Asymmetric Borylative Propargylation of Ketones Catalyzed by a

3 days ago - Pharmaceutical industry productivity was robust in 2018, a year in which the US Food and Drug Administration... SCIENCE CONCENTRATES ...
0 downloads 0 Views 1MB Size
Letter Cite This: Org. Lett. XXXX, XXX, XXX−XXX

pubs.acs.org/OrgLett

Asymmetric Borylative Propargylation of Ketones Catalyzed by a Copper(I) Complex Xu-Cheng Gan and Liang Yin* CAS Key Laboratory of Synthetic Chemistry of Natural Substances, Center for Excellence in Molecular Synthesis, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, 345 Lingling Road, Shanghai 200032, China

Org. Lett. Downloaded from pubs.acs.org by IOWA STATE UNIV on 01/25/19. For personal use only.

S Supporting Information *

ABSTRACT: A copper(I)-catalyzed asymmetric borylative propargylation of simple ketones was disclosed. Additive NaBARF was found to be pivotal to achieve excellent enantioselectivity. This reaction enjoyed advantages of broad substrate scope, good tolerance of functional groups, high diastereo- and enantioselectivities, and reaction robustness. The borylative product served as a suitable cross-coupling partner in a palladium-catalyzed Suzuki−Miyaura reaction. Finally, the synthetic utility of the methodology was showcased by the asymmetric synthesis of a chiral piperidine derivative.

C

good to excellent enantioselectivity with allenylboronate (Scheme 1a).4g Later, the catalytic asymmetric propargylation of ketones was enabled by a copper(I) catalyst with propargylboronate bearing a bulky TMS group on the other side of the alkyne moiety,6c by a silver catalyst with allenylboronate,6e and by a zinc catalyst with allenylboronate (Scheme 1a).6f Moreover, the asymmetric propargylation of

atalytic asymmetric addition of nucleophiles to ketones is one of most direct methods to access the optically active tertiary alcohols, which are prevalent structure units in biologically active compounds.1 However, simple ketones (unactivated) are inert in the presence of relatively weak nucleophiles, which greatly limits the application of ketones in organic synthesis. Moreover, compared to aldehydes, it is more difficult to discriminate between the two non-hydrogen substituents on the carbonyl group and thus to achieve effective enantiofacial control, which renders the asymmetric catalysis with ketones as the electrophiles less efficient. Therefore, developing effective asymmetric catalysis with simple ketones is a challenge and thus highly desirable in organic synthesis.1 The classical methods for the catalytic asymmetric nucleophilic addition to simple ketones employ various preformed metal reagents, such as alkyl or aryl metal reagents (such as Grignard reagent),1b,2 silyl enolates,3 allyl metal species,4 metal acetylides, 5 allenyl or propargyl metal species,4g,6 and other metallic nucleophiles.7 An atomeconomical method is the catalytic asymmetric addition to ketones with nucleophiles generated through deprotonation.8,9 Another efficient method is the asymmetric addition to ketones with a catalytic amount of nucleophiles generated in situ through reactions other than deprotonation, which generally allows the generation of multiple stereogenic carbon centers.10 Among the various catalytic asymmetric addition to ketones, the propargylation has received significant and continuous attention from the chemical community in recent decades due to the exceptionally versatile application of the chiral homopropargyl alcohols in organic synthesis.11 In 2010, Shibasaki and Kanai disclosed an example of copper(I)-catalyzed asymmetric propargylation of ketones in © XXXX American Chemical Society

Scheme 1. Prior Reactions in the Metal-Catalyzed Asymmetric Propargylation of Ketones and Our Reaction

Received: December 7, 2018

A

DOI: 10.1021/acs.orglett.8b03912 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters Table 1. Optimization of the Reaction Conditionsa

ketones of allenylboronate with an organocatalyst was achieved with the assistance of microwave heating.6d In 2011, the Sigman group reported an asymmetric propargylation of ketones with a chiral chromium catalyst by following the Nozaki−Hiyama−Kishi reaction protocol (Scheme 1b).6b Interestingly, the ligand class was designed and optimized by applying three-dimensional free energy relationships correlating steric and electronic effects. 1,3-Enynes found broad synthetic application in organic synthesis.12 They were also utilized in the catalytic asymmetric propargylation of carbonyl compounds. In 2008, Krische uncovered the first example of ruthenium-catalyzed propargylation of aldehydes with 1,3-enynes.13a The parent alcohols of aldehydes can be direct employed by using a transfer hydrogenative C−C coupling strategy. The low diastereoselectivity was further improved by modification of the structure of 1,3-enynes.13b The Krische group also successfully developed a catalytic asymmetric version of the reductive propargylation of aldehyde (or the parent alcohols) with an iridium catalyst in excellent diastereo- and enantioselectivities.13c The scope of 1,3-enynes was further expanded to 2methyl-1,3-enynes by switching the chiral iridium catalyst to a chiral ruthenium catalyst, which led to an array of chiral homopropargyl alcohols containing a gem-dimethyl moiety in excellent enantioselectivity.13e Moreover, 1,3-enynes were employed in the copper(I)-catalyzed asymmetric borylative propargylation of aldehydes by the Hoveyda group.13d In 2016, the Buchwald group developed a copper(I)catalyzed asymmetric reductive propargylation of ketones with 1,3-enynes and (dimethoxy)methylsilane (Scheme 1c).14 The chiral allenylcopper(I) species I were generated by the asymmetric reduction of 1,3-enynes with Cu*H. The allenylcopper(I) species II containing a C-BPin unit could be generated in situ through asymmetric borylation of 1,3enynes with Cu*BPin (Scheme 1d).13d Such a C-BPin moiety would serve as a synthetic handle for further structure elaboration.15 Previously, we reported the catalytic asymmetric addition of such species to fluoroalkyl ketones (activated ketones without acidic α-protons).16 Herein, we report a copper(I)-catalyzed asymmetric three-component reaction of simple ketones (unactivated ketones with acidic α-protons), 1,3-enynes and (BPin)2, which generated chiral homopropargyl alcohols in good to excellent yields with good to excellent stereoselectivity. It should be noted that NaBARF, an additive, was found to be indispensable to achieve excellent enantioselectivity. The three-component reaction of 1,3-enyne (1a), (BPin)2 (2), and acetophenone (3a) was investigated as a model reaction for the optimization of reaction conditions (Table 1). In the presence of 10 mol % of Cu(CH3CN)4PF6, 12 mol % of (R)-BINAP, and 1.5 equiv of LiOtBu, product 4a was obtained in 21% yield with >20/1 dr and 73% ee after oxidation of the C-BPin moiety (entry 1). (R)-SEGPHOS, (R)-DIFLUORPHOS, (R)-DTBM-SEGPHOS, and (R,R)-QUINOXP* did not give improved reaction results (entries 2−5). Ferroceneimbedded ligands, including (S)-(R)-PPFA, (R)-(S)-JOSIPHOS, and (R,RP)-TANIAPHOS, were not effective in the model reaction either (entries 6−8). P,N-Bidentate (S)-iPrPHOX and monodentate (Sa,Sc,Sc)-Feringa ligand were not suitable (entries 9 and 10). Fortunately, the reaction with (R,R)-Ph-BPE gave improved yield with moderate diastereoand enantioselectivities (entry 11). Considering the profound effect of the base on the asymmetric reaction with (BPin)2,16,17

entry 1 2 3 4 5 6 7 8 9 10e 11 12 13 14f 15g

total yieldb (%)

drc

eec (%)

>20/1 >20/1 >20/1

−73d −59d −46d

24 42 24 21

>20/1 >20/1 >20/1 >20/1

−39d 19 0 −17d

LiOtBu LiOtBu

18 15

5/1 >20/1

−22d 15

LiOtBu KOtBu KOMe KOMe KOMe

46 48 51 51 89

8/1 >20/1 >20/1 >20/1 >20/1

53 75 82 95 93

ligand

base

(R)-BINAP (R)-SEGPHOS (R)DIFLUORPHOS (R)-DTBMSEGPHOS (R,R)-QUINOXP* (S)-(R)-PPFA (R)-(S)-JOSIPHOS (R,Rp)TANIAPHOS (S)-iPr-PHOX (Sa,Sc,Sc)-Feringa ligand (R,R)-Ph-BPE (R,R)-Ph-BPE (R,R)-Ph-BPE (R,R)-Ph-BPE (R,R)-Ph-BPE

LiOtBu LiOtBu LiOtBu

21 20 22

LiOtBu

trace

LiOtBu LiOtBu LiOtBu LiOtBu

a 1a, 0.10 mmol; 2, 0.15 mmol; 3a, 0.20 mmol. bDetermined by 1H NMR analysis of reaction crude mixture using Cl2CHCHCl2 as an internal standard. cDetermined by chiral-stationary-phase HPLC analysis. dEnt-4a was obtained. e24 mol % ligand. f12 mol % NaBARF. g5 mol % Cu(CH3CN)4PF6, 6 mol % (R,R)-Ph-BPE, 6 mol % NaBARF, 1.0 equiv 1a, 2.5 euiv 2, 2.5 equiv 3a and 2.0 equiv KOMe were employed.

various bases were studied (for details, see the SI). KOtBu led to 4a in improved stereoselectivity (entry 12). KOMe was found to be the best base in terms of both yield and stereoselectivity (entry 13). The addition of 12 mol % NaBARF, which led to the generation of ionic [(R,R)-PhBPE-Cu]+BARF− complex in situ,18 further improved the enantioselectivity of 4a from 82% to 95% (entry 14). The yield was improved by increasing the amounts of 2, 3a, and KOMe with maintained stereoselectivity in the presence of 5 mol % of copper(I) catalyst (entry 15). Under the optimized reaction conditions, the substrate scope of ketones was evaluated (Scheme 2). As for parasubstituted acetophenones, both electron-withdrawing groups and electron-donating groups were well tolerated (4aa−ak). A variety of homopropargyl alcohols were isolated in good to excellent yields with both excellent diastereo- and enantioselectivities. Moreover, the reaction results were not affected by the position of substituents on the phenyl ring (4al−an). It is particularly noteworthy that various functional groups, such as sulfone (4ac), cyano (4ad and 4am), ester (4ae), bromo (4ah and 4al), and alkoxyl (4ai, 4ak and 4an), were not touched. 2B

DOI: 10.1021/acs.orglett.8b03912 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters Scheme 2. Substrate Scope of Ketonesa

Scheme 3. Substrate Scope of 1,3-Enynesa

a

1a, 0.20 mmol; 2, 0.50 mmol; 3a, 0.50 mmol. Isolated yield reported. Diastereoselectivity and enantioselectivity determined by chiralstationary-phase HPLC analysis. bdr = 5/1. cdr = 3/1.

a

1a, 0.20 mmol; 2, 0.50 mmol; 3a, 0.50 mmol. Isolated yield reported. Diastereoselectivity and enantioselectivity determined by chiralstationary-phase HPLC analysis. bdr = 8/1.

coupled with (BPin)2 and acetophenone efficiently to furnish the products 4ha−oa uniformly in good yields with excellent both diastereo- and enantioselectivities. It is particularly noteworthy that an array of functional groups, such as TBSether (4ia and 4ja), benzyl ether (4ka), ester (4la), tosylate (4ma), sulfonamide (4na), and alkyl chloride (4oa), which can undergo further transformations for structure elaboration, was well tolerated under the present catalytic system. Moreover, it was fortunate to discover that propiophenone (3y), 1-phenyl1-pentanone (3z), 4-phenyl-2-butanone (3aa), and 2-butanone (3ab) were applicable, leading to 4ky, 4kz, 4haa, and 4hab in moderate yields with moderate diastereoselectivity and excellent enantioselectivity. The absolute configurations of these products were assigned temporarily according to the Xray structure of 4ac by analogy. The asymmetric borylative propargylation of ketones could be followed by a Suzuki−Miyaura cross-coupling as shown in Scheme 4. In the presence of 5 mol % of Pd(OAc)2, 10 mol % of Ru-Phos, and 3.0 equiv of Cs2CO3, product 5 was obtained in 73% yield with >20/1 dr and 94% ee. Furthermore, the hydroxyl group in product 4aa was successfully employed as a synthetic handle to introduce functional groups. For example, SN2 reaction with NaN3 afforded product 6 in 70% yield for two steps and SN2 reaction with NaSPh furnished product 7 in 75% yield for two steps (Scheme 4). A C−O bond-forming reaction catalyzed by a copper(I)−diamine complex was also achieved successfully. Product 4al was transformed to chiral tetrahydropyran derivative 8 in 66% yield (Scheme 4). Piperidines with two continuous stereogenic carbon centers (13) were reported to exhibit good analgesic activity.19 The two continuous chiral carbon centers could be accessed by the present methodology. As shown in Scheme 5, in the presence of 0.5 mol % of Cu(CH3CN)4PF6 and 0.6 mol % of (R,R)-Ph-

Acetonaphthone was also a competent substrate to give the product (4ao) in excellent results. Heterocycles are attractive structural units due to their vast presence in pharmaceutically active compounds. However, the heterocycles would potentially coordinate to metal center, which may deactivate the metal catalyst and lead to decreased yield and stereoselectivity. Fortunately, a pyrazole motif in acetophenone did not have a bad effect on both the yield and stereoselectivty (4ap). Moreover, both electron-deficient and electron-rich pyridinyl were not harmful to the reaction results (4aq and 4ar). 2-Acetylbenzo[b]thiophene was also an acceptable substrate (4as). Moreover, α,β-unsaturated ketone showed activity comparable to that of acetophenone (4at). α,β,γ,δ-Unsaturated ketone led to product 4au in moderate yield, possibly due to the side reactions caused by the polyunsaturated carbon−carbon double bonds. More importantly, the present reaction conditions were successfully extended to functionalized acetophenones, such as α-fluoro-, α,α-difluoro-, and α-chloroacetophenones (3v−x). Among these products (4av−ax), 4aw was remarkable as further transformations were enabled by the synthetically versatile alkyl chloride motif. The absolute configuration of 4ac was determined through X-ray crystallography analysis of its single crystals. The absolute configurations of other products were tentatively deduced by analogy. Then the scope of 1,3-enynes was investigated (Scheme 3). Several 1,3-enynes bearing aromatic or heteroaromatic substituents reacted with (BPin)2 and acetophenone smoothly to give the corresponding homopropargyl alcohols uniformly in good to excellent yields with excellent stereoselectivity (4aa−ga). Furthermore, a number of aliphatic 1,3-enynes C

DOI: 10.1021/acs.orglett.8b03912 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters

ology and application of it in the asymmetric synthesis of natural products are currently on going in our laboratory.

Scheme 4. Transformations of the Products



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.orglett.8b03912. Experimental procedures, X-ray diffraction data for 4ac, and spectroscopic data for all new compounds, including 1 H, 19F, and 13C NMR spectra and HPLC charts (PDF) Accession Codes

CCDC 1871853 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Scheme 5. Synthetic Application of Present Methodology



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Liang Yin: 0000-0001-9604-5198 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We gratefully acknowledge financial support from the “Thousand Youth Talents Plan”, the National Natural Science Foundation of China (Nos. 21672235 and 21871287), and the Strategic Priority Research Program of the Chinese Academy of Sciences (No. XDB20000000). Mr. Xian-Liang Wang (a joint graduate student at Shanghai University and Shanghai Institute of Organic Chemistry) is acknowledged for checking of the reproducibility of some experiments.

BPE, the gram-scale reaction of 3w proceeded in equal efficiency as the reaction with 5 mol % of copper(I) catalyst. The SN2 substitution of primary alkyl chloride with NaCN produced 9 in 84% yield. The full saturation of alkyne moiety and the reduction of the cyano group generated 10 in 82% yield. The selective conversion of primary alcohol to mesylate and the subsequent cyclization produced 11 in 83% yield for two steps. Transformation of the N-Boc moiety to the N-Me moiety and the following esterification afforded 12 in 74% yield for two steps. Clearly, our methodology provides a general synthetic approach for preparing chiral trans-13 with various substituents at C3-position. In conclusion, a copper(I)-catalyzed asymmetric borylative propargylation of simple ketones was achieved in good to excellent diastereoselectivity and excellent enantioselectivity. The reaction enjoys broad substrate scope for both ketones and 1,3-enynes. A broad range of functional groups, such as trifluoromethyl, sulfone, nitrile, ester, trifluoromethoxyl, alkoxyl, heterocycles, 1,3-diene, alkyl chloride, difluoromethyl, dimethyl amino, TBS ether, benzyloxy, tosylate, and tosylate amide, was well tolerated under the present reaction conditions. The borylated product was successfully employed in the following Suzuki−Miyaura cross-coupling reaction. Furthermore, the hydroxyl group in the final product served as a synthetic handle for the introduction of functional groups. Finally, the present methodology was applied in the asymmetric synthesis of a chiral piperidine derivative with potential analgesic activity. Further expansion of the method-



REFERENCES

(1) (a) Shibasaki, M.; Kanai, M. Asymmetric synthesis of tertiary alcohols and α-tertiary amines via Cu-catalyzed C−C bond formation to ketones and ketimines. Chem. Rev. 2008, 108, 2853−2873. (b) Collados, J. F.; Solà, R.; Harutyunyan, S. R.; Maciá, B. Catalytic synthesis of enantiopure chiral alcohols via addition of Grignard reagents to carbonyl compounds. ACS Catal. 2016, 6, 1952−1970. (2) For some selected recent examples, see: (a) Madduri, A. V. R.; Harutyunyan, S. R.; Minnaard, A. J. Asymmetric copper-catalyzed addition of Grignard reagents to aryl alkyl ketones. Angew. Chem., Int. Ed. 2012, 51, 3164−3167. (b) Liao, Y.-X.; Xing, C.-H.; Hu, Q.-S. Rhodium(I)Diene-catalyzed addition reactions of arylborons with ketones. Org. Lett. 2012, 14, 1544−1547. (c) Madduri, A. V. R.; Minnaard, A. J.; Harutyunyan, S. R. Copper(I) catalyzed asymmetric 1,2-addition of Grignard reagents to α-methyl substituted α,βunsaturated ketones. Chem. Commun. 2012, 48, 1478−1480. (3) For some selected examples, see: (a) Denmark, S. E.; Fan, Y. Catalytic, enantioselective aldol additions to ketones. J. Am. Chem. Soc. 2002, 124, 4233−4235. (b) Denmark, S. E.; Fan, Y.; Eastgate, M. D. Lewis base catalyzed, enantioselective aldol addition of methyl trichlorosilyl ketene acetal to ketones. J. Org. Chem. 2005, 70, 5235− 5248. (c) Moreau, X.; Bazán-Tejeda; Campagne, J.-M. Catalytic and asymmetric vinylogous Mukaiyama reactions on aliphatic ketones: formal asymmetric synthesis of Taurospongin A. J. Am. Chem. Soc. 2005, 127, 7288−7289. (d) Oisaki, K.; Zhao, D.; Kanai, M.; D

DOI: 10.1021/acs.orglett.8b03912 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters Shibasaki, M. Catalytic enantioselective aldol reaction to ketones. J. Am. Chem. Soc. 2006, 128, 7164−7165. (e) Bae, H. Y.; Höfler, D.; Kaib, P. S. J.; Kasaplar, P.; De, C. K.; Döhring, A.; Lee, S.; Kaupmees, K.; Leito, I.; List, B. Approaching sub-ppm-level asymmetric organocatalysis of a highly challenging and scalable carbon−carbon bond forming reaction. Nat. Chem. 2018, 10, 888−894. (f) Lee, S.; Bae, H. Y.; List, B. Can a ketone be more reactive than an aldehyde? Catalytic asymmetric synthesis of substituted tetrahydrofurans. Angew. Chem., Int. Ed. 2018, 57, 12162−12166. (4) For some selected examples, see: (a) Wada, R.; Oisaki, K.; Kanai, M.; Shibasaki, M. Catalytic enantioselective allylboration of ketones. J. Am. Chem. Soc. 2004, 126, 8910−8911. (b) Wadamoto, M.; Yamamoto, H. Silver-catalyzed asymmetric Sakurai-Hosomi allylation of ketones. J. Am. Chem. Soc. 2005, 127, 14556−14557. (c) Lou, S.; Moquist, P. N.; Schaus, S. E. Asymmetric allylation of ketones catalyzed by chiral diols. J. Am. Chem. Soc. 2006, 128, 12660−12661. (d) Miller, J. J.; Sigman, M. S. Design and synthesis of modular oxazoline ligands for the enantioselective chromium-catalyzed addition of allyl bromide to ketones. J. Am. Chem. Soc. 2007, 129, 2752−2753. (e) Miller, J. J.; Sigman, M. S. Quantitatively correlating the effect of ligand-substituent size in asymmetric catalysis using linear free energy relationships. Angew. Chem., Int. Ed. 2008, 47, 771−774. (f) Barnett, D. S.; Moquist, P. N.; Schaus, S. E. The mechanism and an improved asymmetric allylboration of ketones catalyzed by chiral biphenols. Angew. Chem., Int. Ed. 2009, 48, 8679−8682. (g) Shi, S.-L.; Xu, L.-W.; Oisaki, K.; Kanai, M.; Shibasaki, M. Identification of modular chiral bisphosphines effective for Cu(I)-catalyzed asymmetric allylation and propargylation of ketones. J. Am. Chem. Soc. 2010, 132, 6638−6639. (h) Alam, R.; Vollgraff, T.; Eriksson, L.; Szabó, K. J. Synthesis of adjacent quaternary stereocenters by catalytic asymmetric allylboration. J. Am. Chem. Soc. 2015, 137, 11262−11265. (i) Robbins, D. W.; Lee, K.; Silverio, D. L.; Volkov, A.; Torker, S.; Hoveyda, A. H. Practical and Broadly applicable catalytic enantioselective additions of allyl-B(pin) compounds to ketones and α-ketoesters. Angew. Chem., Int. Ed. 2016, 55, 9610−9614 and references cited therein . (5) For some selected examples, see: (a) Cozzi, P. G. Enantioselective alkynylation of ketones catalyzed by Zn(salen) complexes. Angew. Chem., Int. Ed. 2003, 42, 2895−2898. (b) Lu, G.; Li, X.; Jia, X.; Chan, W. L.; Chan, A. S. C. Enantioselective alkynylation of aromatic ketones catalyzed by chiral camphorsulfonamide ligands. Angew. Chem., Int. Ed. 2003, 42, 5057−5058. (c) Kotani, S.; Kukita, K.; Tanaka, K.; Ichibakase, T.; Nakajima, M. Lithium binaphtholate-catalyzed asymmetric addition of lithium acetylides to carbonyl compounds. J. Org. Chem. 2014, 79, 4817−4825 and references cited therein . (6) For a selected review on catalytic asymmetric propargylation of ketones, see: (a) Ding, C.-H.; Hou, X.-L. Catalytic asymmetric propargylation. Chem. Rev. 2011, 111, 1914−1937. For some selected recent examples on catalytic asymmetric propargylation of ketones, see: (b) Harper, K. C.; Sigman, M. S. Three-Dimensional correlation of steric and electronic free energy relationships guides asymmetric propargylation. Science 2011, 333, 1875−1878. (c) Fandrick, K. R.; Fandrick, D. R.; Reeves, J. T.; Gao, J.; Ma, S.; Li, W.; Lee, H.; Grinberg, N.; Lu, B.; Senanayake, C. H. A general copper-BINAPcatalyzed asymmetric propargylation of ketones with propargyl boronates. J. Am. Chem. Soc. 2011, 133, 10332−10335. (d) Barnett, D. S.; Schaus, S. E. Asymmetric propargylation of ketones using allenylboronates catalyzed by chiral biphenols. Org. Lett. 2011, 13, 4020−4023. (e) Kohn, B. L.; Ichiishi, N.; Jarvo, E. R. Silver-catalyzed allenylation and enantioselective propargylation reactions of ketones. Angew. Chem., Int. Ed. 2013, 52, 4414−4417. (f) Yamashita, Y.; Cui, Y.; Xie, P.; Kobayashi, S. Zinc amide catalyzed regioselective allenylation and propargylation of ketones with allenyl boronate. Org. Lett. 2015, 17, 6042−6045. (7) For some selected recent examples, see: (a) North, M.; OmedesPujol, M.; Williamson, C. Investigation of Lewis acid versus Lewis bse catalysis in asymmetric cyanohydrin synthesis. Chem. - Eur. J. 2010, 16, 11367−11375. (b) Kubota, K.; Osaki, S.; Jin, M.; Ito, H.

Copper(I)-catalyzed enantioselective nucleophilic borylation of aliphatic ketones: synthesis of enantioenriched chiral tertiary αhydroxyboronates. Angew. Chem., Int. Ed. 2017, 56, 6646−6650. (8) For a review on proton-transfer reaction in asymmetric catalysis, see: Kumagai, N.; Shibasaki, M. Recent advances in direct catalytic asymmetric transformations under proton-transfer conditions. Angew. Chem., Int. Ed. 2011, 50, 4760−4772. (9) For some selected examples, see: (a) Yazaki, R.; Kumagai, N.; Shibasaki, M. Direct catalytic asymmetric addition of allyl cyanide to ketones. J. Am. Chem. Soc. 2009, 131, 3195−3197. (b) Benfatti, F.; Yilmaz, S.; Cozzi, P. G. The first catalytic enantioselectivie aldol-type reaction of ethyl diazoacetate to ketones. Adv. Synth. Catal. 2009, 351, 1763−1767. (c) Yoshino, T.; Morimoto, H.; Lu, G.; Matsunaga, S.; Shibasaki, M. Construction of contiguous tetrasubstituted chiral carbon stereocenters via direct catalytic asymmetric aldol reaction of α-isothiocyanato esters with ketones. J. Am. Chem. Soc. 2009, 131, 17082−17083. (d) Yazaki, R.; Kumagai, N.; Shibasaki, M. Direct catalytic asymmetric addition of allyl cyanide to ketones via soft Lewis acid/hard Brønsted base/hard Lewis base catalysis. J. Am. Chem. Soc. 2010, 132, 5522−5531. (e) de la Campa, R.; Ortín, I.; Dixon, D. J. Direct catalytic enantio- and diastereoselective ketone aldol reactions of isocyanoacetates. Angew. Chem., Int. Ed. 2015, 54, 4895−4898. (f) Wei, X.-F.; Xie, X.-W.; Shimizu, Y.; Kanai, M. Copper(I)-catalyzed enantioselective addition of enynes to ketones. J. Am. Chem. Soc. 2017, 139, 4647−4650. (10) For some selected examples, see: (a) Bocknack, B. M.; Wang, L.-C.; Krische, M. J. Desymmetrization of enone-diones via rhodiumcatalyzed diastereo- and enantioselectivie tandem conjugate additionaldol cyclization. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 5421−5424. (b) Lam, H. W.; Joensuu, P. M. Cu(I)-catalyzed reductive aldol cyclizations: diastereo- and enantioselective synthesis of β-hydroxylactones. Org. Lett. 2005, 7, 4225−4228. (c) Deschamp, J.; Chuzel, O.; Hannedouche, J.; Riant, O. Highly diastereo- and enantioselective copper-catalyzed domino reduction/aldol reaction of ketones with methyl acrylate. Angew. Chem., Int. Ed. 2006, 45, 1292−1297. (d) Zhao, D.; Oisaki, K.; Kanai, M.; Shibasaki, M. Dramatic ligand effect in catalytic reductive aldol reaction of allenic esters to ketones. J. Am. Chem. Soc. 2006, 128, 14440−14441. (e) Oisaki, K.; Zhao, D.; Kanai, M.; Shibasaki, M. Catalytic enantioselective alkylative aldol reaction: efficient multicomponent assembly of dialkylzincs, allenic esters, and ketones toward highly functionalized δ-lactones with tetrasubstituted chiral centers. J. Am. Chem. Soc. 2007, 129, 7439− 7443. (f) Saxena, A.; Choi, B.; Lam, H. W. Enantioselective coppercatalyzed reductive coupling of alkenylazaarenes with ketones. J. Am. Chem. Soc. 2012, 134, 8428−8431. (g) Meng, F.; Jang, H.; Jung, B.; Hoveyda, A. H. Cu-catalyzed chemoselective preparation of 2(Pinacolato)boron-substituted allylcopper complexes and their in situ site-, diastereo-, and enantioselective additions to aldehydes and ketones. Angew. Chem., Int. Ed. 2013, 52, 5046−5051. (h) Kawai, J.; Chikkade, P. K.; Shimizu, Y.; Kanai, M. In situ catalytic generation of allylcopper species for asymmetric allylation: toward 1H-isochromene skeletons. Angew. Chem., Int. Ed. 2013, 52, 7177−7180. (i) Chikkade, P. K.; Shimizu, Y.; Kanai, M. Catalytic enantioselective synthesis of 2(2-hydroxyethyl)indole scaffolds via consecutive intramolecular amido-cupration of allenes and asymmetric addition of carbonyl compounds. Chem. Sci. 2014, 5, 1585−1590. (j) Tsai, E. Y.; Liu, R. Y.; Yang, Y.; Buchwald, S. L. A regio- and enantioselective CuH-catalyzed ketone allylation with terminal allenes. J. Am. Chem. Soc. 2018, 140, 2007−2011. (k) Cheng, F.; Lu, W.; Huang, W.; Wen, L.; Li, M.; Meng, F. Cu-catalyzed enantioselective synthesis of tertiary benzylic copper complexes and their in situ addition to carbonyl compounds. Chem. Sci. 2018, 9, 4992−4998 For a selected very recent review referring to this topic, see: . (l) Holmes, M.; Schwartz, L. A.; Krische, M. J. Intermolecular Metal-catalyzed reductive coupling of dienes, allenes, and enynes with carbonyl compounds and imines. Chem. Rev. 2018, 118, 6026−6052. (11) For selected examples on the application of chiral homopropargyl alcohols in organic synthesis, see: (a) Trost, B. M.; Dong, G. Total synthesis of bryostatin 16 using atom-economical and E

DOI: 10.1021/acs.orglett.8b03912 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters chemoselective approaches. Nature 2008, 456, 485−488. (b) Dalby, S. M.; Goodwin-Tindall, J.; Paterson, I. Total synthesis of (−)-Rhizopodin. Angew. Chem., Int. Ed. 2013, 52, 6517−6521. (c) Li, H.; Sheeran, J. W.; Clausen, A. W.; Fang, Y.-Q.; Bio, M. M.; Bader, S. Flow asymmetric propargylation: development of continuous processes for the preparation of a chiral β-amino alcohol. Angew. Chem., Int. Ed. 2017, 56, 9425−9429. (12) For some applications of 1,3-enynes in organic synthesis other than catalytic asymmetric propargylation of carbonyl groups, see: (a) Jang, H.-Y.; Huddleston, R. R.; Krische, M. J. Hydrogen-mediated C−C bond formation: catalytic regio- and stereoselective reductive condensation of α-keto aldehydes and 1,3-enynes. J. Am. Chem. Soc. 2004, 126, 4664−4668. (b) Kong, J.-R.; Cho, C.-W.; Krische, M. J. Hydrogen-mediated reductive coupling of conjugated alkynes with Ethyl (N-sulfinyl)iminoacetates: Synthesis of Unnatural α-amino acids via rhodium-catalyzed C−C bond forming hydrogenation. J. Am. Chem. Soc. 2005, 127, 11269−11276. (c) Kong, J.-R.; Ngai, M.-Y.; Krische, M. J. Highly enantioselective direct reductive coupling of conjugated alkynes and α-ketoesters via rhodium-catalyzed asymmetric hydrogenation. J. Am. Chem. Soc. 2006, 128, 718−719. (d) Komanduri, V.; Krische, M. J. Enantioselective reductive coupling of 1,3-enynes to heterocyclic aromatic aldehydes and ketones via rhodium-catalyzed asymmetric hydrogenation: mechanistic insight into the role of Brønsted acid additives. J. Am. Chem. Soc. 2006, 128, 16448−16449. (e) Waser, J.; González-Gómez, J. C.; Nambu, H.; Huber, P.; Carreira, E. M. Cobalt-catalyzed hydrohydrazination of dienes and enynes: access to allylic and propargylic hydrazides. Org. Lett. 2005, 7, 4249−4252. (f) Cho, C.-W.; Krische, M. J. Enantioselective reductive coupling of alkynes and α-keto aldehydes via rhodium-catalyzed hydrogenation: an approach to bryostatin substructures. Org. Lett. 2006, 8, 891−894. (g) Hong, Y.-T.; Cho, C.W.; Skucas, E.; Krische, M. J. Enantioselective reductive coupling of 1,3-enynes to glyoxalates mediated by hydrogen: asymmetric synthesis of β,γ-unsaturated α-hydroxy esters. Org. Lett. 2007, 9, 3745−3748. (h) Sasaki, Y.; Horita, Y.; Zhong, C.; Sawamura, M.; Ito, H. Copper(I)-catalyzed regioselective monoborylation of 1,3-enynes with an internal triple bond: selective synthesis of 1,3-dienylboronates and 3-alkynylboronates. Angew. Chem., Int. Ed. 2011, 50, 2778−2782. (i) Cheng, J.-K.; Loh, T.-P. Copper- and Cobalt-catalyzed direct coupling of sp3 α-carbon of alcohols with alkenes and hydroperoxides. J. Am. Chem. Soc. 2015, 137, 42−45. (j) Huang, Y.; del Pozo, J.; Torker, S.; Hoveyda, A. H. Enantioselective synthesis of trisubstituted allenyl−B(pin) compounds by phosphine−Cu-catalyzed 1,3-enyne hydroboration Insights regarding stereochemical integrity of Cu− allenyl intermediates. J. Am. Chem. Soc. 2018, 140, 2643−2655. (13) (a) Patman, R. L.; Williams, V. M.; Bower, J. F.; Krische, M. J. Carbonyl propargylation from the alcohol or aldehyde oxidation level employing 1,3-enynes as surrogates to preformed allenylametal regents: a ruthenium-catalyzed C−C bond-forming transfer hydrogenation. Angew. Chem., Int. Ed. 2008, 47, 5220−5223. (b) Geary, L. M.; Leung, J. C.; Krische, M. J. Ruthenium-catalyzed reductive coupling of 1,3-enynes and aldehydes by transfer hydrogenation: antidiastereoselective carbonyl propargylation. Chem. - Eur. J. 2012, 18, 16823−16827. (c) Geary, L. M.; Woo, S. K.; Leung, J. C.; Krische, M. J. Diastereo- and enantioselective iridium-catalyzed carbonyl propargylation from the alcohol or aldehyde oxidation level: 1,3-enynes as allenylmetal equivalents. Angew. Chem., Int. Ed. 2012, 51, 2972−2976. (d) Meng, F.; Haeffner, F.; Hoveyda, A. H. Diastereo- and enantioselective reactions of Bis(pinacolato)diboron, 1,3-enynes, and aldehydes catalyzed by an easily accessible bisphosphine-Cu complex. J. Am. Chem. Soc. 2014, 136, 11304−11307. (e) Nguyen, K. D.; Herkommer, D.; Krische, M. J. Ruthenium-BINAP catalyzed alcohol C−H tert-Prenylation via 1,3-enyne transfer hydrogenation: beyond stoichiometric carbonions in enantioselective carbonyl propargylation. J. Am. Chem. Soc. 2016, 138, 5238−5241. (14) Yang, Y.; Perry, I. B.; Lu, G.; Liu, P.; Buchwald, S. L. Coppercatalyzed asymmetric addition of olefin-derived nucleophiles to ketones. Science 2016, 353, 144−150.

(15) For some selected reviews, see: (a) Doucet, H. Suzuki-Miyaura cross-coupling reactions of alkylboronic acid derivatives or alkyltrifluoroborates with aryl, alkenyl or alkyl halides and triflates. Eur. J. Org. Chem. 2008, 2008, 2013−2020. (b) Collins, B. S. L.; Wilson, C. M.; Myers, E. L.; Aggarwal, V. K. Asymmetric synthesis of secondary and tertiary boronic esters. Angew. Chem., Int. Ed. 2017, 56, 11700− 11733. (16) Gan, X.-C.; Zhang, Q.; Jia, X.-S.; Yin, L. Asymmetric construction of fluoroalkyl tertiary alcohols through a threecomponent reaction of (Bpin)2, 1,3-enynes, and fluoroalkyl ketones catalyzed by a copper(I) complex. Org. Lett. 2018, 20, 1070−1073. (17) (a) Jiang, L.; Cao, P.; Wang, M.; Chen, B.; Wang, B.; Liao, J. Highly diastereo- and enantioselective Cu-catalyzed borylative coupling of 1,3-dienes and aldimines. Angew. Chem., Int. Ed. 2016, 55, 13854−13858. (b) Chen, L.; Zou, X.; Zhao, H.; Xu, S. Coppercatalyzed asymmetric protoboration of β-amidoacrylonitriles and βamidoacrylate esters: an efficient approach to functionalized chiral αamino boronate esters. Org. Lett. 2017, 19, 3676−3679. (18) For some selected examples with beneficial effect of NaBARF in asymmetric transition metal catalysis, see: (a) Shi, W.-J.; Zhang, Q.; Xie, J.-H.; Zhu, S.-F.; Hou, G.-H.; Zhou, Q.-L. Highly enantioselective hydrovinylation of α-alkyl vinylarenes. An approach to the construction of all-carbon quaternary stereocenters. J. Am. Chem. Soc. 2006, 128, 2780−2781. (b) Evans, L. A.; Fey, N.; Harvey, J. N.; Hose, D.; Lloyd-Jones, G. C.; Murray, P.; Orpen, A. G.; Osborne, R.; Owen-Smith, G. J. J.; Purdie, M. Counterintuitive kinetics in TsujiTrost allylation: ion-pair partitioning and implications for asymmetric catalysis. J. Am. Chem. Soc. 2008, 130, 14471−14473. (c) Cao, P.; Deng, C.; Zhou, Y.-Y.; Sun, X.-L.; Zheng, J.-C.; Xie, Z.; Tang, Y. Asymmetric Nazarov reaction catalyzed by chiral tris(oxazoline)/ copper(II). Angew. Chem., Int. Ed. 2010, 49, 4463−4466. (19) McElvain, S. M.; Barnett, M. D. Piperidine derivatives. XXVIII. 1-Methyl-3-alkyl-4-phenyl-4-acyloxypiperidines. J. Am. Chem. Soc. 1956, 78, 3140−3143.

F

DOI: 10.1021/acs.orglett.8b03912 Org. Lett. XXXX, XXX, XXX−XXX