N-Sulfonyl Bisimidazoline Ligands and Their Applications in Pd(II

6 hours ago - A new class of chiral N-sulfonyl bisimidazoline (Bim) ligands have been designed, prepared, and applied in Pd(II)-catalyzed asymmetric ...
0 downloads 0 Views 1MB Size
Letter Cite This: Org. Lett. XXXX, XXX, XXX−XXX

pubs.acs.org/OrgLett

N‑Sulfonyl Bisimidazoline Ligands and Their Applications in Pd(II)Catalyzed Asymmetric Addition toward α‑Tertiary Amines Jia-Yin Wang,† Meng-Wei Li,† Meng-Fan Li,† Wen-Juan Hao,† Guigen Li,*,‡,§ Shu-Jiang Tu,*,† and Bo Jiang*,† Org. Lett. Downloaded from pubs.acs.org by UNIV OF CALIFORNIA SANTA BARBARA on 09/14/18. For personal use only.



School of Chemistry & Materials Science, Jiangsu Key Laboratory of Green Synthetic Chemistry for Functional Materials, Jiangsu Normal University, Xuzhou 221116, P. R. China ‡ Institute of Chemistry & BioMedical Sciences, Collaborative Innovation Center of Chemistry for Life Sciences, Nanjing University, Nanjing 210093, P. R. China § Department of Chemistry and Biochemistry, Texas Tech University, Lubbock, Texas 79409-1061, United States S Supporting Information *

ABSTRACT: A new class of chiral N-sulfonyl bisimidazoline (Bim) ligands have been designed, prepared, and applied in Pd(II)-catalyzed asymmetric addition of arylboronic acids to isatin-derived N-Boc ketimines. The combination of Pd(OCOCF3)2 and N-tosyl Bim ligand shows high catalytic activity and excellent asymmetric induction, enabling asymmetric addition to offer α-tertiary amines with generally good to high yields and excellent enantioselectivity (up to 96% yield, 96% ee). This asymmetric Pd(II) catalysis can tolerate air conditions, providing a practical and operationally simple protocol toward the construction of an enantioenriched α-tertiary stereocenter.

T

have remained unexploited for asymmetric catalysis so far. Inspired by the recent success of chiral bis(imidazolidine)pyridine (PyBidine) ligands in asymmetric catalysis,5 we believed that the Bim ligands with easily tunable functional modifications would generate a new catalyst class for unique activity and asymmetric controls (Figure 1). Meanwhile, it is well-known that chiral quaternary carbon stereocenters exist in a wide assortment of natural products, bioactive substances, pharmaceuticals, and agrochemical ingredients.6 Therefore, the enantioselective construction of these stereocenters has attracted extensive interest in the organic community but poses a particular challenge in asymmetric catalysis due to the inherent steric congestion.7 In fact, the control of chiral α-tertiary amines is also very important owing to the versatile chemical and pharmaceutical applications of chiral amines.8,9 In this regard, asymmetric addition to ketimines has been a straightforward and practical strategy. Numerous organocatalytic and metal-catalyzed asymmetric reactions have been reported for the formation of chiral α-tertiary amines.9,10 Specifically, asymmetric nucleophilic additions to isatin-derived ketimines provide an efficient and direct entry to chiral α-tertiary amines by using various C-,11 O-,12 P-,13 and S-nucleophiles (Scheme 1b).14 Pd(II)-catalyzed asymmetric addition of arylboron to ketimines has attracted considerable attention.15 Recently,

he development of more powerful chiral ligands has been one of the most attractive topics in chemical synthesis, especially for those ligands which form extremely efficient catalytic complexes and result in excellent enantioselectivity and chemical yields.1 A variety of structurally diverse chiral ligands containing phosphines and/or amines have been designed and synthesized, and some of them have exhibited outstanding catalytic performances in numerous enantioselective transformations.2 Among these ligands, chiral bisoxazolines (Box, type I, Figure 1) have been widely utilized for

Figure 1. Design of chiral ligands.

forming transition-metal complexes for many asymmetric reactions.3 However, structural modifications of these ligands are limited by the availability of their backbone starting materials.4 As shown in Figure 1, chiral C2-symmetric bisimidazolines (Bim, type II) showed great structural flexibility and availability in regard to N-protection groups via N-sulfonylation, N-alkylation, and N-acylation to tune the catalytic reactivity and asymmetric induction by catalysts. Surprisingly, the above chiral C2-symmetric bisimidazolines © XXXX American Chemical Society

Received: July 26, 2018

A

DOI: 10.1021/acs.orglett.8b02374 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters Table 1. Optimization of the Reaction Conditionsa

Scheme 1. Profiles of Asymmetric Addition of IsatinDerived Ketimines

the group of Zhang reported an elegant and efficient palladium-catalyzed enantioselective addition of arylboronic acids to isatin-derived N-sulfonyl ketimines with the 2,6dicholobenzyl protecting group, but the Boc-protected counterparts were not applicable for this transformation (Scheme 1b).15a Therefore, the catalytic asymmetric addition of aryl nucleophiles to isatin-derived N-Boc ketimines is still challenging (Scheme 1c).16 We elaborate herein the facile synthesis of modular N-sulfonyl Bim ligands and their successful application in the Pd(II)-catalyzed asymmetric addition of arylboronic acids 2 to isatin-derived N-Boc ketimines 1 for enantioselective construction of α-tertiary amines 3 (Scheme 1d). Initially, the benzyl-protected isatin-derived N-Boc ketimine 1a and phenylboronic acid (2a) were selected as representative substrates to optimize the reaction conditions (Table 1). With Pd(OCOCF3)2 and Ph-Pyox (L1) as a catalytic system, the reaction in the presence of LiOH in toluene at room temperature did not give the desired product 3a (entry 1). A number of other bidentate oxazoline ligands often used in asymmetric catalysis, such as tBu-Pyox (L2), tBu-Pyox (L3), and Box (L3−L5) were next assayed with the aim of promoting the desired addition reaction. However, these chiral ligands did not show any improvement on both the reactivity and enantioselectivity (entries 2−5). Next, we considered using imidazolidine ligands as bidentate nitrogen ligands to investigate this asymmetric addition reaction. The easily available imidazolidine ligands, such as Pyim (L6), Quinim (L7), and PyBim (L8), were then evaluated. Unfortunately, none of the ligands L6−L8 were able to allow a successful reaction (entries 6−8). Afterward, our first synthesized Nsulfonyl-protected Bim ligands L9−L13 were applied in this reaction system. Delightfully, the use of ligands L9−L13 all could drive the conversion of 1a into 3a (entries 9−13), in which the combination of Pd(II) catalyst and N-tosyl Bim ligand L10 exhibited the best catalytic performance in this transformation regarding the reactivity and enantioselectivity, affording the adduct product (R)-3a in 86% yield and 90% ee value (entry 10). The absolute configuration of 3a was confirmed by X-ray diffraction analysis, and its derivatives were

entry

L (mol %)

solvent

yieldb (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22d 23e

L1 (5) L2 (5) L3 (5) L4 (5) L5 (5) L6 (5) L7 (5) L8 (5) L9 (5) L10 (5) L11 (5) L12 (5) L13 (5) L14 (5) L10 (5) L10 (5) L10 (5) L10 (5) L10 (5) L10 (5) L10 (5) L10 (5) L10 (10)

toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene MTBE EA THF DCE benzene PhCl PhCF3 toluene toluene

ND ND ND ND ND ND NR NR 65 86 71 42 58 NR 70 85 79 63 80 55 58 88 70

eec (%)

90 90 90 90 87 85 81 84 68 90 86 79 91 90

a

Reaction conditions: 1a (0.1 mmol), 2a (0.3 mmol), Pd(O2CCF3)2 (5 mol %), ligand (5 mol %), LiOH (0.5 equiv), solvent (1.0 mL), under air conditions, 8 h, at room temperature. bIsolated yield based on substrate 1a. cThe ee value was determined by HPLC. d0 °C. eUse of Pd(O2CCF3)2 (10 mol %).

assigned by analogy (see the Supporting Information). C2linked Bim ligand L14 proved to be ineffective in the asymmetric catalysis (entry 14). The following screening of the solvents, such as methyl tert-butyl ether (MTBE), CH3CO2Et (EA), tetrahydrofuran (THF), 1,2-dichloroethane (DCE), benzene, chlorobenzene (PhCl), and trifluoromethylbenzene (PhCF3), revealed that all these reaction media gave relatively inferior outcomes with respect to reaction yields and/or ee value as compared with toluene (entries 15−21 vs entry 10). Moreover, decreasing the reaction temperature to 0 °C is beneficial to both the yield and the enantioselectivity, and the product 3a was obtained in 88% yield and 91% ee (entry B

DOI: 10.1021/acs.orglett.8b02374 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters Scheme 3. Substrate Scope of Arylboronic Acidsa−c

22). However, the increase in the loading of the catalyst and the ligand did not facilitate the reaction process (entry 23). With the effective catalytic system identified by the screening of bidentate nitrogen ligands in hand, we next evaluated the scope of asymmetric catalysis by varying isatinderived N-Boc ketimines and arylboronic acid components. The representative results are listed in Scheme 2. N-Boc Scheme 2. Substrate Scope of Isatin-Derived N-Boc Ketiminesa−c

a Reaction conditions: 1 (0.1 mmol), arylboronic acids (0.3 mmol), Pd(TFA)2 (5 mol %), ligand (5 mol %), LiOH (0.5 equiv), toluene (1.0 mL), under air conditions, at 0 °C. bIsolated yield based on substrate 1. cThe ee value was determined by chiral HPLC.

methyl functionality reacted efficiently with arylboronic acids 2 with both electronically poor (chloro 2b) and rich (methyl 2c) groups. However, the ee value is excellent for the former product 3m, while an appreciably reduced ee value was observed for the latter 3n. These results indicate that the electronic nature of arylboronic acids seems to impose a remarkable effect on the enantioselectivity. This similar observation also appears in the conversion of ketimine 1d possessing C6 chloro functionality into α-tertiary amine products 3o−u. Arylboronic acids bearing a functional group, such as halogen (e.g., fluoro, chloro, and bromo), ether (PMP = p-methoxyphenyl), and alkyl, proved to be good candidates for the reaction, in which the presence of electron-withdrawing groups is generally required to achieve excellent enantioselectivity (3o−3q vs 3t−3u). As F-containing compounds exhibit important pharmacological properties, we decided to investigate the generality of asymmetric Pd-catalysis by combining with fluoro-containing N-Boc ketimines 1m−o and 4-chlorophenylboronic acid 2b. As expected, the reaction could proceed smoothly, and high levels of stereochemical control were established regardless of their structural features (3v−x). Alternatively, the opposite enantiomer of (S)-3p was formed with 83% yield and 96% ee when the chiral (S,S,S,S)bisimidazole behaved as the bidentate nitrogen ligand. The prepared N-sulfonyl bisimidazoline ligand shows excellent chiral induction, perhaps because this ligand is an anion, rather than a strong base, and most of the negative charge is on the nitrogen. Thus, the ligand is a relatively strong electrondonating ligand, but not strong enough to totally neutralize one positive charge of Pd(II), which will increase the nucleophilic ability of the Pd−Ar species and keep the ability of Pd(II) for activation of ketimine. It is a good balance between electron-withdrawing and electron-donating ligands, leading to the high reactivity of this catalytic system.17

a

Reaction conditions: 1 (0.1 mmol), PhB(OH)2 (0.3 mmol), Pd(TFA)2 (5 mol %), ligand (5 mol %), LiOH (0.5 equiv), toluene (1.0 mL), under air conditions, at 0 °C. bIsolated yield based on substrate 1. cThe ee value was determined by chiral HPLC.

ketimines with diverse functional groups were first investigated in combination with benzoboric acid under the standard conditions. Substituents with different electronic nature residing in different positions of arene ring would be accommodated, confirming the efficiency of asymmetric catalysis, as products 3b−g were generated in 70%−89% yields and between 82% and 94% ee except for 3g. Generally, introduction of a functional groups into C6 or C7 position of isatin-derived ketimines 1 allowed for the addition of 2a to proceed with the better enantioselectivity control in the presence of Pd(II)-Bim system as compared with C4 or C5 substituents (Scheme 2, 3c and 3d vs 3f and 3g). The 4-chloro substituent at the C4 position of indolin-2-one ring 1g seems reluctant to go through this process in which both low yield and poor chiral induction were observed (3g, 43%, 60% ee), maybe due to steric effects. Next, we investigated the effect of the substituent group at the 1-position of the ketimines 1. Isatin-derived ketimines with different functional groups at this nitrogen atom, such as 4-methylbenzyl (1h), 4-bromobenzyl (1i), phenyl (1j), N-Boc (1k), and methyl (1k), were efficiently transformed into the corresponding products with high enantioselectivity (86−94% ee). To broaden the application of this Pd(II)-catalyzed asymmetric addition, N-Boc ketimines with C6 or C7 substituents were adopted to react with arylboronic acids 2 with different substitution patterns, due to substituents at both these positions favor to show higher reactivity and better enantioselectivity (Scheme 3). N-Boc ketimine 1b bearing C7 C

DOI: 10.1021/acs.orglett.8b02374 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters The significant potential of our approach based on the amplification reaction for the synthesis of (R)-3p was amply demonstrated with the excellent enantioselectivity being retained (Scheme 4a). Removal of the Boc protecting group

Bo Jiang: 0000-0003-3878-515X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are grateful for financial support from the NSFC (No 21472071), PAPD of Jiangsu Higher Education Institutions, the Outstanding Youth Fund of JSNU (YQ2015003), and the Robert A. Welch Foundation (D-1361, USA).

Scheme 4. Amplification Reaction and Deprotection of 3p



of 3p provided the free chiral amine product 4 in high yield without loss of stereochemical purity, which is a useful chiral amine precursor in organic chemistry (Scheme 4b). In conclusion, we have established a class of new N-sulfonyl bisimidazoline ligands for asymmetric catalysis by forming Pd(II) catalysts and demonstrated both modularity and versatility. In comparison with known chiral ligands, the ability of N-sulfonyl Bim ligand to easily vary the steric and electronic properties of the N-protecting groups over a wide range is extensively practical. By applying this Pd(II)-N-tosyl Bim catalytic system, asymmetric addition of arylboronic acids to isatin-derived N-Boc ketimines was first realized, enabling a direct and practical protocol toward the highly enantioselective construction of α-tertiary amines. Furthermore, the reaction worked readily under the air conditions, exhibiting a remarkable tolerance to oxygen and making the asymmetric transformation simple and mild. Further expanding the scope and application of ligands for distinct catalysis are currently in process in our laboratory.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.orglett.8b02374. Experimental procedures, characterization data for all products, detailed optimized geometries, and free energies (PDF) Accession Codes

CCDC 1856780−1856781 contain 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.



REFERENCES

(1) (a) Pretot, R.; Pfaltz, A. New ligands for regio- and enantiocontrol in Pd-catalyzed allylic alkylations. Angew. Chem., Int. Ed. 1998, 37, 323−325. (b) Zhang, Z.-M.; Chen, P.; Li, W.; Niu, Y.; Zhao, X.-L.; Zhang, A. New Type of Chiral Sulfinamide Monophosphine Ligands: Stereodivergent Synthesis and Application in Enantioselective Gold(I)-Catalyzed Cycloaddition Reactions. Angew. Chem., Int. Ed. 2014, 53, 4350−4354. (c) Genet, J.-P. Asymmetric Catalytic Hydrogenation. Design of New Ru Catalysts and Chiral Ligands: From Laboratory to Industrial Applications. Acc. Chem. Res. 2003, 36, 908−9918. (d) Pfaltz, A.; Drury, W. J. Design of chiral ligands for asymmetric catalysis: from C2-symmetric P,P- and N,Nligands to sterically and electronically nonsymmetrical P,N-ligands. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 5723−5726. (e) Fischer, C.; Defieber, C.; Suzuki, T.; Carreira, E. M. Readily Available [2.2.2]Bicyclooctadienes as New Chiral Ligands for Ir(I): Catalytic, Kinetic Resolution of Allyl Carbonates. J. Am. Chem. Soc. 2004, 126, 1628− 1629. (f) Wang, Z.-Q.; Feng, C.-G.; Xu, M.-H.; Lin, G.-Q. Design of C2-Symmetric Tetrahydropentalenes as New Chiral Diene Ligands for Highly Enantioselective Rh-Catalyzed Arylation of N-Tosylarylimines with Arylboronic acids. J. Am. Chem. Soc. 2007, 129, 5336− 5337. (g) Sattely, E. S.; Meek, S. J.; Malcolmson, S. J.; Schrock, R. R.; Hoveyda, A. H. Design and Stereoselective Preparation of a New Class of Chiral Olefin Metathesis Catalysts and Application to Enantioselective Synthesis of Quebrachamine: Catalyst Development Inspired by Natural Product Synthesis. J. Am. Chem. Soc. 2009, 131, 943−953. (2) (a) Pietrusiewicz, K. M.; Zablocka, M. Preparation of Scalemic P-Chiral Phosphines and Their Derivatives. Chem. Rev. 1994, 94, 1375−1411. (b) Gladiali, S.; Alberico, E. Asymmetric Transfer Hydrogenation: Chiral Ligands and Applications. Chem. Soc. Rev. 2006, 35, 226−236. (c) Yin, J.; Rainka, M. P.; Zhang, X.-X.; Buchwald, S. L. A Highly Active Suzuki Catalyst for the Synthesis of Sterically Hindered Biaryls: Novel Ligand Coordination. J. Am. Chem. Soc. 2002, 124, 1162−1163. (d) Gómez Arrayás, R.; Adrio, J.; Carretero, J. C. Recent applications of chiral ferrocene ligands in asymmetric catalysis. Angew. Chem., Int. Ed. 2006, 45, 7674−7715. (e) Poyatos, M.; Mata, J. A.; Peris, E. Complexes with Poly(Nheterocyclic carbene) Ligands: Structural Features and Catalytic Applications. Chem. Rev. 2009, 109, 3677−3707. (f) Nugent, T. C.; El-Shazly, M. Chiral amine synthesis. Recent developments and trends for enamide reduction, reductive amination, and imine reduction. Adv. Synth. Catal. 2010, 352, 753−819. (g) Yang, G.; Zhang, W. Renaissance of pyridine-oxazolines as chiral ligands for asymmetric catalysis. Chem. Soc. Rev. 2018, 47, 1783−1810. (3) (a) Pfaltz, A. Chiral semicorrins and related nitrogen heterocycles as ligands in asymmetric catalysis. Acc. Chem. Res. 1993, 26, 339−345. (b) Nishiyama, H.; Itoh, Y.; Matsumoto, H.; Park, S.-B.; Itoh, K. New Chiral Ruthenium Bis(oxazolinyl)pyridine Catalyst. Efficient Asymmetric Cyclopropanation of Olefins with Diazoacetates. J. Am. Chem. Soc. 1994, 116, 2223−2224. (c) Johnson, J. S.; Evans, D. A. Chiral Bis(oxazoline) Copper(II) Complexes: Versatile Catalysts for Enantioselective Cycloaddition, Aldol, Michael, and Carbonyl Ene Reactions. Acc. Chem. Res. 2000, 33, 325−335. (d) Zhao, C.-X.; Duffey, M. O.; Taylor, S. J.; Morken, J. P. Enantioand diastereoselective reductive aldol reactions with iridium-pybox catalysts. Org. Lett. 2001, 3, 1829−1831. (e) Zhou, J.; Fu, G. C.

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. *E-mail: [email protected]. ORCID

Guigen Li: 0000-0002-9312-412X D

DOI: 10.1021/acs.orglett.8b02374 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters

(8) For reviews on synthesis of chiral α-tertiary amines, see: (a) Denissova, I.; Barriault, L. Stereoselective formation of quaternary carbon centers and related functions. Tetrahedron 2003, 59, 10105− 10146. (b) Riant, O.; Hannedouche, J. Asymmetric catalysis for the construction of quaternary carbon centres: nucleophilic addition on ketones and ketimines. Org. Biomol. Chem. 2007, 5, 873−888. (c) Cozzi, P. G.; Hilgraf, R.; Zimmermann, N. Enantioselective Catalytic Formation of Quaternary Stereogenic Centers. Eur. J. Org. Chem. 2007, 2007, 5969−5994. (d) 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. (9) Nugent, T. C. Chiral Amine Synthesis: Methods, Developments and Applications; Wiley-VCH: Weinheim, 2010. (10) For selected examples, see: (a) Kobayashi, S.; Mori, Y.; Fossey, J. S.; Salter, M. M. Catalytic Enantioselective Formation of C−C Bonds by Addition to Imines and Hydrazones: A Ten-Year Update. Chem. Rev. 2011, 111, 2626−2703. (b) Friestad, G. K.; Mathies, A. K. Recent developments in asymmetric catalytic addition to C = N bonds. Tetrahedron 2007, 63, 2541−2569. (c) Vachal, P.; Jacobsen, E. N. Structure-based analysis and optimization of a highly enantioselective catalyst for the Strecker reaction. J. Am. Chem. Soc. 2002, 124, 10012−10014. (d) Wang, J.; Hu, X.; Jiang, J.; Gou, S.; Huang, X.; Liu, X.; Feng, X. Asymmetric Activation of tropos 2,2’-Biphenol with Cinchonine Generates an Effective Catalyst for the Asymmetric Strecker Reaction of N-Tosyl-Protected Aldimines and Ketoimines. Angew. Chem., Int. Ed. 2007, 46, 8468−8470. (e) Hou, Z.; Wang, J.; Liu, X.; Feng, X. Highly Enantioselective Strecker Reaction of Ketoimines Catalyzed by an Organocatalyst from (S)-BINOL and LProlinamide. Chem. - Eur. J. 2008, 14, 4484−4486. (f) He, M.; Bode, J. W. Enantioselective, NHC-Catalyzed Bicyclo-β-Lactam Formation via Direct Annulations of Enals and Unsaturated N-Sulfonyl Ketimines. J. Am. Chem. Soc. 2008, 130, 418−419. (g) Rommel, M.; Fukuzumi, T.; Bode, J. W. Cyclic Ketimines as Superior Electrophiles for NHC-Catalyzed Homoenolate Additions with Broad Scope and Low Catalyst Loadings. J. Am. Chem. Soc. 2008, 130, 17266−17267. (h) Hashimoto, T.; Nakatsu, H.; Yamamoto, K.; Maruoka, K. Chiral Brønsted acid-catalyzed asymmetric trisubstituted aziridine synthesis using α-diazoacyl oxazolidinones. J. Am. Chem. Soc. 2011, 133, 9730−9733. (i) Kano, T.; Song, S.; Kubota, Y.; Maruoka, K. Highly Diastereo- and Enantioselective Mannich Reactions of Synthetically Flexible Ketimines with Secondary Amine Organocatalysts. Angew. Chem. 2012, 124, 1217−1220. (j) Wada, R.; Shibuguchi, T.; Makino, S.; Oisaki, K.; Kanai, M.; Shibasaki, M. Catalytic Enantioselective Allylation of Ketoimines. J. Am. Chem. Soc. 2006, 128, 7687−7691. (k) Fu, P.; Snapper, M. L.; Hoveyda, A. H. Catalytic Asymmetric Alkylations of Ketoimines. Enantioselective Synthesis of N-Substituted Quaternary Carbon Stereogenic Centers by Zr-Catalyzed Additions of Dialkylzinc Reagents to Aryl-, Alkyl-, and Trifluoroalkyl-Substituted Ketoimines. J. Am. Chem. Soc. 2008, 130, 5530−5541. (l) Tran, D. N.; Cramer, N. Enantioselective Rhodium(I)-Catalyzed [3 + 2] Annulations of Aromatic Ketimines Induced by Directed C-H Activations. Angew. Chem. 2011, 123, 11294−11298. (m) Jung, H. H.; Buesking, A. W.; Ellman, J. A. Highly Functional Group Compatible Rh-Catalyzed Addition of Arylboroxines to Activated N-tert-Butanesulfinyl Ketimines. Org. Lett. 2011, 13, 3912−3915. (n) Luo, Y.; Hepburn, H. B.; Chotsaeng, N.; Lam, H. W. Enantioselective Rhodium-Catalyzed Nucleophilic Allylation of Cyclic Imines with Allylboron Reagents. Angew. Chem. 2012, 124, 8434− 8438. (11) For selected examples, see: (a) Zhang, X.; Zhang, J.; Lin, L.; Zheng, H.; Wu, W.; Liu, X.; Feng, X. Chiral N,N’-Dioxide-Zinc(II) Complex-Catalyzed Asymmetric Aza-Friedel-Crafts Reaction of Isatin-Derived Ketimines with Indoles. Adv. Synth. Catal. 2016, 358, 3021−3026. (b) Montesinos-Magraner, M.; Vila, C.; Cantón, R.; Blay, G.; Fernández, I.; Muñoz, M. C.; Pedro, J. R. Organocatalytic Asymmetric Addition of Naphthols and Electron-Rich Phenols to IsatinDerived Ketimines: Highly Enantioselective Construction of Tetrasubstituted Stereocenters. Angew. Chem., Int. Ed. 2015, 54,

Suzuki Cross-Couplings of Unactivated Secondary Alkyl Bromides and Iodides. J. Am. Chem. Soc. 2004, 126, 1340−1341. (f) Desimoni, G.; Faita, G.; Quadrelli, P. Pyridine-2,6-bis(oxazolines), Helpful Ligands for Asymmetric Catalysts. Chem. Rev. 2003, 103, 3119−3154. (g) Zhou, Y.-Y.; Wang, L.-J.; Li, J.; Sun, X.-L.; Tang, Y. Side-ArmPromoted Highly Enantioselective Ring-Opening Reactions and Kinetic Resolution of Donor-Acceptor Cyclopropanes with Amines. J. Am. Chem. Soc. 2012, 134, 9066−9069. (h) Liao, S.; Sun, X.-L.; Tang, Y. Side Arm Strategy for Catalyst Design: Modifying Bisoxazolines for Remote Control of Enantioselection and Related. Acc. Chem. Res. 2014, 47, 2260−2272. (i) Keith, J. M.; Jacobsen, E. N. Asymmetric hydrocyanation of hydrazones catalyzed by lanthanidePYBOX complexes. Org. Lett. 2004, 6, 153−155. (j) Yang, G.; Zhang, W. Renaissance of pyridine-oxazolines as chiral ligands for asymmetric catalysis. Chem. Soc. Rev. 2018, 47, 1783−1810. (4) (a) Desimoni, G.; Faita, G.; Quadrelli, P. Pyridine-2,6bis(oxazolines), Helpful Ligands for Asymmetric Catalysts. Chem. Rev. 2003, 103, 3119−3154. (b) Nishiyama, H. Chiral bis(oxazolinyl) pyridine (Pybox) ligands for asymmetric transition metal catalysis. Adv. Catal. Proc. 1998, 2, 153−188. (c) Bhor, S.; Anilkumar, G.; Tse, M. K.; Klawonn, M.; Döbler, C.; Bitterlich, B.; Grotevendt, A.; Beller, M. Synthesis of a New Chiral N,N,N-Tridentate Pyridinebisimidazoline Ligand Library and Its Application in Ru-Catalyzed Asymmetric Epoxidation. Org. Lett. 2005, 7, 3393−3396. (5) (a) Arai, T.; Mishiro, A.; Yokoyama, N.; Suzuki, K.; Sato, H. Chiral Bis(imidazolidine)pyridine-Cu(OTf)2: Catalytic Asymmetric Endo-Selective [3 + 2] Cycloaddition of Imino Esters with Nitroalkenes. J. Am. Chem. Soc. 2010, 132, 5338−5339. (b) Arai, T.; Matsumura, E.; Masu, H. Bis(imidazolidine)pyridine-NiCl2 Catalyst for Nitro-Mannich Reaction of Isatin-Derived N-Boc Ketimines: Asymmetric Synthesis of Chiral 3-Substituted 3-Amino2-oxindoles. Org. Lett. 2014, 16, 2768−2771. (c) Nakamura, S.; Ohara, M.; Nakamura, Y.; Shibata, N.; Toru, T. Copper-Catalyzed Enantioselective Three-Component Synthesis of Optically Active Propargylamines from Aldehydes, Amines, and Aliphatic Alkynes. Chem. - Eur. J. 2010, 16, 2360−2362. (d) Ohara, M.; Nakamura, S.; Shibata, N. Direct Enantioselective Three-Component KabachnikFields Reaction Catalyzed by Chiral Bis(imidazoline)-Zinc(II) Catalysts. Adv. Synth. Catal. 2011, 353, 3285−3289. (6) (a) Wang, B.; Tu, Y.-Q. Stereoselective construction of quaternary carbon stereocenters via a semipinacol rearrangement strategy. Acc. Chem. Res. 2011, 44, 1207−1222. (b) Newman, D. J.; Cragg, G. M. Natural Products as Sources of New Drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629−661. (c) Feng, J.; Holmes, M.; Krische, M. J. Acyclic Quaternary Carbon Stereocenters via Enantioselective Transition Metal Catalysis. Chem. Rev. 2017, 117, 12564−12580. (d) Yu, J.-S.; Liao, F.-M.; Gao, W.-M.; Liao, K.; Zuo, R.-L.; Zhou, J. Michael Addition Catalyzed by Chiral Secondary Amine Phosphoramide Using Fluorinated Silyl Enol Ethers: Formation of Quaternary Carbon Stereocenters. Angew. Chem., Int. Ed. 2015, 54, 7381−7385. (7) (a) Peterson, E. A.; Overman, L. E. Contiguous stereogenic quaternary carbons: A daunting challenge in natural products synthesis. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 11943−11948. (b) Trost, B. M.; Jiang, C. Catalytic enantioselective construction of all-carbon quaternary stereocenters. Synthesis 2006, 369−396. (c) Hawner, C.; Alexakis, A. Metal-catalyzed asymmetric conjugate addition reaction: formation of quaternary stereocenters. Chem. Commun. 2010, 46, 7295−7306. (d) Das, J. P.; Marek, I. Enantioselective synthesis of all-carbon quaternary stereogenic centers in acyclic systems. Chem. Commun. 2011, 47, 4593−4623. (e) Mitsunuma, H.; Shibasaki, M.; Kanai, M.; Matsunaga, S. Catalytic Asymmetric Total Synthesis of Chimonanthine, Folicanthine, and Calycanthine through Double Michael Reaction of Bisoxindole. Angew. Chem., Int. Ed. 2012, 51, 5217−5221. (f) Trost, B. M.; Osipov, M. Palladium-Catalyzed Asymmetric Construction of Vicinal All-Carbon Quaternary Stereocenters and its Application to the Synthesis of Cyclotryptamine Alkaloids. Angew. Chem., Int. Ed. 2013, 52, 9176−9181. E

DOI: 10.1021/acs.orglett.8b02374 Org. Lett. XXXX, XXX, XXX−XXX

Letter

Organic Letters 6320−6324. (c) Chen, Q.; Xie, L.; Li, Z.; Tang, Y.; Zhao, P.; Lin, L.; Feng, X.; Liu, X. Copper/guanidine-catalyzed asymmetric alkynylation of isatin-derived ketimines. Chem. Commun. 2018, 54, 678−681. (d) Arai, T.; Matsumura, E.; Masu, H. Bis(imidazolidine)pyridineNiCl2 Catalyst for Nitro-Mannich Reaction of Isatin-Derived N-Boc Ketimines: Asymmetric Synthesis of Chiral 3-Substituted 3-Amino-2oxindoles. Org. Lett. 2014, 16, 2768−2771. (e) Holmquist, M.; Blay, G.; Pedro, J. R. Highly enantioselective aza-Henry reaction with isatin N-Boc ketimines. Chem. Commun. 2014, 50, 9309−9312. (f) Liu, Y.L.; Zhou, J. Organocatalytic asymmetric cyanation of isatin derived NBoc ketoimines. Chem. Commun. 2013, 49, 4421−4423. (12) Arai, T.; Tsuchiya, K.; Matsumura, E. PyBidine-NiCl2Catalyzed Asymmetric Addition of Alcohols and Peroxides to Isatin-Derived Ketimines. Org. Lett. 2015, 17, 2416−2419. (13) (a) George, J.; Sridhar, B.; Reddy, B. V. S. First Example of Quinine-Squaramide Catalyzed Enantioselective Addition of Diphenyl Phosphite to Ketimines Derived from Isatins. Org. Biomol. Chem. 2014, 12, 1595−1602. (b) Kumar, A.; Sharma, V.; Kaur, J.; Kumar, V.; Mahajan, S.; Kumar, N.; Chimni, S. S. Cinchona-Derived Thiourea Catalyzed Hydrophosphonylation of Ketimines-An Enantioselective Synthesis of α-Amino Phosphonates. Tetrahedron 2014, 70, 7044−7049. (14) (a) Nakamura, S.; Takahashi, S.; Nakane, D.; Masuda, H. Organocatalytic Enantioselective Addition of Thiols to Ketimines Derived from Isatins. Org. Lett. 2015, 17, 106−109. (b) Bece ņo, C.; Chauhan, P.; Rembiak, A.; Wang, A.; Enders, D. Brønsted AcidCatalyzed Enantioselective Synthesis of Isatin-Derived N,S-Acetals. Adv. Synth. Catal. 2015, 357, 672−676. (15) (a) He, Q.; Wu, L.; Kou, X.; Butt, N.; Yang, G.; Zhang, W. Pd(II)-Catalyzed Asymmetric Addition of Arylboronic Acids to Isatin-Derived Ketimines. Org. Lett. 2016, 18, 288−291. (b) Quan, M.; Wu, L.; Yang, G.; Zhang, W. Pd(II), Ni(II) and Co(II)-catalyzed enantioselective additions of organoboron reagents to ketimines. Chem. Commun. 2018, DOI: 10.1039/C8CC04932G. (c) Quan, M.; Yang, G.; Xie, F.; Gridnev, I. D.; Zhang, W. Pd(II)-Catalyzed Asymmetric Addition of Arylboronic Acids to Cyclic N-sulfonyl Ketimine Esters and a DFT Study of Its Mechanism. Org. Chem. Front. 2015, 2, 398−402. (d) Yang, G.; Zhang, W. A. PalladiumCatalyzed Enantioselective Addition of Arylboronic Acids to Cyclic Ketimines. Angew. Chem., Int. Ed. 2013, 52, 7540−7544. (16) Marques, C. S.; Burke, A. J. Modular Catalytic Synthesis of 3Amino-3-aryl-2-oxindoles: Rh Catalysis with Isatin-Derived N-BocProtected Ketimines. Eur. J. Org. Chem. 2016, 2016, 806−812. (17) Schrapel, D. C.; Peters, R. Exogenous-Base-Free PalladacycleCatalyzed Highly Enantioselective Arylation of Imines with Arylboroxines. Angew. Chem., Int. Ed. 2015, 54, 10289−10293.

F

DOI: 10.1021/acs.orglett.8b02374 Org. Lett. XXXX, XXX, XXX−XXX