Regio- and Enantioselective Rhodium-Catalyzed Allylic Alkylation of

3 hours ago - Highly regio- and enantioselective rhodium-catalyzed allylic alkylation of 1,3-diketones with racemic secondary allylic alcohols is repo...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Article

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 J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b05126 • Publication Date (Web): 25 May 2018 Downloaded from http://pubs.acs.org on May 25, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

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

Supporting Information Placeholder ABSTRACT: Highly regio- and enantioselective rhodium-catalyzed 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 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 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.

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 Ir3 and Cu4. The utilization of rhodium catalytic systems for asymmetric allylic alkylation reaction remains underexplored.5,6-9 In this regard, Evans and coworkers pioneered in rhodium-catalyzed allylic alkylation reactions, and disclosed asymmetric reactions employing enantioenriched chiral allylic substrates via double-inversion process (eq 1, Scheme 1).10 In addition, recent efforts from the Evans group led to elegant Rhcatalyzed asymmetric allylic alkylation of prochiral nucleophiles.11 In 2003, Hayashi et al reported a very impressive asymmetric allylic alkylation with allylic acetates by a RhPhox catalytic system (eq 2, Scheme 1).12 Kazmaier and coworkers reported Rh-catalyzed asymmetric allylic alkylation reactions of chelated enolates with enantioenriched allyl phosphates, where excellent chiral transfer was achieved.13 To be noted, the groups of Nguyen,14 Breit15, Vrieze16 and Guo17 reported asymmetric allylic amination reactions with racemic allylic esters. Recently, the pioneering work from the groups of Breit,18 Dong19 and Kang20 demonstrated that allenes and alkynes are suitable allylic precursors to afford allylic alkylation products via Rh-catalytic system. However, the reactions with aryl substituted substrates generally led to moderate enantioselectivity (eq 3, Scheme 1).18e Considering the known reports are mainly limited with the employment of activated allylic alcohol precursors and allenes, the direct utilization of readily available allylic alcohols will be highly desirable.21 Herein, we report

our results on Rh-(P, olefin) complex catalyzed asymmetric allylic alkylation with secondary racemic allylic alcohols directly.

Scheme 1. Rh-catalyzed asymmetric allylic alkylation 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, with an

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

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 the active catalyst requires the P-olefin binding mode. Finally, the optimal conditions were obtained as following: (2 mol %) [Rh(C2H4)2Cl]2 and (5 mol %) (S)-L1 were stirred for 15 min in Et2O, followed by the addition of substrates and TFA (25 mol %). Table 1. Optimization of the reaction conditionsa

3a/4ac

-

yield (%)b -

-

ee (%)d -

L1

Fe(OTf)2

85

10/1

39

CH2Cl2

L1

TFA

94

> 19/1

71

Rh1

CH2Cl2

L1

PhCO2H

93

> 19/1

67

5

Rh1

CH2Cl2

L1

TsOH

82

> 19/1

69

6

Rh1

CH2Cl2

L1

trace

-

-

7

Rh1

CH2Cl2

L1

(nBuO)2 PO2H TfOH

78

> 19/1

66

8

Rh1

DCE

L1

TFA

90

> 19/1

71

9

Rh1

L1

TFA

93

> 19/1

75

10

Rh1

toluene CHCl3

L1

TFA

23

> 19/1

-

11

Rh1

Et2O

L1

TFA

94

> 19/1

92

12

Rh1

L1

TFA

92

> 19/1

86

13

Rh1

BuOMe DME

L1

TFA

90

> 19/1

89

entry 1

[Rh]

ligand L1

additive

Rh1

solvent CH2Cl2

2e

Rh1

CH2Cl2

3

Rh1

4

t

Page 2 of 8

14

Rh1

THF

L1

TFA

62

> 19/1

35

15

Rh2

Et2O

L1

TFA

95

> 19/1

95

16

Rh3

Et2O

L1

TFA

-

-

-

17

Rh4

Et2O

L1

TFA

-

-

-

18

Rh2

Et2O

L2

TFA

91

> 19/1

86

19

Rh2

Et2O

L3

TFA

trace

-

-

20

Rh2

Et2O

L4

TFA

82

> 19/1

76

a

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

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 electron-donating group (4-Me, 4-OMe, 3-OMe, 2-OMe) at the different positions of phenyl ring led to their desired 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) on the phenyl ring required prolonged reaction time, providing good to excellent yields (88-94%) and enantioselectivity (91-97% ee) (3f-3k). The well 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 2-naphthyl- and bipenylsubstituted 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 2furanyl allyl alcohol proceeded in slightly decreased yield (60%, 87% ee, 3n), likely due to the unstability of substrate under acidic conditions. 2-Benzothiophenyl allylic alcohol was also a suitable substrate, leading to product 3o in 88% yield and 93% ee.

ACS Paragon Plus Environment

Page 3 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

a

a

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

Scheme 2. Scope of the aromatic racemic allylic alcoholsa-d 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 substituents, the yields and enantioselectivity of products were increased (78-89% yields, 92-98% ee, 3p-3s). A variety of aliphatic allylic alcohols which have different substituent in the alkyl chain were allowed to react with acetylacetone (2a), good yields and excellent enantioselectivity were obtained (3t-3u). Of particular note, various generally labile functional groups such as Br, N3, and benzoyl and ether were well compatible, the alkylation products (3v-3y) were obtained in excellent enantioselectivity (93-96% ee). Moreover, cyclopentyl and cyclohexyl substituted substrates also worked well, yielding the alkylation products (3z-3aa) in excellent regio- and enantioselectivity (89-90% ee).

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

Scheme 3. Scope of the aliphatic allylic alcoholsa-d The scope of 1,3-diketones was then examined. The results are summarized in Scheme 4. The reaction of heptane-1,3-diones with alcohol 1a gave product 3ab in 89% yield and 92% ee. However, when the sterically more congested derivative (R = tBu) was tested, the enantioselectivity was decreased (31% ee, 3ad, Scheme 4). The reaction of the 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).

a

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

Scheme 4. Scope of 1,3-diketonesa-d

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

gioselectivity 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).

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 o b c was stirred at 70 C for 30 min. Isolated yield. Determined 1 d by H NMR analysis. Determined by HPLC analysis.

Scheme 5. One-pot protocol of alkylation and deacetylationa-d 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 literature.25

Transformations. To further demonstrate the utility of this method, the transformations of the branched allylated 1,3-diketones were conducted as shown in Scheme 7. Compound 3a could be condensed with hydroxylamine and hydrazines, leading to their corresponding oxazole (6a) and pyrazoles (7a-b), respectively, in excellent yields. There is no erosion of the enantiomeric purity during the transformations.

Scheme 7. Transformations of product 3a Kinetic Resolution Studies. Since the racemic allylic alcohols were used, we became intrigued to investigate the possible kinetic resolution involved. Accordingly, the reactions were quenched at 2 h and 24 h, respectively, as shown in Scheme 8. It was found the ee of the desired product 3a remained excellent and unchanged in both cases (95%) and the allylic alcohol 1a was recovered with 25% and 65% ee, respectively, which suggested the kinetic resolution is occurring (eqs 1-2, Scheme 8). Meanwhile, the reaction between the same equivalent amount 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).

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; The decarboxylation condition: 3b' (0.1 mmol), KOH (0.4 mmol) and EtOH/H2O = 1/1 (3 mL) were added, and the o b c mixture was stirred at 80 C for 1 h. Isolated yield. Dr and 1 d b/l was determined by H NMR analysis. Determined by HPLC analysis.

Scheme 6. Investigation of β-ketoesters and dimethyl malonatea-d

Scheme 8. Kinetic resolution studies

In addition to 1,3-diketones, cyclic β-ketoester 2a' was also well tolerated, furnishing the corresponding α-allyl βketoester compound 3a' in 83% yield with 6/1 dr and the ee for the major diastereoisomer was confirmed as 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 the desired product 3b' could be obtained in 96% yield with excellent re-

Proposed Catalytic Cycle. Based on the experimental observations, a plausible mechanism was proposed in Scheme 9 by utilizing the formation of 3a as an example. In the presence of 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 water and

ACS Paragon Plus Environment

Page 4 of 8

Page 5 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society conjugate base of the acid. Then, the acetylacetone (2a) directly attacks the allyl fragment to deliver η2-allylrhodium(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).

Scheme 9. Proposed catalytic cycle

CONCLUSION In conclusion, we realized a rhodium-catalyzed highly regio- and enantioselective allylic alkylation reaction of 1,3-diketones, using unactivated allylic alcohols as electrophiles for the first time. 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 enantioselectivity for both aryl and aliphatic allylic alcohols warrant the synthetic potentials of this method.

ASSOCIATED CONTENT Supporting Information Experimental procedures and compound characterization data. This material is available free of charge via the internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author E-mail: [email protected]. Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT We thank the National Key R&D Program of China (2016YFA0202900), the National Basic Research Program of China (2015CB856600), NSFC (21332009, 21421091, 21572250), Program of Shanghai Subject Chief Scientist (16XD1404300),

and the CAS (XDB20000000, QYZDYSSWSLH012) for generous financial support. Dr Sheng-Biao Tang thanks Pharmaron for a postdoctoral fellowship.

REFERENCES (1) For selected reviews, see: (a) Trost, B. M.; Van Vranken, D. L. Asymmetric Transition Metal-Catalyzed Allylic Alkylations. Chem. Rev. 1996, 96, 395. (b) Trost, B. M.; Crawley, M. L. Asymmetric Transition-Metal-Catalyzed Allylic Alkylations:  Applications in Total Synthesis. Chem. Rev. 2003, 103, 2921. (c) Lu, Z.; Ma, S. MetalCatalyzed Enantioselective Allylation in Asymmetric Synthesis. Angew. Chem. Int. Ed. 2007, 47, 258. (d) Weaver, J. D.; Recio, A. III.; Grenning, A. J.; Tunge, J. A. Transition Metal-Catalyzed Decarboxylative Allylation and Benzylation Reactions. Chem. Rev. 2011, 111, 1846. (e) Butt, N. A.; Zhang, W. Transition Metal-Catalyzed Allylic Substitution Reactions With Unactivated Allylic Substrates. Chem. Soc. Rev. 2015, 44, 7929. For a book, see: (f) Transition Metal Catalyzed Enantioselective Allylic Substitution in Organic Synthesis (Ed.: Kazmaier, U.), Springer Heidelberg, 2012. (2) For selected reviews, see: (a) Helmchen, G. Enantioselective Palladium-Catalyzed Allylic Substitutions With Asymmetric Chiral Ligands. J. Organomet. Chem. 1999, 576, 203. (b) Tenaglia, A.; Heumann, A. Palladium-Catalyzed Enantioselective Organic Transformations. Angew. Chem. Int. Ed. 1999, 38, 2180. (c) Trost, B. M. Pd Asymmetric Allylic Alkylation (AAA). A Powerful Synthetic Tool. Chem. Pharm. Bull. 2002, 50, 1. (d) Kazmaier, U. Palladium Catalyzed Allylic Alkylations of Nonstabilized Enolates. Curr. Org. Chem. 2003, 7, 317. (e) Trost, B. M. Asymmetric Allylic Alkylation, an Enabling Methodology. J. Org. Chem. 2004, 69, 5813. (f) Jensen, T.; Fristrup, P. Toward Efficient Palladium-Catalyzed Allylic C-H Alkylation. Chem. Eur. J. 2009, 15, 9632. (g) Milhau, L.; Guiry, P. J. Palladium-Catalyzed Enantioselective Allylic Substitution. Top. Organomet. Chem. 2011, 38, 95. (3) For selected reviews, see: (a) Takeuchi, R.; Kezuka, S. IridiumCatalyzed Formation of Carbon-Carbon and Carbon-Heteroatom Bonds. Synthesis 2006, 2006, (b elmchen, ahnz, bon, chelwies, eihofen, Iridium-Catalysed Asymmetricallylic Substitutions. Chem. Commun. 2007, 675. (c) Hartwig, J. F.; Stanley, L. M. Mechanistically Driven Development of Iridium Catalysts for Asymmetric Allylic Substitution. Acc. Chem. Res. 2010, 43, 1461. (d) Hartwig, J. F.; Pouy, M. J. Iridium-Catalyzed Allylic Substitution. Top. Organomet. Chem. 2011, 34, 169. (e) Liu, W.-B.; Xia, J.-B.; You, S.-L. Iridium-catalyzed asymmetric allylic substitutions. Top. Organomet. Chem. 2011, 38, 155. (f) Tosatti, P.; Nelson, A.; Marsden, S. P. Recent Advances and Applications of Iridium-Catalysed Asymmetric Allylic Substitution. Org. Biomol. Chem. 2012, 10, 3147. (g) Helmchen, G. In Molecular Catalysis; Gade, L. H.; Hofmann, P., Ed.; Wiley-VCH: Weinheim, 2014; pp 235-254. (h) Zhuo, C.-X.; Zheng, C.; You, S.-L. Transition-Metal-Catalyzed Asymmetric Allylic Dearomatization Reactions. Acc. Chem. Res. 2014, 47, 2558. (i) Hethcox, J. C.; Shockley, S. E.; Stoltz, B. M. Iridium-Catalyzed Diastereo-, Enantio-, and Regioselective Allylic Alkylation with Prochiral Enolates. ACS Catal. 2016, 6, 6207. (j) Qu, J.; Helmchen, G. Applications of IridiumCatalyzed Asymmetric Allylic Substitution Reactions in TargetOriented Synthesis. Acc. Chem. Res. 2017, 50, 2539. (4) For selected reviews, see: (a) Pineschi, M. Copper-Catalyzed Enantioselective Allylic Alkylation Ring-Opening Reactions of SmallRing Heterocycles with Hard Alkyl Metals. New J. Chem. 2004, 28, 657. (b) Yorimitsu, H.; Oshima, K. Recent Progress in Asymmetric Allylic Substitutions Catalyzed by Chiral Copper Complexes. Angew. Chem. Int. Ed. 2005, 44, 4435. (c) Alexakis, A.; Bäckvall, J. E.; Krause, N.; Pàmies, O.; Diéguez, M. Enantioselective Copper-Catalyzed Conjugate Addition and Allylic Substitution Reactions. Chem. Rev. 2008, 108, 2796. (d) Harutyunyan, S. R.; den Hartog, T.; Geurts, K.; Minnaard, A. J.; Feringa, B. L. Catalytic Asymmetric Conjugate Addition and Allylic Alkylation with Grignard Reagents. Chem. Rev. 2008, 108, 2824. (e) Falciola, C. A.; Alexakis, A. Copper-Catalyzed Asymmetric Allylic Alkylation. Eur. J. Org. Chem. 2008, 2008, 3765. (f) Langlois, J.-B.; Alexakis, A. Copper-Catalyzed Enantioselective Allylic Substitu-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

tion. Top. Organomet. Chem. 2011, 38, 235. (g) Pineschi, M.; Bussolo, V. D. I.; Crotti, P. Copper-Catalyzed Divergent Kinetic Resolution of Racemic Allylic Substrates. Chirality 2011, 23, 703. (h) Hornillos, V.; Gualtierotti, J.-B.; Feringa, B. L. Asymmetric Allylic Substitutions Using Organometallic Reagents. Top. Organomet. Chem. 2016, 58, 1. For a recent book, see: (i) Copper-Catalyzed Asymmetric Synthesis. Alexakis A.; Krause N.; Woodward S. (eds), Wiley, Weinheim, 2014. (5) For selected reviews, see: (a) Oliver, S.; Evans, P. A. TransitionMetal-Catalyzed Allylic Substitution Reactions: Stereoselective Construction of α- and β-Substituted Carbonyl Compounds. Synthesis 2013, 45, 3179. (b) Koschker, P.; Breit, B. Branching Out: RhodiumCatalyzed Allylation with Alkynes and Allenes. Acc. Chem. Res. 2016, 49, 1524. (c) Haydl, A. M.; Breit, B.; Liang, T.; Krische, M. J. Alkynes as Electrophilic or Nucleophilic Allylmetal Precursors in Transition Metal Catalysis. Angew. Chem. Int. Ed. 2017, 56, 11312. (6) For selected reviews on Mo-catalyzed AAA reactions, see: (a) Belda, O.; Moberg, C. Molybdenum-Catalyzed Asymmetric Allylic Alkylations. Acc. Chem. Res. 2004, 37, 159. (b) Moberg, C. Molybdenum-Catalyzed and Tungsten-Catalyzed Enantioselective Allylic Substitutions. Top. Organomet. Chem. 2011, 38, 209. For selected examples, see: (c) Trost, B. M.; Hachiya, I. Asymmetric Molybdenum-Catalyzed Alkylations. J. Am. Chem. Soc. 1998, 120, 1104. (d) Glorius, F.; Pfaltz, A. Enantioselective Molybdenum-Catalyzed Allylic Alkylation Using Chiral Bisoxazoline Ligands. Org. Lett. 1999, 1, 141. (e) Trost, B. M.; Hildbrand, S.; Dogra, K. Regio- and Enantioselective Molybdenum-Catalyzed Alkylations of Polyenyl Esters. J. Am. Chem. Soc. 1999, 121, 10416. (f) Belda, O.; Kaiser, N.-F.; Bremberg, U.; Larhed, M.; Hallberg, A.; Moberg, C. Highly Stereo- and Regioselective Allylations Catalyzed by o− yridylamide Complexes: Electronic and Steric Effects of the Ligand. J. Org. Chem. 2000, 65, 5868. (g) Trost, B. M.; Zhang, Y. Mo-Catalyzed Regio-, Diastereo-, and Enantioselective Allylic Alkylation of 3-Aryloxindoles. J. Am. Chem. Soc. 2007, 129, 14548. (h) Trost, B. M.; Miller, J. R.; Hoffman, C. M. A Highly Enantio- and Diastereoselective Molybdenum-Catalyzed Asymmetric Allylic Alkylation of Cyanoesters. J. Am. Chem. Soc. 2011, 133, 8165. (7) For an early example on Tungsten-catalyzed AAA reactions, see: Lloyd-Jones, G. C.; Pfaltz, A. Chiral Phosphanodihydrooxazoles in Asymmetric Catalysis: Tungsten-Catalyzed Allylic Substitution. Angew. Chem. Int. Ed. 1995, 34, 462. (8) For selected reviews on Ru-catalyzed AAA reactions, see: (a) Bruneau, C.; Renaud, J.; Demerseman, B. Pentamethylcyclopentadienyl-Ruthenium Catalysts for Regio- and Enantioselective Allylation of Nucleophiles. Chem. Eur. J. 2006, 12, 5178. (b) Bruneau, C.; Renaud, J.; Demerseman, B. Ruthenium Catalysts for Selective Nucleophilic Allylic Substitution. Pure Appl. Chem. 2008, 80, 861. (c) Begouin, J.M.; Klein, J.; Weickmann, D.; Plietker, B. Allylic Substitutions Catalyzed by Miscellaneous Metals. Top. Organomet. Chem. 2011, 38, 269. For selected examples, see: (d) Matsushima, Y.; Onitsuka, K.; Kondo, T.; Mitsudo, T. A.; Takahashi, S. Asymmetric Catalysis of PlanarChiral Cyclopentadienylruthenium Complexes in Allylic Amination and Alkylation. J. Am. Chem. Soc. 2001, 123, 10405. (e) Onitsuka, K.; Matsushima, Y.; Takahashi, S. Kinetic Resolution of Allyl Carbonates in Asymmetric Allylic Alkylation Catalyzed by Planar-Chiral Cyclopentadienyl-Ruthenium Complexes. Organometallics 2005, 24, 6472. (f) Bayer, A.; Kazmaier, U. Highly Regioselective RutheniumCatalyzed Allylic Alkylations of Chelated Enolates. Org. Lett. 2010, 12, 4960. (g) Kanbayashi, N.; Hosoda, K.; Kato, M.; Takii, K.; Okamura, T. A.; Onitsuka, K. Enantio- and Diastereoselective Asymmetric Allylic Alkylation Catalyzed by a Planar-chiral Cyclopentadienyl Ruthenium Complex. Chem. Comm. 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. First Highly Enantioselective Allylic Alkylations Catalysed by Platinum Complexes. Chem. Comm. 1999, 913. (b) Blacker, A. J.; Clarke, M. L.; Loft, M. S.; Mahon, M. F.; Humphries, M. E.; Williams, J. M. Platinum-Catalysed Allylic Alkylation: Reactivity, Enantioselectivity, and Regioselectivity. Chem. Eur. J. 2000, 6, 353. (10) (a) Evans, P. A.; Nelson, J. D. Conservation of Absolute Configuration in the Acyclic Rhodium-Catalyzed Allylic Alkylation Reaction:  Evidence for an Enyl (σ + π Organorhodium Intermediate. J. Am. Chem. Soc. 1998, 120, 5581. (b) Evans, P. A.; Kennedy, L. J. Enan-

Page 6 of 8

tiospecific and Regioselective Rhodium-Catalyzed Allylic Alkylation: Diastereoselective Approach to Quaternary Carbon Stereogenic Centers. Org. Lett. 2000, 2, 2213. (c) Evans, P. A.; Robinson, J. E. Regio-and Diastereoselective Tandem Rhodium-Catalyzed Allylic Alkylation/Pauson-Khand Annulation Reactions. J. Am. Chem. Soc. 2001, 123, 4609. (d) Evans, P. A.; Uraguchi, D. Regio- and Enantiospecific Rhodium-catalyzed Arylation of Unsymmetrical Fluorinated Acyclic Allylic Carbonates: Inversion of Absolute Configuration. J. Am. Chem. Soc. 2003, 125, 7158. (e) Evans, P. A.; Leahy, D. K. Regioselective and Enantiospecific Rhodium-Catalyzed Allylic Alkylation Reactions Using Copper (I) Enolates: Synthesis of (-)-Sugiresinol Dimethyl Ether. J. Am. Chem. Soc. 2003, 125, 8974. (f) Evans, P. A.; Lawler, M. J. Regio- and Diastereoselective Rhodium-Catalyzed llylic ubstitution with cyclic α-Alkoxy-Substituted Copper(I) Enolates:  Stereodivergent Approach to 2,3,6-Trisubstituted Dihydropyrans. J. Am. Chem. Soc. 2004, 126, 8642. (g) Evans, P. A.; Clizbe, E. A.; Lawler, M. J.; Oliver, S. Enantioselective Rhodium-Catalyzed Allylic Alkylation of Acyclic α-Alkoxy Aubstituted Ketones Using a Chiral Monodentate Phosphite Ligand. Chem. Sci. 2012, 3, 1835. (h) Evans, P. A.; Oliver, S.; Chae, J. Rhodium-Catalyzed Allylic Substitution with an Acyl Anion Equivalent: Stereospecific Construction of Acyclic Quaternary Carbon Stereogenic Centers. J. Am. Chem. Soc. 2012, 134, 19314. (i) Evans, P. A.; Oliver, S. Regio- and Enantiospecific Rhodium-Catalyzed Allylic Substitution with an Acyl Anion Equivalent. Org. Lett. 2013, 15, 5626. (j) Turnbull, B. W. H.; Oliver, S.; Evans, P. A. Stereospecific Rhodium-Catalyzed Allylic Substitution with Alkenyl Cyanohydrin Pronucleophiles: Construction of Acyclic Quaternary ubstituted α,β-Unsaturated Ketones. J. Am. Chem. Soc. 2015, 137, 15374. (k) Turnbull, B. W. H.; Chae, J.; Oliver, S.; Evans, P. A. Regio- and Stereospecific Rhodium-Catalyzed Allylic Alkylation with an Acyl Anion Equivalent: an Approach to Acyclic α-ternary β, γUnsaturated Aryl Ketones. Chem. Sci. 2017, 8, 4001. (11) (a) Turnbull, B. W. H.; Evans, P. A. Enantioselective RhodiumCatalyzed Allylic Substitution with a Nitrile Anion: Construction of Acyclic Quaternary Carbon Stereogenic Centers. J. Am. Chem. Soc. 2015, 137, 6156. (b) Wright, T. B.; Evans, P. A. Enantioselective Rhodium-Catalyzed llylic lkylation of rochiral α,α-Disubstituted Aldehyde Enolates for the Construction of Acyclic Quaternary Stereogenic Centers. J. Am. Chem. Soc. 2016, 138, 15303. (12) Hayashi, T.; Okada, A.; Suzuka, T.; Kawatsura, M. High Enantioselectivity in Rhodium-Catalyzed Allylic Alkylation of 1Substituted 2-Propenyl Acetates. Org. Lett. 2003, 5, 1713. (13) Kazmaier, U.; Stolz, D. Regio- and Stereoselective RhodiumCatalyzed Allylic Alkylations of Chelated Enolates. Angew. Chem. Int. Ed. 2006, 45, 3072. (14) For selected examples, see: (a) Arnold, J. S.; Nguyen, H. M. Rhodium-Catalyzed Dynamic Kinetic Asymmetric Transformations of Racemic Tertiary Allylic Trichloroacetimidates with Anilines. J. Am. Chem. Soc. 2012, 134, 8380. (b) Arnold, J. S.; Cizio, G. T.; Heitz, D. R.; Nguyen, H. M. Rhodium-Catalyzed Regio- and Enantioselective Amination of Racemic Secondary Allylic Trichloroacetimidates with N-methyl Anilines. Chem. Comm. 2012, 48, 11531. (c) Arnold, J. S.; Nguyen, H. M. Rhodium-Catalyzed Asymmetric Amination of Allylic Trichloroacetimidates. Synthesis 2013, 45, 2101. (d) Arnold, J. S.; Mwenda, E. T.; Nguyen, H. M. Rhodium-Catalyzed Sequential Allylic Amination and Olefin Hydroacylation Reactions: Enantioselective Synthesis of Seven-Membered Nitrogen Heterocycles. Angew. Chem. Int. Ed. 2014, 53, 3688. (e) Mwenda, E. T.; Nguyen, H. M. Enantioselective Synthesis of 1, 2-Diamines Containing Tertiary and Quaternary Centers through Rhodium-Catalyzed DYKAT of Racemic Allylic Trichloroacetimidates. Org. Lett. 2017, 19, 4814. (15) Li, C.; Breit, B. Rhodium-Catalyzed Dynamic Kinetic Asymmetric Allylation of Phenols and 2-Hydroxypyridines. Chem. Eur. J. 2016, 22, 14655. (b) Zhou, Y.; Breit, B. Rhodium-Catalyzed Asymmetric NH Functionalization of Quinazolinones with Allenes and Allylic Carbonates: The First Enantioselective Formal Total Synthesis of (-)Chaetominine. Chem. Eur. J. 2017, 23, 18156. (16) Vrieze, D. C.; Hoge, G. S.; Hoerter, P. Z.; Van Haitsma, J. T.; Samas, B. M. A Highly Enantioselective Allylic Amination Reaction Using a Commercially Available Chiral Rhodium Catalyst: Resolution of Racemic Allylic Carbonates. Org. Lett. 2009, 11, 3140.

ACS Paragon Plus Environment

Page 7 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society (17) Liang, L.; Xie, M.-S.; Qin, T.; Zhu, M.; Qu, G.-R.; Guo, H.-M. Regio- and Enantioselective Synthesis of Chiral Pyrimidine Acyclic Nucleosides via Rhodium-Catalyzed Asymmetric Allylation of Pyrimidines. Org. Lett. 2017, 19, 5212. (18) For selected examples, see: (a) Li, C.; Breit, B. RhodiumCatalyzed Chemo- and egioselective ecarboxylative ddition of βKetoacids to Allenes: Efficient Construction of Tertiary and Quaternary Carbon Centers. J. Am. Chem. Soc. 2014, 136, 862. (b) Beck, T. M.; Breit, B. Regioselective Rhodium-Catalyzed Addition of 1, 3Dicarbonyl Compounds to Terminal Alkynes. Org. Lett. 2016, 18, 124. (c) Li, C.; Grugel, C. P.; Breit, B. Rhodium-Catalyzed Chemo- and Regioselective Decarboxylative Addition of β-Ketoacids to Alkynes. Chem. Comm. 2016, 52, 5840. (d) Beck, T. M.; Breit, B. Regioselective Rhodium-Catalyzed Addition of β-Keto Esters, β-Keto Amides, and 1, 3-Diketones to Internal Alkynes. Eur. J. Org. Chem. 2016, 2016, 5839. (e) Beck, T. M.; Breit, B. Regio- and Enantioselective RhodiumCatalyzed Addition of 1, 3-Diketones to Allenes: Construction of Asymmetric Tertiary and Quaternary All Carbon Centers. Angew. Chem. Int. Ed. 2017, 56, 1903. (19) For selected examples, see: (a) Chen, Q.-A.; Chen, Z.; Dong, V. M. Rhodium-Catalyzed Enantioselective Hydroamination of Alkynes with Indolines. J. Am. Chem. Soc. 2015, 137, 8392. (b) Cruz, F. A.; Chen, Z.; Kurtoic, S. I.; Dong, V. M. Tandem Rh-Catalysis: Decarboxylative β-Keto acid and Alkyne Cross-Coupling. Chem. Commun. 2016, 5836. (c) Cruz, F. A.; Zhu, Y.; Tercenio, Q. D.; Shen, Z.; Dong, V. M. Alkyne Hydroheteroarylation: Enantioselective Coupling of Indoles and Alkynes via Rh-Hydride Catalysis. J. Am. Chem. Soc. 2017, 139, 10641. (d) Chen, Z.; Dong, V. M. Enantioselective Semireduction of Allenes. Nat. Comm. 2017, 8, 784. (20) (a) Zheng, W.-F.; Xu, Q.-J.; Kang, Q. Rhodium/Lewis Acid Catalyzed Regioselective Addition of 1,3-Dicarbonyl Compounds to Internal Alkynes. Organometallics 2017, 36, 2323. (b) Bora, P. P.; Sun, G.-J.; Zheng, W.-F.; Kang, Q. Rh/Lewis Acid Catalyzed Regio-, Diastereo- and Enantioselective Addition of 2-Acyl Imidazoles with Allenes. Chin. J. Chem. 2018, 36, 20. (21) For selected reviews, see: (a) Bandini, M. Allylic Alcohols: Sustainable Sources for Catalytic Enantioselective Alkylation Reactions. Angew. Chem. Int. Ed. 2011, 50, 994. (b) Bandini, M.; Cera, G.; Chiarucci, M. Catalytic Enantioselective Alkylations with Allylic Alcohols. Synthesis 2012, 44, 504. (c) Sundararaju, B.; Achard, M.; Bruneau, C. Transition Metal Catalyzed Nucleophilic Allylic Substitution: Activation of Allylic Alcohols via π-Allylic Species. Chem. Soc. Rev. 2012, 41, 4467. (22) Defieber, C.; Ariger, M. A.; Moriel, P.; Carreira, E. M. IridiumCatalyzed Synthesis of Primary Allylic Amines from Allylic Alcohols: Sulfamic Acid as Ammonia Equivalent. Angew. Chem. Int. Ed. 2007, 46, 3139. (23) (a) Roggen, M.; Carreira, E. M. Enantioselective Allylic Etherification: Selective Coupling of Two Unactivated Alcohols. Angew. Chem. Int. Ed. 2011, 50, 5568. (b) Lafrance, M.; Roggen, M.; Carreira, E. M. Direct, Enantioselective Iridium-Catalyzed Allylic Amination of Racemic Allylic Alcohols. Angew. Chem. Int. Ed. 2012, 51, 3470. (c) Liu, W.-B.; Reeves, C. M.; Stoltz, B. M. Enantio-, Diastereo-, and Regioselective Iridium-Catalyzed Asymmetric Allylic Alkylation of Acyclic βKetoesters. J. Am. Chem. Soc. 2013, 135, 17298. (24) (a) Fujimoto, K.; Maekawa, H.; Matsubara, Y.; Nishiguchi, I. Electrochemical Deacetylation of 1,3-Dicarbonyl Compounds. Chem. Lett. 1996, 25, 143. (b) Kanojia, S. V.; Chatterjee, S.; Gamre, S.; Chattopadhyay, S.; Sharma, A. Asymmetric Synthesis of the Constitutive C22-carboxylic Acid of Macroviracin A. Tetrahedron 2015, 71, 1732. (25) Hamilton, J. Y.; Sarlah, D.; Carreira, E. M. Iridium-Catalyzed Enantioselective Allyl-alkene Coupling. J. Am. Chem. Soc. 2014, 136, 3006.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 8 of 8

8