Abstract
The design of a single complicated chiral ligand to well-promote each step of an asymmetric cascade reaction is sometimes a formidable challenge in transition metal catalysis. In this work, a highly regio- and enantioselective Ni-catalysed migratory hydroarylation relay process has been achieved with the combination of two simple ligands, one which accomplishes chain-walking and the other causing asymmetric arylation. This formal asymmetric C(sp3)−H arylation provides direct access to a wide range of structurally diverse chiral 1,1-diarylalkanes, a structural unit found in a number of bioactive molecules. The value of this strategy was further demonstrated by the Ni-catalysed migratory asymmetric 1,3-arylboration.
Subject terms: Synthetic chemistry methodology, Asymmetric catalysis
Migratory alkene isomerizations and cross-coupling reactions are both possible under nickel catalysis, but usually require different conditions. Here the authors show a combined protocol to isomerize a double bond and then, via an in-situ exchange of ligands, perform an enantioselective C(sp2)–C(sp3) cross-coupling.
Introduction
The synergistic combination of olefin isomerization and cross-coupling, nickel-catalysed1–3 migratory hydrofunctionalization4–9 of olefins has emerged as a complementary and versatile platform that can realize a wide variety of remote C(sp3)−H functionalizations10–26. So far, much progress has been made in the NiH-catalysed asymmetric ipso-hydrofunctionalization of alkenes27–50. However, designing a chiral ligand for nickel that could promote both chain-walking and subsequent regio- and enantioselective reductive coupling at a remote position remains a formidable challenge (Fig. 1a).
Overcoming the limitation of single metal catalysis [ML], multimetallic catalysis [MALA & MBLB] provides a complementary strategy that has been well exploited. The alternative reaction using multiligands has remained largely underexplored. Pioneering work in this area has shown that multiligand based binary ligand complexes [MLALB] could be formed and are more reactive catalysts51–53. A catalyst mixture [MLA & MLB] formed by multiligands was also used by Buchwald et al.54 to broaden the substrate scope of a system using a single catalyst [MLA]. Instead of designing a complex ligand, another unique mode of employing multiligands is to undergo multiligand relay catalysis via dynamic ligand exchange in cases where a single ligand fails to selectively or efficiently promote all the steps of the transformation. In this attractive but largely underexplored pathway, each ligand is only required to promote partial steps of the catalytic cycle. This concept was preliminarily explored by White’s group in the Pd-catalysed non-asymmetric allylic oxidation system while using stoichiometric amount of benzoquinone or DMSO as another ligand for functionalization step55–57. Very recently, Fu et al.58 and Mauleón et al.59 demonstrated that the exchange of dynamic mutiligands with metals could be used to sequentially promote different steps in Cu-catalysed relay reactions. At the same time, our laboratory60 successfully introduced this concept to the nickel-catalysed asymmetric migratory hydrofunctionalization process, wherein the entire catalytical cycle could be subsequently promoted by an achiral chain-walking ligand and a structurally simple chiral asymmetric coupling ligand27–44 (Fig. 1b). Pursuing this theme, we hypothesized that migratory asymmetric hydroarylation which consists of an achiral ligand (L) promoting chain-walking and a chiral ligand (L*) promoting asymmetric arylation at benzylic position61–65 could be possible, leading to the facile synthesis of an enantioenriched 1,1-diarylalkane, a biologically active pharmacophore (Fig. 1c).
In this work, we describe a highly regio- and enantioselective Ni-catalysed migratory hydroarylation relay process enabled by a multiligand relay catalysis strategy. By synergistic combination of a simple ligand for chain-walking and a known ligand for asymmetric arylation, a wide variety of enantioenriched 1,1-diarylalkanes can be rapidly obtained under mild conditions.
Results and discussion
Reaction design and optimization
Our initial investigation focused on the enantioselective remote hydroarylation of 4-phenyl-1-butene (1a) with 4-iodoanisole (2a) (Fig. 2). It was found that NiCl2 ∙ glyme (glyme = ethylene glycol dimethyl ether) and the synergistic combination of a chain-walking ligand (L, 2,9-dimethyl-1,10-phenanthroline) and an asymmetric arylation ligand (L*, (4 R,4′R)-1,1′-bis(3-(tert-butyl)-phenyl)-4,4′-di(heptan-4-yl)-4,4′,5,5′-tetrahydro-1H,1′H-2,2′-biimidazole)31 with DMMS (dimethoxymethylsilane) could afford the desired migratory product 1-methoxy-4-(1-phenylbutyl)-benzene (3a) as a single regioisomer in 84% isolated yield and with 95% enantiomeric excess (ee) (entry 1). Control experiments revealed that these two ligands are both essential for simultaneous control of the regio- and stereochemistry, and poor regioselectivity was observed in the absence of the chain-walking ligand (entries 2–4). Importantly, lowering the loading of the chain-walking ligand to 0.4 mol% had little impact on the overall performance (entry 5). Increasing the chain-walking ligand loading and decreasing the arylation ligand loading led however to a moderate decrease of both the yield and the enantioselectivity (entry 6). An alternative chain-walking ligand (L1) was found to be competent but less effective than L (entry 7). Evaluation of arylation ligands revealed that L* provided the highest ee (entry 1 vs. entries 8–10). Diminished yields were obtained when using other nickel sources (entry 11) or employing other silanes (entry 12) or replacing K3PO4 ∙ H2O as base (entry 13). The use of a single solvent led to either diminished yield (entry 14) or almost complete failure of the reaction (entry 15). Notably, a slightly reduced yield was obtained when a less reactive aryl bromide was used (entry 16).
Substrate scope
With the well-established conditions in hand, the scope and generality of the reaction were evaluated. As illustrated in Fig. 3, a wide variety of aryl and heteroaryl iodides or bromides are tolerated. In general, the less reactive aryl bromides resulted in a slightly decreased yield (3a, 3b, 3f, 3h, 3k, and 3t). The reaction proceeded well with both electron-rich (3b–3e) and electron-withdrawing aryl halides (3f–3q). A variety of functional groups are readily accommodated, including ethers (3a, 3e–3g, 3i, and 3s), esters (3b, 3k), a Boc carbamate (3c), an amide (3d), a trifluoromethyl group (3j), a nitrile (3l), an aryl fluoride (3m), and an aryl chloride (3n). Notably, under these exceptionally mild reaction conditions, sensitive functional groups such as aryl triflate (3o) commonly used for subsequent complementary cross-coupling, and readily reduced aldehydes (3p) and ketones (3q) were all unaffected. Heterocycles such as indole (3r) and pyridine (3s and 3t) are also competent coupling partners. However, o-substituted (hetero)aryl halides gave lower yields under current conditions.
We next explored the scope of alkenes (Fig. 4). As shown in Fig. 4a, a wide range of terminal aliphatic alkenes bearing a remote aryl (4b–4n, 4s, 4t) or heteroaryl (4o–4r) group undergo asymmetric migratory hydroarylation smoothly, regardless of the chain length between the C=C bond and the remote aryl group. In general, a slightly increased loading of the chain-walking ligand is beneficial for alkene substrates with a long chain (4m). A variety of substituents on the remote aromatic ring, including both electron-donating (4c–4e) and electron-withdrawing (4f–4l) substituents, are all well-tolerated. Notably, the 1,1-disubstituted alkene (4n) is also a suitable substrate, although the migratory product was obtained in decreased yield. As shown in Fig. 4b, unactivated internal alkenes are also suitable substrates (4u–4a’). Both E (4y, 4z) and Z (4a’) alkenes, as well as E/Z mixtures (4u–4x) were suitable substrates. In addition, a range of different substituents at the other terminus of the alkyl chain, even with a heteroatomic substituent (4x–4z), were all well-tolerated, and arylation at benzylic position was still preferred.
Application
Benefiting from the chain-walking catalysis, isomeric mixtures of olefins could be used directly to produce the enantioenriched product (4m) in a regioconvergent fashion (Fig. 5a). The current multiple ligand catalysis could also be applied to migratory difunctionalization of alkenes66–70. As shown in Fig. 5b, a nickel-catalysed 1,3-arylboration reaction was carried out using a racemic ligand (L2)66 to promote NiBpin insertion/chain-walking together with the asymmetric arylation ligand (L*) which promotes asymmetric arylation. The desired migratory chiral products (7a–7h) were obtained with high regioselectivity and excellent ee.
Mechanistic investigation
A series of mechanistic experiments were carried out to understand the relay process. As shown in Fig. 6a, isomeric mixtures of olefins could be produced during the reaction process, an observation that is consistent with a fast chain-walking step. As shown in Fig. 6b, in the absence of cross-coupling partner, control experiment revealed that the alkene isomerization could also proceed smoothly with dual ligands or with a single chain-walking ligand. These results indicate that chain-walking precedes arylation without the participation of cross-coupling partner. In contrast, only a very small quantity of isomerized alkenes was observed while using arylation ligand L* alone. This observation is consistent with the conclusion that the chain-walking process is mainly promoted by achiral chain-walking ligand L. To probe whether hydronickellation is the enantio-determining step, isotopic labelling experiments of a cyclic styrene (1b’) substrate were carried out with deuteropinacolborane (Fig. 6c). If the migratory insertion of NiD into styrene is the enantio-determining step, a diastereomerically pure product (4b’-D) should be obtained. However, a diastereomeric mixture was observed with asymmetric arylation ligand (L*) alone or with dual ligand (L/L*), and the diastereomeric ratio decreased while increasing the amount of achiral chain-walking ligand L. This observation indicates that the hydronickellation is not likely to be the enantio-determining step.
To better understand the reaction mechanism, we conducted a DFT study of the reaction pathways shown in Fig. 7 (see Supplementary Data 1). Previous computational studies have clarified that achiral Ni(II)H controls chain-walking process and gives the desired internal alkylnickel(II)20,60, which undergoes transmetallation with Ni(I)H to generate the corresponding alkylnickel(I)60. Here we started from achiral alkylnickel(I) INT1-a and investigated the asymmetric arylation process. The direct oxidative addition of INT1-a with aryl iodide via TS1-a requires an activation free energy of 20.4 kcal/mol. Alternatively, INT1-a can first exchange its achiral bipyridine ligand with chiral bisimidazoline ligand to form either INT1-R or INT1-S. Although this process is endergonic by about 6 kcal/mol, the subsequent Ni(I)/Ni(III) oxidative additions involving chiral ligand via TS1-R and TS1-S are much easier. The overall barrier for the formation of chiral Ni(III) species INT2-R is 16.2 kcal/mol via TS1-R, which is 4.2 kcal/mol lower than that for the oxidative addition of achiral alkylnickel(I) via TS1-a (20.4 kcal/mol). For two competing enantioselective transition states TS1-R and TS1-S, there is a 1.8 kcal/mol preference toward (R)-intermediate, corresponding to a 91% ee value of the final product at 298 K. In TS1-R, a less sterically demanding phenyl is placed close to the bulky isopropyl group of ligand, while in TS1-S, it is the ethyl group that generates steric repulsions with the bulky isopropyl group (highlighted in pink, Fig. 7). Meanwhile, the unreacted INT1-S can be rapidly converted to INT1-R via a reversible ligand exchange, transmetallation and alkene insertion process (see Supplementary Fig. 5). After the irreversible formation of chiral Ni(III) intermediate INT2-R, the following reductive elimination via TS2-R to yield product is very facile with a barrier of only 4.1 kcal/mol, which is much lower than the possible homolysis via TS3-R62. Therefore, the chiral bisimidazoline ligand controlled the Ni(I)/Ni(III) oxidative addition with aryl iodide, which is the rate- and enantioselectivity-determining step in the asymmetric arylation.
In conclusion, by the combination of a chain-walking ligand and an asymmetric arylation ligand used in nickel chemistry, we have developed a highly enantioselective remote hydroarylation protocol. This mild, general and attractive route allows rapid access to an array of enantioenriched 1,1-diarylalkanes. Further application of this multiligand relay catalysis to simplify the chiral ligand design in nickel and other metal-catalysed multistep asymmetric reactions are currently in progress.
Methods
General procedure (A) for the regio- and enantioselective C(sp3)−H arylation
In a nitrogen-filled glove box, to an oven-dried 8 mL screw-cap vial equipped with a magnetic stir bar was added NiCl2·glyme (2.2 mg, 5.0 mol%), KF (23.2 mg, 2.0 equiv), L* (7.2 mg, 6.0 mol%), L (0.42 mg, 0.2 mL stock solution, 2.1 mg/mL in toluene), anhydrous toluene (0.60 mL) and DMPU (0.20 mL). The mixture was stirred for 15 min at rt, at which time 4-phenyl-1-butene (30 μL, 0.20 mmol), 4-iodoanisole (94.0 mg, 0.40 mmol) and DMMS (49.3 μL, 0.40 mmol) were added to the resulting mixture in this order. The tube was sealed with a teflon-lined screw cap, removed from the glove box and the reaction was stirred at rt (22 ~ 26 °C) for up to 24 h (the mixture was stirred at 750 rpm, ensuring that the base was uniformly suspended). After the reaction was complete, the reaction mixture was directly filtered through a short pad of silica gel (using EtOAc in petroleum ether) to give the crude product. n-Dodecane (20 μL) was added as an internal standard for GC analysis. The product was purified by chromatography on silica gel for each substrate. The yields reported are the average of at least two experiments, unless otherwise indicated. The enantiomeric excesses (% ee) were determined by HPLC analysis using chiral stationary phases.
Supplementary information
Acknowledgements
Support was provided by NSFC (92156004, 21822105, 22001118), NSF of Jiangsu Province (BK20190281, BK20211555, BK20201245, BK20200300), the Fundamental Research Funds for the Central Universities (020514380253, 020514380263), programs for high-level entrepreneurial and innovative talents introduction of Jiangsu Province (group program), and Open Research Fund of School of Chemistry and Chemical Engineering, Henan Normal University. We thank the High Performance Computing Center (HPCC) of Nanjing University for doing the numerical calculations in this paper on its blade cluster system.
Author contributions
S.Z. designed and supervised the project. Y.H., H.S., Y.Z. and Y.W. performed and analysed the experiments. J.M. and Y.L. designed, carried out and analysed the computational experiments. All authors co-wrote the manuscript, analysed the data, discussed the results, commented on the manuscript, and approved the final version of the manuscript.
Peer review
Peer review information
Nature Communications thanks the anonymous reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.
Data availability
The authors declare that the main data supporting the findings of this study, including experimental procedures and compound characterization, are available within the article and its supplementary information files, and also are available from the corresponding author. CCDC 2122000 contains the supplementary crystallographic data for 4t. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.
Competing interests
The authors declare the following competing interest(s): S.Z., Y.Z., and Y.H. are inventors on a patent application number CN202110888015.6 which is based on multiligand relay catalysis.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Yong Liang, Email: yongliang@nju.edu.cn.
You Wang, Email: wangyou@nju.edu.cn.
Shaolin Zhu, Email: shaolinzhu@nju.edu.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-022-30006-2.
References
- 1.Tasker SZ, Standley EA, Jamison TF. Recent advances in homogeneous nickel catalysis. Nature. 2014;509:299–309. doi: 10.1038/nature13274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Fu GC. Transition-metal catalysis of nucleophilic substitution reactions: a radical alternative to SN1 and SN2 processes. ACS Cent. Sci. 2017;3:692–700. doi: 10.1021/acscentsci.7b00212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Poremba KE, Dibrell SE, Reisman SE. Nickel-catalyzed enantioselective reductive cross-coupling reactions. ACS Catal. 2020;10:8237–8246. doi: 10.1021/acscatal.0c01842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Larionov E, Li H, Mazet C. Well-defined transition metal hydrides in catalytic isomerizations. Chem. Commun. 2014;50:9816–9826. doi: 10.1039/C4CC02399D. [DOI] [PubMed] [Google Scholar]
- 5.Vasseur A, Bruffaerts J, Marek I. Remote functionalization through alkene isomerization. Nat. Chem. 2016;8:209–219. doi: 10.1038/nchem.2445. [DOI] [PubMed] [Google Scholar]
- 6.Sommer H, Juliá-Hernández F, Martin R, Marek I. Walking metals for remote functionalization. ACS Cent. Sci. 2018;4:153–165. doi: 10.1021/acscentsci.8b00005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Janssen-Müller D, Sahoo B, Sun S-Z, Martin R. Tackling remote sp3 C−H functionalization via Ni-catalyzed “chain-walking” reactions. Isr. J. Chem. 2020;60:195–206. doi: 10.1002/ijch.201900072. [DOI] [Google Scholar]
- 8.Wang X-X, Lu X, Li Y, Wang J-W, Fu Y. Recent advances in nickel-catalyzed reductive hydroalkylation and hydroarylation of electronically unbiased alkenes. Sci. China Chem. 2020;63:1586–1600. doi: 10.1007/s11426-020-9838-x. [DOI] [Google Scholar]
- 9.He Y, Chen J, Jiang X, Zhu S. Enantioselective NiH-catalyzed reductive hydrofunctionalization of alkenes. Chin. J. Chem. 2022;40:651–661. doi: 10.1002/cjoc.202100763. [DOI] [Google Scholar]
- 10.He Y, Cai Y, Zhu S. Mild and regioselective benzylic C–H functionalization: Ni-catalyzed reductive arylation of remote and proximal olefins. J. Am. Chem. Soc. 2017;139:1061–1064. doi: 10.1021/jacs.6b11962. [DOI] [PubMed] [Google Scholar]
- 11.Juliá-Hernández F, Moragas T, Cornella J, Martin R. Remote carboxylation of halogenated aliphatic hydrocarbons with carbon dioxide. Nature. 2017;545:84–88. doi: 10.1038/nature22316. [DOI] [PubMed] [Google Scholar]
- 12.Gaydou M, Moragas T, Juliá-Hernández F, Martin R. Site-selective catalytic carboxylation of unsaturated hydrocarbons with CO2 and water. J. Am. Chem. Soc. 2017;139:12161–12164. doi: 10.1021/jacs.7b07637. [DOI] [PubMed] [Google Scholar]
- 13.Chen F, et al. Remote migratory cross-electrophile coupling and olefin hydroarylation reactions enabled by in situ generation of NiH. J. Am. Chem. Soc. 2017;139:13929–13935. doi: 10.1021/jacs.7b08064. [DOI] [PubMed] [Google Scholar]
- 14.Zhou F, Zhu J, Zhang Y, Zhu S. NiH-catalyzed reductive relay hydroalkylation: a strategy for the remote C(sp3)–H alkylation of alkenes. Angew. Chem. Int. Ed. 2018;57:4058–4062. doi: 10.1002/anie.201712731. [DOI] [PubMed] [Google Scholar]
- 15.Sun S-Z, Börjesson M, Martin-Montero R, Martin R. Site-selective Ni-catalyzed reductive coupling of α-haloboranes with unactivated olefins. J. Am. Chem. Soc. 2018;140:12765–12769. doi: 10.1021/jacs.8b09425. [DOI] [PubMed] [Google Scholar]
- 16.Zhang Y, Xu X, Zhu S. Nickel-catalysed selective migratory hydrothiolation of alkenes and alkynes with thiols. Nat. Commun. 2019;10:1752. doi: 10.1038/s41467-019-09783-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Sun S-Z, Romano C, Martin R. Site-selective catalytic deaminative alkylation of unactivated olefins. J. Am. Chem. Soc. 2019;141:16197–16201. doi: 10.1021/jacs.9b07489. [DOI] [PubMed] [Google Scholar]
- 18.Zhang Y, Han B, Zhu S. Rapid access to highly functionalized alkylboronates via NiH-catalyzed remote hydroarylation of boron-containing alkenes. Angew. Chem. Int. Ed. 2019;58:13860–13864. doi: 10.1002/anie.201907185. [DOI] [PubMed] [Google Scholar]
- 19.Qian D, Hu X. Ligand-controlled regiodivergent hydroalkylation of pyrrolines. Angew. Chem. Int. Ed. 2019;58:18519–18523. doi: 10.1002/anie.201912629. [DOI] [PubMed] [Google Scholar]
- 20.Li YQ, et al. Reaction scope and mechanistic insights of nickel-catalyzed migratory Suzuki-Miyaura cross-coupling. Nat. Commun. 2020;11:417. doi: 10.1038/s41467-019-14016-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Gao Y, et al. Visible-light-induced nickel-catalyzed cross-coupling with alkylzirconocenes from unactivated alkenes. Chem. 2020;6:675–688. doi: 10.1016/j.chempr.2019.12.010. [DOI] [Google Scholar]
- 22.Yang C, Gao Y, Bai S, Jiang C, Qi X. Chemoselective cross-coupling of gem-borazirconocene alkanes with aryl halides. J. Am. Chem. Soc. 2020;142:11506–11513. doi: 10.1021/jacs.0c03821. [DOI] [PubMed] [Google Scholar]
- 23.Zhang Y, He J, Song P, Wang Y, Zhu S. Ligand-enabled NiH-catalyzed migratory hydroamination: chain walking as a strategy for regiodivergent/regioconvergent remote sp3 C–H amination. CCS Chem. 2021;3:2259–2268. doi: 10.31635/ccschem.020.202000490. [DOI] [Google Scholar]
- 24.Lu X, et al. Practical carbon–carbon bond formation from olefins through nickel-catalyzed reductive olefin hydrocarbonation. Nat. Commun. 2016;7:11129. doi: 10.1038/ncomms11129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Shevick SL, Obradors C, Shenvi RA. Mechanistic interrogation of Co/Ni-dual catalyzed hydroarylation. J. Am. Chem. Soc. 2018;140:12056–12068. doi: 10.1021/jacs.8b06458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Bera S, Hu X. Nickel-catalyzed regioselective hydroalkylation and hydroarylation of alkenyl boronic esters. Angew. Chem. Int. Ed. 2019;58:13854–13859. doi: 10.1002/anie.201907045. [DOI] [PubMed] [Google Scholar]
- 27.Wang Z, Yin H, Fu GC. Catalytic enantioconvergent coupling of secondary and tertiary electrophiles with olefins. Nature. 2018;563:379–383. doi: 10.1038/s41586-018-0669-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Zhou F, Zhang Y, Xu X, Zhu S. NiH-catalyzed remote asymmetric hydroalkylation of alkenes with racemic α-bromo amides. Angew. Chem. Int. Ed. 2019;58:1754–1758. doi: 10.1002/anie.201813222. [DOI] [PubMed] [Google Scholar]
- 29.He S-J, et al. Nickel-catalyzed enantioconvergent reductive hydroalkylation of olefins with α-heteroatom phosphorus or sulfur alkyl electrophiles. J. Am. Chem. Soc. 2020;142:214–221. doi: 10.1021/jacs.9b09415. [DOI] [PubMed] [Google Scholar]
- 30.Yang Z-P, Fu GC. Convergent catalytic asymmetric synthesis of esters of chiral dialkyl carbinols. J. Am. Chem. Soc. 2020;142:5870–5875. doi: 10.1021/jacs.0c01324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.He Y, Liu C, Yu L, Zhu S. Enantio- and regioselective NiH-catalyzed reductive hydroarylation of vinylarenes with aryl iodides. Angew. Chem. Int. Ed. 2020;59:21530–21534. doi: 10.1002/anie.202010386. [DOI] [PubMed] [Google Scholar]
- 32.Bera S, Mao R, Hu X. Enantioselective C(sp3)–C(sp3) cross-coupling of non-activated alkyl electrophiles via nickel hydride catalysis. Nat. Chem. 2021;13:270–277. doi: 10.1038/s41557-020-00576-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Shi L, Xing L-L, Hu W-B, Shu W. Regio- and enantioselective Ni-catalyzed formal hydroalkylation, hydrobenzylation and hydropropargylation of acrylamides to α-tertiary amides. Angew. Chem. Int. Ed. 2021;60:1599–1604. doi: 10.1002/anie.202011339. [DOI] [PubMed] [Google Scholar]
- 34.Cuesta-Galisteo S, Schörgenhumer J, Wei X, Merino E, Nevado C. Nickel-catalyzed asymmetric synthesis of α-arylbenzamides. Angew. Chem. Int. Ed. 2021;60:1605–1609. doi: 10.1002/anie.202011342. [DOI] [PubMed] [Google Scholar]
- 35.Liu J, Gong H, Zhu S. Nickel-catalyzed, regio- and enantioselective benzylic alkenylation of olefins with alkenyl bromide. Angew. Chem. Int. Ed. 2021;60:4060–4064. doi: 10.1002/anie.202012614. [DOI] [PubMed] [Google Scholar]
- 36.He Y, Song H, Chen J, Zhu S. NiH-catalyzed asymmetric hydroarylation of N-acyl enamines: practical access to chiral benzylamines. Nat. Commun. 2021;12:638. doi: 10.1038/s41467-020-20888-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Qian D, Bera S, Hu X. Chiral alkyl amine synthesis via catalytic enantioselective hydroalkylation of enecarbamates. J. Am. Chem. Soc. 2021;143:1959–1967. doi: 10.1021/jacs.0c11630. [DOI] [PubMed] [Google Scholar]
- 38.Wang J-W, et al. Catalytic asymmetric reductive hydroalkylation of enamides and enecarbamates to chiral aliphatic amines. Nat. Commun. 2021;12:1313. doi: 10.1038/s41467-021-21600-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Wang X-X, et al. NiH-catalyzed reductive hydrocarbonation of enol esters and ethers. CCS Chem. 2021;3:727–737. [Google Scholar]
- 40.Wang S, et al. Enantioselective access to chiral aliphatic amines and alcohols via Ni-catalyzed hydroalkylations. Nat. Commun. 2021;12:2771. doi: 10.1038/s41467-021-22983-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Jiang X, et al. Nickel-catalysed migratory hydroalkynylation and enantioselective hydroalkynylation of olefins with bromoalkynes. Nat. Commun. 2021;12:3792. doi: 10.1038/s41467-021-24094-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Zhou F, Zhu S. Catalytic asymmetric hydroalkylation of α,β-unsaturated amides enabled by regio-reversed and enantiodifferentiating syn-hydronickellation. ACS Catal. 2021;11:8766–8773. doi: 10.1021/acscatal.1c02299. [DOI] [Google Scholar]
- 43.Chen J, Zhu S. Nickel-catalyzed multicomponent coupling: synthesis of α-chiral ketones by reductive hydrocarbonylation of alkenes. J. Am. Chem. Soc. 2021;143:14089–14096. doi: 10.1021/jacs.1c07851. [DOI] [PubMed] [Google Scholar]
- 44.Meng L, Yang J, Duan M, Wang Y, Zhu S. Facile synthesis of chiral arylamines, alkylamines and amides by enantioselective NiH-catalyzed hydroamination. Angew. Chem. Int. Ed. 2021;60:23584–23589. doi: 10.1002/anie.202109881. [DOI] [PubMed] [Google Scholar]
- 45.Cheng L, Li M-M, Xiao L-J, Xie J-H, Zhou Q-L. Nickel(0)-catalyzed hydroalkylation of 1,3-dienes with simple ketones. J. Am. Chem. Soc. 2018;140:11627–11630. doi: 10.1021/jacs.8b09346. [DOI] [PubMed] [Google Scholar]
- 46.Chen Y-G, et al. Nickel-catalyzed enantioselective hydroarylation and hydroalkenylation of styrenes. J. Am. Chem. Soc. 2019;141:3395–3399. doi: 10.1021/jacs.8b13524. [DOI] [PubMed] [Google Scholar]
- 47.Lv X-Y, Fan C, Xiao L-J, Xie J-H, Zhou Q-L. Ligand-enabled Ni-catalyzed enantioselective hydroarylation of styrenes and 1,3-dienes with arylboronic acids. CCS Chem. 2019;1:328–334. doi: 10.31635/ccschem.019.20190026. [DOI] [Google Scholar]
- 48.Zhang W-B, Yang X-T, Ma J-B, Su Z-M, Shi S-L. Regio- and enantioselective C–H cyclization of pyridines with alkenes enabled by a nickel/N-heterocyclic carbene catalysis. J. Am. Chem. Soc. 2019;141:5628–5634. doi: 10.1021/jacs.9b00931. [DOI] [PubMed] [Google Scholar]
- 49.Marcum JS, Taylor TR, Meek SJ. Enantioselective synthesis of functionalized arenes by nickel-catalyzed site-selective hydroarylation of 1,3-dienes with aryl boronates. Angew. Chem. Int. Ed. 2020;59:14070–14075. doi: 10.1002/anie.202004982. [DOI] [PubMed] [Google Scholar]
- 50.Yu R, Rajasekar S, Fang X. Enantioselective nickel-catalyzed migratory hydrocyanation of nonconjugated dienes. Angew. Chem. Int. Ed. 2020;59:21436–21441. doi: 10.1002/anie.202008854. [DOI] [PubMed] [Google Scholar]
- 51.Gennari C, Piarulli U. Combinatorial libraries of chiral ligands for enantioselective catalysis. Chem. Rev. 2003;103:3071–3100. doi: 10.1021/cr020058r. [DOI] [PubMed] [Google Scholar]
- 52.Ding K. Synergistic effect of binary component ligands in chiral catalyst library engineering for enantioselective reactions. Chem. Commun. 2008;44:909–921. doi: 10.1039/b710668h. [DOI] [PubMed] [Google Scholar]
- 53.Reetz MT. Combinatorial transition-metal catalysis: mixing monodentate ligands to control enantio-, diastereo-, and regioselectivity. Angew. Chem. Int. Ed. 2008;47:2556–2588. doi: 10.1002/anie.200704327. [DOI] [PubMed] [Google Scholar]
- 54.Fors BP, Buchwald SL. A multiligand based Pd catalyst for C−N cross-coupling reactions. J. Am. Chem. Soc. 2010;132:15914–15917. doi: 10.1021/ja108074t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Chen MS, Prabagaran N, Labenz NA, White MC. Serial ligand catalysis: a highly selective allylic C−H oxidation. J. Am. Chem. Soc. 2005;127:6970–6971. doi: 10.1021/ja0500198. [DOI] [PubMed] [Google Scholar]
- 56.Fraunhoffer KJ, Prabagaran N, Sirois LE, White MC. Macrolactonization via hydrocarbon oxidation. J. Am. Chem. Soc. 2006;128:9032–9033. doi: 10.1021/ja063096r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Young AJ, White MC. Allylic C–H alkylation of unactivated α-olefins: serial ligand catalysis resumed. Angew. Chem. Int. Ed. 2011;50:6824–6827. doi: 10.1002/anie.201101654. [DOI] [PubMed] [Google Scholar]
- 58.Chen C, Peters JC, Fu GC. Photoinduced copper-catalysed asymmetric amidation via ligand cooperativity. Nature. 2021;596:250–256. doi: 10.1038/s41586-021-03730-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kim-Lee S-H, Mauleón P, Arrayás RG, Carretero JC. Dynamic multiligand catalysis: a polar to radical crossover strategy expands alkyne carboboration to unactivated secondary alkyl halides. Chem. 2021;7:2212–2226. doi: 10.1016/j.chempr.2021.06.002. [DOI] [Google Scholar]
- 60.Zhang Y, et al. A relay catalysis strategy for enantioselective nickel-catalyzed migratory hydroarylation forming chiral α-aryl alkylboronates. Chem. 2021;7:3171–3188. doi: 10.1016/j.chempr.2021.10.015. [DOI] [Google Scholar]
- 61.Do H-Q, Chandrashekar ERR, Fu GC. Nickel/bis(oxazoline)-catalyzed asymmetric Negishi arylations of racemic secondary benzylic electrophiles to generate enantioenriched 1,1-diarylalkanes. J. Am. Chem. Soc. 2013;135:16288–16291. doi: 10.1021/ja408561b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Gutierrez O, Tellis JC, Primer DN, Molander GA, Kozlowski MC. Nickel-catalyzed cross-coupling of photoredox-generated radicals: uncovering a general manifold for stereoconvergence in nickel-catalyzed cross-couplings. J. Am. Chem. Soc. 2015;137:4896–4899. doi: 10.1021/ja513079r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Poremba KE, Kadunce NT, Suzuki N, Cherney AH, Reisman SE. Nickel-catalyzed asymmetric reductive cross-coupling to access 1,1-diarylalkanes. J. Am. Chem. Soc. 2017;139:5684–5687. doi: 10.1021/jacs.7b01705. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Cheng X, Lu H, Lu Z. Enantioselective benzylic C–H arylation via photoredox and nickel dual catalysis. Nat. Commun. 2019;10:3549. doi: 10.1038/s41467-019-11392-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Anthony D, Lin Q, Baudet J, Diao T. Ni-catalyzed asymmetric reductive diarylation of vinylarenes. Angew. Chem. Int. Ed. 2019;58:3198–3202. doi: 10.1002/anie.201900228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Wang W, et al. Migratory arylboration of unactivated alkenes enabled by nickel catalysis. Angew. Chem. Int. Ed. 2019;58:4612–4616. doi: 10.1002/anie.201814572. [DOI] [PubMed] [Google Scholar]
- 67.Li Y, et al. Nickel-catalyzed 1,1-alkylboration of electronically unbiased terminal alkenes. Angew. Chem. Int. Ed. 2019;58:8872–8876. doi: 10.1002/anie.201903890. [DOI] [PubMed] [Google Scholar]
- 68.Wang W, Ding C, Yin G. Catalyst-controlled enantioselective 1,1-arylboration of unactivated olefins. Nat. Catal. 2020;3:951–958. doi: 10.1038/s41929-020-00523-8. [DOI] [Google Scholar]
- 69.McCarty BJ, Tang W. A dancing nickel in asymmetric catalysis: enantioselective synthesis of boronic esters by 1,1-addition to terminal alkenes. Green. Synth. Catal. 2021;2:1–3. doi: 10.1016/j.gresc.2020.12.004. [DOI] [Google Scholar]
- 70.Ding C, Ren Y, Sun C, Long J, Yin G. Regio- and stereoselective alkylboration of endocyclic olefins enabled by nickel catalysis. J. Am. Chem. Soc. 2021;143:20027–20034. doi: 10.1021/jacs.1c09214. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The authors declare that the main data supporting the findings of this study, including experimental procedures and compound characterization, are available within the article and its supplementary information files, and also are available from the corresponding author. CCDC 2122000 contains the supplementary crystallographic data for 4t. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.