Cobalt-Catalyzed Electrochemical Enantioselective Reductive Cross-Coupling of Organohalides
2024; Chinese Chemical Society; Linguagem: Inglês
10.31635/ccschem.024.202403939
ISSN2096-5745
AutoresShi-Shuo Xu, Hui Qiu, Pei‐Pei Xie, Zhen‐Hua Wang, Xiu Wang, Chao Zheng, Shu‐Li You, Tian‐Sheng Mei,
Tópico(s)Radical Photochemical Reactions
ResumoOpen AccessCCS ChemistryRESEARCH ARTICLES2 Apr 2024Cobalt-Catalyzed Electrochemical Enantioselective Reductive Cross-Coupling of Organohalides Shi-Shuo Xu†, Hui Qiu†, Pei-Pei Xie†, Zhen-Hua Wang†, Xiu Wang, Chao Zheng, Shu-Li You and Tian-Sheng Mei Shi-Shuo Xu† Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 , Hui Qiu† Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 , Pei-Pei Xie† Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 , Zhen-Hua Wang† Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 , Xiu Wang Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 , Chao Zheng *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] E-mail Address: [email protected] Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 , Shu-Li You *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] E-mail Address: [email protected] Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 and Tian-Sheng Mei *Corresponding authors: E-mail Address: [email protected] E-mail Address: [email protected] E-mail Address: [email protected] Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032 https://doi.org/10.31635/ccschem.024.202403939 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesDownload CitationsTrack Citations ShareFacebookTwitterLinked InEmail Transition metal-catalyzed asymmetric reductive coupling of two electrophiles has emerged as a powerful tool for carbon–carbon bond formation; however, the enantioselective reductive cross-coupling of two aryl halides has not been developed and remains a challenge. Herein, we demonstrated the enantioselective reductive cross-coupling of two aryl halides by combining cobalt catalysis with electrochemistry to afford axially chiral biaryls with good chemoselectivities (cross-coupling: homo-coupling > 20:1) and enantioselectivities (up to 94% ee). This strategy could be extended to the asymmetric cross-couplings of aryl and vinyl halides to provide axially chiral non-C2-symmetric compounds with excellent chemoselectivities (cross-coupling: homo-coupling > 20:1) and enantioselectivities (up to 99% ee). Density functional theory calculations and control experiments suggested that high chemo- and enantioselectivities could be realized by selective cathodic reduction of an aryl cobalt(III) species from the first oxidative addition owing to the different redox potentials. Download figure Download PowerPoint Introduction Biaryl atropisomers are prevalent in biologically active natural products, chiral ligands, material, and organocatalysts; therefore, tremendous efforts have been devoted to the synthesis of axially chiral biaryls.1,2 Among these protocols,3 the catalytic asymmetric Suzuki–Miyaura cross-coupling reaction is an attractive tool for constructing atropisomers, where precious palladium catalysts and aryl metal reagents are typically employed (Figure 1a).4–8 Alternatively, catalytic reductive cross-electrophile coupling reactions9–11 have emerged as a powerful strategy for forming C–C bonds because low cost metals, such as nickel, are typically employed and the use of organometallic reagents is avoided. Accordingly, several researchers, including Reisman and coworkers,12–14 Doyle and coworkers,15,16 and others, have made considerable contributions to the development of enantioselective variants.17–20 However, catalytic enantioselective reductive coupling for accessing axially chiral biaryls is currently limited to the homo-coupling of aryl electrophiles.21–24 The enantioselective reductive cross-coupling of two different aryl electrophiles has not been achieved yet, presumably owing to the similar reactivity of the two C(sp2)-electrophiles. This is because both aryl electrophiles can compete for oxidative addition to the catalyst, resulting in mixtures of cross-coupling and homo-coupling biaryls (Figure 1b).25 To solve the aforementioned chemoselectivity problem of reductive cross-coupling for preparing unsymmetrical biaryls, researchers have performed numerous studies on the selective activation of aryl electrophiles to obtain the desired cross-selectivity.26 For instance, Weix and coworkers27,28 reported the multimetallic-catalyzed cross-coupling of two different aryl electrophiles in which nickel and palladium separately activate different substrates, resulting in excellent chemoselectivity. Ye and coworkers29 developed a cross-coupling method for aromatic and heteroaromatic halides by employing a zirconaaziridine complex as a shuttle for dual palladium-catalyzed processes. Despite significant progress in this field,30–33 accessing the axially chiral unsymmetrical biaryls via direct reductive cross-coupling of two electrophiles remains a challenge. Figure 1 | Asymmetric reductive coupling of aryl halides. Download figure Download PowerPoint Electrochemistry-enabled catalytic reductive homo-couplings of aryl bromides were demonstrated by Jennings et al.34 in 1976, where a nickel catalyst was reactivated by cathodic reduction. Later, Gosmini et al.35 reported electrochemical cross-couplings between aryl halides to prepare unsymmetrical biaryls by cobalt catalyst. Recently, Yang and coworkers36 illustrated that cobalt shows unique reactivity toward aryl bromides and vinyl triflates, delivering axially chiral phosphine oxide compounds with good chemo- and enantioselectivity. The past decade has witnessed a renaissance in synthetic organic electrochemistry37–40 because it allows researchers to control the redox chemistry with rare precision for traditional organic chemistry.41–47 Recently, our group developed enantioselective homo-couplings of aryl bromides by combining nickel catalysis with electrochemistry to afford various binaphthols (BINOLs) with good yields and enantioselectivities.22 Building on these studies, we envisioned that chemo- and enantio-selective reductive cross-couplings of two different aryl halides could be realized by selective cathodic reduction of an aryl cobalt(III) species owing to the different redox potentials (Figure 1c). First, a weakly coordinated group, such as carbonyl, will facilitate the oxidative addition of cobalt to electron-deficient aryl bromide 1a in the presence of electron-rich aryl iodide 2a, providing intermediates B and C with a similar reaction rate (vide infra). The aryl moiety serves as a ligand for the metal center, which directly determines its redox potential. Thus, by slightly changing the electronic properties of the aryl halides, one of the oxidative addition intermediates B could be preferentially reduced to D and participate in the next elementary steps. This largely eliminates the contribution of intermediate C for coupling reactions. Third, the resulting aryl cobalt D would be preferred over oxidative addition to aryl iodide 2a because it is typically more reactive than aryl bromide 1a. After reductive elimination from the biaryl cobalt F, chiral cross-coupled biaryls could be formed in the presence of chiral ligands. Experimental Methods In a glove box, an oven-dried electrochemical cell (Electrasyn 2.0) with a stir bar was charged with aryl bromide 1a (0.3 mmol, 1.0 equiv), aryl iodide 2a (0.6 mmol, 2.0 equiv), CoI2 (0.045 mmol, 15 mol %), ligand ( L1) (0.09 mmol, 30 mol %), NaI (0.3 mmol, 1.0 equiv), 4Å molecular sieve (MS) (70 mg), and 5 mL of dimethylacetamide (DMAc). The tube was installed with stainless steel as a cathode and an iron rod as a sacrificial anode. The mixture was stirred at room temperature for 30 min and electrolyzed under a constant current of 1 mA until complete consumption of the starting materials was monitored by thin layer chromatography (TLC) (about 24 h). The resulting mixture was diluted with ethyl acetate (80 mL), and 1 M HCl was added to the solution. The organic layer was washed by saturated NH4Cl (20 mL×3), filtered, and evaporated. The crude product was purified by flash column chromatography to afford the desired product. More experimental procedures and a photographic guide for Co-catalyzed enantioselective electrochemical cross-couplings are provided in the Supporting Information. Results and Discussion Considering these points, we evaluated the hypothesis using different aryl halides with CoI2 as the catalyst and indane-fused pyrox22 as the chiral ligand under constant current electrolysis at 1 mA in an undivided cell (Table 1). Initially, two aryl bromides bearing electron-donating groups were chosen as substrates (entry 1), and no desired cross-coupling products were formed even if one of the aryl bromides was changed to an aryl iodide (entry 2). The dehalogenative protonation product of the aryl iodide and unreacted aryl bromides were the main components, indicating that the cobalt catalyst has low reactivity toward electron-rich aryl bromides. When one of the electron-rich aryl bromides was replaced by an electron-deficient one, the cross-coupling products were obtained in 24%–35% yields with a relatively low ratio of 3 to 4 (entries 3–6). However, good enantioselectivity was achieved when we used the combination of 1-bromo-2-(benzyl)-naphthoate or 1-iodo-2-(acetyl)-naphthalene with 1-bromo-2-(benzyloxy)-naphthalene (entries 4 and 6). By further increasing the reactivity of 2, a higher chemoselectivity of the cross-coupling was obtained with lower enantioselectivity (entry 7). To our delight, the combination of 1-bromo-2-(benzyl)-naphthoate with 1-iodo-2-(benzyloxy)-naphthalene resulted in the cross-coupling product 3a in a 72% yield with a 20 : 1 ratio of cross-coupling to homo-coupling, with 91% ee (entry 8). Notably, the ester substituent on compound 1 shifted from position 2 to 4, and the homo-coupling product of 1 was formed in 90% yield according to nuclear magnetic resonance (NMR) spectroscopy without resulting in cross-coupling (entry 9). This highlights that the position of the ester moiety is crucial for the second oxidative addition, whereas the protective benzyl group is preferred for obtaining high enantioselectivity because the ee value of the cross-coupling product decreased when benzyl was replaced by a methyl group, although the chemoselectivity for cross-coupling was excellent (entry 10). Thus, π•••π interactions or the steric repulsion between the benzyl groups of the substrates and the indanyl groups of the pyrox ligand may play an important role in enantioselectivity control (vide infra). Under identical reaction conditions, nickel catalysts resulted in products with low selectivity and a lower yield (entry 11). Notably, inferior selectivity was observed when the electric current was replaced by a metal reductant (entry 12). In addition, the indane-fused pyrox with a 6-cyclohexyl substituent in the pyridine ring was found to be the most optimal of the various alkyl groups (Table 1). Table 1 | Reaction Developmenta,b aReaction conditions: 1 (0.3 mmol, 1.0 equiv), 2 (0.6 mmol, 2.0 equiv), CoI2 (15 mol %), L1 (30 mol %), 4 Å MS (70 mg), and NaI (0.3 mmol, 1.0 equiv) in DMAc (5 mL) at 25 °C for 24 h in an undivided cell. bYields were determined via 1H NMR spectroscopy using mesitylene as the internal standard. Enantioselectivities were determined by high performance liquid chromatography (HPLC) analysis with a chiral stationary phase. The ratio of 3a: 4a was determined via 1H NMR spectroscopy. cIsolated yield. dNiBr2•glyme was used in lieu of CoI2. eMn (5 equiv) was used in lieu of electric current. Substrate scope of the electroreductive cross-coupling reaction With the optimized reaction conditions in hand, we investigated the substrate scope of the electroreductive cross-coupling reaction (Figure 2). To our delight, naphthalene bearing a variety of functional groups, such as alkyl, ether, alkene, and ester, is well tolerated under the electrochemical reductive cross-coupling conditions, affording the desired unsymmetrical axially chiral biaryls in good yields and with high enantioselectivity ( 3a– 3p). In addition, the structure of 3i was unambiguously determined by X-ray analysis and has an (R)-configuration. Figure 2 | Cross-couplings of aryl halides. Enantioselectivities were determined by HPLC analysis with a chiral stationary phase. Values in parentheses represent the ratio of cross-coupled biaryl (3): homo-coupled biaryl (4), as determined via 1H NMR spectroscopy. Download figure Download PowerPoint Axially chiral non-biaryl compounds have received less attention, although they are frequently used as chiral ligands in organic synthesis.48 This is primarily due to the lack of efficient synthetic methods.49 This encouraged us to extend the catalytic aryl–aryl cross-coupling system to aryl–vinyl for the synthesis of axially chiral non-C2-symmetric compounds. After slightly modifying the reaction conditions ( Supporting Information Tables S4–S8), we found that L5 was efficient for cobalt-catalyzed cross-coupling of 2-benzyloxy-1-iodonaphthalene ( 2a) with benzyl 1-chloro-6-methoxy-3,4-dihydron aphthalene-2-carboxylate ( 6a), and the desired product 7a was obtained in 88% isolated yield and 95% ee (Figure 3). Next, we investigated the substrate scope of the electroreductive cross-coupling reaction. First, we investigated the influence of different alkenyl halogenates on the reaction. Good yields and high ee values were obtained with phenyl, alkyl, halogen, ester, and alkenyl substituents at the 6 positions of the alkenyl halogenate ( 7a– 7i). High yields were obtained for either the electron-withdrawing or electron-donating substituents at the 5 and 7 positions ( 7j– 7r). Both the ee value and yield of the product decreased when the six-membered ring of the substrate was replaced by a larger seven-membered ring ( 7s). We also investigated the influence of different aryl iodides on the reaction. Alkyl, alkoxyl, phenyl, halogen, and ester were well accommodated under the electrochemical conditions, affording 7w– 7al with good yields and enantioselectivities. When the benzyl ether was replaced by methyl ether ( 7al), the yield of the desired product increased, but the ee value decreased, which indicated that bulky sterically hindered protecting groups were required for the high stereoselectivity. The absolute R configuration of 7al was unambiguously verified by X-ray crystallographic analysis (Figure 3). Figure 3 | Cross-couplings of aryl halides and vinyl halides. Enantioselectivities were determined by HPLC analysis with a chiral stationary phase. Values in parentheses represent the ratio of cross-coupled product: homo-coupled biaryl, as determined via 1H NMR spectroscopy. aCoI2 (15 mol %) and ligand (30 mol %) were used in lieu of CoI2 (10 mol %), ligand (20 mol %), b6a (0.2 mmol), and 2a (0.4 mmol). Download figure Download PowerPoint To gain further insight into this electrochemical reductive coupling reaction, we prepared [CoI(Pybox)2]I complex 8, the structure of which was also unambiguously assigned via X-ray crystallographic analysis (Figure 4a). Complex 8 manifests two reductive peaks in its cyclic voltammogram (−1.55 and −1.69 V vs Fc+/Fc0), which might be related to mono- and bis-ligated Co(II) species, respectively (Figure 4b and Supporting Information Figure S4).50,51 By contrast, the reduction potentials of substrate 1a and 2a are approximately −2.26 and −2.64 V versus Fc+/Fc0 ( Supporting Information Figure S1), which indicates that the Co(II) catalyst is more readily reduced than aryl halides in this electrochemical reductive coupling reaction. Furthermore, the addition of 1a and 2a to a DMAc solution of complex 8 resulted in the loss of the oxidation peak (−1.42 V vs Fc+/Fc0), which is indicative of a Co(II)/Co(I) redox couple and suggests that the oxidative addition of 1a or 2a to Co(I) readily occurs (Figure 4b, pink and blue lines, respectively). The addition of 1a resulted in a greater increase in the maximum reduction peak of cobalt compared with that of 2a (Figure 4b), which suggested that the reactivity of 1a is higher than that of 2a toward the first oxidative addition. The ee of the ligand with the product suggests a linear relationship (Figure 4c), revealing that a single ligand and one cobalt species may be involved in the enantioselectivity-determining step. Moreover, Co(I) species was in-situ prepared by the reaction of CoI2 with L1 in the presence of Mn power (Figure 4d), and when 1a and 2a were added to the solution and reacted for half an hour. Then the reaction mixture was quenched with 1 M HCl solution, affording the protonated product in 13% and 9% yields, respectively (Figure 4e). Besides, the reaction could be easily scaled up, affording 3a in 74% yield with any loss of enantioselectivity (Figure 4f). We have demonstrated the synthetic utility of this chemistry by the preparation of important compounds 3x as shown in Figure 4g, which can be used as ligands or catalysts for asymmetric transformations.52 It is worth noting that 3x was obtained by only two steps starting from easily available aryl halides, while five steps were required for the synthesis of 3x with the traditional method.53 Figure 4 | Mechanistic studies and synthetic application. Download figure Download PowerPoint Density functional theory calculation Density functional theory (DFT) calculations were performed using model substrates 1-bromo-2-(methyl)-naphthoate and 1-iodo-2-(methoxy)-naphthalene to determine the reaction profile and the origins of chemoselectivity (Figure 5a,b). Starting from the Co(I) catalyst in which two chiral pyrox ligands and one iodide ion bonded to the Co(I) center ( Cat, 0.0 kcal/mol), the oxidative addition of both aryl bromide and aryl iodide could proceed smoothly via the corresponding transition states, TS1 (23.7 kcal/mol) and TS2 (24.6 kcal/mol), respectively. The relatively small energy difference (0.9 kcal/mol) between these two transition states might be attributed to two reasons. One is the favorable chelation effect of the ester group in TS1, and the other is the disfavorable geometric distortion caused by SN2-type C–I bond cleavage in TS2 (See Supporting Information Figure S11 for details). Therefore, the initial oxidative addition step may not be chemoselective. After further substitution of one pyrox ligand with one iodide ion, two different neutral Co(III) complexes ( INT2 and INT4) were formed. The thermodynamics of the following two-electron reductions leading to aryl–Co(I) complexes ( INT2 to INT3 and INT4 to INT5) were evaluated. Both processes are highly exergonic, and the Gibbs free energy change from INT2 to INT3 was more negative than that from INT4 to INT5 by 7.7 kcal/mol. This corresponds to a potential difference of approximate 0.17 V upon applying the Nernst equation. Our calculations suggested the more favorable formation of aryl–Co(I) complex INT3 via chemoselective electroreduction, which set the foundation for the desired cross-coupling reaction. The second oxidative addition of INT3 (0.0 kcal/mol) with aryl iodide or aryl bromide was then considered. The oxidative addition of aryl iodide in the transition state TS3 (13.7 kcal/mol) was more favorable than the oxidative addition transition state TS3′ (15.7 kcal/mol) with another aryl bromide, which was necessary for the homo-coupling. Since the geometries of the core structure of TS3 and TS3′ are similar, the energy difference between these two transition states was probably due to the different bond dissociation energies of C–Br and C–I bonds (see Supporting Information Figure S13 for details). The dissociation of one iodide anion from the resulting biaryl–Co(III) complex INT6 (5.2 kcal/mol) was highly exergonic, resulting in the cationic biaryl–Co(III) complex INT7 (−13.7 kcal/mol) in which both the ester group of the original aryl bromide and the methoxy group of the original aryl iodide chelated to the Co center. We assumed the free conformational fluctuation of both aryl groups on the Co center in INT7, and the final reductive elimination step via TS4 (−3.4 kcal/mol) to the Co(I) complex INT8 (−32.8 kcal/mol) was determined to be irreversible, with the establishment of the enantioselectivity for the entire reaction. In addition, a comprehensive conformational search was performed for the reductive elimination step using the substrates 1-bromo-2-(benzyl)-naphthoate and 1-iodo-2-(benzyloxy)-naphthalene. The Gibbs free energy difference between the lowest energy transition states leading to either enantiomer of the products ( TS4- R and TS4- S) was calculated to be 1.5 kcal/mol, which well reproduced the experimentally observed enantioselectivity favoring the formation of the (R)-product. The multiple attractive C–H•••π interactions, π•••π interactions, and/or steric repulsion between the benzyl groups of the substrates and the indanyl and cyclohexyl groups of the pyrox ligand all contributed to the chiral discrimination process. Figure 5 | DFT calculations. All the calculations were performed at the (U)M06/def2-TZVP (SMD, DMAc)//(U)B3LYP-D3(BJ)/def2-SVP (gas) level of theory. The relative Gibbs free energies are reported in kcal/mol and the bond distances in Å. Download figure Download PowerPoint Conclusion In summary, highly chemoselective and enantioselective cross-coupling of two aryl halides was achieved by selective reduction of the aryl cobalt(III) intermediates resulting from oxidative addition via an electrochemical approach. We anticipate that the merger of asymmetric cobalt catalysis and electrochemistry will inspire the development of other enantioselective reductive cross-coupling reactions to complement existing syntheses. Supporting Information Supporting Information is available and includes additional experimental details: experimental procedures, characterization of new compounds, and spectroscopic data (PDF) crystallographic data for 3i (CCDC 2122020) (CIF), 7am (CCDC 2214972) (CIF), and 8 (CCDC 2122017) (CIF). Conflict of Interest There is no conflict of interest to report. Funding Information This work was financially supported by the National Key R&D Program of China (grant no. 2021YFA1500100), the Strategic Priority Research Program of the Chinese Academy of Sciences (grant no. XDB0610000), the NSF of China (grant nos. 21821002 and 22361142834), and Natural Science Foundation of Ningbo (grant no. 2023J035). References 1. Cheng J. K.; Xiang S.-H.; Li S.; Ye L.; Tan B.Recent Advances in Catalytic Asymmetric Construction of Atropisomers.Chem. Rev.2021, 121, 4805–4902. Google Scholar 2. Collins B. S. L.; Kistemaker J. C. M.; Otten E.; Feringa B. L.A Chemically Powered Unidirectional Rotary Molecular Motor Based on a Palladium Redox Cycle.Nat. Chem.2016, 8, 860–866. Google Scholar 3. Cherney A. H.; Kadunce N. T.; Reisman S. E.Enantioselective and Enantiospecific Transition-Metal-Catalyzed Cross-Coupling Reactions of Organometallic Reagents To Construct C–C Bonds.Chem. Rev.2015, 115, 9587–9652. Google Scholar 4. Yang H.; Sun J.; Gu W.; Tang W.Enantioselective Cross-Coupling for Axially Chiral Tetra-ortho-Substituted Biaryls and Asymmetric Synthesis of Gossypol.J. Am. Chem. Soc.2020, 142, 8036–8043. Google Scholar 5. Shen D.; Xu Y.; Shi S.-L.A Bulky Chiral N-Heterocyclic Carbene Palladium Catalyst Enables Highly Enantioselective Suzuki-Miyaura Cross-Coupling Reactions for the Synthesis of Biaryl Atropisomers.J. Am. Chem. Soc.2019, 141, 14938–14945. Google Scholar 6. Shen X.; Jones G. O.; Watson D. A.; Bhayana B.; Buchwald S. L.Enantioselective Synthesis of Axially Chiral Biaryls by the Pd-Catalyzed Suzuki–Miyaura Reaction: Substrate Scope and Quantum Mechanical Investigations.J. Am. Chem. Soc.2010, 132, 11278–11287. Google Scholar 7. Yin J.; Buchwald S. L.A Catalytic Asymmetric Suzuki Coupling for the Synthesis of Axially Chiral Biaryl Compounds.J. Am. Chem. Soc.2000, 122, 12051–12052. Google Scholar 8. Cammidge A. N.; Crepy K. V. L.The First Asymmetric Suzuki Cross-Coupling Reaction.Chem. Commun.2000, 1723–1724. Google Scholar 9. Liu J.; Ye Y., Sessler J. L.; Gong H.Cross-Electrophile Couplings of Activated and Sterically Hindered Halides and Alcohol Derivatives.Acc. Chem. Res.2020, 53, 1833–1845. Google Scholar 10. Weix D. J.Methods and Mechanisms for Cross-Electrophile Coupling of Csp2 Halides with Alkyl Electrophiles.Acc. Chem. Res.2015, 48, 1767–1775. Google Scholar 11. Knappke C. E. I.; Knappke C. E. I.; Grupe S.; Gärtner D.; Corpet M.; Gosmini C.; von Wangelin A. J.Reductive Cross-Coupling Reactions Between Two Electrophiles.Chem.–Eur. J.2014, 20, 6828–6842. Google Scholar 12. Turro R. F.; Wahlman J. L.H.; Tong Z. J.; Chen X.; Yang M.; Chen E. P.; Hong X.; Hadt R. G.; Houk K. N.; Yang Y.-F.; Reisman S. E.Mechanistic Investigation of Ni-Catalyzed Reductive Cross-Coupling of Alkenyl and Benzyl Electrophiles.J. Am. Chem. Soc.2023, 145, 14705–14715. Google Scholar 13. DeLano T. J.; Reisman S. E.Enantioselective Electroreductive Coupling of Alkenyl and Benzyl Halides via Nickel Catalysis.ACS Catal.2019, 9, 6751–6754. Google Scholar 14. Cherney A. H.; Kadunce N. T.; Reisman S. E.Catalytic Asymmetric Reductive Acyl Cross-Coupling: Synthesis of Enantioenriched Acyclic α,α-Disubstituted Ketones.J. Am. Chem. Soc.2013, 135, 7442–7445. Google Scholar 15. Lau S. H.; Borden M. A.; Steiman T. J.; Wang L. S.; Parasram M.; Doyle A. G.Ni/Photoredox-Catalyzed Enantioselective Cross-Electrophile Coupling of Styrene Oxides with Aryl Iodides.J. Am. Chem. Soc.2021, 143, 15873–15881. Google Scholar 16. Woods B. P.; Orlandi M.; Huang C.-Y.; Sigman M. S.; Doyle A. G.Nickel-Catalyzed Enantioselective Reductive Cross-Coupling of Styrenyl Aziridines.J. Am. Chem. Soc.2017, 139, 5688–5691. Google Scholar 17. Pan Q., Ping Y., Kong W.Nickel-Catalyzed Ligand-Controlled Selective Reductive Cyclization/Cross-Couplings.Acc. Chem. Res.2023, 56, 515–535. Crossref, Google Scholar 18. Tu H.-Y.; Wang F.; Huo L.; Li Y.; Zhu S.; Zhao X.; Li H.; Qing F.-L.; Chu L.Enantioselective Three-Component Fluoroalkylarylation of Unactivated Olefins Through Nickel-Catalyzed Cross-Electrophile Coupling.J. Am. Chem. Soc.2020, 142, 9604–9611. Google Scholar 19. Tian Z.-X.; Qiao J.-B.; Xu G.-L.; Pang X.; Qi L.; Ma W.-Y.; Zhao Z.-Z.; Duan J.; Du Y.-F.; Su P.; Liu X.-Y.; Shu X.-Z.Highly Enantioselective Cross-Electrophile Aryl-Alkenylation of Unactivated Alkenes.J. Am. Chem. Soc.2019, 141, 7637–7643. Google Scholar 20. Zhao Y.; Weix D. J.Enantioselective Cross-Coupling of meso-Epoxides with Aryl Halides.J. Am. Chem. Soc.2015, 137, 3237–3240. Google Scholar 21. Chen W.-W.; Zhao Q.; Xu M.-H.; Lin G.-Q.Nickel-Catalyzed Asymmetric Ullmann Coupling for the Synthesis of Axially Chiral Tetra-ortho-Substituted Biaryl Dials.Org. Lett.2010, 12, 1072–1075. Google Scholar 22. 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. Google Scholar 23. Zuo Z.; Kim R. S.; Watson D. A.Synthesis of Axially Chiral 2,2′-Bisphosphobiarenes via a Nickel-Catalyzed Asymmetric Ullmann Coupling: General Access to Privileged Chiral Ligands without Optical Resolution.J. Am. Chem. Soc.2021, 143, 1328–1333. Google Scholar 24. Perveen S.; Zhang S.; Wang L.; Song P.; Ouyang Y.; Jiao J.; Duan X.-H.; Li P.Synthesis of Axially Chiral Biaryls via Enantioselective Ullmann Coupling of ortho-Chlorinated Aryl Aldehydes Enabled by a Chiral 2,2′-Bipyridine Ligand.Angew. Chem. Int. Ed.2022, 61, e202212108. Google Scholar 25. Everson D. A.; Weix D. J.Cross-Electrophile Coupling: Principles of Reactivity and Selectivity.J. Org. Chem.2014, 79, 4793–4798. Google Scholar 26. Duan A.; Xiao F.; Lan Y.; Niu L.Mechanistic Views and Computational Studies on Transition-Metal-Catalyzed Reductive Coupling Reactions.Chem. Soc. Rev.2022, 51, 9986–10015. Google Scholar 27. Ackerman L. K. G.; Lovell M. M.; Weix D. J.Multimetallic Catalysed Cross-Coupling of Aryl Bromides with Aryl Triflates.Nature2015, 524, 454–457. Google Scholar 28. Ackerman-Biegasiewicz L. K. G.; Kariofillis S. K.; Weix D. J.Multimetallic-Catalyzed C–C Bond-Forming Reactions: From Serendipity to Strategy.J. Am. Chem. Soc.2023, 145, 6596–6614. Google Scholar 29. Wu T.-F.; Zhang Y-J.; Fu Y.; Liu F-J.; Tang J-T.; Liu P.; Toste F. D.; Ye B.Zirconium-Redox-Shuttled Cross-Electrophile Coupling of Aromatic and Heteroaromatic Halides.Chem2021, 7, 1963–1974. Google Scholar 30. Amatore M.; Gosmini C.Efficient Cobalt-Catalyzed Formation of Unsymmetrical Biaryl Compounds and Its Application in the Synthesis of a Sartan Intermediate.Angew. Chem. Int. Ed.2008, 47, 2089–2092. Google Scholar 31. Tang J.; Liu L. L.; Yang S.; Cong X.; Luo M.; Zeng X.Chemoselective Cross-Coupling between Two Different and Unactivated C(aryl)–O Bonds Enabled by Chromium Catalysis.J. Am. Chem. Soc.2020, 142, 7715–7720. Google Scholar 32. Mirabi B.; Marchese A. D.; Lautens M.Nickel-Catalyzed Reductive Cross-Coupling of Heteroaryl Chlorides and Aryl Chlorides.ACS Catal.2021, 11, 12785–12793. Google Scholar 33. Nohira I.; Chatani N.Nickel-Catalyzed Cross-Electrophile Coupling between C(sp2)–F and C(sp2)–Cl Bonds by the Reaction of ortho-Fluoro-Aromatic Amides with Aryl Chlorides.ACS Catal.2021, 11, 4644–4649. Google Scholar 34. Jennings P. W.; Pillsbury D. G.; Hall J. L.; Brice V. T.Carbon-Carbon Bond Formation via Organometallic Electrochemistry.J. Org. Chem.1976, 41, 719–722. Google Scholar 35. Gomes P.; Fillon H.; Gosmini C.; Labbe E.; Perichon J.Synthesis of Unsymmetrical Biaryls by Electroreductive Cobalt-Catalyzed Cross-Coupling of Aryl Halides.Tetrahedron2002, 58, 8417–8424. Google Scholar 36. Zhang X.; Wang J.; Yang S.-D.Enantioselective Cobalt-Catalyzed Reductive Cross-Coupling for the Synthesis of Axially Chiral Phosphine–Olefin Ligands.ACS Catal.2021, 11, 14008–14015. Google Scholar 37. Yan M.; Kawamata Y.; Baran P. S.Synthetic Organic Electrochemical Methods Since 2000: On the Verge of a Renaissance.Chem. Rev.2017, 117, 13230–13319. Google Scholar 38. Novaes L. F. T.; Novaes L.F. T.; Liu J.; Shen Y.; Lu L.; Meinhardt J. M.; Lin S.Electrocatalysis as an Enabling Technology for Organic Synthesis.Chem. Soc. Rev.2021, 50, 7941–8002. Crossref, Google Scholar 39. Malapit C. A.; Prater M. B.; Cabrera-Pardo J. R.; Li M.; Pham T. D.; McFadden T. P.; Blank S.; Minteer S. D.Advances on the Merger of Electrochemistry and Transition Metal Catalysis for Organic Synthesis.Chem. Rev.2022, 122, 3180–3218. Google Scholar 40. Liu Y.; Li P.; Wang Y.; Qiu Y.Electroreductive Cross-Electrophile Coupling (eXEC) Reactions.Angew. Chem. Int. Ed.2023, 62, e202306679. Google Scholar 41. Badalyan A.; Stahl S. S.Cooperative Electrocatalytic Alcohol Oxidation with Electron-Proton-Transfer Mediators.Nature2016, 535, 406–410. Google Scholar 42. Mo Y.; Lu Z. H.; Rughoobur G.; Patil P.; Gershenfeld N.; Akinwande A. I.; Buchwald S. L.; Jensen K.Microfluidic Electrochemistry for Single-Electron Transfer Redox-Neutral Reactions.Science2020, 368, 1352–1357. Google Scholar 43. Dong X.; Roeckl J. L.; Waldvogel S. R.; Morandi B.Merging Shuttle Reactions and Paired Electrolysis for Reversible Vicinal Dihalogenations.Science2021, 371, 507–514. Google Scholar 44. Hamby T. B.; LaLama M. J.; Sevov C. S.Controlling Ni Redox States by Dynamic Ligand Exchange for Electroreductive Csp3–Csp2 Coupling.Science2022, 376, 410–416. Google Scholar 45. Harwood S. J.; Palkowitz M. D.; Gannett C. N.; Perez P.; Yao Z.; Sun L. J.; Abruña H. D.; Anderson S. L.; Baran P. S.Modular Terpene Synthesis Enabled by Mild Electrochemical Couplings.Science2022, 375, 745–752. Google Scholar 46. Zhang W.; Lu L. X.; Zhang W.; Wang Y.; Ware S. D.; Mondragon J.; Rein J.; Strotman N.; Lehnherr D.; See K. A.; Lin S.Electrochemically Driven Cross-Electrophile Coupling of Alkyl Halides.Nature2022, 604, 292–297. Google Scholar 47. Hu X.; Cheng-Sánchez I.; Cuesta-Galisteo S.; Nevado C.Nickel-Catalyzed Enantioselective Electrochemical Reductive Cross-Coupling of Aryl Aziridines with Alkenyl Bromides.J. Am. Chem. Soc.2023, 145, 6270–6279. Google Scholar 48. Egami H.; Asada J.; Sato K.; Hashizume D.; Kawato Y. J.; Hamashima Y.Asymmetric Fluorolactonization with a Bifunctional Hydroxyl Carboxylate Catalyst.J. Am. Chem. Soc.2015, 137, 10132–10135. Google Scholar 49. Meyers A. I.; Lutomski K. A.Enantioselective Synthesis of Binaphthyls via Nucleophilic Aromatic Substitution on Chiral Oxazolines.J. Am. Chem. Soc.1982, 104, 879–881. Google Scholar 50. Sandford C.; Fries L. R.; Ball T. E.; Minteer S. D.; Sigman M. S.Mechanistic Studies into the Oxidative Addition of Co(I) Complexes: Combining Electroanalytical Techniques with Parameterization.J. Am. Chem. Soc.2019, 141, 18877–18889. Google Scholar 51. Hickey D. P.; Sandford C.; Rhodes Z.; Gensch T.; Fries L. R.; Sigman M. S.; Minteer S. D.Investigating the Role of Ligand Electronics on Stabilizing Electrocatalytically Relevant Low-Valent Co(I) Intermediates.J. Am. Chem. Soc.2019, 141, 1382–1392. Google Scholar 52. Yoshino T.; Matsunaga S.Chiral Carboxylic Acid Assisted Enantioselective C–H Activation with Achiral CpxMIII (M = Co, Rh, Ir) Catalysts.ACS Catal.2021, 11, 6455–6466. Google Scholar 53. Helmbrecht S. L.; Schlüter J.; Blazejak M.; Hintermann L.Axially Chiral 1,1'-Binaphthyl-2-Carboxylic Acid (BINA-Cox) as Ligands for Titanium-Catalyzed Asymmetric Hydroalkoxylation.Eur. J. Org. Chem.2020, 2020, 2062–2076. Google Scholar Previous articleNext article FiguresReferencesRelatedDetails Issue AssignmentNot Yet AssignedSupporting Information Copyright & Permissions© 2024 Chinese Chemical SocietyKeywordsreductive cross-couplingsasymmetric catalysissynthetic organic electrochemistrycobalt catalysisbiaryl atropisomers Downloaded 1,516 times PDF downloadLoading ...
Referência(s)