe mit Hülfe einer elektrischen Säule entwickelt wird. J. Prakt. Chem. 1847, 41, 137–139. 10.1002/prac.18470410118. [DOI] [Google Scholar]
  10. a Lingane J. J.; Swain C. G.; Fields M. Polarographically controlled syntheses, with particular reference to organic chemistry. J. Am. Chem. Soc. 1943, 65, 1348–1353. 10.1021/ja01247a024. [DOI] [Google Scholar]; b Hickling A. Studies in electrode polarisation. Part IV.—The automatic control of the potential of a working electrode. Trans. Faraday Soc. 1942, 38, 27–33. 10.1039/TF9423800027. [DOI] [Google Scholar]
  11. a Bard A. J.; Zoski C. G. Voltammetry Retrospective. Anal. Chem. 2000, 72, 346 A–352 A. 10.1021/ac002791t. [DOI] [PubMed] [Google Scholar]; b Heinze J. Cyclic Voltammetry—“Electrochemical Spectroscopy”. New Analytical Methods (25). Angew. Chem., Int. Ed. Engl. 1984, 23, 831–847. 10.1002/anie.198408313. [DOI] [Google Scholar]; c Randles J. E. A cathode ray polarograph. Part II.—The current-voltage curves. Trans. Faraday Soc. 1948, 44, 327–338. 10.1039/TF9484400327. [DOI] [Google Scholar]; d Heyrovský J. Electrolysis with a Dropping Mercury Cathode. Philos. Mag. 1923, 45, 303–315. 10.1080/14786442308634117. [DOI] [Google Scholar]; e Heyrovský J. Elektrolysa se rtufovou kapkovou kathodou. Chem. Listy 1922, 16, 256–264. [DOI] [PubMed] [Google Scholar]
  12. Simons J. The electrochemical process for the production of fluorocarbons. J. Electrochem. Soc. 1949, 95, 47–67. 10.1149/1.2776733. [DOI] [Google Scholar]
  13. a Baizer M. M. Recent developments in organic synthesis by electrolysis. Tetrahedron 1984, 40, 935–969. 10.1016/S0040-4020(01)91232-3. [DOI] [Google Scholar]; b Baizer M. M. Electrolytic Reductive Coupling: I. Acrylonitrile. J. Electrochem. Soc. 1964, 111, 215. 10.1149/1.2426086. [DOI] [Google Scholar]
  14. a Leech M. C.; Garcia A. D.; Petti A.; Dobbs A. P.; Lam K. Organic electrosynthesis: from academia to industry. React. Chem. Eng. 2020, 5, 977–990. 10.1039/D0RE00064G. [DOI] [Google Scholar]; b Cardoso D. S. P.; Šljukić B.; Santos D. M. F.; Sequeira C. A. C. Organic Electrosynthesis: From Laboratorial Practice to Industrial Applications. Org. Process Res. Dev. 2017, 21, 1213–1226. 10.1021/acs.oprd.7b00004. [DOI] [Google Scholar]
  15. a Llorente M. J.; Nguyen B. H.; Kubiak C. P.; Moeller K. D. Paired Electrolysis in the Simultaneous Production of Synthetic Intermediates and Substrates. J. Am. Chem. Soc. 2016, 138, 15110–15113. 10.1021/jacs.6b08667. [DOI] [PubMed] [Google Scholar]; b Ibanez J. G.; Frontana-Uribe B. A.; Vasquez-Medrano R. Paired electrochemical processes: overview, systematization, selection criteria, design strategies, and projection. J. Mex. Chem. Soc. 2016, 60, 247–260. [Google Scholar]; c Botte G. G. Electrochemical manufacturing in the chemical industry. Electrochem. Soc. Interface 2014, 23, 49–55. 10.1149/2.F04143if. [DOI] [Google Scholar]; d Paddon C. A.; Atobe M.; Fuchigami T.; He P.; Watts P.; Haswell S. J.; Pritchard G. J.; Bull S. D.; Marken F. Towards paired and coupled electrode reactions for clean organic microreactor electrosyntheses. J. Appl. Electrochem. 2006, 36, 617–634. 10.1007/s10800-006-9122-2. [DOI] [Google Scholar]
  16. Yoshida J.-i.; Murata T.; Isoe S. Electrochemical oxidation of organosilicon compounds I. Oxidative cleavage of carbon-silicon bond in allylsilanes and benzylsilanes. Tetrahedron Lett. 1986, 27, 3373–3376. 10.1016/S0040-4039(00)84799-1. [DOI] [Google Scholar]
  17. Le Blanc M. Über einen Versuch zur Demonstration des elektrolytischen Lösungsdruckes. Z. Elektrochem. Angew. Phys. Chem. 1900, 7, 287–290. 10.1002/bbpc.19000072003. [DOI] [Google Scholar]
  18. a Steckhan E. Organic syntheses with electrochemically regenerable redox systems. Top. Curr. Chem. 1987, 142, 1–69. 10.1007/3-540-17871-6_11. [DOI] [Google Scholar]; b Steckhan E. Indirect Electroorganic Syntheses—A Modern Chapter of Organic Electrochemistry. Angew. Chem., Int. Ed. Engl. 1986, 25, 683–701. 10.1002/anie.198606831. [DOI] [Google Scholar]
  19. a Francke R.; Little R. D. Redox catalysis in organic electrosynthesis: basic principles and recent developments. Chem. Soc. Rev. 2014, 43, 2492–2521. 10.1039/c3cs60464k. [DOI] [PubMed] [Google Scholar]; b Little R. D.; Schwaebe M. K. Reductive cyclizations at the cathode. Top. Curr. Chem. 1997, 185, 1–48. 10.1007/3-540-61454-0_69. [DOI] [Google Scholar]; c Sowell C. G.; Wolin R. L.; Little R. D. Electroreductive cyclization reactions. Stereoselection, creation of quaternary centers in bicyclic frameworks, and a formal total synthesis of quadrone. Tetrahedron Lett. 1990, 31, 485–488. 10.1016/0040-4039(90)87014-Q. [DOI] [Google Scholar]; d Little R. D.; Fox D. P.; Van Hijfte L.; Dannecker R.; Sowell G.; Wolin R. L.; Moens L.; Baizer M. M. Electroreductive cyclization. Ketones and aldehydes tethered to α,β-unsaturated esters (nitriles). Fundamental investigations. J. Org. Chem. 1988, 53, 2287–2294. 10.1021/jo00245a029. [DOI] [Google Scholar]
  20. a Schäfer H. J. Contributions of organic electrosynthesis to green chemistry. C. R. Chim. 2011, 14, 745–765. 10.1016/j.crci.2011.01.002. [DOI] [Google Scholar]; b Schäfer H. J. Anodic and Cathodic C–C-Bond Formation. Angew. Chem., Int. Ed. Engl. 1981, 20, 911–934. 10.1002/anie.198109111. [DOI] [Google Scholar]
  21. a Lund H. A century of organic electrochemistry. J. Electrochem. Soc. 2002, 149, S21–S33. 10.1149/1.1462037. [DOI] [Google Scholar]; b Iversen P. E.; Lund H. Electrolytic generation of strong bases I. Wittig reaction. Tetrahedron Lett. 1969, 10, 3523–3524. 10.1016/S0040-4039(01)88438-0. [DOI] [Google Scholar]; c Lund H.; Lunde P. Quaternization Reactions. II. Pyridazines. Acta Chem. Scand. 1967, 21, 1067–1080. 10.3891/acta.chem.scand.21-1067. [DOI] [Google Scholar]
  22. a Gieshoff T.; Kehl A.; Schollmeyer D.; Moeller K. D.; Waldvogel S. R. Insights into the mechanism of anodic N–N bond formation by dehydrogenative coupling. J. Am. Chem. Soc. 2017, 139, 12317–12324. 10.1021/jacs.7b07488. [DOI] [PubMed] [Google Scholar]; b Xu H.-C.; Campbell J. M.; Moeller K. D. Cyclization reactions of anode-generated amidyl radicals. J. Org. Chem. 2014, 79, 379–391. 10.1021/jo402623r. [DOI] [PubMed] [Google Scholar]; c Moeller K. D. Synthetic applications of anodic electrochemistry. Tetrahedron 2000, 56, 9527–9554. 10.1016/S0040-4020(00)00840-1. [DOI] [Google Scholar]
  23. a Amatore C.; Cammoun C.; Jutand A. Pd(OAc)2/p-Benzoquinone-Catalyzed Anaerobic Electrooxidative Homocoupling of Arylboronic Acids, Arylboronates and Aryltrifluoroborates in DMF and/or Water. Eur. J. Org. Chem. 2008, 4567–4570. 10.1002/ejoc.200800631. [DOI] [Google Scholar]
  24. Jutand A. Contribution of electrochemistry to organometallic catalysis. Chem. Rev. 2008, 108, 2300–2347. 10.1021/cr068072h. [DOI] [PubMed] [Google Scholar]
  25. a Yoshida J.-i.; Kataoka K.; Horcajada R.; Nagaki A. Modern strategies in electroorganic synthesis. Chem. Rev. 2008, 108, 2265–2299. 10.1021/cr0680843. [DOI] [PubMed] [Google Scholar]; b Yoshida J.-i.; Suga S.; Suzuki S.; Kinomura N.; Yamamoto A.; Fujiwara K. Direct Oxidative Carbon–Carbon Bond Formation Using the “Cation Pool” Method. 1. Generation of Iminium Cation Pools and Their Reaction with Carbon Nucleophiles. J. Am. Chem. Soc. 1999, 121, 9546–9549. 10.1021/ja9920112. [DOI] [Google Scholar]; The list of names is not intended to be comprehensive and possible omissions are unintended.
  26. a Gandeepan P.; Finger L. H.; Meyer T. H.; Ackermann L. 3d metallaelectrocatalysis for resource economical syntheses. Chem. Soc. Rev. 2020, 49, 4254–4272. 10.1039/D0CS00149J. [DOI] [PubMed] [Google Scholar]; b Jiao K.-J.; Xing Y.-K.; Yang Q.-L.; Qiu H.; Mei T.-S. Site-Selective C–H Functionalization via Synergistic Use of Electrochemistry and Transition Metal Catalysis. Acc. Chem. Res. 2020, 53, 300–310. 10.1021/acs.accounts.9b00603. [DOI] [PubMed] [Google Scholar]; c Röckl J. L.; Pollok D.; Franke R.; Waldvogel S. R. A Decade of Electrochemical Dehydrogenative C–C-Coupling of Aryls. Acc. Chem. Res. 2020, 53, 45–61. 10.1021/acs.accounts.9b00511. [DOI] [PubMed] [Google Scholar]; d Meyer T. H.; Finger L. H.; Gandeepan P.; Ackermann L. Resource Economy by Metallaelectrocatalysis: Merging Electrochemistry and C–H Activation. Trends Chem. 2019, 1, 63–76. 10.1016/j.trechm.2019.01.011. [DOI] [Google Scholar]; e Yuan Y.; Lei A. Electrochemical oxidative cross-coupling with hydrogen evolution reactions. Acc. Chem. Res. 2019, 52, 3309–3324. 10.1021/acs.accounts.9b00512. [DOI] [PubMed] [Google Scholar]; f Xiong P.; Xu H.-C. Chemistry with Electrochemically Generated N-Centered Radicals. Acc. Chem. Res. 2019, 52, 3339–3350. 10.1021/acs.accounts.9b00472. [DOI] [PubMed] [Google Scholar]; g Elsherbini M.; Wirth T. Electroorganic Synthesis under Flow Conditions. Acc. Chem. Res. 2019, 52, 3287–3296. 10.1021/acs.accounts.9b00497. [DOI] [PubMed] [Google Scholar]; h Shrestha A.; Lee M.; Dunn A. L.; Sanford M. S. Palladium-catalyzed C–H bond acetoxylation via electrochemical oxidation. Org. Lett. 2018, 20, 204–207. 10.1021/acs.orglett.7b03559. [DOI] [PMC free article] [PubMed] [Google Scholar]; i Tang S.; Liu Y.; Lei A. Electrochemical Oxidative Cross-coupling with Hydrogen Evolution: A Green and Sustainable Way for Bond Formation. Chem. 2018, 4, 27–45. 10.1016/j.chempr.2017.10.001. [DOI] [Google Scholar]; j 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. 10.1021/ja5013323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Liu J.; Lu L.; Wood D.; Lin S. New Redox Strategies in Organic Synthesis by Means of Electrochemistry and Photochemistry. ACS Cent. Sci. 2020, 6, 1317–1340. 10.1021/acscentsci.0c00549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Samanta R. C.; Meyer T. H.; Siewert I.; Ackermann L. Renewable resources for sustainable metallaelectro-catalysed C–H activation. Chem. Sci. 2020, 11, 8657–8670. 10.1039/D0SC03578E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. a Tian C.; Meyer T. H.; Stangier M.; Dhawa U.; Rauch K.; Finger L. H.; Ackermann L. Cobaltaelectro-catalyzed C–H activation for resource-economical molecular syntheses. Nat. Protoc. 2020, 15, 1760–1774. 10.1038/s41596-020-0306-8. [DOI] [PubMed] [Google Scholar]; b Yan M.; Kawamata Y.; Baran P. S. Synthetic Organic Electrochemistry: Calling All Engineers. Angew. Chem., Int. Ed. 2018, 57, 4149–4155. 10.1002/anie.201707584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. a 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. 10.1021/acs.accounts.9b00539. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Waldvogel S. R.; Lips S.; Selt M.; Riehl B.; Kampf C. J. Electrochemical Arylation Reaction. Chem. Rev. 2018, 118, 6706–6765. 10.1021/acs.chemrev.8b00233. [DOI] [PubMed] [Google Scholar]; c 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. 10.1039/b512308a. [DOI] [PubMed] [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. 10.1021/acs.chemrev.7b00397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. a Waldvogel S. R.; Janza B. Renaissance of electrosynthetic methods for the construction of complex molecules. Angew. Chem., Int. Ed. 2014, 53, 7122–7123. 10.1002/anie.201405082. [DOI] [PubMed] [Google Scholar]; b Mihelcic J.; Moeller K. D. Anodic cyclization reactions: The total synthesis of alliacol A. J. Am. Chem. Soc. 2003, 125, 36–37. 10.1021/ja029064v. [DOI] [PubMed] [Google Scholar]
  33. Lorion M. M.; Maindan K.; Kapdi A. R.; Ackermann L. Heteromultimetallic catalysis for sustainable organic syntheses. Chem. Soc. Rev. 2017, 46, 7399–7420. 10.1039/C6CS00787B. [DOI] [PubMed] [Google Scholar]
  34. a Yu Y.; Guo P.; Zhong J.-S.; Yuan Y.; Ye K.-Y. Merging photochemistry with electrochemistry in organic synthesis. Org. Chem. Front. 2020, 7, 131–135. 10.1039/C9QO01193E. [DOI] [Google Scholar]; b Capaldo L.; Quadri L. L.; Ravelli D. Merging Photocatalysis with Electrochemistry: The Dawn of a new Alliance in Organic Synthesis. Angew. Chem., Int. Ed. 2019, 58, 17508–17510. 10.1002/anie.201910348. [DOI] [PubMed] [Google Scholar]; c Zhang L.; Liardet L.; Luo J.; Ren D.; Grätzel M.; Hu X. Photoelectrocatalytic arene C–H amination. Nat. Catal. 2019, 2, 366–373. 10.1038/s41929-019-0231-9. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Wang F.; Stahl S. S. Merging Photochemistry with Electrochemistry: Functional-Group Tolerant Electrochemical Amination of C(sp3)–H Bonds. Angew. Chem., Int. Ed. 2019, 58, 6385–6390. 10.1002/anie.201813960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Ackermann L. Metalla-electrocatalyzed C–H Activation by Earth-Abundant 3d Metals and Beyond. Acc. Chem. Res. 2020, 53, 84–104. 10.1021/acs.accounts.9b00510. [DOI] [PubMed] [Google Scholar]
  36. a Manabe S.; Wong C. M.; Sevov C. S. Direct and Scalable Electroreduction of Triphenylphosphine Oxide to Triphenylphosphine. J. Am. Chem. Soc. 2020, 142, 3024–3031. 10.1021/jacs.9b12112. [DOI] [PubMed] [Google Scholar]; b Elias J. S.; Costentin C.; Nocera D. G. Direct Electrochemical P(V) to P(III) Reduction of Phosphine Oxide Facilitated by Triaryl Borates. J. Am. Chem. Soc. 2018, 140, 13711–13718. 10.1021/jacs.8b07149. [DOI] [PubMed] [Google Scholar]; c Heinze J.Cathodic Reactions of Hydrocarbons, Olefins, and Aromatic Compounds. Encyclopedia of Electrochemistry; Wiley-VCH, 2007; pp 95–119. [Google Scholar]; d Eberson L.; Schäfer H. J.. Organic Electrochemistry; Springer, 1971; Vol. 34, pp 1–182. [Google Scholar]
  37. a Chang X.; Zhang Q.; Guo C. Asymmetric Electrochemical Transformations. Angew. Chem., Int. Ed. 2020, 59, 12612–12622. 10.1002/anie.202000016. [DOI] [PubMed] [Google Scholar]; b Lin Q.; Luo S. Tailoring radicals by asymmetric electrochemical catalysis. Org. Chem. Front. 2020, 7, 2997–3000. 10.1039/D0QO00803F. [DOI] [Google Scholar]; c Phillips A. M. F.; Pombeiro A. J. L. Electrochemical asymmetric synthesis of biologically active substances. Org. Biomol. Chem. 2020, 18, 7026–7055. 10.1039/D0OB01425G. [DOI] [PubMed] [Google Scholar]; d Lin Q.; Li L.; Luo S. Asymmetric electrochemical catalysis. Chem. - Eur. J. 2019, 25, 10033–10044. 10.1002/chem.201901284. [DOI] [PubMed] [Google Scholar]; e Ghosh M.; Shinde V. S.; Rueping M. A review of asymmetric synthetic organic electrochemistry and electrocatalysis: concepts, applications, recent developments and future directions. Beilstein J. Org. Chem. 2019, 15, 2710–2746. 10.3762/bjoc.15.264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. a Kärkäs M. D. Electrochemical strategies for C–H functionalization and C–N bond formation. Chem. Soc. Rev. 2018, 47, 5786–5865. 10.1039/C7CS00619E. [DOI] [PubMed] [Google Scholar]; b Wang H.; Gao X.; Lv Z.; Abdelilah T.; Lei A. Recent advances in oxidative R1–H/R2–H cross-coupling with hydrogen evolution via photo-/electrochemistry: focus review. Chem. Rev. 2019, 119, 6769–6787. 10.1021/acs.chemrev.9b00045. [DOI] [PubMed] [Google Scholar]; c Wiebe A.; Gieshoff T.; Möhle S.; Rodrigo E.; Zirbes M.; Waldvogel S. R. Electrifying Organic Synthesis. Angew. Chem., Int. Ed. 2018, 57, 5594–5619. 10.1002/anie.201711060. [DOI] [PMC free article] [PubMed] [Google Scholar]; and references cited therein.
  39. a Costentin C.; Savéant J.-M. Concepts and tools for mechanism and selectivity analysis in synthetic organic electrochemistry. Proc. Natl. Acad. Sci. U. S. A. 2019, 116, 11147–11152. 10.1073/pnas.1904439116. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Oh Y.; Hu X. Organic molecules as mediators and catalysts for photocatalytic and electrocatalytic CO2 reduction. Chem. Soc. Rev. 2013, 42, 2253–2261. 10.1039/C2CS35276A. [DOI] [PubMed] [Google Scholar]; c Amatore C.; Cammoun C.; Jutand A. Electrochemical Recycling of Benzoquinone in the Pd/Benzoquinone-Catalyzed Heck-Type Reactions from Arenes. Adv. Synth. Catal. 2007, 349, 292–296. 10.1002/adsc.200600389. [DOI] [Google Scholar]
  40. a Nutting J. E.; Rafiee M.; Stahl S. S. Tetramethylpiperidine N-Oxyl (TEMPO), Phthalimide N-Oxyl (PINO), and Related N-Oxyl Species: Electrochemical Properties and Their Use in Electrocatalytic Reactions. Chem. Rev. 2018, 118, 4834–4885. 10.1021/acs.chemrev.7b00763. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Ogibin Y. N.; Elinson M. N.; Nikishin G. I. Mediator oxidation systems in organic electrosynthesis. Russ. Chem. Rev. 2009, 78, 89–140. 10.1070/RC2009v078n02ABEH003886. [DOI] [Google Scholar]
  41. a Shono T.; Ikeda A. Electroorganic chemistry. X. Anodic allylic substitution. J. Am. Chem. Soc. 1972, 94, 7892–7898. 10.1021/ja00777a036. [DOI] [Google Scholar]; b Shono T.; Kosaka T. Organic synthesis by electrolysis III anodic allylic substitution. Tetrahedron Lett. 1968, 9, 6207–6208. 10.1016/S0040-4039(00)70831-8. [DOI] [Google Scholar]
  42. a Ueda C.; Noyama M.; Ohmori H.; Masui M. Reactivity of phthalimide-N-oxyl: a kinetic study. Chem. Pharm. Bull. 1987, 35, 1372–1377. 10.1248/cpb.35.1372. [DOI] [Google Scholar]; b Masui M.; Hosomi K.; Tsuchida K.; Ozaki S. Electrochemical oxidation of olefins using N-hydroxyphthalimide as a mediator. Chem. Pharm. Bull. 1985, 33, 4798–4802. 10.1248/cpb.33.4798. [DOI] [Google Scholar]; c Masui M.; Hara S.; Ueshima T.; Kawaguchi T.; Ozaki S. Anodic oxidation of compounds having benzylic or allylic carbon and α-carbon to hetero atom using N-hydroxyphthalimide as a mediator. Chem. Pharm. Bull. 1983, 31, 4209–4211. 10.1248/cpb.31.4209. [DOI] [Google Scholar]
  43. a Horn E. J.; Rosen B. R.; Chen Y.; Tang J.; Chen K.; Eastgate M. D.; Baran P. S. Scalable and sustainable electrochemical allylic C–H oxidation. Nature 2016, 533, 77–81. 10.1038/nature17431. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Cai Y.; Koshino N.; Saha B.; Espenson J. H. Kinetics of self-decomposition and hydrogen atom transfer reactions of substituted phthalimide N-oxyl radicals in acetic acid. J. Org. Chem. 2005, 70, 238–243. 10.1021/jo048418t. [DOI] [PubMed] [Google Scholar]
  44. Kawamata Y.; Yan M.; Liu Z.; Bao D.-H.; Chen J.; Starr J. T.; Baran P. S. Scalable, Electrochemical Oxidation of Unactivated C–H Bonds. J. Am. Chem. Soc. 2017, 139, 7448–7451. 10.1021/jacs.7b03539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Wang F.; Rafiee M.; Stahl S. S. Electrochemical Functional-Group-Tolerant Shono-type Oxidation of Cyclic Carbamates Enabled by Aminoxyl Mediators. Angew. Chem., Int. Ed. 2018, 57, 6686–6690. 10.1002/anie.201803539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Qiu Y.; Stangier M.; Meyer T. H.; Oliveira J. C. A.; Ackermann L. Iridium-Catalyzed Electrooxidative C–H Activation by Chemoselective Redox-Catalyst Cooperation. Angew. Chem., Int. Ed. 2018, 57, 14179–14183. 10.1002/anie.201809611. [DOI] [PubMed] [Google Scholar]
  47. Yang Q.-L.; Xing Y.-K.; Wang X.-Y.; Ma H.-X.; Weng X.-J.; Yang X.; Guo H.-M.; Mei T.-S. Electrochemistry-Enabled Ir-Catalyzed Vinylic C–H Functionalization. J. Am. Chem. Soc. 2019, 141, 18970–18976. 10.1021/jacs.9b11915. [DOI] [PubMed] [Google Scholar]
  48. Francke R.; Little R. D. Electrons and Holes as Catalysts in Organic Electrosynthesis. ChemElectroChem 2019, 6, 4373–4382. 10.1002/celc.201900432. [DOI] [Google Scholar]
  49. Heard D. M.; Lennox A. J. Electrode materials in modern organic electrochemistry. Angew. Chem., Int. Ed. 2020, 59, 18866–18884. 10.1002/anie.202005745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. a Quideau S.; Pouységu L.; Deffieux D. Chemical and electrochemical oxidative activation of arenol derivatives for carbon-carbon bond formation. Curr. Org. Chem. 2004, 8, 113–148. 10.2174/1385272043486016. [DOI] [Google Scholar]; b Yamamura S.; Nishiyama S. Anodic Oxidation of Phenols Towards the Synthesis of Bioactive Natural Products. Synlett 2002, 533–543. 10.1055/s-2002-22696. [DOI] [Google Scholar]; c Eickhoff H.; Jung G.; Rieker A. Oxidative phenol coupling—tyrosine dimers and libraries containing tyrosyl peptide dimers. Tetrahedron 2001, 57, 353–364. 10.1016/S0040-4020(00)00942-X. [DOI] [Google Scholar]; d Mori K.; Takahashi M.; Yamamura S.; Nishiyama S. Anodic oxidation of monohalogenated phenols. Tetrahedron 2001, 57, 5527–5532. 10.1016/S0040-4020(01)00478-1. [DOI] [Google Scholar]; e Nilsson A.; Ahlberg E. Electrolytically Initiated Selective Aliphatic Hydrogen Exchange in 4-Isopropylanisole. Acta Chem. Scand. 1988, 42, 278–383. [Google Scholar]; f Iguchi M.; Nishiyama A.; Terada Y.; Yamamura S. Anodic Oxidation of 4-Allyl-2-Methoxyphenol Syntheses of Demethoxyasatone And Demethoxyisoasatone. Chem. Lett. 1978, 7, 451–454. 10.1246/cl.1978.451. [DOI] [Google Scholar]
  51. a Lips S.; Waldvogel S. R. Use of Boron-Doped Diamond Electrodes in Electro-Organic Synthesis. ChemElectroChem 2019, 6, 1649–1660. 10.1002/celc.201801620. [DOI] [Google Scholar]; b Malkowsky I. M.; Griesbach U.; Pütter H.; Waldvogel S. R. Unexpected Highly Chemoselective Anodic ortho-Coupling Reaction of 2,4-Dimethylphenol on Boron-Doped Diamond Electrodes. Eur. J. Org. Chem. 2006, 4569–4572. 10.1002/ejoc.200600466. [DOI] [Google Scholar]; c Kirste A.; Schnakenburg G.; Stecker F.; Fischer A.; Waldvogel S. R. Anodic Phenol–Arene Cross-Coupling Reaction on Boron-Doped Diamond Electrodes. Angew. Chem., Int. Ed. 2010, 49, 971–975. 10.1002/anie.200904763. [DOI] [PubMed] [Google Scholar]; d Kirste A.; Nieger M.; Malkowsky I. M.; Stecker F.; Fischer A.; Waldvogel S. R. ortho-Selective Phenol-Coupling Reaction by Anodic Treatment on Boron-Doped Diamond Electrode Using Fluorinated Alcohols. Chem. - Eur. J. 2009, 15, 2273–2277. 10.1002/chem.200802556. [DOI] [PubMed] [Google Scholar]
  52. Siu J. C.; Fu N.; Lin S. Catalyzing Electrosynthesis: A Homogeneous Electrocatalytic Approach to Reaction Discovery. Acc. Chem. Res. 2020, 53, 547–560. 10.1021/acs.accounts.9b00529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Song L.; Fu N.; Ernst B. G.; Lee W. H.; Frederick M. O.; Di Stasio R. A. Jr.; Lin S. Dual electrocatalysis enables enantioselective hydrocyanation of conjugated alkenes. Nat. Chem. 2020, 12, 747–754. 10.1038/s41557-020-0469-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. a Crossley S. W.; Barabe F.; Shenvi R. A. Simple, chemoselective, catalytic olefin isomerization. J. Am. Chem. Soc. 2014, 136, 16788–16791. 10.1021/ja5105602. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Li G.; Han A.; Pulling M. E.; Estes D. P.; Norton J. R. Evidence for formation of a Co–H bond from (H2O)2Co(dmgBF2)2 under H2: application to radical cyclizations. J. Am. Chem. Soc. 2012, 134, 14662–14665. 10.1021/ja306037w. [DOI] [PubMed] [Google Scholar]
  55. Wang F.; Chen P.; Liu G. Copper-catalyzed radical relay for asymmetric radical transformations. Acc. Chem. Res. 2018, 51, 2036–2046. 10.1021/acs.accounts.8b00265. [DOI] [PubMed] [Google Scholar]
  56. a Hari D. P.; Caramenti P.; Waser J. Cyclic Hypervalent Iodine Reagents: Enabling Tools for Bond Disconnection via Reactivity Umpolung. Acc. Chem. Res. 2018, 51, 3212–3225. 10.1021/acs.accounts.8b00468. [DOI] [PubMed] [Google Scholar]; b Yang F.; Rauch K.; Kettelhoit K.; Ackermann L. Aldehyde-Assisted Ruthenium(II)-Catalyzed C–H Oxygenations. Angew. Chem., Int. Ed. 2014, 53, 11285–11288. 10.1002/anie.201405647. [DOI] [PubMed] [Google Scholar]; c Liu W.; Ackermann L. Ortho- and Para-Selective Ruthenium-Catalyzed C(sp2)–H Oxygenations of Phenol Derivatives. Org. Lett. 2013, 15, 3484–3486. 10.1021/ol401535k. [DOI] [PubMed] [Google Scholar]; d Yang F.; Ackermann L. Ruthenium-Catalyzed C–H Oxygenation on Aryl Weinreb Amides. Org. Lett. 2013, 15, 718–720. 10.1021/ol303520h. [DOI] [PubMed] [Google Scholar]; e Thirunavukkarasu V. S.; Ackermann L. Ruthenium-Catalyzed C–H Bond Oxygenations with Weakly Coordinating Ketones. Org. Lett. 2012, 14, 6206–6209. 10.1021/ol302956s. [DOI] [PubMed] [Google Scholar]; f Thirunavukkarasu V. S.; Hubrich J.; Ackermann L. Ruthenium-Catalyzed Oxidative C(sp2)–H Bond Hydroxylation: Site-Selective C–O Bond Formation on Benzamides. Org. Lett. 2012, 14, 4210–4213. 10.1021/ol3018819. [DOI] [PubMed] [Google Scholar]; g Richardson R. D.; Wirth T. Hypervalent iodine goes catalytic. Angew. Chem., Int. Ed. 2006, 45, 4402–4404. 10.1002/anie.200601817. [DOI] [PubMed] [Google Scholar]; h Zhdankin V. V.; Stang P. J. Recent developments in the chemistry of polyvalent iodine compounds. Chem. Rev. 2002, 102, 2523–2584. 10.1021/cr010003+. [DOI] [PubMed] [Google Scholar]
  57. a Elsherbini M.; Winterson B.; Alharbi H.; Folgueiras-Amador A. A.; Génot C.; Wirth T. Continuous-Flow Electrochemical Generator of Hypervalent Iodine Reagents: Synthetic Applications. Angew. Chem., Int. Ed. 2019, 58, 9811–9815. 10.1002/anie.201904379. [DOI] [PubMed] [Google Scholar]; b Haupt J. D.; Berger M.; Waldvogel S. R. Electrochemical fluorocyclization of N-allylcarboxamides to 2-oxazolines by hypervalent iodine mediator. Org. Lett. 2019, 21, 242–245. 10.1021/acs.orglett.8b03682. [DOI] [PubMed] [Google Scholar]; c Broese T.; Francke R. Electrosynthesis using a recyclable mediator–electrolyte system based on ionically tagged phenyl iodide and 1,1,1,3,3,3-hexafluoroisopropanol. Org. Lett. 2016, 18, 5896–5899. 10.1021/acs.orglett.6b02979. [DOI] [PubMed] [Google Scholar]; d Inoue K.; Ishikawa Y.; Nishiyama S. Synthesis of Tetrahydropyrroloiminoquinone Alkaloids Based on Electrochemically Generated Hypervalent Iodine Oxidative Cyclization. Org. Lett. 2010, 12, 436–439. 10.1021/ol902566p. [DOI] [PubMed] [Google Scholar]
  58. Massignan L.; Tan X.; Meyer T. H.; Kuniyil R.; Messinis A. M.; Ackermann L. C–H Oxygenation Reactions Enabled by Dual Catalysis with Electrogenerated Hypervalent Iodine Species and Ruthenium Complexes. Angew. Chem., Int. Ed. 2020, 59, 3184–3189. 10.1002/anie.201914226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Bu Q.; Kuniyil R.; Shen Z.; Gońka E.; Ackermann L. Insights into Ruthenium(II/IV)-Catalyzed Distal C–H Oxygenation by Weak Coordination. Chem. - Eur. J. 2020, 26, 16450–16454. 10.1002/chem.202003062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Barham J. P.; König B. Synthetic Photoelectrochemistry. Angew. Chem., Int. Ed. 2020, 59, 11732–11747. 10.1002/anie.201913767. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Yan H.; Hou Z.-W.; Xu H.-C. Photoelectrochemical C–H Alkylation of Heteroarenes with Organotrifluoroborates. Angew. Chem., Int. Ed. 2019, 58, 4592–4595. 10.1002/anie.201814488. [DOI] [PubMed] [Google Scholar]
  62. a Lai X. L.; Shu X. M.; Song J.; Xu H. C. Electrophotocatalytic decarboxylative C–H functionalization of heteroarenes. Angew. Chem., Int. Ed. 2020, 59, 10626–10632. 10.1002/anie.202002900. [DOI] [PubMed] [Google Scholar]; b Xu P.; Chen P.-Y.; Xu H.-C. Scalable Photoelectrochemical Dehydrogenative Cross-Coupling of Heteroarenes with Aliphatic C–H Bonds. Angew. Chem., Int. Ed. 2020, 59, 14275–14280. 10.1002/anie.202005724. [DOI] [PubMed] [Google Scholar]
  63. a Huang H.; Lambert T. H. Electrophotocatalytic SNAr Reactions of Unactivated Aryl Fluorides at Ambient Temperature and Without Base. Angew. Chem., Int. Ed. 2020, 59, 658–662. 10.1002/anie.201909983. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Huang H.; Strater Z. M.; Lambert T. H. Electrophotocatalytic C–H functionalization of ethers with high regioselectivity. J. Am. Chem. Soc. 2020, 142, 1698–1703. 10.1021/jacs.9b11472. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Huang H.; Strater Z. M.; Rauch M.; Shee J.; Sisto T. J.; Nuckolls C.; Lambert T. H. Electrophotocatalysis with a Trisaminocyclopropenium Radical Dication. Angew. Chem., Int. Ed. 2019, 58, 13318–13322. 10.1002/anie.201906381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Kim H.; Kim H.; Lambert T. H.; Lin S. Reductive electrophotocatalysis: merging electricity and light to achieve extreme reduction potentials. J. Am. Chem. Soc. 2020, 142, 2087–2092. 10.1021/jacs.9b10678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Cowper N. G. W.; Chernowsky C. P.; Williams O. P.; Wickens Z. K. Potent Reductants via Electron-Primed Photoredox Catalysis: Unlocking Aryl Chlorides for Radical Coupling. J. Am. Chem. Soc. 2020, 142, 2093–2099. 10.1021/jacs.9b12328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Qiu Y.; Scheremetjew A.; Finger L. H.; Ackermann L. Electrophotocatalytic Undirected C–H Trifluoromethylations of (Het) Arenes. Chem. - Eur. J. 2020, 26, 3241–3246. 10.1002/chem.201905774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Zhang W.; Carpenter K. L.; Lin S. Electrochemistry Broadens the Scope of Flavin Photocatalysis: Photoelectrocatalytic Oxidation of Unactivated Alcohols. Angew. Chem., Int. Ed. 2020, 59, 409–417. 10.1002/anie.201910300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. a Diccianni J. B.; Diao T. Mechanisms of nickel-catalyzed cross-coupling reactions. Trends Chem. 2019, 1, 830–844. 10.1016/j.trechm.2019.08.004. [DOI] [Google Scholar]; b Tasker S. Z.; Standley E. A.; Jamison T. F. Recent advances in homogeneous nickel catalysis. Nature 2014, 509, 299–309. 10.1038/nature13274. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Hu X. Nickel-catalyzed cross coupling of non-activated alkyl halides: a mechanistic perspective. Chem. Sci. 2011, 2, 1867–1886. 10.1039/c1sc00368b. [DOI] [Google Scholar]
  69. 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. 10.1002/chem.202001318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Li C.; Kawamata Y.; Nakamura H.; Vantourout J. C.; Liu Z.; Hou Q.; Bao D.; Starr J. T.; Chen J.; Yan M.; Baran P. S. Electrochemically Enabled, Nickel-Catalyzed Amination. Angew. Chem., Int. Ed. 2017, 56, 13088–13093. 10.1002/anie.201707906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Kawamata Y.; Vantourout J. C.; Hickey D. P.; Bai P.; Chen L.; Hou Q.; Qiao W.; Barman K.; Edwards M. A.; Garrido-Castro A. F.; deGruyter J. N.; Nakamura H.; Knouse K.; Qin C.; Clay K. J.; Bao D.; Li C.; Starr J. T.; Garcia-Irizarry C.; Sach N.; White H. S.; Neurock M.; Minteer S. D.; Baran P. S. Electrochemically Driven, Ni-Catalyzed Aryl Amination: Scope, Mechanism, and Applications. J. Am. Chem. Soc. 2019, 141, 6392–6402. 10.1021/jacs.9b01886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. a Sambiagio C.; Schönbauer D.; Blieck R.; Dao-Huy T.; Pototschnig G.; Schaaf P.; Wiesinger T.; Zia M. F.; Wencel-Delord J.; Besset T.; et al. A comprehensive overview of directing groups applied in metal-catalysed C–H functionalisation chemistry. Chem. Soc. Rev. 2018, 47, 6603–6743. 10.1039/C8CS00201K. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Gandeepan P.; Ackermann L. Transient directing groups for transformative C–H activation by synergistic metal catalysis. Chem. 2018, 4, 199–222. 10.1016/j.chempr.2017.11.002. [DOI] [Google Scholar]; c He J.; Wasa M.; Chan K. S. L.; Shao Q.; Yu J.-Q. Palladium-Catalyzed Transformations of Alkyl C–H Bonds. Chem. Rev. 2017, 117, 8754–8786. 10.1021/acs.chemrev.6b00622. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Park Y.; Kim Y.; Chang S. Transition metal-catalyzed C–H amination: Scope, mechanism, and applications. Chem. Rev. 2017, 117, 9247–9301. 10.1021/acs.chemrev.6b00644. [DOI] [PubMed] [Google Scholar]; e Ma W.; Gandeepan P.; Li J.; Ackermann L. Recent advances in positional-selective alkenylations: removable guidance for twofold C–H activation. Org. Chem. Front. 2017, 4, 1435–1467. 10.1039/C7QO00134G. [DOI] [Google Scholar]; f Ye B.; Cramer N. Chiral cyclopentadienyls: enabling ligands for asymmetric Rh(III)-catalyzed C–H functionalizations. Acc. Chem. Res. 2015, 48, 1308–1318. 10.1021/acs.accounts.5b00092. [DOI] [PubMed] [Google Scholar]; g Wencel-Delord J.; Glorius F. C–H bond activation enables the rapid construction and late-stage diversification of functional molecules. Nat. Chem. 2013, 5, 369–375. 10.1038/nchem.1607. [DOI] [PubMed] [Google Scholar]; h Arockiam P. B.; Bruneau C.; Dixneuf P. H. Ruthenium (II)-catalyzed C–H bond activation and functionalization. Chem. Rev. 2012, 112, 5879–5918. 10.1021/cr300153j. [DOI] [PubMed] [Google Scholar]; i McMurray L.; O’Hara F.; Gaunt M. J. Recent developments in natural product synthesis using metal-catalysed C–H bond functionalisation. Chem. Soc. Rev. 2011, 40, 1885–1898. 10.1039/c1cs15013h. [DOI] [PubMed] [Google Scholar]; j Ackermann L. Carboxylate-Assisted Transition-Metal-Catalyzed C–H Bond Functionalizations: Mechanism and Scope. Chem. Rev. 2011, 111, 1315–1345. 10.1021/cr100412j. [DOI] [PubMed] [Google Scholar]; k Lyons T. W.; Sanford M. S. Palladium-catalyzed ligand-directed C–H functionalization reactions. Chem. Rev. 2010, 110, 1147–1169. 10.1021/cr900184e. [DOI] [PMC free article] [PubMed] [Google Scholar]; l Ackermann L.; Vicente R.; Kapdi A. R. Transition-metal-catalyzed direct arylation of (hetero) arenes by C–H bond cleavage. Angew. Chem., Int. Ed. 2009, 48, 9792–9826. 10.1002/anie.200902996. [DOI] [PubMed] [Google Scholar]; m Lewis J. C.; Bergman R. G.; Ellman J. A. Direct functionalization of nitrogen heterocycles via Rh-catalyzed C–H bond activation. Acc. Chem. Res. 2008, 41, 1013–1025. 10.1021/ar800042p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. a Yang Q.-L.; Fang P.; Mei T.-S. Recent Advances in Organic Electrochemical C–H Functionalization. Chin. J. Chem. 2018, 36, 338–352. 10.1002/cjoc.201700740. [DOI] [Google Scholar]; b Sauermann N.; Meyer T. H.; Qiu Y.; Ackermann L. Electrocatalytic C–H Activation. ACS Catal. 2018, 8, 7086–7103. 10.1021/acscatal.8b01682. [DOI] [Google Scholar]
  74. Sauermann N.; Meyer T. H.; Tian C.; Ackermann L. Electrochemical Cobalt-Catalyzed C–H Oxygenation at Room Temperature. J. Am. Chem. Soc. 2017, 139, 18452–18455. 10.1021/jacs.7b11025. [DOI] [PubMed] [Google Scholar]
  75. a Sauermann N.; Mei R.; Ackermann L. Electrochemical C–H Amination by Cobalt Catalysis in a Renewable Solvent. Angew. Chem., Int. Ed. 2018, 57, 5090–5094. 10.1002/anie.201802206. [DOI] [PubMed] [Google Scholar]; b Gao X.; Wang P.; Zeng L.; Tang S.; Lei A. Cobalt(II)-catalyzed electrooxidative C–H amination of arenes with alkylamines. J. Am. Chem. Soc. 2018, 140, 4195–4199. 10.1021/jacs.7b13049. [DOI] [PubMed] [Google Scholar]
  76. a Tian C.; Massignan L.; Meyer T. H.; Ackermann L. Electrochemical C–H/N–H Activation by Water-Tolerant Cobalt Catalysis at Room Temperature. Angew. Chem., Int. Ed. 2018, 57, 2383–2387. 10.1002/anie.201712647. [DOI] [PubMed] [Google Scholar]; b Mei R.; Sauermann N.; Oliveira J. C. A.; Ackermann L. Electroremovable Traceless Hydrazides for Cobalt-Catalyzed Electro-Oxidative C–H/N–H Activation with Internal Alkynes. J. Am. Chem. Soc. 2018, 140, 7913–7921. 10.1021/jacs.8b03521. [DOI] [PubMed] [Google Scholar]; c Tang S.; Wang D.; Liu Y.; Zeng L.; Lei A. Cobalt-catalyzed electrooxidative C–H/N–H [4 + 2] annulation with ethylene or ethyne. Nat. Commun. 2018, 9, 798. 10.1038/s41467-018-03246-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. a Mei R.; Ma W.; Zhang Y.; Guo X.; Ackermann L. Cobaltaelectro-Catalyzed Oxidative C–H/N–H Activation with 1,3-Diynes by Electro-Removable Hydrazides. Org. Lett. 2019, 21, 6534–6538. 10.1021/acs.orglett.9b02463. [DOI] [PubMed] [Google Scholar]; b Meyer T. H.; Oliveira J. C. A.; Sau S. C.; Ang N. W. J.; Ackermann L. Electrooxidative Allene Annulations by Mild Cobalt-Catalyzed C–H Activation. ACS Catal. 2018, 8, 9140–9147. 10.1021/acscatal.8b03066. [DOI] [Google Scholar]
  78. a Sau S. C.; Mei R.; Struwe J.; Ackermann L. Cobaltaelectro-Catalyzed C–H Activation with Carbon Monoxide or Isocyanides. ChemSusChem 2019, 12, 3023–3027. 10.1002/cssc.201900378. [DOI] [PubMed] [Google Scholar]; b Zeng L.; Li H.; Tang S.; Gao X.; Deng Y.; Zhang G.; Pao C.-W.; Chen J.-L.; Lee J.-F.; Lei A. Cobalt-catalyzed electrochemical oxidative C–H/N–H carbonylation with hydrogen evolution. ACS Catal. 2018, 8, 5448–5453. 10.1021/acscatal.8b00683. [DOI] [Google Scholar]
  79. Meyer T. H.; Oliveira J. C. A.; Ghorai D.; Ackermann L. Insights into Cobalta(III/IV/II)-Electrocatalysis: Oxidation-Induced Reductive Elimination for Twofold C–H Activation. Angew. Chem., Int. Ed. 2020, 59, 10955–10960. 10.1002/anie.202002258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. a Kim J.; Shin K.; Jin S.; Kim D.; Chang S. Oxidatively Induced Reductive Elimination: Exploring the Scope and Catalyst Systems with Ir, Rh, and Ru Complexes. J. Am. Chem. Soc. 2019, 141, 4137–4146. 10.1021/jacs.9b00364. [DOI] [PubMed] [Google Scholar]; b Li L.; Brennessel W. W.; Jones W. D. An Efficient Low-Temperature Route to Polycyclic Isoquinoline Salt Synthesis via C–H Activation with [Cp*MCl2]2 (M = Rh, Ir). J. Am. Chem. Soc. 2008, 130, 12414–12419. 10.1021/ja802415h. [DOI] [PubMed] [Google Scholar]
  81. a Khake S. M.; Chatani N. Nickel-Catalyzed C–H Functionalization Using A Non-directed Strategy. Chem. 2020, 6, 1056–1081. 10.1016/j.chempr.2020.04.005. [DOI] [Google Scholar]; b Harry N. A.; Saranya S.; Ujwaldev S. M.; Anilkumar G. Recent advances and prospects in nickel-catalyzed C–H activation. Catal. Sci. Technol. 2019, 9, 1726–1743. 10.1039/C9CY00009G. [DOI] [Google Scholar]; c Pototschnig G.; Maulide N.; Schnürch M. Direct Functionalization of C–H Bonds by Iron, Nickel, and Cobalt Catalysis. Chem. - Eur. J. 2017, 23, 9206–9232. 10.1002/chem.201605657. [DOI] [PubMed] [Google Scholar]
  82. Zhang S.-K.; Samanta R. C.; Sauermann N.; Ackermann L. Nickel-Catalyzed Electrooxidative C–H Amination: Support for Nickel(IV). Chem. - Eur. J. 2018, 24, 19166–19170. 10.1002/chem.201805441. [DOI] [PubMed] [Google Scholar]
  83. Zhang S.-K.; Struwe J.; Hu L.; Ackermann L. Nickela-electrocatalyzed C–H Alkoxylation with Secondary Alcohols: Oxidation-Induced Reductive Elimination at Nickel(III). Angew. Chem., Int. Ed. 2020, 59, 3178–3183. 10.1002/anie.201913930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. a Li Z.-l.; Sun K.-k.; Cai C. Nickel-catalyzed regioselective C–H oxygenation: new routes for versatile C–O bond formation. Org. Chem. Front. 2019, 6, 637–642. 10.1039/C8QO01274A. [DOI] [Google Scholar]; b Arun V.; Mahanty K.; De Sarkar S. Nickel-Catalyzed Dehydrogenative Couplings. ChemCatChem 2019, 11, 2243–2259. 10.1002/cctc.201900254. [DOI] [Google Scholar]; c Castro L. C. M.; Chatani N. Nickel Catalysts/N,N′-Bidentate Directing Groups: An Excellent Partnership in Directed C–H Activation Reactions. Chem. Lett. 2015, 44, 410–421. 10.1246/cl.150024. [DOI] [Google Scholar]
  85. Samanta R. C.; Struwe J.; Ackermann L. Nickela-electrocatalyzed Mild C–H Alkylations at Room Temperature. Angew. Chem., Int. Ed. 2020, 59, 14154–14159. 10.1002/anie.202004958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. a Shang R.; Ilies L.; Nakamura E. Iron-Catalyzed C–H Bond Activation. Chem. Rev. 2017, 117, 9086–9139. 10.1021/acs.chemrev.6b00772. [DOI] [PubMed] [Google Scholar]; b Cera G.; Ackermann L. Iron-Catalyzed C–H Functionalization Processes. Top. Curr. Chem. 2016, 374, 57. 10.1007/s41061-016-0059-6. [DOI] [PubMed] [Google Scholar]; c Gu Q.; Al Mamari H. H.; Graczyk K.; Diers E.; Ackermann L. Iron-Catalyzed C(sp2)–H and C(sp3)–H Arylation by Triazole Assistance. Angew. Chem., Int. Ed. 2014, 53, 3868–3871. 10.1002/anie.201311024. [DOI] [PubMed] [Google Scholar]
  87. Zhu C.; Stangier M.; Oliveira J. C. A.; Massignan L.; Ackermann L. Iron-Electrocatalyzed C–H Arylations: Mechanistic Insights into Oxidation-Induced Reductive Elimination for Ferraelectrocatalysis. Chem. - Eur. J. 2019, 25, 16382–16389. 10.1002/chem.201904018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Boddie T. E.; Carpenter S. H.; Baker T. M.; DeMuth J. C.; Cera G.; Brennessel W. W.; Ackermann L.; Neidig M. L. Identification and Reactivity of Cyclometalated Iron(II) Intermediates in Triazole-Directed Iron-Catalyzed C–H Activation. J. Am. Chem. Soc. 2019, 141, 12338–12345. 10.1021/jacs.9b05269. [DOI] [PubMed] [Google Scholar]
  89. Yang L.; Steinbock R.; Scheremetjew A.; Kuniyil R.; Finger L. H.; Messinis A. M.; Ackermann L. Azaruthena(II)-bicyclo[3.2.0]heptadiene: Key Intermediate for Ruthenaelectro(II/III/I)-catalyzed Alkyne Annulations. Angew. Chem., Int. Ed. 2020, 59, 11130–11135. 10.1002/anie.202000762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. a Nareddy P.; Jordan F.; Brenner-Moyer S. E.; Szostak M. Ruthenium(II)-Catalyzed Regioselective C–H Arylation of Cyclic and N,N-Dialkyl Benzamides with Boronic Acids by Weak Coordination. ACS Catal. 2016, 6, 4755–4759. 10.1021/acscatal.6b01360. [DOI] [Google Scholar]; b Bruneau C.; Dixneuf P. H. Ruthenium (II)-Catalysed Functionalisation of C–H Bonds with Alkenes: Alkenylation versus Alkylation. Top. Organomet. Chem. 2015, 55, 137–188. 10.1007/3418_2015_134. [DOI] [Google Scholar]; c Ackermann L. Carboxylate-Assisted Ruthenium-Catalyzed Alkyne Annulations by C–H/Het–H Bond Functionalizations. Acc. Chem. Res. 2014, 47, 281–295. 10.1021/ar3002798. [DOI] [PubMed] [Google Scholar]; d Arockiam P. B.; Bruneau C.; Dixneuf P. H. Ruthenium(II)-Catalyzed C–H Bond Activation and Functionalization. Chem. Rev. 2012, 112, 5879–5918. 10.1021/cr300153j. [DOI] [PubMed] [Google Scholar]
  91. Qiu Y.; Zhu C.; Stangier M.; Struwe J.; Ackermann L. Rhodaelectro-Catalyzed C–H and C–C Activation. CCS Chem. 2020, 2, 1529–1552. 10.31635/ccschem.020.202000365. [DOI] [Google Scholar]
  92. a Noël T.; Cao Y.; Laudadio G. The Fundamentals Behind the Use of Flow Reactors in Electrochemistry. Acc. Chem. Res. 2019, 52, 2858–2869. 10.1021/acs.accounts.9b00412. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Pletcher D.; Green R. A.; Brown R. C. D. Flow Electrolysis Cells for the Synthetic Organic Chemistry Laboratory. Chem. Rev. 2018, 118, 4573–4591. 10.1021/acs.chemrev.7b00360. [DOI] [PubMed] [Google Scholar]; c Atobe M.; Tateno H.; Matsumura Y. Applications of Flow Microreactors in Electrosynthetic Processes. Chem. Rev. 2018, 118, 4541–4572. 10.1021/acs.chemrev.7b00353. [DOI] [PubMed] [Google Scholar]
  93. Kong W.-J.; Finger L. H.; Messinis A. M.; Kuniyil R.; Oliveira J. C. A.; Ackermann L. Flow Rhodaelectro-Catalyzed Alkyne Annulations by Versatile C–H Activation: Mechanistic Support for Rhodium(III/IV). J. Am. Chem. Soc. 2019, 141, 17198–17206. 10.1021/jacs.9b07763. [DOI] [PubMed] [Google Scholar]
  94. Kong W.-J.; Shen Z.; Finger L. H.; Ackermann L. Electrochemical Access to Aza-Polycyclic Aromatic Hydrocarbons: Rhoda-Electrocatalyzed Domino Alkyne Annulations. Angew. Chem., Int. Ed. 2020, 59, 5551–5556. 10.1002/anie.201914775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Wu Z.-J.; Su F.; Lin W.; Song J.; Wen T.-B.; Zhang H.-J.; Xu H.-C. Scalable Rhodium(III)-Catalyzed Aryl C–H Phosphorylation Enabled by Anodic Oxidation Induced Reductive Elimination. Angew. Chem., Int. Ed. 2019, 58, 16770–16774. 10.1002/anie.201909951. [DOI] [PubMed] [Google Scholar]
  96. Qiu Y.; Scheremetjew A.; Ackermann L. Electro-Oxidative C–C Alkenylation by Rhodium(III) Catalysis. J. Am. Chem. Soc. 2019, 141, 2731–2738. 10.1021/jacs.8b13692. [DOI] [PubMed] [Google Scholar]
  97. Kronenwetter H.; Husek J.; Etz B.; Jones A.; Manchanayakage R. Electrochemical pinacol coupling of aromatic carbonyl compounds in a [BMIM][BF4]–H2O mixture. Green Chem. 2014, 16, 1489–1495. 10.1039/c3gc41641k. [DOI] [Google Scholar]
  98. Kise N.; Agui S.; Morimoto S.; Ueda N. Electroreductive acylation of aromatic ketones with acylimidazoles. J. Org. Chem. 2005, 70, 9407–9410. 10.1021/jo051498w. [DOI] [PubMed] [Google Scholar]
  99. Shono T.; Masuda H.; Murase H.; Shimomura M.; Kashimura S. Facile electroreduction of methyl esters and N,N-dimethylamides of aliphatic carboxylic acids to primary alcohols. J. Org. Chem. 1992, 57, 1061–1063. 10.1021/jo00030a003. [DOI] [Google Scholar]
  100. a Ishifune M.; Yamashita H.; Kera Y.; Yamashita N.; Hirata K.; Murase H.; Kashimura S. Electroreduction of aromatics using magnesium electrodes in aprotic solvents containing alcoholic proton donors. Electrochim. Acta 2003, 48, 2405–2409. 10.1016/S0013-4686(03)00259-7. [DOI] [Google Scholar]; b Swenson K. E.; Zemach D.; Nanjundiah C.; Kariv-Miller E. Birch reductions of methoxyaromatics in aqueous solution. J. Org. Chem. 1983, 48, 1777–1779. 10.1021/jo00158a042. [DOI] [Google Scholar]; c Benkeser R. A.; Kaiser E. M. An Electrochemical Method of Reducing Aromatic Compounds Selectively to Dihydro or Tetrahydro Products. J. Am. Chem. Soc. 1963, 85, 2858–2859. 10.1021/ja00901a047. [DOI] [Google Scholar]
  101. Peters B. K.; Rodriguez K. X.; Reisberg S. H.; Beil S. B.; Hickey D. P.; Kawamata Y.; Collins M.; Starr J.; Chen L.; Udyavara S.; Klunder K.; Gorey T. J.; Anderson S. L.; Neurock M.; Minteer S. D.; Baran P. S. Scalable and safe synthetic organic electroreduction inspired by Li-ion battery chemistry. Science 2019, 363, 838–845. 10.1126/science.aav5606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. a Burkart M. D.; Hazari N.; Tway C. L.; Zeitler E. L. Opportunities and Challenges for Catalysis in Carbon Dioxide Utilization. ACS Catal. 2019, 9, 7937–7956. 10.1021/acscatal.9b02113. [DOI] [Google Scholar]; b Francke R.; Schille B.; Roemelt M. Homogeneously Catalyzed Electroreduction of Carbon Dioxide—Methods, Mechanisms, and Catalysts. Chem. Rev. 2018, 118, 4631–4701. 10.1021/acs.chemrev.7b00459. [DOI] [PubMed] [Google Scholar]
  103. a Liu Q.; Wu L.; Jackstell R.; Beller M. Using carbon dioxide as a building block in organic synthesis. Nat. Commun. 2015, 6, 5933. 10.1038/ncomms6933. [DOI] [PubMed] [Google Scholar]; b Aresta M.; Dibenedetto A.; Angelini A. Catalysis for the Valorization of Exhaust Carbon: from CO2 to Chemicals, Materials, and Fuels. Technological Use of CO2. Chem. Rev. 2014, 114, 1709–1742. 10.1021/cr4002758. [DOI] [PubMed] [Google Scholar]
  104. a Fujihara T.; Tsuji Y. Carboxylation Reactions Using Carbon Dioxide as the C1 Source via Catalytically Generated Allyl Metal Intermediates. Front. Chem. 2019, 7, 430. 10.3389/fchem.2019.00430. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Tortajada A.; Juliá-Hernández F.; Börjesson M.; Moragas T.; Martin R. Transition-Metal-Catalyzed Carboxylation Reactions with Carbon Dioxide. Angew. Chem., Int. Ed. 2018, 57, 15948–15982. 10.1002/anie.201803186. [DOI] [PubMed] [Google Scholar]; c Luan Y.-X.; Ye M. Transition metal-mediated or catalyzed hydrocarboxylation of olefins with CO2. Tetrahedron Lett. 2018, 59, 853–861. 10.1016/j.tetlet.2018.01.035. [DOI] [Google Scholar]; d Chen Y.-G.; Xu X.-T.; Zhang K.; Li Y.-Q.; Zhang L.-P.; Fang P.; Mei T.-S. Transition-Metal-Catalyzed Carboxylation of Organic Halides and Their Surrogates with Carbon Dioxide. Synthesis 2018, 50, 35–48. 10.1055/s-0036-1590908. [DOI] [Google Scholar]; e Cai X.; Xie B. Direct carboxylative reactions for the transformation of carbon dioxide into carboxylic acids and derivatives. Synthesis 2013, 45, 3305–3324. 10.1055/s-0033-1340061. [DOI] [Google Scholar]; f Tsuji Y.; Fujihara T. Carbon dioxide as a carbon source in organic transformation: carbon–carbon bond forming reactions by transition-metal catalysts. Chem. Commun. 2012, 48, 9956–9964. 10.1039/c2cc33848c. [DOI] [PubMed] [Google Scholar]
  105. a Liu Q.; Wu L.; Fleischer I.; Selent D.; Franke R.; Jackstell R.; Beller M. Development of a ruthenium/phosphite catalyst system for domino hydroformylation–reduction of olefins with carbon dioxide. Chem. - Eur. J. 2014, 20, 6809–6809. 10.1002/chem.201400435. [DOI] [PubMed] [Google Scholar]; b Wu L.; Liu Q.; Fleischer I.; Jackstell R.; Beller M. Ruthenium-catalysed alkoxycarbonylation of alkenes with carbon dioxide. Nat. Commun. 2014, 5, 3091. 10.1038/ncomms4091. [DOI] [PubMed] [Google Scholar]; c Lescot C.; Nielsen D. U.; Makarov I. S.; Lindhardt A. T.; Daasbjerg K.; Skrydstrup T. Efficient fluoride-catalyzed conversion of CO2 to CO at room temperature. J. Am. Chem. Soc. 2014, 136, 6142–6147. 10.1021/ja502911e. [DOI] [PubMed] [Google Scholar]; d Hayashi Y.; Kita S.; Brunschwig B. S.; Fujita E. Involvement of a Binuclear Species with the Re–C(O)O–Re Moiety in CO2 Reduction Catalyzed by Tricarbonyl Rhenium(I) Complexes with Diimine Ligands: Strikingly Slow Formation of the Re–Re and Re–C(O)O–Re Species from Re (dmb)(CO)3S (dmb = 4, 4′-Dimethyl-2,2′-bipyridine, S = Solvent). J. Am. Chem. Soc. 2003, 125, 11976–11987. 10.1021/ja035960a. [DOI] [PubMed] [Google Scholar]; e Tominaga K.-i.; Sasaki Y. Ruthenium complex-catalyzed hydroformylation of alkenes with carbon dioxide. Catal. Commun. 2000, 1, 1–3. 10.1016/S1566-7367(00)00006-6. [DOI] [Google Scholar]
  106. a Cui X.; Dai X.; Zhang Y.; Deng Y.; Shi F. Methylation of amines, nitrobenzenes and aromatic nitriles with carbon dioxide and molecular hydrogen. Chem. Sci. 2014, 5, 649–655. 10.1039/C3SC52676C. [DOI] [Google Scholar]; b Li Y.; Fang X.; Junge K.; Beller M. A general catalytic methylation of amines using carbon dioxide. Angew. Chem., Int. Ed. 2013, 52, 9568–9571. 10.1002/anie.201301349. [DOI] [PubMed] [Google Scholar]; c Jacquet O.; Frogneux X.; Das Neves Gomes C.; Cantat T. CO2 as a C1-building block for the catalytic methylation of amines. Chem. Sci. 2013, 4, 2127–2131. 10.1039/c3sc22240c. [DOI] [Google Scholar]; d Beydoun K.; vom Stein T.; Klankermayer J.; Leitner W. Ruthenium-Catalyzed Direct Methylation of Primary and Secondary Aromatic Amines Using Carbon Dioxide and Molecular Hydrogen. Angew. Chem., Int. Ed. 2013, 52, 9554–9557. 10.1002/anie.201304656. [DOI] [PubMed] [Google Scholar]; e Li Y.; Sorribes I.; Yan T.; Junge K.; Beller M. Selective methylation of amines with carbon dioxide and H2. Angew. Chem., Int. Ed. 2013, 52, 12156–12160. 10.1002/anie.201306850. [DOI] [PubMed] [Google Scholar]
  107. a Pal T. K.; De D.; Bharadwaj P. K. Metal–organic frameworks for the chemical fixation of CO2 into cyclic carbonates. Coord. Chem. Rev. 2020, 408, 213173. 10.1016/j.ccr.2019.213173. [DOI] [Google Scholar]; b Azzouz R.; Moreno V. C.; Herasme-Grullon C.; Levacher V.; Estel L.; Ledoux A.; Derrouiche S.; Marsais F.; Bischoff L. Efficient Conversion of Epoxides into Carbonates with CO2 and a Single Organocatalyst: Laboratory and Kilogram-Scale Experiments. Synlett 2020, 31, 183–188. 10.1055/s-0039-1691405. [DOI] [Google Scholar]; c Sakakura T.; Kohno K. The synthesis of organic carbonates from carbon dioxide. Chem. Commun. 2009, 1312–1330. 10.1039/b819997c. [DOI] [PubMed] [Google Scholar]; d Coates G. W.; Moore D. R. Discrete metal-based catalysts for the copolymerization of CO2 and epoxides: discovery, reactivity, optimization, and mechanism. Angew. Chem., Int. Ed. 2004, 43, 6618–6639. 10.1002/anie.200460442. [DOI] [PubMed] [Google Scholar]
  108. Matthessen R.; Fransaer J.; Binnemans K.; De Vos D. E. Electrocarboxylation: towards sustainable and efficient synthesis of valuable carboxylic acids. Beilstein J. Org. Chem. 2014, 10, 2484–2500. 10.3762/bjoc.10.260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. a Medvedev J. J.; Medvedeva X. V.; Li F.; Zienchuk T. A.; Klinkova A. Electrochemical CO2 Fixation to α-Methylbenzyl Bromide in Divided Cells with Nonsacrificial Anodes and Aqueous Anolytes. ACS Sustainable Chem. Eng. 2019, 7, 19631–19639. 10.1021/acssuschemeng.9b04647. [DOI] [Google Scholar]; b Wu L.-X.; Sun Q.-L.; Yang M.-P.; Zhao Y.-G.; Guan Y.-B.; Wang H.; Lu J.-X. Highly Efficient Electrocatalytic Carboxylation of 1-Phenylethyl Chloride at Cu Foam Cathode. Catalysts 2018, 8, 273. 10.3390/catal8070273. [DOI] [Google Scholar]
  110. Chen B.-L.; Zhu H.-W.; Xiao Y.; Sun Q.-L.; Wang H.; Lu J.-X. Asymmetric electrocarboxylation of 1-phenylethyl chloride catalyzed by electrogenerated chiral [CoI(salen)]–complex. Electrochem. Commun. 2014, 42, 55–59. 10.1016/j.elecom.2014.02.009. [DOI] [Google Scholar]
  111. Jiao K.-J.; Li Z.-M.; Xu X.-T.; Zhang L.-P.; Li Y.-Q.; Zhang K.; Mei T.-S. Palladium-catalyzed reductive electrocarboxylation of allyl esters with carbon dioxide. Org. Chem. Front. 2018, 5, 2244–2248. 10.1039/C8QO00507A. [DOI] [Google Scholar]
  112. Ang N. W. J.; Oliveira J. C. A.; Ackermann L. Electroreductive Cobalt-Catalyzed Carboxylation: Cross-Electrophile Electrocoupling with Atmospheric CO2. Angew. Chem., Int. Ed. 2020, 59, 12842–12847. 10.1002/anie.202003218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Bortnikov E. O.; Semenov S. N. Coupling of Alternating Current to Transition-Metal Catalysis: Examples of Nickel-Catalyzed Cross-Coupling. J. Org. Chem. 2021, 86, 782–793. 10.1021/acs.joc.0c02350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Wills A. G.; Poole D. L.; Alder C. M.; Reid M. A mechanistic and cautionary case study on the use of alternating potential in electrochemical reactions. ChemElectroChem 2020, 7, 2771–2776. 10.1002/celc.202000648. [DOI] [Google Scholar]
  115. Rodrigo S.; Um C.; Mixdorf J. C.; Gunasekera D.; Nguyen H. M.; Luo L. Alternating Current Electrolysis for Organic Electrosynthesis: Trifluoromethylation of (Hetero)arenes. Org. Lett. 2020, 22, 6719–6723. 10.1021/acs.orglett.0c01906. [DOI] [PubMed] [Google Scholar]
  116. a Yutthalekha T.; Wattanakit C.; Lapeyre V.; Nokbin S.; Warakulwit C.; Limtrakul J.; Kuhn A. Asymmetric synthesis using chiral-encoded metal. Nat. Commun. 2016, 7, 12678. 10.1038/ncomms12678. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Arnaboldi S.; Benincori T.; Cirilli R.; Kutner W.; Magni M.; Mussini P. R.; Noworyta K.; Sannicolo F. Inherently chiral electrodes: the tool for chiral voltammetry. Chem. Sci. 2015, 6, 1706–1711. 10.1039/C4SC03713H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Arnaboldi S.; Magni M.; Mussini P. R. Enantioselective selectors for chiral electrochemistry and electroanalysis: Stereogenic elements and enantioselection performance. Curr. Opin. Electrochem. 2018, 8, 60–72. 10.1016/j.coelec.2018.01.002. [DOI] [Google Scholar]
  118. Seebach D.; Oei H. A. Mechanism of electrochemical pinacolization. The first asymmetric synthesis in a chiral medium. Angew. Chem., Int. Ed. Engl. 1975, 14, 634–636. 10.1002/anie.197506342. [DOI] [Google Scholar]
  119. Fu N.; Song L.; Liu J.; Shen Y.; Siu J. C.; Lin S. New Bisoxazoline Ligands Enable Enantioselective Electrocatalytic Cyanofunctionalization of Vinylarenes. J. Am. Chem. Soc. 2019, 141, 14480–14485. 10.1021/jacs.9b03296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Li Z.-L.; Fang G.-C.; Gu Q.-S.; Liu X.-Y. Recent advances in copper-catalysed radical-involved asymmetric 1,2-difunctionalization of alkenes. Chem. Soc. Rev. 2020, 49, 32–48. 10.1039/C9CS00681H. [DOI] [PubMed] [Google Scholar]
  121. a Tanaka H.; Kuroboshi M.; Takeda H.; Kanda H.; Torii S. Electrochemical asymmetric epoxidation of olefins by using an optically active Mn-salen complex. J. Electroanal. Chem. 2001, 507, 75–81. 10.1016/S0022-0728(01)00387-4. [DOI] [Google Scholar]; b Torii S.; Liu P.; Tanaka H. Electrochemical Os-catalyzed asymmetric dihydroxylation of olefins with Sharpless’ ligand. Chem. Lett. 1995, 24, 319–320. 10.1246/cl.1995.319. [DOI] [Google Scholar]
  122. a Wang D.; Zhu N.; Chen P.; Lin Z.; Liu G. Enantioselective Decarboxylative Cyanation Employing Cooperative Photoredox Catalysis and Copper Catalysis. J. Am. Chem. Soc. 2017, 139, 15632–15635. 10.1021/jacs.7b09802. [DOI] [PubMed] [Google Scholar]; b Zhang W.; Wang F.; McCann S. D.; Wang D.; Chen P.; Stahl S. S.; Liu G. Enantioselective cyanation of benzylic C–H bonds via copper-catalyzed radical relay. Science 2016, 353, 1014–1018. 10.1126/science.aaf7783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. DeLano T. J.; Reisman S. E. Enantioselective Electroreductive Coupling of Alkenyl and Benzyl Halides via Nickel Catalysis. ACS Catal. 2019, 9, 6751–6754. 10.1021/acscatal.9b01785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Qiu H.; Shuai B.; Wang Y.-Z.; Liu D.; Chen Y.-G.; Gao P.-S.; Ma H.-X.; Chen S.; Mei T.-S. Enantioselective Ni-Catalyzed Electrochemical Synthesis of Biaryl Atropisomers. J. Am. Chem. Soc. 2020, 142, 9872–9878. 10.1021/jacs.9b13117. [DOI] [PubMed] [Google Scholar]
  125. Dhawa U.; Tian C.; Wdowik T.; Oliveira J. C. A.; Hao J.; Ackermann L. Enantioselective Pallada-Electrocatalyzed C–H Activation by Transient Directing Groups: Expedient Access to Helicenes. Angew. Chem., Int. Ed. 2020, 59, 13451–13457. 10.1002/anie.202003826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. a Muchez L.; De Vos D. E.; Kim M. Sacrificial Anode-Free Electrosynthesis of α-Hydroxy Acids via Electrocatalytic Coupling of Carbon Dioxide to Aromatic Alcohols. ACS Sustainable Chem. Eng. 2019, 7, 15860–15864. 10.1021/acssuschemeng.9b04612. [DOI] [Google Scholar]; b Chaussard J.; Folest J.-C.; Nedelec J.-Y.; Perichon J.; Sibille S.; Troupel M. Use of Sacrificial Anodes in Electrochemical Functionalization of Organic Halides. Synthesis 1990, 1990, 369–381. 10.1055/s-1990-26880. [DOI] [Google Scholar]
  127. Frontana-Uribe B. A.; Little R. D.; Ibanez J. G.; Palma A.; Vasquez-Medrano R. Organic electrosynthesis: a promising green methodology in organic chemistry. Green Chem. 2010, 12, 2099–2119. 10.1039/c0gc00382d. [DOI] [Google Scholar]
  128. Dong X.; Roeckl J. L.; Waldvogel S. R.; Morandi B. Merging shuttle reactions and paired electrolysis for reversible vicinal dihalogenations. Science 2021, 371, 507–514. 10.1126/science.abf2974. [DOI] [PubMed] [Google Scholar]
  129. Shen T.; Lambert T. H. Electrophotocatalytic diamination of vicinal C–H bonds. Science 2021, 371, 620–626. 10.1126/science.abf2798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. a Leow W. R.; Lum Y.; Ozden A.; Wang Y.; Nam D.-H.; Chen B.; Wicks J.; Zhuang T.-T.; Li F.; Sinton D.; Sargent E. H. Chloride-mediated selective electrosynthesis of ethylene and propylene oxides at high current density. Science 2020, 368, 1228–1233. 10.1126/science.aaz8459. [DOI] [PubMed] [Google Scholar]; b Zhang N.; Zheng F.; Huang B.; Ji Y.; Shao Q.; Li Y.; Xiao X.; Huang X. Exploring Bi2Te3 Nanoplates as Versatile Catalysts for Electrochemical Reduction of Small Molecules. Adv. Mater. 2020, 32, 1906477. 10.1002/adma.201906477. [DOI] [PubMed] [Google Scholar]; c Zhang W.; Lai W.; Cao R. Energy-Related Small Molecule Activation Reactions: Oxygen Reduction and Hydrogen and Oxygen Evolution Reactions Catalyzed by Porphyrin- and Corrole-Based Systems. Chem. Rev. 2017, 117, 3717–3797. 10.1021/acs.chemrev.6b00299. [DOI] [PubMed] [Google Scholar]
  131. Shin K.; Park Y.; Baik M.-H.; Chang S. Iridium-catalysed arylation of C–H bonds enabled by oxidatively induced reductive elimination. Nat. Chem. 2018, 10, 218–224. 10.1038/nchem.2900. [DOI] [PubMed] [Google Scholar]
  132. a Chen H.; Prater M. B.; Cai R.; Dong F.; Chen H.; Minteer S. D. Bioelectrocatalytic Conversion from N2 to Chiral Amino Acids in a H2/α-Keto Acid Enzymatic Fuel Cell. J. Am. Chem. Soc. 2020, 142, 4028–4036. 10.1021/jacs.9b13968. [DOI] [PubMed] [Google Scholar]; b Weng Y.; Song C.; Chiang C.-W.; Lei A. Single electron transfer-based peptide/protein bioconjugations driven by biocompatible energy input. Commun. Chem. 2020, 3, 171. 10.1038/s42004-020-00413-x. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Song C.; Liu K.; Wang Z.; Ding B.; Wang S.; Weng Y.; Chiang C.-W.; Lei A. Electrochemical oxidation induced selective tyrosine bioconjugation for the modification of biomolecules. Chem. Sci. 2019, 10, 7982–7987. 10.1039/C9SC02218J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. a Chen H.; Simoska O.; Lim K.; Grattieri M.; Yuan M.; Dong F.; Lee Y. S.; Beaver K.; Weliwatte S.; Gaffney E. M.; Minteer S. D. Fundamentals, Applications, and Future Directions of Bioelectrocatalysis. Chem. Rev. 2020, 120, 12903–12993. 10.1021/acs.chemrev.0c00472. [DOI] [PubMed] [Google Scholar]; b Yang J.; Cai F.; Desbois N.; Huang L.; Gros C. P.; Bolze F.; Fang Y.; Wang S.; Xu H.-J. Synthesis, spectroscopic characterization, one and two-photon absorption properties and electrochemistry of π-expanded BODIPYs dyes. Dyes Pigm. 2020, 175, 108173. 10.1016/j.dyepig.2019.108173. [DOI] [Google Scholar]; c Chen H.; Dong F.; Minteer S. D. The progress and outlook of bioelectrocatalysis for the production of chemicals, fuels and materials. Nat. Catal. 2020, 3, 225–244. 10.1038/s41929-019-0408-2. [DOI] [Google Scholar]; d Ruff A.; Conzuelo F.; Schuhmann W. Bioelectrocatalysis as the basis for the design of enzyme-based biofuel cells and semi-artificial biophotoelectrodes. Nat. Catal. 2020, 3, 214–224. 10.1038/s41929-019-0381-9. [DOI] [Google Scholar]; e Abdellaoui S.; Macazo F. C.; Cai R.; De Lacey A. L.; Pita M.; Minteer S. D. Enzymatic Electrosynthesis of Alkanes by Bioelectrocatalytic Decarbonylation of Fatty Aldehydes. Angew. Chem., Int. Ed. 2018, 57, 2404–2408. 10.1002/anie.201712890. [DOI] [PubMed] [Google Scholar]
  134. a Peterson B. M.; Lin S.; Fors B. P. Electrochemically Controlled Cationic Polymerization of Vinyl Ethers. J. Am. Chem. Soc. 2018, 140, 2076–2079. 10.1021/jacs.8b00173. [DOI] [PubMed] [Google Scholar]; b Qi M.; Dong Q.; Wang D.; Byers J. A. Electrochemically Switchable Ring-Opening Polymerization of Lactide and Cyclohexene Oxide. J. Am. Chem. Soc. 2018, 140, 5686–5690. 10.1021/jacs.8b02171. [DOI] [PubMed] [Google Scholar]; c Wang Y.; Fantin M.; Park S.; Gottlieb E.; Fu L.; Matyjaszewski K. Electrochemically Mediated Reversible Addition–Fragmentation Chain-Transfer Polymerization. Macromolecules 2017, 50, 7872–7879. 10.1021/acs.macromol.7b02005. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Magenau A. J. D.; Strandwitz N. C.; Gennaro A.; Matyjaszewski K. Electrochemically Mediated Atom Transfer Radical Polymerization. Science 2011, 332, 81–84. 10.1126/science.1202357. [DOI] [PubMed] [Google Scholar]

Articles from ACS Central Science are provided here courtesy of American Chemical Society

RESOURCES