and Enantioselective Rhodium-Catalyzed Allylic ... - ACS Publications

May 25, 2018 - (g) Pineschi, M.; Bussolo, V. D. I.; Crotti, P. Chirality 2011, 23,. 703. .... F. A.; Zhu, Y.; Tercenio, Q. D.; Shen, Z.; Dong, V. M. J...
0 downloads 0 Views 1MB Size
Article Cite This: J. Am. Chem. Soc. 2018, 140, 7737−7742

pubs.acs.org/JACS

Regio- and Enantioselective Rhodium-Catalyzed Allylic Alkylation of Racemic Allylic Alcohols with 1,3-Diketones Sheng-Biao Tang,† Xiao Zhang,† Hang-Fei Tu,† and Shu-Li You*,†,‡ †

State Key Laboratory of Organometallic Chemistry, Center for Excellence in Molecular Synthesis, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, 345 Lingling Lu, Shanghai 200032, China ‡ Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 300072, China

Downloaded via UNIV OF TOLEDO on June 29, 2018 at 13:49:26 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Highly regio- and enantioselective rhodiumcatalyzed allylic alkylation of 1,3-diketones with racemic secondary allylic alcohols is reported. In the presence of a Rh-catalyst derived from the Carreira (P, olefin)-ligand and TFA as an additive, chiral branched α-allylated 1,3-diketones could be obtained in good to excellent yields, with excellent regio- and enantioselectivity (b/l > 19/1, 86−98% ee). The direct utilization of allyl alcohols as electrophiles represents an improvement from the viewpoint of an atom economy. Both aryl- and aliphatic-substituted allyl alcohols are suitable substrates with excellent reaction outcomes. This reaction features mild conditions, broad substrate scope, and readily available substrates.



system (eq 2, Scheme 1).12 Kazmaier and co-workers reported Rh-catalyzed AAA reactions of chelated enolates with enantioenriched allyl phosphates, where excellent chiral transfer was achieved.13 To be noted, the groups of Nguyen,14 Breit,15 Vrieze,16 and Guo17 reported asymmetric allylic amination reactions with racemic allylic esters. Recently, pioneering work from the groups of Breit,18 Dong,19 and Kang20 demonstrated that allenes and alkynes are suitable allylic precursors to afford allylic alkylation products via a Rh-catalytic system. However, the reactions with aryl-substituted substrates generally led to moderate enantioselectivity (eq 3, Scheme 1).18e Considering that the known reports are mainly limited to employment of activated allylic precursors and allenes, direct utilization of readily available allylic alcohols will be highly desirable.21 Herein, we report our results on a Rh-(P, olefin) complex-catalyzed AAA with secondary racemic allylic alcohols directly.

INTRODUCTION

Transition-metal-catalyzed asymmetric allylic alkylation (AAA) reactions provide a straightforward and reliable method to construct the allylic stereocenters.1 However, the transition metals employed for AAA reactions mainly focused on Pd,2 Ir,3 and Cu.4 Utilization of rhodium catalytic systems for AAA reaction remains underexplored.5−9 In this regard, Evans and co-workers pioneered rhodium-catalyzed allylic alkylation reactions and disclosed asymmetric reactions employing enantioenriched chiral allylic substrates via a double-inversion process (eq 1, Scheme 1).10 In addition, recent efforts from the Evans group led to elegant Rh-catalyzed AAA of prochiral nucleophiles.11 In 2003, Hayashi et al. reported a very impressive AAA with allylic acetates by a Rh-Phox catalytic Scheme 1. Rh-Catalyzed AAA of Allylic Esters or Addition of Allenes



RESULTS AND DISCUSSION Optimization of the Reaction Conditions. We began our study by testing the reaction between phenyl vinyl carbinol (1a) and acetylacetone (2a) with the Rh catalyst derived from [Rh(cod)Cl]2 and the Carreira P/olefin ligand22 (Table 1). In the presence of 2 mol % of [Rh(cod)Cl]2 and 5 mol % of (S)L1, the reaction did not occur in CH2Cl2 (entry 1). With 100 mol % Fe(OTf)2 as an additive, the alkylation products (3a/4a) were obtained in 10/1 b/l regioselectivity, with the isolation of 3a in 39% ee (entry 2). To our delight, with TFA as an additive, the desired product 3a was obtained in 94% yield and 71% ee, Received: May 16, 2018 Published: May 25, 2018 © 2018 American Chemical Society

7737

DOI: 10.1021/jacs.8b05126 J. Am. Chem. Soc. 2018, 140, 7737−7742

Article

Journal of the American Chemical Society Table 1. Optimization of the Reaction Conditionsa

entry

[Rh]

solvent

ligand

additive

yield (%)b

3a/4ac

ee (%)d

1 2e 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh1 Rh2 Rh3 Rh4 Rh2 Rh2 Rh2

CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 DCE toluene CHCl3 Et2O t BuO-Me DME THF Et2O Et2O Et2O Et2O Et2O Et2O

L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L1 L2 L3 L4

Fe(OTf)2 TFA PhCO2H TsOH (nBuO)2PO2H TfOH TFA TFA TFA TFA TFA TFA TFA TFA TFA TFA TFA TFA TFA

85 94 93 82 trace 78 90 93 23 94 92 90 62 95

10/1 >19/1 >19/1 >19/1

39 71 67 69

>19/1 >19/1 >19/1 >19/1 >19/1 >19/1 >19/1 >19/1 >19/1

66 71 75 92 86 89 35 95

>19/1

86

>19/1

76

91 trace 82

a

General conditions: 1a (0.4 mmol), 2a (0.2 mmol), [Rh] (2 mol %), ligand (5 mol %), and additive (25 mol %) in solvent (2.5 mL) at room temperature. bYield of isolated product. cDetermined by 1H NMR analysis. dDetermined by HPLC analysis. eFe(OTf)2 (100 mol %) was used. DME = 1,2-dimethoxyethane. [Rh(nbd)Cl]2 = norbornadiene rhodium(I) chloride dimer.

products in 90−95% yields, b/l > 19/1, and 87−95% ee (3a− 3e) in 10−24 h. The substrates with an electron-withdrawing group (CF3, F, Cl, Br, I, CO2Me) on the phenyl ring required a prolonged reaction time, providing good to excellent yields (86−94%) and enantioselectivity (91−97% ee) (3f−3k). The good tolerance with Cl, Br, and I groups, which are known to undergo diverse Pd-mediated transformations, shows the advantage of Rh catalysis over Pd catalysis and offers easy handles for further modification. Moreover, the reactions of 2naphthyl- and bipenyl-substituted allyl alcohols proceeded smoothly, leading to the isolation of 3l and 3m in 85% yield with 91% ee and in 88% yield and 87% ee, respectively. The reaction of 2-furanyl allyl alcohol proceeded in slightly decreased yield (60%, 87% ee, 3n), likely due to the instability of the substrate under acidic conditions. 2-Benzothiophenyl allylic alcohol was also a suitable substrate, leading to product 3o in 88% yield and 93% ee. Next, aliphatic allylic alcohols were also explored (Scheme 3), which were known as challenging substrates in allylic substitution reactions catalyzed by other metals.23 Under the slightly modified conditions, [Rh(C2H4)2Cl]2 (3 mol %)/(S)L1 (7 mol %) and TFA (100 mol %), the reactions of aliphatic allyl alcohols could provide their corresponding alkylation products in high to excellent selectivities and good yields (Scheme 3). With substrates bearing longer alkyl chain

with an excellent regioselectivity (b/l > 19/1) (entry 3). Next, several Brønsted acids such as benzoic acid, TsOH, (nBuO)2PO2H, and TfOH were tested, and none of them gave better results (entries 4−7). Furthermore, the solvents were examined (entries 8−14), and the reaction in Et2O led to excellent yield and high regio- and enantioselectivity (94% yield, b/l > 19/1, 92% ee, entry 11). Different rhodium precursors such as [Rh(C2H4)2Cl]2, [Rh(nbd)Cl]2, and [Rh(cod)OTf]2 were examined in Et2O (entries 15−17). The catalyst derived from [Rh(C2H4)2Cl]2 gave the best results (95% yield, b/l > 19/1, 95% ee, entry 15). Further examination of ligands L2−L4 afforded no better results (entries 18−20). Notably, the loss of reactivity with the catalyst derived from ligand L3 without an olefin moiety indicated that the active catalyst requires the P−olefin binding mode. Finally, the optimal conditions were obtained as follows: [Rh(C2H4)2Cl]2 (2 mol %) and (S)-L1 (5 mol %) were stirred for 15 min in Et2O, followed by the addition of substrates and TFA (25 mol %). Reaction Scope. With the optimized reaction conditions in hand, we next investigated the substrate scope (Scheme 2). Various racemic aromatic allylic alcohols were allowed to react with acetylacetone (2a). The substrates bearing an electrondonating group (4-Me, 4-OMe, 3-OMe, 2-OMe) at the different positions of the phenyl ring led to their desired 7738

DOI: 10.1021/jacs.8b05126 J. Am. Chem. Soc. 2018, 140, 7737−7742

Article

Journal of the American Chemical Society

groups such as Br, N3, and benzoyl and ether were very compatible, and the alkylation products (3v−3y) were obtained in excellent enantioselectivity (93−97% ee). Moreover, cyclopentyl- and cyclohexyl-substituted substrates also worked well, yielding the alkylation products (3z−3aa) in excellent regioand enantioselectivity (89−90% ee). The scope of 1,3-diketones was then examined. The results are summarized in Scheme 4. The reaction of 1,3-diones with

Scheme 2. Scope of the Aromatic Racemic Allylic Alcoholsa−d

Scheme 4. Scope of 1,3-Diketonesa−d

a

General conditions: 1a (0.4 mmol), 2 (0.2 mmol), [Rh(C2H4)2Cl]2 (2 mol %), L1 (5 mol %), and TFA (25 mol %) in Et2O (2.5 mL) at room temperature. bIsolated yield. cDetermined by 1H NMR analysis. d Determined by HPLC analysis. eGram scale: 1a (14 mmol), 2a (7 mmol), [Rh(C2H4)2Cl]2 (3 mol %), L1 (7 mol %), and TFA (25 mol %) in Et2O (80 mL) at room temperature for 72 h.

a

General conditions: 1 (0.4 mmol), 2a (0.2 mmol), [Rh(C2H4)2Cl]2 (2 mol %), L1 (5 mol %), and TFA (25 mol %) in Et2O (2.5 mL) at room temperature. bIsolated yield. cDetermined by 1H NMR analysis. d Determined by HPLC analysis.

alcohol 1a gave product 3ab and 3ac in 89% yield with 92% ee and 82% yield with 86% ee, respectively. However, when the sterically more congested derivative (R = tBu) was tested, the enantioselectivity was decreased (31% ee, 3ad, Scheme 4). The reaction of 1,3-diphenyl-1,3-propanedione afforded the branched product 3ae in 96% yield and 88% ee. Excellent enantioselectivity (87−93% ee) was obtained for the reactions of α-branched 1,3-diketones (3af−3ah). Furthermore, the reaction was performed on a gram-scale to give the isolated product in 75% yield (1.14 g) and 94% ee (3a, Scheme 4). The synthetically valuable unsaturated ketones could be obtained from alcohols in a one-pot protocol including sequential alkylation and deacetylation18e,24 (Scheme 5). The enantiomeric purity of the deacetylation products was well preserved from the alkylation products (5a−5e). The absolute configuration of product 5a was determined as R by comparing the optical rotation with that reported in the literature.25 In addition to 1,3-diketones, cyclic β-ketoester 2a′ was also well tolerated, furnishing the corresponding α-allyl β-ketoester compound 3a′ in 83% yield with 6/1 dr, and the ee for the major diastereoisomer was confirmed to be 96% (eq 1, Scheme 6). However, the reaction based on acyclic β-ketoester 2b′ did not proceed at all under the previously optimized conditions. Finally, we found that the desired product 3b′ could be obtained in 96% yield with excellent regioselectivity and 1.2/1 dr by changing the acid promoter to one equivalent amount of Yb(OTf)3. The subsequent decarboxylation occurred to generate 5a in 62% yield and 70% ee with an opposite configuration (S) (eq 2, Scheme 6). Similarly, dimethyl malonate was proved to be a viable participate, leading to the alkylation product 3c′ in good yield, excellent regioselectivity, and moderate enantioselectivity (b/l > 19/1, 89% yield, 65% ee, eq 3, Scheme 6).

Scheme 3. Scope of the Aliphatic Allylic Alcoholsa−d

a

General conditions: 1 (0.4 mmol), 2a (0.2 mmol), [Rh(C2H4)2Cl]2 (3 mol %), L1 (7 mol %), and TFA (100 mol %) in Et2O (2.5 mL) at room temperature. bIsolated yield. cDetermined by 1H NMR analysis. d Determined by HPLC analysis.

substituents, the yields and enantioselectivity of products were increased (78−89% yields, 92−98% ee, 3p−3s). A variety of aliphatic allylic alcohols that have different substituents in the alkyl chain were allowed to react with acetylacetone (2a), and good yields and excellent enantioselectivity were obtained (3t− 3u). Of particular note, various generally labile functional 7739

DOI: 10.1021/jacs.8b05126 J. Am. Chem. Soc. 2018, 140, 7737−7742

Article

Journal of the American Chemical Society Scheme 5. One-Pot Protocol of Alkylation and Deacetylationa−d

Scheme 8. Kinetic Resolution Studies

a

General conditions: 1 (0.4 mmol), 2 (0.2 mmol), [Rh(C2H4)2Cl]2 (2 mol %), L1 (5 mol %), and TFA (25 mol %) in Et2O (2.5 mL) at room temperature for 24−36 h; then, KOH (0.8 mmol) and EtOH (5 mL) were added, and the mixture was stirred at 70 °C for 30 min. b Isolated yield. cDetermined by 1H NMR analysis. dDetermined by HPLC analysis.

alcohol 1a was recovered with 25 and 65% ee, respectively, which suggested that the kinetic resolution is occurring (eqs 1 and 2, Scheme 8). Meanwhile, the reaction between the same equivalent amounts of 1a and 2a was tested, and 3a was obtained in 82% yield and 87% ee, indicating an overall dynamic kinetic resolution process for the product formation (eq 3, Scheme 8). Proposed Catalytic Cycle. On the basis of the experimental observations, a plausible mechanism was proposed in Scheme 9 by utilizing the formation of 3a as an

Scheme 6. Investigation of β-Ketoesters and Dimethyl Malonatea−d

Scheme 9. Proposed Catalytic Cycle

a

General conditions: 1a (0.4 mmol), 2′ (0.2 mmol), [Rh(C2H4)2Cl]2 (2 mol %), L1 (5 mol %), and TFA (25 mol %) or Yb(OTf)3 (100 mol %) in Et2O (2.5 mL) at room temperature; decarboxylation condition: 3b′ (0.1 mmol), KOH (0.4 mmol), and EtOH/H2O = 1/1 (3 mL) were added, and the mixture was stirred at 80 °C for 1 h. b Isolated yield. cThe dr and b/l were determined by 1H NMR analysis. d Determined by HPLC analysis.

example. In the presence of a catalytic amount of acid additive, the rhodium(I) species (A) undergoes oxidative addition to afford the corresponding (η3-allyl)rhodium(III) species (B) with the release of one molecule of water and a conjugate base of the acid. Then, the acetylacetone (2a) directly attacks the allyl fragment to deliver (η2-allyl)rhodium(I) species (C). The substitution of product (3a) by allylic alcohol (1a) completes the catalytic cycle, and species A is regenerated for the next catalytic cycle (Scheme 9).

Transformations. To further demonstrate the utility of this method, the transformations of the branched allylated 1,3diketones were conducted as shown in Scheme 7. Compound Scheme 7. Transformations of Product 3a



CONCLUSION In conclusion, we realized a rhodium-catalyzed highly regioand enantioselective allylic alkylation reaction of 1,3-diketones, using unactivated allylic alcohols as electrophiles for the first time. A catalytic amount of TFA as the additive plays a key role for activating the alcohol and achieving excellent enantioselective control. Both racemic aromatic and aliphatic allylic alcohols were well tolerated in this reaction, affording the alkylation products in good to excellent yields with excellent regio- and enantioselectivity. A wide range of functional groups are compatible with the reaction conditions. The unsaturated ketones were obtained from alcohols in a one-pot protocol as well. The utilization of readily available allylic alcohols as the starting materials and uniformly high regio- and enantiose-

3a could be condensed with hydroxylamine and hydrazines, leading to their corresponding oxazole (6a) and pyrazoles (7a, 7b), respectively, in excellent yields. There is no erosion of the enantiomeric purity during the transformations. Kinetic Resolution Studies. Because the racemic allylic alcohols were used, we became intrigued to investigate the possible kinetic resolution involved. Accordingly, the reactions were quenched at 2 and 24 h, respectively, as shown in Scheme 8. It was found that the ee of the desired product 3a remained excellent and unchanged in both cases (95%) and the allylic 7740

DOI: 10.1021/jacs.8b05126 J. Am. Chem. Soc. 2018, 140, 7737−7742

Article

Journal of the American Chemical Society

Catalyzed Asymmetric Synthesis; Alexakis, A., Krause, N., Woodward, S., Eds; Wiley: Weinheim, 2014. (5) For selected reviews, see: (a) Oliver, S.; Evans, P. A. Synthesis 2013, 45, 3179. (b) Koschker, P.; Breit, B. Acc. Chem. Res. 2016, 49, 1524. (c) Haydl, A. M.; Breit, B.; Liang, T.; Krische, M. J. Angew. Chem., Int. Ed. 2017, 56, 11312. (6) For selected reviews on Mo-catalyzed AAA reactions, see: (a) Belda, O.; Moberg, C. Acc. Chem. Res. 2004, 37, 159. (b) Moberg, C. Top. Organomet. Chem. 2011, 38, 209. For selected examples, see: (c) Trost, B. M.; Hachiya, I. J. Am. Chem. Soc. 1998, 120, 1104. (d) Glorius, F.; Pfaltz, A. Org. Lett. 1999, 1, 141. (e) Trost, B. M.; Hildbrand, S.; Dogra, K. J. Am. Chem. Soc. 1999, 121, 10416. (f) Belda, O.; Kaiser, N.-F.; Bremberg, U.; Larhed, M.; Hallberg, A.; Moberg, C. J. Org. Chem. 2000, 65, 5868. (g) Trost, B. M.; Zhang, Y. J. Am. Chem. Soc. 2007, 129, 14548. (h) Trost, B. M.; Miller, J. R.; Hoffman, C. M. A. J. Am. Chem. Soc. 2011, 133, 8165. (7) For an early example on tungsten-catalyzed AAA reactions, see: Lloyd-Jones, G. C.; Pfaltz, A. Angew. Chem., Int. Ed. Engl. 1995, 34, 462. (8) For selected reviews on Ru-catalyzed AAA reactions, see: (a) Bruneau, C.; Renaud, J.; Demerseman, B. Chem. - Eur. J. 2006, 12, 5178. (b) Bruneau, C.; Renaud, J.; Demerseman, B. Pure Appl. Chem. 2008, 80, 861. (c) Begouin, J.-M.; Klein, J.; Weickmann, D.; Plietker, B. Top. Organomet. Chem. 2011, 38, 269. For selected examples, see: (d) Matsushima, Y.; Onitsuka, K.; Kondo, T.; Mitsudo, T. A.; Takahashi, S. J. Am. Chem. Soc. 2001, 123, 10405. (e) Onitsuka, K.; Matsushima, Y.; Takahashi, S. Organometallics 2005, 24, 6472. (f) Bayer, A.; Kazmaier, U. Org. Lett. 2010, 12, 4960. (g) Kanbayashi, N.; Hosoda, K.; Kato, M.; Takii, K.; Okamura, T. A.; Onitsuka, K. Chem. Commun. 2015, 51, 10895. (9) For selected examples on platinum-catalyzed AAA reactions, see: (a) Blacker, A. J.; Clark, M. L.; Williams, M. J.; Loft, M. S. Chem. Commun. 1999, 913. (b) Blacker, A. J.; Clarke, M. L.; Loft, M. S.; Mahon, M. F.; Humphries, M. E.; Williams, J. M. Chem. - Eur. J. 2000, 6, 353. (10) (a) Evans, P. A.; Nelson, J. D. J. Am. Chem. Soc. 1998, 120, 5581. (b) Evans, P. A.; Kennedy, L. J. Org. Lett. 2000, 2, 2213. (c) Evans, P. A.; Robinson, J. E. J. Am. Chem. Soc. 2001, 123, 4609. (d) Evans, P. A.; Uraguchi, D. J. Am. Chem. Soc. 2003, 125, 7158. (e) Evans, P. A.; Leahy, D. K. J. Am. Chem. Soc. 2003, 125, 8974. (f) Evans, P. A.; Lawler, M. J. J. Am. Chem. Soc. 2004, 126, 8642. (g) Evans, P. A.; Clizbe, E. A.; Lawler, M. J.; Oliver, S. Chem. Sci. 2012, 3, 1835. (h) Evans, P. A.; Oliver, S.; Chae, J. J. Am. Chem. Soc. 2012, 134, 19314. (i) Evans, P. A.; Oliver, S. Org. Lett. 2013, 15, 5626. (j) Turnbull, B. W. H.; Oliver, S.; Evans, P. A. J. Am. Chem. Soc. 2015, 137, 15374. (k) Turnbull, B. W. H.; Chae, J.; Oliver, S.; Evans, P. A. Chem. Sci. 2017, 8, 4001. (11) (a) Turnbull, B. W. H.; Evans, P. A. J. Am. Chem. Soc. 2015, 137, 6156. (b) Wright, T. B.; Evans, P. A. J. Am. Chem. Soc. 2016, 138, 15303. (12) Hayashi, T.; Okada, A.; Suzuka, T.; Kawatsura, M. Org. Lett. 2003, 5, 1713. (13) Kazmaier, U.; Stolz, D. Angew. Chem., Int. Ed. 2006, 45, 3072. (14) For selected examples, see: (a) Arnold, J. S.; Nguyen, H. M. J. Am. Chem. Soc. 2012, 134, 8380. (b) Arnold, J. S.; Cizio, G. T.; Heitz, D. R.; Nguyen, H. M. Chem. Commun. 2012, 48, 11531. (c) Arnold, J. S.; Nguyen, H. M. Synthesis 2013, 45, 2101. (d) Arnold, J. S.; Mwenda, E. T.; Nguyen, H. M. Angew. Chem., Int. Ed. 2014, 53, 3688. (e) Mwenda, E. T.; Nguyen, H. M. Org. Lett. 2017, 19, 4814. (15) Li, C.; Breit, B. Chem. - Eur. J. 2016, 22, 14655. (b) Zhou, Y.; Breit, B. Chem. - Eur. J. 2017, 23, 18156. (16) Vrieze, D. C.; Hoge, G. S.; Hoerter, P. Z.; Van Haitsma, J. T.; Samas, B. M. Org. Lett. 2009, 11, 3140. (17) Liang, L.; Xie, M.-S.; Qin, T.; Zhu, M.; Qu, G.-R.; Guo, H.-M. Org. Lett. 2017, 19, 5212. (18) For selected examples, see: (a) Li, C.; Breit, B. J. Am. Chem. Soc. 2014, 136, 862. (b) Beck, T. M.; Breit, B. Org. Lett. 2016, 18, 124. (c) Li, C.; Grugel, C. P.; Breit, B. Chem. Commun. 2016, 52, 5840.

lectivity for both aryl and aliphatic allylic alcohols warrant the synthetic potentials of this method.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.8b05126. Experimental procedures and compound characterization data (PDF)



AUTHOR INFORMATION

Corresponding Author

*[email protected]. ORCID

Shu-Li You: 0000-0003-4586-8359 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank the National Key R&D Program of China (2016YFA0202900), the National Basic Research Program of China (2015CB856600), NSFC (21332009, 21572252), Program of Shanghai Subject Chief Scientist (16XD1404300), and the CAS (XDB20000000, QYZDY-SSW-SLH012) for generous financial support. S.-B.T. thanks Pharmaron for a postdoctoral fellowship.



REFERENCES

(1) For selected reviews, see: (a) Trost, B. M.; Van Vranken, D. L. Chem. Rev. 1996, 96, 395. (b) Trost, B. M.; Crawley, M. L. Chem. Rev. 2003, 103, 2921. (c) Lu, Z.; Ma, S. Angew. Chem., Int. Ed. 2008, 47, 258. (d) Weaver, J. D.; Recio, A., III.; Grenning, A. J.; Tunge, J. A. Chem. Rev. 2011, 111, 1846. (e) Butt, N. A.; Zhang, W. Chem. Soc. Rev. 2015, 44, 7929. For a book, see: (f) Transition Metal Catalyzed Enantioselective Allylic Substitution. In Organic Synthesis; Kazmaier, U., Ed.; Springer: Heidelberg, 2012. (2) For selected reviews, see: (a) Helmchen, G. J. Organomet. Chem. 1999, 576, 203. (b) Tenaglia, A.; Heumann, A. Angew. Chem., Int. Ed. 1999, 38, 2180. (c) Trost, B. M. Chem. Pharm. Bull. 2002, 50, 1. (d) Kazmaier, U. Curr. Org. Chem. 2003, 7, 317. (e) Trost, B. M. J. Org. Chem. 2004, 69, 5813. (f) Jensen, T.; Fristrup, P. Chem. - Eur. J. 2009, 15, 9632. (g) Milhau, L.; Guiry, P. J. Top. Organomet. Chem. 2011, 38, 95. (3) For selected reviews, see: (a) Takeuchi, R.; Kezuka, S. Synthesis 2006, 2006, 3349. (b) Helmchen, G.; Dahnz, A.; Dübon, P.; Schelwies, M.; Weihofen, R. Chem. Commun. 2007, 675. (c) Hartwig, J. F.; Stanley, L. M. Acc. Chem. Res. 2010, 43, 1461. (d) Hartwig, J. F.; Pouy, M. J. Top. Organomet. Chem. 2011, 34, 169. (e) Liu, W.-B.; Xia, J.-B.; You, S.-L. Top. Organomet. Chem. 2011, 38, 155. (f) Tosatti, P.; Nelson, A.; Marsden, S. P. Org. Biomol. Chem. 2012, 10, 3147. (g) Helmchen, G. In Molecular Catalysis; Gade, L. H., Hofmann, P., Eds.; Wiley-VCH: Weinheim, 2014; pp 235−254. (h) Zhuo, C.-X.; Zheng, C.; You, S.-L. Acc. Chem. Res. 2014, 47, 2558. (i) Hethcox, J. C.; Shockley, S. E.; Stoltz, B. M. ACS Catal. 2016, 6, 6207. (j) Qu, J.; Helmchen, G. Acc. Chem. Res. 2017, 50, 2539. (4) For selected reviews, see: (a) Pineschi, M. New J. Chem. 2004, 28, 657. (b) Yorimitsu, H.; Oshima, K. Angew. Chem., Int. Ed. 2005, 44, 4435. (c) Alexakis, A.; Bäckvall, J. E.; Krause, N.; Pàmies, O.; Diéguez, M. Chem. Rev. 2008, 108, 2796. (d) Harutyunyan, S. R.; den Hartog, T.; Geurts, K.; Minnaard, A. J.; Feringa, B. L. Chem. Rev. 2008, 108, 2824. (e) Falciola, C. A.; Alexakis, A. Eur. J. Org. Chem. 2008, 2008, 3765. (f) Langlois, J.-B.; Alexakis, A. Top. Organomet. Chem. 2011, 38, 235. (g) Pineschi, M.; Bussolo, V. D. I.; Crotti, P. Chirality 2011, 23, 703. (h) Hornillos, V.; Gualtierotti, J.-B.; Feringa, B. L. Top. Organomet. Chem. 2016, 58, 1. For a recent book, see: (i) Copper7741

DOI: 10.1021/jacs.8b05126 J. Am. Chem. Soc. 2018, 140, 7737−7742

Article

Journal of the American Chemical Society (d) Beck, T. M.; Breit, B. Eur. J. Org. Chem. 2016, 2016, 5839. (e) Beck, T. M.; Breit, B. Angew. Chem., Int. Ed. 2017, 56, 1903. (19) For selected examples, see: (a) Chen, Q.-A.; Chen, Z.; Dong, V. M. J. Am. Chem. Soc. 2015, 137, 8392. (b) Cruz, F. A.; Chen, Z.; Kurtoic, S. I.; Dong, V. M. Chem. Commun. 2016, 52, 5836. (c) Cruz, F. A.; Zhu, Y.; Tercenio, Q. D.; Shen, Z.; Dong, V. M. J. Am. Chem. Soc. 2017, 139, 10641. (d) Chen, Z.; Dong, V. M. Nat. Commun. 2017, 8, 784. (20) (a) Zheng, W.-F.; Xu, Q.-J.; Kang, Q. Organometallics 2017, 36, 2323. (b) Bora, P. P.; Sun, G.-J.; Zheng, W.-F.; Kang, Q. Chin. J. Chem. 2018, 36, 20. (21) For selected reviews, see: (a) Bandini, M. Angew. Chem., Int. Ed. 2011, 50, 994. (b) Bandini, M.; Cera, G.; Chiarucci, M. Synthesis 2012, 44, 504. (c) Sundararaju, B.; Achard, M.; Bruneau, C. Chem. Soc. Rev. 2012, 41, 4467. (22) Defieber, C.; Ariger, M. A.; Moriel, P.; Carreira, E. M. Angew. Chem., Int. Ed. 2007, 46, 3139. (23) (a) Roggen, M.; Carreira, E. M. Angew. Chem., Int. Ed. 2011, 50, 5568. (b) Lafrance, M.; Roggen, M.; Carreira, E. M. Angew. Chem., Int. Ed. 2012, 51, 3470. (c) Liu, W.-B.; Reeves, C. M.; Stoltz, B. M. J. Am. Chem. Soc. 2013, 135, 17298. (24) (a) Fujimoto, K.; Maekawa, H.; Matsubara, Y.; Nishiguchi, I. Chem. Lett. 1996, 25, 143. (b) Kanojia, S. V.; Chatterjee, S.; Gamre, S.; Chattopadhyay, S.; Sharma, A. Tetrahedron 2015, 71, 1732. (25) Hamilton, J. Y.; Sarlah, D.; Carreira, E. M. J. Am. Chem. Soc. 2014, 136, 3006.

7742

DOI: 10.1021/jacs.8b05126 J. Am. Chem. Soc. 2018, 140, 7737−7742