Electrocatalytic C–H Activation - ACS Catalysis (ACS Publications)

Jun 18, 2018 - (1−3) When thermal reactions are performed, electricity is initially transferred in ... The palladium(0) intermediate 10 is then reox...
0 downloads 0 Views 6MB Size
Review Cite This: ACS Catal. 2018, 8, 7086−7103

pubs.acs.org/acscatalysis

Electrocatalytic C−H Activation Nicolas Sauermann, Tjark H. Meyer, Youai Qiu, and Lutz Ackermann*

Downloaded via UNIV OF WINNIPEG on July 7, 2018 at 05:09:54 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Institut für Organische und Biomolekulare Chemie, Georg-August-Universität Göttingen, Tammannstraße 2, 37077 Göttingen, Germany ABSTRACT: C−H activation has emerged as a transformative tool in molecular synthesis, but until recently oxidative C−H activations have largely involved the use of stoichiometric amounts of expensive and toxic metal oxidants, compromising the overall sustainable nature of C−H activation chemistry. In sharp contrast, electrochemical C−H activation has been identified as a more efficient strategy that exploits storable electricity in place of byproduct-generating chemical reagents. Thus, transition-metal catalysts were shown to enable versatile C−H activation reactions in a sustainable manner. While palladium catalysis set the stage for C(sp2)−H and C(sp3)−H functionalizations by N-containing directing groups, rhodium and ruthenium catalysts allowed the use of weakly coordinating amides and acids. In contrast to these precious 4d transition metals, the recent year has witnessed the emergence of versatile cobalt catalysts for C−H oxygenations, C−H nitrogenations, and C−C-forming [4+2] alkyne annulations. Thereby, the use of toxic and expensive silver(I) oxidants was prevented, improving the environmentally benign nature of C−H activation catalysis. Herein, we summarize the recent major advances in organometallic activations of otherwise inert C−H bonds by electrocatalysis through May 2018. KEYWORDS: C−H activation, cobalt, electrochemistry, electrosynthesis, mechanism, pyridines, transition metal photo-induced organic transformations.9−17 However, the direct use of daylight continues to be challenging because of the strong variations in the accessible energy during day and night time. Hence, solar energy is usually first converted into storable electrical energy. In addition, photo-redox catalysis often required precious transition-metal catalysts for strong C−H bond activation,18,19 and largely employed costly reagents as the sacrificial redox mediators.9−17 In sharp contrast, the direct use of storable electricity compares favorably, and arguably represents an attractive means for enabling chemical reactions, with particular potential for redox transformations (Figure 1).20,21 In the past decade, transition-metal-catalyzed C−H activation has emerged as a powerful tool in molecular sciences, which significantly reduces the number of synthetic operations.22−26 At the same time, C−H activation prevents the formation of undesired byproducts, thereby improving the overall footprint of organic syntheses. Yet, despite indisputable advances in the field, redoxidative C−H activation processes were until recently largely limited to stoichiometric quantities of expensive or toxic chemical reagents, such as copper(II) and silver(I) salts, as the sacrificial oxidants, compromising the overall sustainable nature of C−H activation (Scheme 1a). In addition, metal-free, innate transformations were predominantly realized with activated substrates in terms of

1. INTRODUCTION The use of alternative forms of energy, including wind power, solar energy, and hydropower (Figure 1), to establish novel molecular transformations bears great potential for establishing sustainable organic syntheses.1−3 When thermal reactions are performed, electricity is initially transferred in an inefficient manner into thermal energy due to dissipative processes.4−7 As a consequence, recent years have witnessed a renaissance8 of

Received: April 29, 2018 Revised: June 15, 2018 Published: June 18, 2018

Figure 1. Gross electricity generation in Germany in April of 2018, in TWh,20 and overall energy mix.21 © XXXX American Chemical Society

7086

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Scheme 1. (a) Stoichiometric Oxidants, (b) Electrochemical C−H Activation, and (c) Comparison of Bond Dissociation Energies27−30

nucleophilicity, oxidation potential, and bond dissociation energies (Scheme 1c).27−30 Since the early reports by Kolbe in the 1840s,31,32 the enormous potential of electro-organic synthesis has been repeatedly exploited.33−42 Hence, electro-transformations of activated substrates have been explored since the 1960s and have proven instrumental for recent elegant developments in electro-organic synthesis, exploiting the innate reactivity profile of a given organic molecule. Hence, electrochemical, metal-free functionalizations have gained considerable recent momentum due to major contributions from Waldvogel,43−49 Baran,50−54 Yoshida,55−61 and Xu,62−68 among others.69−75 While those findings significantly broadened the viable scope of electrosynthesis, these transformations are based on the inherent reactivity and exploit the oxidation or radical formation in somewhat preactivated positions. In contrast, the transformation of strong C−H bonds with a bond dissociation energy BDEC−H > 113.5 kcal/mol30 and pKa > 4476 is challenging under metal-free conditions (Scheme 1c). Here, the enabling merger of metal-catalyzed C−H activation with electrocatalysis constitutes the key to success for the development of environmentally benign syntheses beyond the naturally encoded innate reactivities. Thus, electrochemistry holds great potential to replace costly stoichiometric redox reagents in organometallic C−H activation by userfriendly cost-effective electricity (Figure 2).77 Herein, we summarize the recent merger of electrocatalysis with organometallic C−H activation through Spring 2018.78

Figure 2. Costs of common oxidants and reductants.

Scheme 2. Electrocatalytic Fujiwara−Moritani Reaction

cant potential of merging palladium-catalyzed C−H activation with electrosynthesis. The arylation of acrylates 2 and styrenes 3 was proposed to be initiated by a base-assisted83 C−H activation manifold, followed by a migratory insertion of the double bond into the palladium−carbon bond.79 The subsequent β-elimination liberates the final product 5/6, and reductive elimination of acetic acid generates the key palladium(0) species 10. The palladium(0) intermediate 10 is then reoxidized by BQ 4. The thus formed hydroquinone 11 is itself reoxidized at the carbon anode, while the reduction of protons to molecular hydrogen takes place at the nickel cathode (Scheme 3). In 2009, Kakiuchi disclosed a palladium-catalyzed halogenation using electrochemical C−H functionalization by palladium catalysis (Scheme 4).84 In contrast to the previous Fujiwara−Moritani approach by Amatore and Jutand,79 here

2. PALLADIUM-CATALYZED TRANSFORMATIONS An early example of the combination of palladium-catalyzed C−H activation and electrochemistry was elegantly designed by Amatore and Jutand in 2007.79 In their pioneering contribution, a palladium(II/0)-catalyzed Fujiwara−Moritanitype80−82 C−H alkenylation reaction was devised (Scheme 2),79 using co-catalytic amounts of p-benzoquinone (4, BQ) as a redox mediator.79 Thus, the formed palladium(0) species was not directly oxidized at the electrode surface, but rather the electron shuttle BQ (4) was interacting with the anode. This example illustrated that palladium C−H activation catalysis is compatible with electrosynthesis, while indicating the signifi7087

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Scheme 3. Proposed Catalytic Cycle for the Electrochemical Fujiwara−Moritani Reaction

Scheme 5. Proposed Catalytic Cycle for the Electrochemical C−X Formation

Scheme 6. (a) Electrochemical Palladium-Catalyzed Iodination and (b) One-Pot Suzuki−Miyaura Coupling Scheme 4. Electrochemical Halogenation of Phenylpyridines 12a

a

PdBr2 (2.0 mol %) and 2M HBr(aq) used.

oxidation of a palladium(0) intermediate was not occurring.84 Instead, the electricity was solely required for the generation of the electrophilic Cl+ cation from mineral acid. This presents an elegant and cost-efficient alternative to the use of reactive halogenation reagents, such as N-halosuccinimides, which drastically improves the efficiency of direct halogenation.84 In contrast to the potentiostatic approach by Amatore and Jutand,79 Kakiuchi used constant current electrolysis (CCE),84 which resulted in an arguably overall more user-friendly protocol. The catalyst’s working mode was proposed to follow the pathway of palladium(II)-catalyzed C−H halogenations by initial coordination through the Lewis-basic nitrogen, followed by a proximity-induced electrophilic C−H palladation (Scheme 5). The cyclometalated intermediate 15 then forms a C−X bond with the electrochemically generated halonium ion, which finally liberates the desired product 13 by ligand exchange at the palladium center. Later, the C−H halogenation strategy was nicely extended to enable C−H iodinations (Scheme 6a).85 Here, elemental iodine proved amenable as the iodonium precursor of choice, while potassium iodide was also found to be a competent iodonium source. Additionally, the authors demonstrated the power of this approach by palladium(0)-catalyzed Suzuki− Miyaura coupling with the thus-formed products in a one-pot

fashion, thereby highlighting the versatility of the halogenated arenes 13 as transient functional groups86 in organic synthesis (Scheme 6b). In addition to C−H halogenations, direct C−H oxygenations87,88 have been of interest in electrochemically guided metal-catalyzed C−H activation. In 2013, the group of Budnikova found that palladium complexes are competent catalysts in the perfluoroalkoxylation of phenylpyridines 12 (Scheme 7a).89 While the scope of this C−H functionalization appeared to be thus far limited to perfluoroalkylated acids 18, the authors provided invaluable information on the reaction mechanism, including the isolation of key intermediates (Scheme 7b) as well as informative cyclic voltammetry (CV) studies. Based on their findings on electrochemical C−H halogenations,84,85 Kakiuchi devised reaction conditions for the homodimerization of phenylpyridines 12 in the presence of stoichiometric or co-catalytic amounts of iodine (Scheme 8).90 It is noteworthy that a remarkable regioselectivity was achieved with meta-substituted arenes 12, which cannot be matched by chemical oxidation conditions with oxone.91 Also, the reaction 7088

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Scheme 7. (a) Palladium-Catalyzed Electrochemical C−H Perfluorooxygenation and (b) Stoichiometric Studies

Scheme 9. Electrochemical C−H Chlorination of Benzamides 23 by Bidentate Assistance

Scheme 8. Homodimerization of Phenylpyridines 12

chlorination of the quinoline in C-5 and C-7 position led to mixtures of multichlorinated products.100 The practical relevance of the electrochemical chlorination of phenylpyridines 12 and benzamides 23 was mirrored by an elegant synthesis of vismodegib (27),100 a marketed drug for cancer treatment (Scheme 9b).105,106 To gain further insights into the reaction mechanism, cyclometalated palladium(II) complex 28 was isolated and shown to undergo electrophilic chlorination under the electrochemical conditions in a stoichiometric fashion (Scheme 9c).100 This finding is suggestive of the complex 28 being a kinetically relevant, on-cycle intermediate of the electrocatalytic transformation. Thus far, electrochemical C−H activation was severely limited to C(sp2)−H transformations. In sharp contrast, the use of electricity in palladium-catalyzed C(sp3)−H oxygenation was achieved by Mei (Scheme 10),107 whichto the best of our knowledgerepresented the first C(sp3)−H activation by electrochemical transition-metal catalysis. Mei achieved the desired reaction with Pd(OAc)2 as the catalyst in a divided cell setup, using carboxylic acids as the solvent and the corresponding sodium salts as the base. The broad scope as to the substitution pattern at the oxime 29 is noteworthy. Valuable functional groups, such as ester, nitrile, and halides, were also well tolerated. Likewise, the enantiomerically enriched oxazoline 32 underwent the C−H oxygenation without racemization of the stereogenic center. The reaction mechanism was proposed to initiate by a proximity-induced base-assisted C−H metalation at the neighboring primary C(sp3)−H bond (Scheme 11). The

featured a good tolerance of valuable functional groups, such as ester and bromide substituents.90 More recently, a C−H phosphorylation was disclosed by Budnikova by means of electrocatalysis.92,93 The generated products are of considerable interest for their coordination abilities and their potential application as ligands. In these mechanistic studies, crucial intermediates could be isolated and were shown to be competent in stoichiometric settings. Moreover, detailed CV studies provided strong support for the involvement of a palladium(IV) species within the catalytic cycle. Budnikova also reported on the perfluoroalkoxylation of phenylpyridines 12 by means of electrochemistry, during which considerable amounts of perfluoroalkylated product were observed as a byproduct at higher currents.89 The same group disclosed an improved protocol, and perfluoroalkyl bromides as well as perfluoroalkylcarboxylic acids 18 proved to be viable substrates likewise.94−96 This approach was suggested not to be limited to palladium catalysis.94,97−99 Until recently, electrochemical palladium-catalyzed C−H activation was largely restricted to strongly coordinating phenylpyridines 12 and electron-rich anilides 1. In contrast, Kakiuchi established C−H chlorinations with synthetically useful electron-withdrawing benzamides 23 by bidentate chelation assistance (Scheme 9a).100 Here, the bidentate 8aminoquinoline directing group (8-AQ)101−104 had to be adjusted, because otherwise side reactions of undesired 7089

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Scheme 10. Palladium-Catalyzed Primary C(sp3)−H Oxygenation

Scheme 12. Palladium-Catalyzed C−H Alkylation

Scheme 13. Proposed Catalytic Cycle for the Electrochemical C−H Methylation of Oximes 36

Scheme 11. Plausible Mechanism for the PalladiumCatalyzed C(sp3)−H Oxygenation

Pd(IV) species. This key step could take place either by transmetalation under anodic oxidation conditions with MeBF3K (37) or by the attack of a radical generated in situ from the same reagent at the anode surface. Finally, the product 38 is obtained by the oxidation-induced reductive elimination. Moreover, Mei established the palladium-catalyzed C−H benzoylation of oximes 36 (Scheme 14).108 Yet, the reaction thus far featured 2-oxo-2-phenylacetic acid (41) as the sole coupling partner. Based on his previous finding on C(sp3)−H oxygenation,107 Mei also developed the oxygenation of aromatic C−H bonds with a related catalytic system using oximes 36 as the directing group (Scheme 15a).109 The reaction temperature could be slightly decreased. Ortho-substituted arenes 36 performed somewhat less efficiently. A range of valuable functional groups was tolerated, and likewise alkenes 43 were identified as viable substrates. Independently, Sanford found an electrochemical C−H oxygenation (Scheme 15b), broadening the substrate scope to include quinoline 45 and pyrazine 46 directing groups.110 Electrophilic functionalities, such as esters and nitro

intermediate 33 is then likely oxidized at the anode to palladium(IV) complex 34, which can subsequently undergo reductive elimination to form the C−O bond, followed by a ligand exchange to liberate the desired oxygenated product 30. Besides kinetic isotope effect (KIE) studies (kH/kD ≈ 3.7 in parallel experiments), the authors also conducted detailed CV experiments of their reaction to substantiate the proposed palladium(IV) intermediate 34. Mei also discovered a transformative C−C formation in terms of oxidative C−H methylation with methyltrifluoroborates 37 (Scheme 12a).108 The authors disclosed a wide oxime 36 scope, while the C−C-forming manifold seemed mostly applicable to methylations. A KIE of kH/kD ≈ 1.2 was determined, and the cyclometalated complex 39 was isolated and unambiguously characterized (Scheme 12b). The complex 39 was further shown to be a competent catalyst for the C−H functionalization reaction. With this information in hand, the authors proposed a catalytic cycle (Scheme 13), which was suggested to proceed via base-assisted metalation83 of the oxime substrate 36, followed by the generation of a Pd(III) or 7090

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Scheme 14. Palladium-Catalyzed Oxidative Benzoylation

Scheme 16. Electrochemical Rhodium-Catalyzed C−H Alkenylation

Scheme 15. Palladium-Catalyzed C−H Oxygenation

benzoic acids 51, benzamide 53 and pyrimidylindole 55 were shown to be competent substrates likewise (Scheme 17). Scheme 17. Electrochemical Rhodium-Catalyzed C−H Alkenylation of Benzamide 53 and Indole 55

The rhodium catalyst’s mode of action was investigated in detail (Scheme 18).125 A competition experiment revealed the Scheme 18. Key Mechanistic Findings substituents, were well tolerated, while oximes 29 showed a reduced efficacy under these reaction conditions.

3. RHODIUM-CATALYZED C−H ACTIVATION Rhodium(III)-catalyzed C−H transformations111,112 have proven instrumental for the oxidative synthesis of carbo- and heterocycles, with key contributions by Miura and Fagnou, among others.113−118 Despite major advances,119−124 oxidative rhodium-catalyzed C−H transformations were largely restricted to the use of copper(II) or silver(I) salts as the stoichiometric oxidants. In contrast, Ackermann recently reported on the first electrochemical C−H activation by rhodium catalysis to realize cross-dehydrogenative alkenylations with weakly coordinating benzoic acids 51 (Scheme 16).125 The versatility of the electrochemical rhodium-catalyzed C− H alkenylation includes sensitive ketones and esters.125 It is noteworthy that more sterically hindered ortho-substituted arenes 51 proved to be viable substrates as well. In addition to

slight preference for more electron-rich benzoic acids 51b, being in agreement with a base-assisted internal electrophilic substitution (BIES) mechanism.126−131 Furthermore, in studies with [D]4-MeOD as the solvent, deuteration in the ortho positions of the re-isolated benzoic acid [D]n-51e as well as in the product [D]n-52e was noted.125 In agreement with 7091

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis this finding, a KIE was not observed. Further, the reaction was studied by CV in terms of its microscopic reverse, and calculations on the oxidation potential of the proposed intermediate 61 were performed. Based on these studies, a plausible catalytic cycle was put forward (Scheme 19).125 After the formation of rhodium

Scheme 20. Ruthenium-Catalyzed Electrochemical Indole Synthesis

Scheme 19. Proposed Catalytic Cycle for the RhodiumCatalyzed Electrochemical Alkenylation

Scheme 21. Ruthenium-Catalyzed Electrochemical Isoquinoline Synthesis biscarboxylate complex 57, a facile C−H activation leads to the formation of five-membered rhodacycle 58. Coordination of the alkene 2 and insertion into the rhodium−carbon bond lead to intermediate 60, which after β-hydride elimination and reductive elimination forms the proposed rhodium(I) sandwich complex 61. Ligand exchange and anodic oxidation of the rhodium center regenerate the active catalyst 57 and furnish the desired phthalide 52.

Scheme 22. Key Mechanistic Findings

4. RUTHENIUM-CATALYZED C−H ACTIVATION Ruthenium catalysts feature unique chemoselectivities132−142 for sustainable C−H activation processes.143−146 Intriguing studies on the use of ruthenium catalysts in electrochemical C−H activation were recently reported by Xu using a ruthenium complex in an alkyne annulation by pyrimidylanilines 62,147 which was previously achieved with chemical oxidants (Scheme 20).148 The substrate scope of the electrooxidative reaction was generally broad, tolerating a range of valuable functional groups in various positions.147 Additionally, unsymmetrical internal alkynes 63 were converted with good levels of regiocontrol. Furthermore, preliminary findings indicated that a ruthenium-catalyzed [4+2] annulation of benzylamine 65 to isoquinolines 66 is viable, albeit as of yet with lower catalytic efficiency (Scheme 21). Key mechanistic findings by Xu included a reversible H/D exchange in the product 64 as well as the re-isolated starting material 62 (Scheme 22).147 Additionally, a preference for the more electron-rich substrates 62 was observed in competition experiments. An inhibitory effect of high concentrations of the alkyne 63 was observed, which was rationalized by the reversible formation of an off-cycle ruthenium−alkyne complex 69. Finally, a reaction with stoichiometric amounts of ruthenium complex 67 in the absence of electricity led to

comparable yields, suggesting that electricity is needed to regenerate the active ruthenium(II) catalyst. Hence, the mechanism depicted in Scheme 23 was proposed, commencing with the in situ formation of the active ruthenium(II) carboxylate complex 68. Substrate 62 coordination and subsequent C−H activation lead to the sixmembered ruthenacycle intermediate 71, which can then undergo migratory insertion with the alkyne 63. From the formed intermediate 73, indole 64 is liberated by reductive elimination, followed by anodic oxidation of the ruthenium(0) intermediate 74. The inhibition by the alkyne 63 is rationalized by an off-cycle ruthenium alkyne complex 69, formed by 7092

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Scheme 23. Plausible Mechanism for the Electrochemical Ruthenium-Catalyzed Indole Synthesis

Scheme 25. Ruthenium-Catalyzed Electrochemical C−H/ N−H Functionalization and CV Studies

coordination of the alkyne 63 to the ruthenium(II)carboxylate 68. Concurrently, Ackermann independently developed a ruthenium-catalyzed synthesis of isocoumarins 75 (Scheme 24).149 Particularly, the use of weakly coordinating benzoic Scheme 24. Ruthenium-Catalyzed Electrochemical C−H/ O−H Functionalization

revealing a significant influence of the acid concentration150,151 on the oxidation potential.149 Based on these findings, a plausible catalytic cycle was collaborated (Scheme 26), which includes first the formation of the five-membered ruthenacycle 78 by BIES C−H Scheme 26. Plausible Mechanism for the RutheniumCatalyzed C−H/O−H Annulation

acids 51 is worthy of note,143 featuring for the first time electrochemical C−H activation of substrates being devoid of strong N-coordination. The ruthenium-catalyzed electro-oxidative C−H activation was characterized by high catalytic efficacy and good functional group tolerance that included chloride, bromide, esters, and nitriles.149 Moreover, benzamides 53 were found to be viable substrates, yielding the corresponding isoquinolones 76 (Scheme 25a). Mechanistic studies revealed significant deuterium incorporation in the presence of an isotopically labeled cosolvent and a preference for electron-rich benzoic acids 51, both of which render a BIES-type126−131 C−H activation likely to be operative (Scheme 25b). Additionally, the proposed ruthenium(0) intermediate 77 was independently prepared, and its key oxidation was probed by CV, 7093

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis metalation,126−131 followed by migratory insertion of the alkyne 63.149 The resulting seven-membered ruthenacycle 80 undergoes reductive elimination to generate the ruthenium(0) sandwich complex 77, which upon anodic oxidation liberates the desired product 75 and regenerates the active ruthenium species 68.

The electrochemical cobalt-catalyzed C−H oxygenation proved viable on electron-rich as well as electron-deficient benzamides 81 using various alcohols 84 (Scheme 28).177 Scheme 28. Cobalt-Catalyzed C−H Oxygenation of Benzamides 81

5. COBALT-CATALYZED C−H FUNCTIONALIZATIONS While notable recent progress has been achieved using precious 4d and 5d transition metals in C−H activation, earth-abundant base metals have emerged as cost-effective alternatives in the field (Figure 3).102,103,152−160 While

Valuable functional groups, such as tertiary amines, thioether, ester, ketone, nitrile, bromo, and iodo substituents, were fully tolerated, being a strong testament to the robust nature of the cobalt-catalyzed electrochemical C−H activation regime (Scheme 29).177 It is especially noteworthy that the

Figure 3. Costs of common metals in C−H activation.176

Scheme 29. Electrochemical C−H Oxygenation of Alkenes 89

reductive couplings using 3d metals had been reported earlier,161−168 electrochemical C−H activation had been limited to expensive noble metal catalysts. In sharp contrast, Ackermann identified cobalt complexes as the catalyst of choice in electrochemical C−H activation. While cobaltcatalyzed transformations had previously been realized with chemical oxidants,153,156 usually169 super-stoichiometric amounts of expensive silver(I) or copper(II) salts were needed as terminal oxidants to facilitate these C−H transformations.170−175 In 2017, Ackermann reported on the first electrochemical C−H activation by cost-effective176 cobalt catalysis.177 Here, the cobalt-catalyzed C−H oxygenation of benzamides 81 was realized, using the bidentate pyridine-N-oxide group, which proved to be essential for high catalytic performance, while the 8-AQ was shown to be inferior (Scheme 27).

smooth conversion of oxidation-sensitive benzylic alcohols as well as of citronellol proved viable. Besides aromatic benzamides 81, also alkenes 89 could be C−H oxygenated in a diastereoselective fashion, further highlighting the versatile nature of the cobalt-catalyzed electrochemical C−H oxygenation. Detailed mechanistic studies revealed an inherent preferred reactivity of the more electron-rich arenes 81 within competition experiments (Scheme 30a).177 A minor KIE of kH/kD ≈ 1.05 was calculated, indicating a facile C−H cleavage (Scheme 30b). By detailed CV studies, Ackermann and coworkers showed that the direct oxidation of the cobalt catalyst occurred at a potential of 1.19 VSCE in MeOH (Scheme 30c). This finding provided support for a single electron transfer

Scheme 27. Directing Group Power in C−H Oxygenation

7094

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Scheme 30. Summary of Mechanistic Findings for the Electrochemical Cobalt-Catalyzed C−H Oxygenation

Scheme 32. Electrochemical C−H/N−H Annulation of Benzamides 81

alkenes as amenable substrates. Furthermore, a variety of alkynes 63 could be used, ranging from diversely decorated phenylacetylenes to alkyl-substituted alkynes, with excellent levels of regiocontrol. The reaction also proceeded in nontoxic H2O180 as the sole reaction medium (Scheme 33a), avoiding the use of organic

(SET)178 oxidation of the cobalt salt, followed by an organometallic C−H activation event. Based on these mechanistic findings, a plausible catalytic cycle was proposed (Scheme 31),177 commencing with the

Scheme 33. Electrochemical Alkyne Annulation in H2O

Scheme 31. Plausible Catalytic Cycle of the Electrochemical Cobalt-Catalyzed Oxygenation

electrochemical oxidation of the cobalt(II) salt to generate the active cobalt(III) catalyst 91, along with the subsequent C−H metalation of the benzamide 81. The C−O formation occurs in a chelation-assisted fashion, and proto-demetalation releases the final product 94. The versatile cobalt-catalyzed C−H activation catalyst was not limited to C−O construction. Indeed, Ackermann highlighted that challenging C−C and C−N bonds could be formed by earth-abundant base metal-catalyzed C−H activation with electricity as the sole oxidant. The proof of concept was provided by a cobalt-catalyzed synthesis of isoquinolones 95 (Scheme 32).179 The versatility of the cobalt catalysis regime was reflected by benzamides, heterocycles, and

solvents in an environmentally benign fashion.179 Furthermore, detailed mechanistic studies unraveled a preference of the more electron-rich substrate for both coupling partners, being in good agreement with a BIES-type126−131 mechanism (Scheme 33b). The mechanism of this C−C/N−C formation was proposed to proceed via chelation-assisted C−H metalation, followed by the regioselective insertion of the alkyne 63 into the C−Co bond, to generate the seven-membered intermediate 98.179 The product is then liberated by reductive elimination 7095

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis followed by SET anodic oxidation to regenerate the active cobalt catalyst 96 (Scheme 34).

Scheme 36. Plausible Catalytic Cycle for the Electrochemical C−H Activation/Annulation Using Ethylene (99)

Scheme 34. Proposed Mechanism for the Cobalt-Catalyzed C−H Activation/[4+2]Annulation

Scheme 37. Electrochemical Cobalt-Catalyzed C−H Amination

This approach was thereafter extended by Lei to include the challenging annulation of gaseous ethylene (99) and ethyne (100) (Scheme 35).181 The obtained yields were found largely Scheme 35. Electrochemical C−H Activation/Annulation Using Ethylene (99) or Ethyne (100)

Kinetic experiments revealed a facile C−H cleavage, and CV studies supported a SET-type oxidation of cobalt(II) to cobalt(III). A reaction profile was elaborated by detailed in situ ReactIR technology (Scheme 38a) and clearly showed the absence of a significant initiation period. Furthermore, the formation of molecular hydrogen as the sole byproduct by cathodic reduction was confirmed by headspace gas chromatographic (GC) analysis (Scheme 38b).182 Hence, a proposed catalytic cycle featured the generation of the active catalyst 109 by anodic oxidation, along with chelation-assisted C−H metalation and C−N formation by reductive elimination (Scheme 39).182 In an independent report, Lei disclosed a similar protocol for electrochemical C−H amination, using the 8-AQ177 directing group (Scheme 40).184 Although an elevated reaction temperature of 65 °C and a higher catalyst loading of 20

moderate to good, and the general catalytic system featured minor variations from the Ackermann conditions.179 The report illustrated detailed kinetic analysis, indicating the reaction rate as only being dependent on the operating current. Thereby, a plausible reaction mechanism was proposed (Scheme 36).181 The electrochemical cobalt-catalyzed C−N formation was not limited to an intramolecular179 scenario. Ackermann disclosed the C−H amination of benzamides 81 by electrochemical cobalt catalysis (Scheme 37).182 This report not only highlighted the first electrochemical C−H amination but also described the first use of a renewable biomass-derived solvent in electrocatalysis.183 The amination proceeded smoothly with secondary amines 107 and tolerated aryl halides, as well as esters and various heterocyclic substrates. 7096

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis

mol % were used, the catalyst performance was inferior.184 Notably, ortho-substituted benzamides 23 were efficiently converted despite their increased steric bulk.184 A plausible catalytic cycle was proposed, based on KIE experiments, which revealed a KIE of kH/kD ≈ 1.13 for intermolecular reactions and 1.70 for the intramolecular reaction (Scheme 41).184

Scheme 38. Kinetic Profile and Headspace GC Analysis

Scheme 41. KIE Experiments for the Electrochemical Cobalt-Catalyzed C−H Amination

Scheme 39. Proposed Catalytic Cycle for the CobaltCatalyzed C−H Amination

Chronoamperometric studies suggested an initial formation of cobalt(III). The reaction is initiated either by the formation of the active cobalt(III) species by anodic oxidation or by an oxidation of the cobalt(II)−substrate complex, both of which result in the same intermediate 115. The intermediate 115 is proposed to undergo base-assisted C−H cobaltation, resulting in the formation of complex 116. From this species 116, formation of the C−N bond is possible by reductive elimination, generating cobalt(I), which is subsequently reoxidized at the anode (Scheme 42).184 Scheme 42. Plausible Mechanism for the Electrochemical Cobalt-Catalyzed Amination of Quinolinamides 23

Scheme 40. Electrochemical Cobalt-Catalyzed Amination of Quinolinyl Amides 23

6. CONCLUSION Electrochemical metal-catalyzed C−H activation has emerged as a powerful tool in molecular synthesis. Thus, palladium catalysts set the stage for electro-oxidative Fujiwara−Moritani reactions as well as C−H oxygenations, methylations, and halogenations. An early report featured the use of benzo7097

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis

(6) Riekert, L. The efficiency of energy-utilization in chemical processes. Chem. Eng. Sci. 1974, 29, 1613−1620. (7) Denbigh, K. G. The second-law efficiency of chemical processes. Chem. Eng. Sci. 1956, 6, 1−9. (8) Ciamician, G. The Photochemistry of the Future. Science 1912, 36, 385−394. (9) Sharma, U. K.; Gemoets, H. P. L.; Schröder, F.; Noël, T.; Van der Eycken, E. V. Merger of Visible-Light Photoredox Catalysis and C−H Activation for the Room-Temperature C-2 Acylation of Indoles in Batch and Flow. ACS Catal. 2017, 7, 3818−3823. (10) Fabry, D. C.; Rueping, M. Merging Visible Light Photoredox Catalysis with Metal Catalyzed C−H Activations: On the Role of Oxygen and Superoxide Ions as Oxidants. Acc. Chem. Res. 2016, 49, 1969−1979. (11) Ghosh, I.; Marzo, L.; Das, A.; Shaikh, R.; König, B. Visible Light Mediated Photoredox Catalytic Arylation Reactions. Acc. Chem. Res. 2016, 49, 1566−1577. (12) Karkas, 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. (13) Ravelli, D.; Protti, S.; Fagnoni, M. Carbon-Carbon Bond Forming Reactions via Photogenerated Intermediates. Chem. Rev. 2016, 116, 9850−9913. (14) Shaw, M. H.; Twilton, J.; MacMillan, D. W. Photoredox Catalysis in Organic Chemistry. J. Org. Chem. 2016, 81, 6898−6926. (15) Romero, N. A.; Nicewicz, D. A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116, 10075−10166. (16) Skubi, K. L.; Blum, T. R.; Yoon, T. P. Dual Catalysis Strategies in Photochemical Synthesis. Chem. Rev. 2016, 116, 10035−10074. (17) Qin, Q.; Jiang, H.; Hu, Z.; Ren, D.; Yu, S. Functionalization of C−H Bonds by Photoredox Catalysis. Chem. Rec. 2017, 17, 754−774. (18) Ghosh, I.; Ghosh, T.; Bardagi, J. I.; König, B. Reduction of aryl halides by consecutive visible light-induced electron transfer processes. Science 2014, 346, 725−728. (19) Hari, D. P.; König, B. Synthetic applications of eosin Y in photoredox catalysis. Chem. Commun. 2014, 50, 6688−6699. (20) SMARD, Bundesnetzagentur. Electricity Market Data, https:// www.smard.de/home, accessed on April 23, 2018. (21) Federal Ministry for Economic Affairs and Energy, Bundesministerium für Wirtschaft und Energie. Renewable Energy, https://www.bmwi.de/Redaktion/EN/Dossier/renewable-energy. html, accessed April 19, 2018. (22) Gandeepan, P.; Ackermann, L. Transient Directing Groups for Transformative C−H Activation by Synergistic Metal Catalysis. Chem. 2018, 4, 199−222. (23) (a) Park, Y.; Kim, Y.; Chang, S. Transition Metal-Catalyzed C− H Amination: Scope, Mechanism, and Applications. Chem. Rev. 2017, 117, 9247−9301. (b) 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. (c) Jerhaoui, S.; Poutrel, P.; Djukic, J. P.; Wencel-Delord, J.; Colobert, F. Stereospecific C−H activation as a key step for the asymmetric synthesis of various biologically active cyclopropanes. Org. Chem. Front. 2018, 5, 409−414. (d) Segawa, Y.; Maekawa, T.; Itami, K. Synthesis of extended pi-systems through C−H activation. Angew. Chem., Int. Ed. 2015, 54, 66−81. (e) 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. (f) Hickman, A. J.; Sanford, M. S. High-valent organometallic copper and palladium in catalysis. Nature 2012, 484, 177−185. (g) 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. (h) Baudoin, O. Transition metal-catalyzed arylation of unactivated C(sp3)−H bonds. Chem. Soc. Rev. 2011, 40, 4902−4911. (i) 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. (j) Davies, H. M.; Manning, J. R. Catalytic C−H functionalization by metal carbenoid and nitrenoid insertion. Nature 2008, 451, 417−424.

quinone as redox mediator in palladium(II/0) catalysis, while subsequent studies have illustrated the power of palladium(III) and/or palladium(IV) intermediates for electro-oxidative C−H activation catalysis. Despite significant advances in C(sp2)−H and C(sp3)−H activation, the palladium catalysts were restricted to strongly N-coordinating directing groups. This limitation was recently successfully addressed by means of rhodium(III) and cost-effective ruthenium(II) catalysis. These strategies enabled the use of challenging, weakly coordinating amides and acids. Until very recently, electrochemical C−H activation was realized by precious 4d transition-metal catalysts. However, very recent findings have indicated the unique potential of inexpensive, earth-abundant cobalt catalysis. Hence, C−H oxygenations, C−H nitrogenations, and [4+2] alkyne annulations by C−H/N−H functionalization proved viable in a most user-friendly, undivided cell setup. The chemoselectivity and sustainability of the cobalt-catalyzed C− H activation were, among others, reflected by C−H functionalizations in nontoxic H2O and biomass-derived reaction media, respectively. Thereby, the use of stoichiometric amounts of chemical oxidants, such as expensive and toxic silver(I) salts, could be prevented. Given the sustainable nature of electrochemical C−H activation, notable further progress is expected in this rapidly evolving research area. Thus, further advances using base metals other than cobalt161,162 and stereoselective185−191 C−H transformations are particularly in high demand. The same holds true for addressing the catalyst performance, the viable solvents, suitable supporting electrolytes, and the development of effective scale-up protocols.192



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Lutz Ackermann: 0000-0001-7034-8772 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Generous support by the DFG (Gottfried-Wilhelm-Leibniz award) and the European Research Council under the Seventh Framework Program of the European Community (FP720072013, ERC Grant Agreement No. 307535) is gratefully acknowledged.



REFERENCES

(1) Lanzafame, P.; Abate, S.; Ampelli, C.; Genovese, C.; Passalacqua, R.; Centi, G.; Perathoner, S. Beyond Solar Fuels: Renewable EnergyDriven Chemistry. ChemSusChem 2017, 10, 4409−4419. (2) Schlögl, R. Chemistry’s role in regenerative energy. Angew. Chem., Int. Ed. 2011, 50, 6424−6426. (3) Schlögl, R. The role of chemistry in the energy challenge. ChemSusChem 2010, 3, 209−222. (4) Gronnow, M. J.; White, R. J.; Clark, J. H.; Macquarrie, D. J. Energy Efficiency in Chemical Reactions: A Comparative Study of Different Reaction Techniques. Org. Process Res. Dev. 2005, 9, 516− 518. (5) Mata, T. M.; Smith, R. L.; Young, D. M.; Costa, C. A. V. Evaluating the environmental friendliness, economics and energy efficiency of chemical processes: heat integration. Clean Technol. Environ. Policy 2003, 5, 302−309. 7098

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis

Electrochemical C−H Amination. Chem. - Eur. J. 2017, 23, 12096− 12099. (45) 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. (46) Wiebe, A.; Schollmeyer, D.; Dyballa, K. M.; Franke, R.; Waldvogel, S. R. Selective Synthesis of Partially Protected Nonsymmetric Biphenols by Reagent- and Metal-Free Anodic CrossCoupling Reaction. Angew. Chem., Int. Ed. 2016, 55, 11801−11805. (47) Lips, S.; Wiebe, A.; Elsler, B.; Schollmeyer, D.; Dyballa, K. M.; Franke, R.; Waldvogel, S. R. Synthesis of meta-Terphenyl-2,2″−diols by Anodic C−C Cross-Coupling Reactions. Angew. Chem., Int. Ed. 2016, 55, 10872−10876. (48) Elsler, B.; Wiebe, A.; Schollmeyer, D.; Dyballa, K. M.; Franke, R.; Waldvogel, S. R. Source of Selectivity in Oxidative Cross-Coupling of Aryls by Solvent Effect of 1,1,1,3,3,3-Hexafluoropropan-2-ol. Chem. - Eur. J. 2015, 21, 12321−12325. (49) Kirste, A.; Elsler, B.; Schnakenburg, G.; Waldvogel, S. R. Efficient Anodic and Direct Phenol-Arene C,C Cross-Coupling: The Benign Role of Water or Methanol. J. Am. Chem. Soc. 2012, 134, 3571−3576. (50) 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. (51) 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. (52) 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. (53) 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. (54) O’Brien, A. G.; Maruyama, A.; Inokuma, Y.; Fujita, M.; Baran, P. S.; Blackmond, D. G. Radical C−H Functionalization of Heteroarenes under Electrochemical Control. Angew. Chem., Int. Ed. 2014, 53, 11868−11871. (55) Hayashi, R.; Shimizu, A.; Yoshida, J.-i. The Stabilized Cation Pool Method: Metal- and Oxidant-Free Benzylic C−H/Aromatic C− H Cross-Coupling. J. Am. Chem. Soc. 2016, 138, 8400−8403. (56) Morofuji, T.; Shimizu, A.; Yoshida, J.-i. Heterocyclization Approach for Electrooxidative Coupling of Functional Primary Alkylamines with Aromatics. J. Am. Chem. Soc. 2015, 137, 9816− 9819. (57) Morofuji, T.; Shimizu, A.; Yoshida, J.-i. Direct C−N Coupling of Imidazoles with Aromatic and Benzylic Compounds via Electrooxidative C−H Functionalization. J. Am. Chem. Soc. 2014, 136, 4496− 4499. (58) Morofuji, T.; Shimizu, A.; Yoshida, J.-i. Electrochemical C−H Amination: Synthesis of Aromatic Primary Amines via N-Arylpyridinium Ions. J. Am. Chem. Soc. 2013, 135, 5000−5003. (59) Ashikari, Y.; Shimizu, A.; Nokami, T.; Yoshida, J.-i. Halogen and Chalcogen Cation Pools Stabilized by DMSO. Versatile Reagents for Alkene Difunctionalization. J. Am. Chem. Soc. 2013, 135, 16070− 16073. (60) Morofuji, T.; Shimizu, A.; Yoshida, J. i. Metal- and ChemicalOxidant-Free C−H/C−H Cross-Coupling of Aromatic Compounds: The Use of Radical-Cation Pools. Angew. Chem., Int. Ed. 2012, 51, 7259−7262. (61) Ashikari, Y.; Nokami, T.; Yoshida, J.-i. Integrated Electrochemical−Chemical Oxidation Mediated by Alkoxysulfonium Ions. J. Am. Chem. Soc. 2011, 133, 11840−11843. (62) Xiong, P.; Xu, H.-H.; Song, J.; Xu, H.-C. Electrochemical Difluoromethylarylation of Alkynes. J. Am. Chem. Soc. 2018, 140, 2460−2464. (63) Zhao, H. B.; Liu, Z. J.; Song, J.; Xu, H. C. Reagent-Free C−H/ N−H Cross-Coupling: Regioselective Synthesis of N-Heteroaro-

(k) Bergman, R. G. Organometallic chemistry: C−H activation. Nature 2007, 446, 391−393 and cited references. (24) He, J.; Wasa, M.; Chan, K. S. L.; Shao, Q.; Yu, J.-Q. PalladiumCatalyzed Transformations of Alkyl C−H Bonds. Chem. Rev. 2017, 117, 8754−8786. (25) Hickman, A. J.; Sanford, M. S. High-valent organometallic copper and palladium in catalysis. Nature 2012, 484, 177−185. (26) Chen, X.; Engle, K. M.; Wang, D. H.; Yu, J.-Q. Palladium(II)Catalyzed C−H Activation/C−C Cross-Coupling Reactions: Versatility and Practicality. Angew. Chem., Int. Ed. 2009, 48, 5094−5115. (27) (a) Mayr, H.; Ofial, A. R. Philicities, Fugalities, and Equilibrium Constants. Acc. Chem. Res. 2016, 49, 952−965. (b) Mayr, H.; Ofial, A. R. Do general nucleophilicity scales exist? J. Phys. Org. Chem. 2008, 21, 584−595. (28) Li, C.; Hoffman, M. Z. One-Electron Redox Potentials of Phenols in Aqueous Solution. J. Phys. Chem. B 1999, 103, 6653−6656. (29) (a) Freeman, D. B.; Furst, L.; Condie, A. G.; Stephenson, C. R. Functionally Diverse Nucleophilic Trapping of Iminium Intermediates Generated Utilizing Visible Light. Org. Lett. 2012, 14, 94−97. (b) Coniglio, A.; Galli, C.; Gentili, P.; Vadala, R. Hydrogen atom abstraction from C−H bonds of benzylamides by the aminoxyl radical BTNO: a kinetic study. Org. Biomol. Chem. 2009, 7, 155−160. (c) Luo, Y.-R. Comprehensive Handbook of Chemical Bond Energies, 1st ed.; CRC Press: Boca Raton, FL, 2007. (30) Davico, G. E.; Bierbaum, V. M.; DePuy, C. H.; Ellison, G. B.; Squires, R. R. The C−H Bond Energy of Benzene. J. Am. Chem. Soc. 1995, 117, 2590−2599. (31) Kolbe, H. Untersuchungen über die Elektrolyse organischer Verbindungen. Liebigs. Ann. Chem. 1849, 69, 257−294. (32) Kolbe, H. Zersetzung der Valeriansäure durch den elektrischen Strom. Liebigs Ann. Chem. 1848, 64, 339−341. (33) Wiebe, A.; Gieshoff, T.; Möhle, S.; Rodrigo, E.; Zirbes, M.; Waldvogel, S. R. Electrifying Organic Synthesis. Angew. Chem., Int. Ed. 2018, 57, 5594−5619. (34) Möhle, S.; Zirbes, M.; Rodrigo, E.; Gieshoff, T.; Wiebe, A.; Waldvogel, S. R. Modern Electrochemical Aspects for the Synthesis of Value-Added Organic Products. Angew. Chem., Int. Ed. 2018, 57, 6018−6041. (35) 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. (36) Feng, R.; Smith, J. A.; Moeller, K. D. Anodic Cyclization Reactions and the Mechanistic Strategies That Enable Optimization. Acc. Chem. Res. 2017, 50, 2346−2352. (37) 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. (38) Alfonso-Súarez, P.; Kolliopoulos, A. V.; Smith, J. P.; Banks, C. E.; Jones, A. M. An experimentalist’s guide to electrosynthesis: the Shono oxidation. Tetrahedron Lett. 2015, 56, 6863−6867. (39) Jones, A. M.; Banks, C. E. The Shono-type electroorganic oxidation of unfunctionalised amides. Carbon−carbon bond formation via electrogenerated N-acyliminium ions. Beilstein J. Org. Chem. 2014, 10, 3056−3072. (40) Francke, R.; Little, R. D. Redox catalysis in organic electrosynthesis: basic principles and recent developments. Chem. Soc. Rev. 2014, 43, 2492−2521. (41) Yoshida, J.-i.; Kataoka, K.; Horcajada, R.; Nagaki, A. Modern Strategies in Electroorganic Synthesis. Chem. Rev. 2008, 108, 2265− 2299. (42) Jutand, A. Contribution of Electrochemistry to Organometallic Catalysis. Chem. Rev. 2008, 108, 2300−2347. (43) Wiebe, A.; Lips, S.; Schollmeyer, D.; Franke, R.; Waldvogel, S. R. Single and Twofold Metal- and Reagent-Free Anodic C−C CrossCoupling of Phenols with Thiophenes. Angew. Chem., Int. Ed. 2017, 56, 14727−14731. (44) Wesenberg, L. J.; Herold, S.; Shimizu, A.; Yoshida, J. i.; Waldvogel, S. R. New Approach to 1,4-Benzoxazin-3-ones by 7099

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis matics from Biaryl Aldehydes and NH3. Angew. Chem., Int. Ed. 2017, 56, 12732−12735. (64) Xiong, P.; Xu, H.-H.; Xu, H.-C. Metal- and Reagent-Free Intramolecular Oxidative Amination of Tri- and Tetrasubstituted Alkenes. J. Am. Chem. Soc. 2017, 139, 2956−2959. (65) Wu, Z. J.; Xu, H. C. Synthesis of C3-Fluorinated Oxindoles through Reagent-Free Cross-Dehydrogenative Coupling. Angew. Chem., Int. Ed. 2017, 56, 4734−4738. (66) Folgueiras-Amador, A. A.; Qian, X. Y.; Xu, H. C.; Wirth, T. Catalyst- and Supporting-Electrolyte-Free Electrosynthesis of Benzothiazoles and Thiazolopyridines in Continuous Flow. Chem. - Eur. J. 2018, 24, 487−491. (67) Zhao, H. B.; Hou, Z. W.; Liu, Z. J.; Zhou, Z. F.; Song, J.; Xu, H. C. Amidinyl Radical Formation through Anodic N−H Bond Cleavage and Its Application in Aromatic C−H Bond Functionalization. Angew. Chem., Int. Ed. 2017, 56, 587−590. (68) Hou, Z. W.; Mao, Z. Y.; Zhao, H. B.; Melcamu, Y. Y.; Lu, X.; Song, J.; Xu, H. C. Electrochemical C−H/N−H Functionalization for the Synthesis of Highly Functionalized (Aza)indoles. Angew. Chem., Int. Ed. 2016, 55, 9168−9172. (69) Wang, P.; Tang, S.; Huang, P.; Lei, A. Electrocatalytic OxidantFree Dehydrogenative C−H/S−H Cross-Coupling. Angew. Chem., Int. Ed. 2017, 56, 3009−3013. (70) Liu, K.; Tang, S.; Huang, P.; Lei, A. External oxidant-free electrooxidative [3+2] annulation between phenol and indole derivatives. Nat. Commun. 2017, 8, 775. (71) 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. (72) Jiang, Y.-Y.; Wang, Q.-Q.; Liang, S.; Hu, L.-M.; Little, R. D.; Zeng, C.-C. Electrochemical Oxidative Amination of Sodium Sulfinates: Synthesis of Sulfonamides Mediated by NH4I as a Redox Catalyst. J. Org. Chem. 2016, 81, 4713−4719. (73) Francke, R.; Little, R. D. Optimizing electron transfer mediators based on arylimidazoles by ring fusion: synthesis, electrochemistry, and computational analysis of 2-aryl-1-methylphenanthro[9,10-d]imidazoles. J. Am. Chem. Soc. 2014, 136, 427−435. (74) (a) Nielsen, M. F.; Batanero, B.; Löhl, T.; Schäfer, H. J.; Würthwein, E. U.; Fröhlich, R. Enantioselective Cathodic Reduction of 4-Methylcoumarin: Dependence of Selectivity on Reaction Conditions and Investigation of the Mechanism. Chem. - Eur. J. 1997, 3, 2011−2024. (b) Klotz-Berendes, B.; Schäfer, H. J.; Grehl, M.; Fröhlich, R. Diastereoselective Coupling of Anodically Generated Radicals Bearing Chiral Amide Groups. Angew. Chem., Int. Ed. Engl. 1995, 34, 189−191. (75) Qiu, Y.; Struwe, J.; Meyer, T. H.; Oliveira, J. C. A.; Ackermann, L. Catalyst- and Reagent-free Electrochemical Azole C−H Amination. Chem. - Eur. J. 2018, DOI: 10.1002/chem.201802832. (76) Shen, K.; Fu, Y.; Li, J.-N.; Liu, L.; Guo, Q.-X. What are the pKa values of C−H bonds in aromatic heterocyclic compounds in DMSO? Tetrahedron 2007, 63, 1568−1576. (77) Average prices calculated on the basis of the largest units available at ABCR, Sigma-Aldrich, and Fisher Scientific. (78) (a) Yang, Q.-L.; Fang, P.; Mei, T.-S. Recent Advances in Organic Electrochemical C−H Functionalization. Chin. J. Chem. 2018, 36, 338−352. (b) Jiao, K.-J.; Zhao, C.-Q.; Fang, P.; Mei, T.-S. Palladium catalyzed C−H functionalization with electrochemical oxidation. Tetrahedron Lett. 2017, 58, 797−802. (c) Dudkina, Y. B.; Gryaznova, T. V.; Sinyashin, O. G.; Budnikova, Y. H. Ligand-directed electrochemical functionalization of C(sp2)−H bonds in the presence of the palladium and nickel compounds. Russ. Chem. Bull. 2015, 64, 1713−1725. (79) 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.

(80) Fujiwara, Y.; Moritani, I.; Danno, S.; Asano, R.; Teranishi, S. Aromatic substitution of olefins. VI. Arylation of olefins with palladium(II) acetate. J. Am. Chem. Soc. 1969, 91, 7166−7169. (81) Fujiwara, Y.; Moritani, I.; Matsuda, M.; Teranishi, S. Aromatic substitution of styrene-palladium chloride complex. II effect of metal acetate. Tetrahedron Lett. 1968, 9, 633−636. (82) Moritani, I.; Fujiwara, Y. Aromatic substitution of styrenepalladium chloride complex. Tetrahedron Lett. 1967, 8, 1119−1122. (83) Ackermann, L. Carboxylate-assisted transition-metal-catalyzed C−H bond functionalizations: mechanism and scope. Chem. Rev. 2011, 111, 1315−1345. (84) Kakiuchi, F.; Kochi, T.; Mutsutani, H.; Kobayashi, N.; Urano, S.; Sato, M.; Nishiyama, S.; Tanabe, T. Palladium-Catalyzed Aromatic C−H Halogenation with Hydrogen Halides by Means of Electrochemical Oxidation. J. Am. Chem. Soc. 2009, 131, 11310−11311. (85) Aiso, H.; Kochi, T.; Mutsutani, H.; Tanabe, T.; Nishiyama, S.; Kakiuchi, F. Catalytic Electrochemical C−H Iodination and One-Pot Arylation by ON/OFF Switching of Electric Current. J. Org. Chem. 2012, 77, 7718−7724. (86) Voskressensky, L.; Golantsov, N.; Maharramov, A. Recent Advances in Bromination of Aromatic and Heteroaromatic Compounds. Synthesis 2016, 48, 615−643. (87) Thirunavukkarasu, V. S.; Kozhushkov, S. I.; Ackermann, L. C− H nitrogenation and oxygenation by ruthenium catalysis. Chem. Commun. 2014, 50, 29−39. (88) Dick, A. R.; Sanford, M. S. Transition metal catalyzed oxidative functionalization of carbon−hydrogen bonds. Tetrahedron 2006, 62, 2439−2463. (89) Dudkina, Y. B.; Mikhaylov, D. Y.; Gryaznova, T. V.; Tufatullin, A. I.; Kataeva, O. N.; Vicic, D. A.; Budnikova, Y. H. Electrochemical Ortho Functionalization of 2-Phenylpyridine with Perfluorocarboxylic Acids Catalyzed by Palladium in Higher Oxidation States. Organometallics 2013, 32, 4785−4792. (90) Saito, F.; Aiso, H.; Kochi, T.; Kakiuchi, F. Palladium-Catalyzed Regioselective Homocoupling of Arenes Using Anodic Oxidation: Formal Electrolysis of Aromatic Carbon−Hydrogen Bonds. Organometallics 2014, 33, 6704−6707. (91) Hull, K. L.; Lanni, E. L.; Sanford, M. S. Highly Regioselective Catalytic Oxidative Coupling Reactions: Synthetic and Mechanistic Investigations. J. Am. Chem. Soc. 2006, 128, 14047−14049. (92) Gryaznova, T.; Dudkina, Y.; Khrizanforov, M.; Sinyashin, O.; Kataeva, O.; Budnikova, Y. Electrochemical properties of diphosphonate-bridged palladacycles and their reactivity in arene phosphonation. J. Solid State Electrochem. 2015, 19, 2665−2672. (93) Grayaznova, T. V.; Dudkina, Y. B.; Islamov, D. R.; Kataeva, O. N.; Sinyashin, O. G.; Vicic, D. A.; Budnikova, Y. H. Pyridine-directed palladium-catalyzed electrochemical phosphonation of C(sp2)−H bond. J. Organomet. Chem. 2015, 785, 68−71. (94) Khrizanforov, M.; Strekalova, S.; Khrizanforova, V.; Grinenko, V.; Kholin, K.; Kadirov, M.; Burganov, T.; Gubaidullin, A.; Gryaznova, T.; Sinyashin, O.; Xu, L.; Vicic, D. A.; Budnikova, Y. Iron-catalyzed electrochemical C−H perfluoroalkylation of arenes. Dalton Trans. 2015, 44, 19674−19681. (95) Mikhaylov, D.; Gryaznova, T.; Dudkina, Y.; Khrizanphorov, M.; Latypov, S.; Kataeva, O.; Vicic, D. A.; Sinyashin, O. G.; Budnikova, Y. Electrochemical nickel-induced fluoroalkylation: synthetic, structural and mechanistic study. Dalton Trans. 2012, 41, 165−172. (96) Dudkina, Y. B.; Mikhaylov, D. Y.; Gryaznova, T. V.; Sinyashin, O. G.; Vicic, D. A.; Budnikova, Y. H. MII/MIII-Catalyzed orthoFluoroalkylation of 2-Phenylpyridine. Eur. J. Org. Chem. 2012, 2114− 2117. (97) Khrizanforov, M. N.; Strekalova, S. O.; Grinenko, V. V.; Khrizanforova, V. V.; Gryaznova, T. V.; Budnikova, Y. H. Fe and Nicatalyzed electrochemical perfluoroalkylation of C−H bonds of coumarins. Russ. Chem. Bull. 2017, 66, 1446−1449. (98) Khrizanforov, M. N.; Fedorenko, S. V.; Strekalova, S. O.; Kholin, K. V.; Mustafina, A. R.; Zhilkin, M. Y.; Khrizanforova, V. V.; Osin, Y. N.; Salnikov, V. V.; Gryaznova, T. V.; Budnikova, Y. H. A Ni(III) complex stabilized by silica nanoparticles as an efficient 7100

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis nanoheterogeneous catalyst for oxidative C−H fluoroalkylation. Dalton Trans. 2016, 45, 11976−11982. (99) Dudkina, Y. B.; Mikhailov, D. Y.; Gryaznova, T. V.; Fattakhov, S. G.; Budnikova, Y. G.; Sinyashin, O. G. Electrocatalytic fluoroalkylation of olefins. Nickel-catalyzed polyfluoroalkylation of allylisocyanurates. Russ. Chem. Bull. 2013, 62, 2362−2366. (100) Konishi, M.; Tsuchida, K.; Sano, K.; Kochi, T.; Kakiuchi, F. Palladium-Catalyzed ortho-Selective C−H Chlorination of Benzamide Derivatives under Anodic Oxidation Conditions. J. Org. Chem. 2017, 82, 8716−8724. (101) Kommagalla, Y.; Chatani, N. Cobalt(II)-catalyzed C−H functionalization using an N,N′-bidentate directing group. Coord. Chem. Rev. 2017, 350, 117−135. (102) Daugulis, O.; Roane, J.; Tran, L. D. Bidentate, Monoanionic Auxiliary-Directed Functionalization of Carbon−Hydrogen Bonds. Acc. Chem. Res. 2015, 48, 1053−1064. (103) 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. (104) Zaitsev, V. G.; Shabashov, D.; Daugulis, O. Highly Regioselective Arylation of sp3 C−H Bonds Catalyzed by Palladium Acetate. J. Am. Chem. Soc. 2005, 127, 13154−13155. (105) Proctor, A. E.; Thompson, L. A.; O’Bryant, C. L. Vismodegib: an inhibitor of the Hedgehog signaling pathway in the treatment of basal cell carcinoma. Ann. Pharmacother. 2014, 48, 99−106. (106) Dessinioti, C.; Plaka, M.; Stratigos, A. J. Vismodegib for the treatment of basal cell carcinoma: results and implications of the ERIVANCE BCC trial. Future Oncol. 2014, 10, 927−936. (107) Yang, Q.-L.; Li, Y.-Q.; Ma, C.; Fang, P.; Zhang, X.-J.; Mei, T.S. Palladium-Catalyzed C(sp3)−H Oxygenation via Electrochemical Oxidation. J. Am. Chem. Soc. 2017, 139, 3293−3298. (108) Ma, C.; Zhao, C.-Q.; Li, Y.-Q.; Zhang, L.-P.; Xu, X.-T.; Zhang, K.; Mei, T.-S. Palladium-catalyzed C−H activation/C−C crosscoupling reactions via electrochemistry. Chem. Commun. 2017, 53, 12189−12192. (109) Li, Y.-Q.; Yang, Q.-L.; Fang, P.; Mei, T.-S.; Zhang, D. Palladium-Catalyzed C(sp2)−H Acetoxylation via Electrochemical Oxidation. Org. Lett. 2017, 19, 2905−2908. (110) Shrestha, A.; Lee, M.; Dunn, A. L.; Sanford, M. S. PalladiumCatalyzed C−H Bond Acetoxylation via Electrochemical Oxidation. Org. Lett. 2018, 20, 204−207. (111) Colby, D. A.; Tsai, A. S.; Bergman, R. G.; Ellman, J. A. Rhodium Catalyzed Chelation-Assisted C−H Bond Functionalization Reactions. Acc. Chem. Res. 2012, 45, 814−825. (112) Colby, D. A.; Bergman, R. G.; Ellman, J. A. RhodiumCatalyzed C−C Bond Formation via Heteroatom-Directed C−H Bond Activation. Chem. Rev. 2010, 110, 624−655. (113) Sperger, T.; Sanhueza, I. A.; Kalvet, I.; Schoenebeck, F. Computational Studies of Synthetically Relevant Homogeneous Organometallic Catalysis Involving Ni, Pd, Ir, and Rh: An Overview of Commonly Employed DFT Methods and Mechanistic Insights. Chem. Rev. 2015, 115, 9532−9586. (114) Song, G.; Li, X. Substrate Activation Strategies in Rhodium(III)-Catalyzed Selective Functionalization of Arenes. Acc. Chem. Res. 2015, 48, 1007−1020. (115) Shin, K.; Kim, H.; Chang, S. Transition-Metal-Catalyzed C−N Bond Forming Reactions Using Organic Azides as the Nitrogen Source: A Journey for the Mild and Versatile C−H Amination. Acc. Chem. Res. 2015, 48, 1040−1052. (116) Kuhl, N.; Schröder, N.; Glorius, F. Formal SN-Type Reactions in Rhodium(III)-Catalyzed C−H Bond Activation. Adv. Synth. Catal. 2014, 356, 1443−1460. (117) Satoh, T.; Miura, M. Oxidative coupling of aromatic substrates with alkynes and alkenes under rhodium catalysis. Chem. - Eur. J. 2010, 16, 11212−11222. (118) Fagnou, K.; Lautens, M. Rhodium-Catalyzed Carbon−Carbon Bond Forming Reactions of Organometallic Compounds. Chem. Rev. 2003, 103, 169−196.

(119) Lu, Y.; Wang, H. W.; Spangler, J. E.; Chen, K.; Cui, P. P.; Zhao, Y.; Sun, W. Y.; Yu, J. Q. Rh(III)-catalyzed C−H olefination of N-pentafluoroaryl benzamides using air as the sole oxidant. Chem. Sci. 2015, 6, 1923−1927. (120) Archambeau, A.; Rovis, T. Rhodium(III)-Catalyzed Allylic C(sp3)−H Activation of Alkenyl Sulfonamides: Unexpected Formation of Azabicycles. Angew. Chem., Int. Ed. 2015, 54, 13337−13340. (121) Zhang, G.; Yu, H.; Qin, G.; Huang, H. Rh-catalyzed oxidative C−H activation/annulation: converting anilines to indoles using molecular oxygen as the sole oxidant. Chem. Commun. 2014, 50, 4331−4334. (122) Zhang, G.; Yang, L.; Wang, Y.; Xie, Y.; Huang, H. An efficient Rh/O2 catalytic system for oxidative C−H activation/annulation: evidence for Rh(I) to Rh(III) oxidation by molecular oxygen. J. Am. Chem. Soc. 2013, 135, 8850−8853. (123) Yu, D. G.; Suri, M.; Glorius, F. Rh(III)/Cu(II)-cocatalyzed synthesis of 1H-indazoles through C−H amidation and N-N bond formation. J. Am. Chem. Soc. 2013, 135, 8802−8805. (124) Stuart, D. R.; Alsabeh, P.; Kuhn, M.; Fagnou, K. Rhodium(III)-Catalyzed Arene and Alkene C−H Bond Functionalization Leading to Indoles and Pyrroles. J. Am. Chem. Soc. 2010, 132, 18326− 18339. (125) Qiu, Y.; Kong, W.-J.; Struwe, J.; Sauermann, N.; Rogge, T.; Scheremetjew, A.; Ackermann, L. Electrooxidative Rhodium-Catalyzed C−H/C−H Acctivation: Electricity as Oxidant for CrossDehydrogenative Alkenylation. Angew. Chem., Int. Ed. 2018, 57, 5828−5832. (126) Tan, E.; Quinonero, O.; de Orbe, M. E.; Echavarren, A. M. Broad-Scope Rh-Catalyzed Inverse-Sonogashira Reaction Directed by Weakly Coordinating Groups. ACS Catal. 2018, 8, 2166−2172. (127) Zell, D.; Bursch, M.; Muller, V.; Grimme, S.; Ackermann, L. Full Selectivity Control in Cobalt(III)-Catalyzed C−H Alkylations by Switching of the C−H Activation Mechanism. Angew. Chem., Int. Ed. 2017, 56, 10378−10382. (128) Wang, H.; Moselage, M.; González, M. J.; Ackermann, L. Selective Synthesis of Indoles by Cobalt(III)-Catalyzed C−H/N−O Functionalization with Nitrones. ACS Catal. 2016, 6, 2705−2709. (129) Santrač, D.; Cella, S.; Wang, W.; Ackermann, L. PalladiumCatalyzed C−H Arylation of Amides by Triazole Assistance. Eur. J. Org. Chem. 2016, 2016, 5429−5436. (130) Mei, R.; Loup, J.; Ackermann, L. Oxazolinyl-Assisted C−H Amidation by Cobalt(III) Catalysis. ACS Catal. 2016, 6, 793−797. (131) Ma, W.; Mei, R.; Tenti, G.; Ackermann, L. Ruthenium(II)catalyzed oxidative C−H alkenylations of sulfonic acids, sulfonyl chlorides and sulfonamides. Chem. - Eur. J. 2014, 20, 15248−15251. (132) Korvorapun, K.; Kaplaneris, N.; Rogge, T.; Warratz, S.; Stückl, A. C.; Ackermann, L. Sequential meta-/ortho-C−H Functionalizations by One-Pot Ruthenium(II/III) Catalysis. ACS Catal. 2018, 8, 886− 892. (133) Fumagalli, F.; Warratz, S.; Zhang, S. K.; Rogge, T.; Zhu, C.; Stuckl, A. C.; Ackermann, L. Arene-Ligand-Free Ruthenium(II/III) Manifold for meta-C−H Alkylation: Remote Purine Diversification. Chem. - Eur. J. 2018, 24, 3984−3988. (134) Warratz, S.; Burns, D. J.; Zhu, C.; Korvorapun, K.; Rogge, T.; Scholz, J.; Jooss, C.; Gelman, D.; Ackermann, L. meta-C−H Bromination on Purine Bases by Heterogeneous Ruthenium Catalysis. Angew. Chem., Int. Ed. 2017, 56, 1557−1560. (135) Ruan, Z.; Zhang, S. K.; Zhu, C.; Ruth, P. N.; Stalke, D.; Ackermann, L. Ruthenium(II)-Catalyzed meta C−H Mono- and Difluoromethylations by Phosphine/Carboxylate Cooperation. Angew. Chem., Int. Ed. 2017, 56, 2045−2049. (136) Paterson, A. J.; Heron, C. J.; McMullin, C. L.; Mahon, M. F.; Press, N. J.; Frost, C. G. α-Halo carbonyls enable meta selective primary, secondary and tertiary C−H alkylations by ruthenium catalysis. Org. Biomol. Chem. 2017, 15, 5993−6000. (137) Li, J.; Korvorapun, K.; De Sarkar, S.; Rogge, T.; Burns, D. J.; Warratz, S.; Ackermann, L. Ruthenium(II)-catalysed remote C−H alkylations as a versatile platform to meta-decorated arenes. Nat. Commun. 2017, 8, 15430. 7101

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis (138) Leitch, J. A.; McMullin, C. L.; Paterson, A. J.; Mahon, M. F.; Bhonoah, Y.; Frost, C. G. Ruthenium-Catalyzed para-Selective C−H Alkylation of Aniline Derivatives. Angew. Chem., Int. Ed. 2017, 56, 15131−15135. (139) Kumar, N. Y.; Bechtoldt, A.; Raghuvanshi, K.; Ackermann, L. Ruthenium(II)-Catalyzed Decarboxylative C−H Activation: Versatile Routes to meta-Alkenylated Arenes. Angew. Chem., Int. Ed. 2016, 55, 6929−6932. (140) Paterson, A. J.; St John-Campbell, S.; Mahon, M. F.; Press, N. J.; Frost, C. G. Catalytic meta-selective C−H functionalization to construct quaternary carbon centres. Chem. Commun. 2015, 51, 12807−12810. (141) Yang, F.; Rauch, K.; Kettelhoit, K.; Ackermann, L. Aldehydeassisted ruthenium(II)-catalyzed C−H oxygenations. Angew. Chem., Int. Ed. 2014, 53, 11285−11288. (142) Hofmann, N.; Ackermann, L. meta-Selective C−H bond alkylation with secondary alkyl halides. J. Am. Chem. Soc. 2013, 135, 5877−5884. (143) De Sarkar, S.; Liu, W.; Kozhushkov, S.; Ackermann, L. Weakly Coordinating Directing Groups for Ruthenium(II)-Catalyzed C−H Activation. Adv. Synth. Catal. 2014, 356, 1461−1479. (144) Kozhushkov, S. I.; Ackermann, L. Ruthenium-catalyzed direct oxidative alkenylation of arenes through twofold C−H bond functionalization. Chem. Sci. 2013, 4, 886−896. (145) Ackermann, L.; Novák, P.; Vicente, R.; Hofmann, N. Ruthenium-Catalyzed Regioselective Direct Alkylation of Arenes with Unactivated Alkyl Halides through C−H Bond Cleavage. Angew. Chem., Int. Ed. 2009, 48, 6045−6048. (146) Ackermann, L. Catalytic Arylations with Challenging Substrates: From Air-Stable HASPO Preligands to Indole Syntheses and C−H-Bond Functionalizations. Synlett 2007, 2007, 507−526. (147) Xu, F.; Li, Y.-J.; Huang, C.; Xu, H.-C. Ruthenium-Catalyzed Electrochemical Dehydrogenative Alkyne Annulation. ACS Catal. 2018, 8, 3820−3824. (148) Ackermann, L.; Lygin, A. V. Cationic Ruthenium(II) Catalysts for Oxidative C−H/N−H Bond Functionalizations of Anilines with Removable Directing Group: Synthesis of Indoles in Water. Org. Lett. 2012, 14, 764−767. (149) Qiu, Y.; Tian, C.; Massignan, L.; Rogge, T.; Ackermann, L. Electrooxidative Ruthenium-Catalyzed C−H/O−H Annulation by Weak O-Coordination. Angew. Chem., Int. Ed. 2018, 57, 5818−5822. (150) Bechtoldt, A.; Tirler, C.; Raghuvanshi, K.; Warratz, S.; Kornhaass, C.; Ackermann, L. Ruthenium Oxidase Catalysis for SiteSelective C−H Alkenylations with Ambient O2 as the Sole Oxidant. Angew. Chem., Int. Ed. 2016, 55, 264−267. (151) Warratz, S.; Kornhaass, C.; Cajaraville, A.; Niepotter, B.; Stalke, D.; Ackermann, L. Ruthenium(II)-catalyzed C−H activation/ alkyne annulation by weak coordination with O2 as the sole oxidant. Angew. Chem., Int. Ed. 2015, 54, 5513−5517. (152) Hu, Y.; Zhou, B.; Wang, C. Inert C−H Bond Transformations Enabled by Organometallic Manganese Catalysis. Acc. Chem. Res. 2018, 51, 816−827. (153) Yoshino, T.; Matsunaga, S. Pentamethylcyclopentadienyl)cobalt(III)-Catalyzed C−H Bond Functionalization: From Discovery to Unique Reactivity and Selectivity. Adv. Synth. Catal. 2017, 359, 1245−1262. (154) Yamaguchi, J.; Muto, K.; Itami, K. Nickel-Catalyzed Aromatic C−H Functionalization. Top. Curr. Chem. 2016, 374, 55. (155) Wei, D.; Zhu, X.; Niu, J.-L.; Song, M.-P. High-Valent-CobaltCatalyzed C−H Functionalization Based on Concerted MetalationDeprotonation and Single-Electron-Transfer Mechanisms. ChemCatChem 2016, 8, 1242−1263. (156) Moselage, M.; Li, J.; Ackermann, L. Cobalt-Catalyzed C−H Activation. ACS Catal. 2016, 6, 498−525. (157) Liu, W.; Ackermann, L. Manganese-Catalyzed C−H Activation. ACS Catal. 2016, 6, 3743−3752. (158) Cera, G.; Ackermann, L. Iron-Catalyzed C−H Functionalization Processes. Top. Curr. Chem. 2016, 374, 57.

(159) Hirano, K.; Miura, M. Recent Advances in Copper-mediated Direct Biaryl Coupling. Chem. Lett. 2015, 44, 868−873. (160) Daugulis, O.; Kulkarni, A. Direct Conversion of CarbonHydrogen into Carbon-Carbon Bonds by First-Row Transition-Metal Catalysis. Synthesis 2009, 4087−4109. (161) Fu, N.; Sauer, G. S.; Saha, A.; Loo, A.; Lin, S. Metal-catalyzed electrochemical diazidation of alkenes. Science 2017, 357, 575−579. (162) Fu, N.; Sauer, G. S.; Lin, S. Electrocatalytic Radical Dichlorination of Alkenes with Nucleophilic Chlorine Sources. J. Am. Chem. Soc. 2017, 139, 15548−15553. (163) Gomes, P.; Gosmini, C.; Périchon, J. Cobalt-Catalyzed Direct Electrochemical Cross-Coupling between Aryl or Heteroaryl Halides and Allylic Acetates or Carbonates. J. Org. Chem. 2003, 68, 1142− 1145. (164) Gomes, P.; Gosmini, C.; Périchon, J. Cobalt-catalyzed electrochemical vinylation of aryl halides using vinylic acetates. Tetrahedron 2003, 59, 2999−3002. (165) Gomes, P.; Gosmini, C.; Nédélec, J.-Y.; Périchon, J. Electrochemical vinylation of aryl and vinyl halides with acrylate esters catalyzed by cobalt bromide. Tetrahedron Lett. 2002, 43, 5901− 5903. (166) Le Gall, E.; Gosmini, C.; Nédélec, J.-Y.; Périchon, J. Cobaltcatalyzed electrochemical cross-coupling of functionalized phenyl halides with 4-chloroquinoline derivatives. Tetrahedron Lett. 2001, 42, 267−269. (167) Gosmini, C.; Nédélec, J. Y.; Périchon, J. Electrochemical cross-coupling between functionalized aryl halides and 2-chloropyrimidine or 2-chloropyrazine catalyzed by nickel 2,2′-bipyridine complex. Tetrahedron Lett. 2000, 41, 201−203. (168) Gomes, P.; Gosmini, C.; Nédélec, J.-Y.; Périchon, J. Cobalt bromide as catalyst in electrochemical addition of aryl halides onto activated olefins. Tetrahedron Lett. 2000, 41, 3385−3388. (169) Mei, R.; Wang, H.; Warratz, S.; Macgregor, S. A.; Ackermann, L. Cobalt-Catalyzed Oxidase C−H/N−H Alkyne Annulation: Mechanistic Insights and Access to Anticancer Agents. Chem. - Eur. J. 2016, 22, 6759−6763. (170) Nguyen, T. T.; Grigorjeva, L.; Daugulis, O. Cobalt-Catalyzed Coupling of Benzoic Acid C−H Bonds with Alkynes, Styrenes, and 1,3-Dienes. Angew. Chem., Int. Ed. 2018, 57, 1688−1691. (171) Du, C.; Li, P. X.; Zhu, X.; Suo, J. F.; Niu, J. L.; Song, M. P. Mixed Directing-Group Strategy: Oxidative C−H/C−H Bond Arylation of Unactivated Arenes by Cobalt Catalysis. Angew. Chem., Int. Ed. 2016, 55, 13571−13575. (172) Zhang, L. B.; Hao, X. Q.; Liu, Z. J.; Zheng, X. X.; Zhang, S. K.; Niu, J. L.; Song, M. P. Cobalt(II)-Catalyzed C-H Alkynylation/ Annulation with Terminal Alkynes: Selective Access to 3-Methyleneisoindolin-1-one. Angew. Chem., Int. Ed. 2015, 54, 10012−10015. (173) Zhang, J.; Chen, H.; Lin, C.; Liu, Z.; Wang, C.; Zhang, Y. Cobalt-Catalyzed Cyclization of Aliphatic Amides and Terminal Alkynes with Silver-Cocatalyst. J. Am. Chem. Soc. 2015, 137, 12990− 12996. (174) Zhang, L. B.; Hao, X. Q.; Zhang, S. K.; Liu, Z. J.; Zheng, X. X.; Gong, J. F.; Niu, J. L.; Song, M. P. Cobalt-Catalyzed C(sp2)-H Alkoxylation of Aromatic and Olefinic Carboxamides. Angew. Chem., Int. Ed. 2015, 54, 272−275. (175) Grigorjeva, L.; Daugulis, O. Cobalt-Catalyzed, Aminoquinoline-Directed C(sp2)−H Bond Alkenylation by Alkynes. Angew. Chem., Int. Ed. 2014, 53, 10209−10212. (176) InvestmentMine. World Mining News, www.infomine.com/ investment/, accessed April 19, 2018. (177) 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. (178) Kochi, J. K.; Tang, R. T.; Bernath, T. Mechanisms of aromatic substitution. Role of cation-radicals in the oxidative substitution of arenes by cobalt(III). J. Am. Chem. Soc. 1973, 95, 7114−7123. (179) Tian, C.; Massignan, L.; Meyer, T. H.; Ackermann, L. Electrochemical C−H/N−H Activation by Water-Tolerant Cobalt 7102

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103

Review

ACS Catalysis Catalysis at Room Temperature. Angew. Chem., Int. Ed. 2018, 57, 2383−2387. (180) Simon, M. O.; Li, C. J. Green chemistry oriented organic synthesis in water. Chem. Soc. Rev. 2012, 41, 1415−1427. (181) 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. (182) 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. (183) Santoro, S.; Ferlin, F.; Luciani, L.; Ackermann, L.; Vaccaro, L. Biomass-derived solvents as effective media for cross-coupling reactions and C−H functionalization processes. Green Chem. 2017, 19, 1601−1612. (184) 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. (185) Saint-Denis, T. G.; Zhu, R.-Y.; Chen, G.; Wu, Q.-F.; Yu, J.-Q. Enantioselective C(sp3)−H bond activation by chiral transition metal catalysts. Science 2018, 359, eaao4798 DOI: 10.1126/science.aao4798. (186) Newton, C. G.; Wang, S. G.; Oliveira, C. C.; Cramer, N. Catalytic Enantioselective Transformations Involving C−H Bond Cleavage by Transition-Metal Complexes. Chem. Rev. 2017, 117, 8908−8976. (187) Gao, D.-W.; Gu, Q.; Zheng, C.; You, S.-L. Synthesis of Planar Chiral Ferrocenes via Transition-Metal-Catalyzed Direct C−H Bond Functionalization. Acc. Chem. Res. 2017, 50, 351−365. (188) Fu, N.; Li, L.; Yang, Q.; Luo, S. Catalytic Asymmetric Electrochemical Oxidative Coupling of Tertiary Amines with Simple Ketones. Org. Lett. 2017, 19, 2122−2125. (189) Pedroni, J.; Cramer, N. TADDOL-based phosphorus(III)ligands in enantioselective Pd(0)-catalysed C−H functionalisations. Chem. Commun. 2015, 51, 17647−17657. (190) Jensen, K. L.; Franke, P. T.; Nielsen, L. T.; Daasbjerg, K.; Jørgensen, K. A. Anodic Oxidation and Organocatalysis: Direct Regioand Stereoselective Access to meta-Substituted Anilines by αArylation of Aldehydes. Angew. Chem., Int. Ed. 2010, 49, 129−133. (191) Minato, D.; Arimoto, H.; Nagasue, Y.; Demizu, Y.; Onomura, O. Asymmetric electrochemical oxidation of 1,2-diols, aminoalcohols, and aminoaldehydes in the presence of chiral copper catalyst. Tetrahedron 2008, 64, 6675−6683. (192) (a) Pletcher, D.; Green, R. A.; Brown, R. C. D. Flow Electrosynthesis Cells for the Synthetic Organic Chemistry Laboratory. Chem. Rev. 2018, 118, 4573−4591. (b) Atobe, M.; Tateno, H.; Matsumura, Y. Applications of Flow Microreactors in Electrosynthetic Processes. Chem. Rev. 2018, 118, 4541−4572. (c) Folgueiras-Amador, A. A.; Philipps, K.; Guilbaud, S.; Poelakker, J.; Wirth, T. An Easy-toMachine Electrochemical Flow Microreactor: Efficient Synthesis of Isoindolone and Flow Functionalization. Angew. Chem., Int. Ed. 2017, 56, 15446−15450. (d) Gütz, C.; Stenglein, A.; Waldvogel, S. R. Highly Modular Flow Cell for Electroorganic Synthesis. Org. Process Res. Dev. 2017, 21, 771−778.

7103

DOI: 10.1021/acscatal.8b01682 ACS Catal. 2018, 8, 7086−7103