Cu(I)-Catalyzed Chemoselective and ... - ACS Publications

Jul 27, 2016 - Jing-Yan Zhu , Wu-Lin Yang , Yang-Zi Liu , Shao-Jing Shang ... Xiao-Xue Sun , Can Li , Ying-Ying He , Zi-Qi Zhu , Guang-Jian Mei , Feng...
0 downloads 0 Views 375KB Size
Subscriber access provided by SUNY DOWNSTATE

Letter

Cu(I)-Catalyzed Chemo- and Stereoselective [3+3] Cycloaddition of Azomethine Ylides with 2-Indolylnitroethylenes: Facile Access to Highly Substituted Tetrahydro-#-Carbolines Wu-Lin Yang, Chun-Yan Li, Wen-Jing Qin, Fei-Fei Tang, Xingxin Yu, and Wei-Ping Deng ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.6b01596 • Publication Date (Web): 27 Jul 2016 Downloaded from http://pubs.acs.org on July 27, 2016

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 free 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 accessible to all readers and 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.

ACS Catalysis 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 6

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 Catalysis

Cu(I)-Catalyzed Chemo- and Stereoselective [3+3] Cycloaddition of Azomethine Ylides with 2-Indolylnitroethylenes: Facile Access to Highly Substituted Tetrahydro-γ-Carbolines Wu-Lin Yang, Chun-Yan Li, Wen-Jing Qin, Fei-Fei Tang, Xingxin Yu, and Wei-Ping Deng* School of Pharmacy and Shanghai Key Laboratory of New Drug Design, East China University of Science and Technology, 130 Meilong Road, Shanghai 200237, China

ABSTRACT: A Cu(I)-catalyzed asymmetric [3+3] cycloaddition of azomethine ylides with 2-indolylnitroethylenes is described. Challenges confronted in this reaction include chemoselectivity between [3+2] cycloaddition and [3+3] cycloaddition, reactivity and stereoselectivity of intramolecular Friedel-Crafts reaction, and enantioselective control for constructing tetrahydro-γ-carbolines bearing multiple stereocenters. In the presence of copper(I)/Ph-Phosferrox complex, an array of chiral tetrahydro-γ-carboline compounds was generally obtained in moderate to high yields (up to 92%) with moderate to high chemoselectivities (up to 94:6 c.r.) and high level of stereoselectivities (up to >98:2 d.r., >99% ee in most cases). KEYWORDS: asymmetric catalysis, chemoselectivity, cycloaddition, azomethine ylide, tetrahydro-γ-carboline

Tetrahydro-carboline skeletons have long been identified as a “privileged scaffold” in a great many of natural bioactive products and pharmaceuticals.1 Among the isomeric compounds, both tetrahydro-β-carboline2 and tetrahydro-γcarboline3 derivatives are endowed with a wide range of important biological activities. For example, Tadalafil was marketed for treating erectile dysfunction (ED). Its analogue, tetrahydro-γ-carboline surrogate, are also potent inhibitor of PDE5 (Figure 1).3b In contrast to a variety of known synthetic routes to chiral tetrahydro-β-carboline derivatives,4 limited methods have been developed for the direct construction of stereochemically enriched tetrahydro-γ-carboline frameworks.5 Existing methods to synthesize chiral tetrahydro-γcarboline compounds include classical resolution,5a Pdcatalyzed enantioselective intramolecular allylic alkylation,5b enantioselective iso-Pictet-Spengler reaction,5c and diastereoselective cyclization of chiral precursors.5d,e To the best of our knowledge, the straightforward and direct route to tetrahydro-γ-carboline compounds containing three stereocenters has not been reported. Therefore, development of efficient catalytic asymmetric method for the preparation of highly substituted tetrahydro-γ-carboline derivatives is still an unmet need with potentially wide utilities due to their prominence in natural products and bioactive molecules.3 The catalytic asymmetric 1,3-dipolar cycloaddition of azomethine ylides with dipolarophiles is one of the most powerful tools for the construction of optically active nitrogenous heterocycles.6 In the past decade, a great deal of efforts has been focused on the development of asymmetric [3+2] cycloaddition of azomethine ylides with a variety of electron-deficient

O H N

Me Rich assortment of synthetic N routes to chiral THBCs MeO

N H

Tetrahydro- -Carbolines (THBCs)

Tadalafil

O

O

Tetrahydro- -Carbolines (THGCs)

H

O

OMe

Reserpine

O

Tadalafil analogue

OR

H MeO2C

Me N N

HN

H

Tetrahydro-Carboline Skeletons

Isomer

O H

N N H H

O

O

Limited methods for constructing chiral THGCs

HO N H

H

N H CHO CO2Me Ervatamine A

Figure 1. Natural products and pharmaceuticals containing tetrahydrocarboline skeletons.

alkenes, which have been widely applied for the preparation of structurally diverse chiral pyrrolidines.7 In conjunction with our continuing efforts in the cycloaddition of azomethine ylides,8 as well as being inspired by pioneering works in [3+3] cycloaddition of azomethine ylides,9 we envisioned that the above-mentioned chiral tetrahydro-γ-carboline scaffold could be retrosynthetically formed through an asymmetric [3+3] cycloaddition of azomethine ylides with 2indolylnitroethylenes10 serving as novel potentially dipolarophiles (Scheme 1). It is noteworthy that several challenges are associated with this strategy, such as: 1) chemoselectivity between [3+2] cycloaddition and [3+3] cycloaddition; 2) reactivity and stereoselectivity of Friedel-Crafts reaction of indole to imine; 3) construction of tetrahydro-γ-carbolines bearing mul-

ACS Paragon Plus Environment

ACS Catalysis

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

tiple stereocenters. Nevertheless, the key point of this strategy relies on the stability of zwitterionic intermediates I derived from the Michael addition step (Scheme 1). Normally, it proceeds via intramolecular Mannich cyclization to form the pyrrolidine skeleton, thus completing [3 + 2] cycloaddition (Scheme 1, path a).11 However, if the intermediates I is stable enough,11b,e,h,12 the reaction could be interrupted and provides a Michael adduct that could tend towards the intramolecular Friedel-Crafts reaction of a regenerated electron-rich indole to imine (Scheme 1, path b),13 thus affording tetrahydro-γcarboline product. Herein, we demonstrate the first catalytic asymmetric [3+3] cycloaddition of azomethine ylides to 2indolylnitroethylenes, affording highly substituted tetrahydroγ-carboline derivatives in moderate to high yields, moderate to high chemoselectivities (up to 94:6 c.r.), and excellent level of stereoselectivities (up to >98:2 d.r., >99% ee in most cases).

lowering the temperature to −40 °C led to a slight improvement in chemoselectivity (88:12 c.r.) and diastereoselectivity (93:7 d.r.) (Table 1, entry 11). Interestingly, the same level of dr and ee values were observed for both [3+3] and [3+2] adducts under the same reaction condition (Table 1, entries 6 and 11).When the catalyst loading was reduced to 5 mol %, the [3+3] cycloadduct 3a was obtained with the similar excellent stereoselectivity (Table 1, entry 14). Considering the yield and reaction time, 10 mol % catalyst loading was employed for the subsequent experiments. The absolute configuration of 3a was determined as (1R,3S,4R) by X-ray diffraction analysis (see SI). Table 1. Optimization of Reaction Conditionsa

We began our study by choosing 2-indolylnitroethylene 1a and imino ester 2a as model substrates to test the feasibility of this new [3+3] cycloaddition strategy for the synthesis of structurally and biologically important chiral tetrahydro-γcarboline derivatives. Initially, planar-chiral ferrocene P,Nligand (Ph-Phosferrox L1) was employed with Cu(CH3CN)4BF4 as catalytic metal salt, along with 20 mol % of Et3N as base in THF at room temperature. As expected, a normal [3+2] cycloadduct 4a was obtained as exclusive product in 97% yield (Table 1, entry 1) (The absolute configuration of 4a was determined as (2S,3S,4R,5S) by X-ray diffraction analysis, see SI for details). Next, in order to interrupt the normal [3+2] cycloaddition, we try to lower the reaction temperature. Pleasingly, the desired [3+3] cycloadduct 3a was obtained in 67% yield with moderate chemoselectivity (68:32 c.r.) and diastereoselectivity (87:13 d.r.) and excellent enantioselectivity (>99% ee) at −20 °C (Table 1, entry 2). Further lowering the temperature to −30 °C resulted in a significant improvement of chemoselectivity (84:16) without loss of diastereoselectivity and enantioselectivity (Table 1, entry 3). Unfortunately, further screening of P,N-ligands containing different substitutions on the oxazoline ring did not show any better results (Table 1, entries 4−6). Screening of other Cu(I) salts as metal sources showed that both Cu(CH3CN)4ClO4 and Cu(CH3CN)4PF6 gave [3+3] cycloadduct 3a as major product with excellent stereoselectivities (Table 1, entries 7−8). However, Cu(II) and Ag(I) salts afforded the [3+2] cycloadduct 4a as exclusive product in 60% and 90% yield, respectively (Table 1, entries 9−10). Among the tested metal salts, Cu(CH3CN)4PF6 gave the best result in terms of chemoselectivity (86:14 c.r.) and stereoselectivity (92:8 d.r., >99% ee). Further optimization with respect to solvent did not give any better results (see SI). Further

1e

CuBF4/L1

2f 3 4 5

CuBF4/L1 CuBF4/L1 CuBF4/L2 CuBF4/L3

6

CuBF4/L4

56

58:42

7 8

CuClO4/L1 CuPF6/L1 Cu(OAc)2/ L1

72 83 99 97 98 93 (93)g >99 >99

2 2 2 2

78:22 86:14

87:13 91:9 74:26 86:14 86:14 (86:14)g 92:8 92:8

-

-

-

4

-

-

-

2

93:7 (93:7)g 91:9 91:9 93:7

>99 (>99)g >99 >99 99

10

AgOAc/L1

11h

CuPF6/L1

85

88:12

i

CuPF6/L1 CuPF6/L1 CuPF6/L1

84 82 83

88:12 88:12 88:12

12 13j 14h,k

d.r.c

2 2 2

6 8 12 24

a All reactions were carried out with 0.1 mmol of 1a and 0.12 mmol of 2a in 1.5 mL of THF at −30 °C, CuBF4 = [Cu(CH3CN)4BF4], CuClO4 = [Cu(CH3CN)4ClO4], CuPF6 = [Cu(CH3CN)4PF6]. bIsolated yield of 3a and its diastereomer, isolated yield of [3+2] cycloadduct 4a is given in parentheses, N.P. = no product. cDetermined by 1H NMR spectroscopy. dDetermined by chiral HPLC analysis, the ee referred to the major diastereomer. eReaction conducted at room temperature. fReaction conducted at −20 °C. g The dr or ee values of [3+2] cycloadduct 4a. hReaction conducted at −40 °C. iReaction conducted at −50 °C. jReaction conducted at −60 °C. k5 mol % catalyst was used.

After optimization of the reaction conditions, the substrate scope for catalytic asymmetric [3+3] cycloaddition was investigated. Initially, we probed the reaction scope with regard to the variation on substitutions of indole ring. Other protecting group of 2-indolylnitroethylene, such as allyl or benzyl (Bn), was also compatible in the transformation, providing [3+3] cycloadducts 3b and 3c in good yields with excellent stereoselectivities, albeit only in moderate chemoselectivities (Table 2, entries 2−4). When indole ring bearing a weak electron-deficient substituent (Br), the reaction afforded the [3+3] cycloadduct 3d in excellent stereoselectivity (Table 2, entry 5), albeit with diminished chemoselectivity, probably due to reduced nucleophilicity of indole ring.13b In contrast, 2indolylnitroethylenes 1 bearing electron-rich groups gave the corresponding [3+3] cycloadducts in high yields (78−86%)

ACS Paragon Plus Environment

Page 3 of 6

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 Catalysis

with high chemoselectivities (84:16−89:11 c.r.) and excellent stereoselectivities (93:7−96:4 d.r., 97−>99% ee) (Table 2, entries 6−9). Table 2. Substrate Scope of 2-Indolylnitroethylenes 1a

1a

3

1a

4

1a

5e

1a

6e

1f

e

1f

8e

1f

7 c.r.c

d.r.c

ee (%)d

H/Me (1a)

88:12

93:7

>99

H/Allyl (1b)

72 (3b)

78:22

96:4

>99

9e

1f

10e

1f

11e

1f

e

12

1f

13e

1f

R1/R2

1 2 3

H/Bn (1c)

65 (3c)

68:32

96:4

>99

4e

H/Bn (1c)

62 (3c)

67:33

96:4

>99

5 6

5-Br/Me (1d) 5-Me/Me (1e)

29 (3d) 83 (3e)

32:68 85:15

96:4 93:7

>99 >99

7

5-MeO/Me (1f)

86 (3f)

89:11

93:7

>99

8

6-MeO/Me (1g)

81 (3g)

84:16

96:4

97

e

6-MeO/Me (1g)

78 (3g)

84:16

96:4

97

14e,f

a

All reactions were carried out with 0.2 mmol of 1 and 0.24 mmol of 2a in 3.0 mL of THF at −40 °C. bIsolated yield of 3 and its diastereomer. cDetermined by 1H NMR spectroscopy. d Determined by chiral HPLC analysis, the ee referred to the major diastereomer. e5 mol % catalyst was used.

Subsequently, the potential of this [3+3] cycloaddition with a variety of azomethine ylides was further evaluated. Azomethine ylides 2 bearing electron-deficient (Table 3, entries 2−4), electronically-neutral (Table 3, entry 1), and electron-rich groups (Table 3, entries 5−11) on the aryl ring all reacted with 2-indolylnitroethylenes 1 smoothly affording the corresponding [3+3] cycloadducts in good yields (54−92%), high diastereoselectivities (93:7−>98:2 d.r.), and excellent enantioselectivities (>99% ee). Notablely, azomethine ylides 2h bearing ortho-methyl substituent on the phenyl ring only delivered moderate chemoselectivity (58:42 c.r.), presumably due to the steric encumbrance. Remarkably, the heteroaryl substituted azomethine ylides 2l−m derived from 2-furylaldehyde and 2-thenaldehyde were also tolerated in this reaction and provided the corresponding [3+3] cycloadducts 3s−t with high chemoselectivities (85:15−89:11 c.r.) and diastereoselectivities (92:8−96:4 d.r.), and excellent enantioselectivities (99% ee). When alkyl substituted azomethine ylide 2n was applied in this transformation, strong base 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) was necessary to promote the prior Michael addition step. Unfortunately, due to the less reactive Friedel-Crafts reaction of aliphatic imine,13a [3+2] cycloadduct was isolated as sole product in 87% yield. It should be mentioned that, in all cases, the [3+3] cycloadducts and [3+2] cycloadducts can be easily separated by using flash column chromatography. Table 3. Substrate Scope of Azomethine Ylides 2a

entry

1a

2

yield (%)b 85 (3a)

entry

9

1

1

R

yield (%)b

c.r.c

d.r.c

ee (%)d

1f

Ph (2b) m-ClC6H4 (2c) p-FC6H4 (2d) p-CF3C6H4 (2e) p-MeC6H4 (2f) p-MeC6H4 (2f) m-MeC6H4 (2g) o-MeC6H4 (2h) p-MeOC6H4 (2i) p-tBuC6H4 (2j) 2-naphthyl (2k) 2-furyl (2l) 2-thienyl (2m) cyclohexyl (2n)

82 (3h)

87:13

93:7

>99

84 (3i)

88:12

94:6

>99

80 (3j)

83:17

94:6

>99

73 (3k)

75:25

93:7

>99

83 (3l)

84:16

96:4

>99

85 (3m)

86:14

97:3

>99

85 (3n)

86:14

96:4

>99

54 (3o)

58:42

>98:2

>99

83 (3p)

84:16

>98:2

>99

92 (3q)

94:6

98:2

>99

84 (3r)

88:12

97:3

>99

83 (3s)

85:15

92:8

99

88 (3t)

89:11

96:4

>99

N.P. (87)g

-

-

-

a

All reactions were carried out with 0.2 mmol of 1 and 0.24 mmol of 2 in 3.0 mL of THF at −40 °C. bIsolated yield of 3 and its diastereomer, N.P. = no product. cDetermined by 1H NMR spectroscopy. dDetermined by chiral HPLC analysis, the ee referred to the major diastereomer. eReaction conducted at −30 °C. f20 mol% DBU was used. gIsolated yield of [3+2] cycloadduct.

To further evaluate the synthetic utility of this process, the [3+3] cycloaddition between 1a and 2a was performed on a gram scale, and compound 3a was obtained in 81% yield and >99% ee (Scheme 2a). The nitro group was reduced using sodium borohydride and nickel chloride without loss of stereochemical integrity (Scheme 2b).

Scheme 2. Demonstration of Synthetic Utility A plausible stepwise mechanism is proposed to rationalize the chemoselectivity and stereoselectivity of this [3+3] cycloaddition (Scheme 3). The in situ formed azomethine ylide is coordinated to the Cu complex A leading to the catalytically active species B. Due to the steric effect of bulky phenyl group in the oxazoline ring and PPh2 group of the Phosferrox ligand, the initial Michael addition of species B to 2-indolylnitroethylene 1 via Si face attack (C) generates the zwitterionic intermediate D (Michael adduct can be obtained, see SI for details).11b Since the carbanion adjacent to nitro group is stable enough at low temperature,11b,e,h,12 intramolecular Mannich cyclization is suppressed (path a). Consequently, the intramolecular Friedel-Crafts reaction of indole to imine moiety is preferential (path b), and subsequent

ACS Paragon Plus Environment

ACS Catalysis

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

protonation step gives the [3+3] cycloadduct 3 with regeneration of the catalyst. Notably, based on our observation of same level of stereoselectivities for both [3+3] and [3+2] cycloadducts (Table 1, entries 6 and 11), we speculate that the Michael addition should be the crucial step for the stereochemistry, and a potential kinetic resolution of intermediate D in the intramolecular Friedel-Crafts reaction was not observed. Nevertheless, a precise understanding of the chemoselectivity between [3+2] cycloaddition and [3+3] cycloaddition awaits further study.

Scheme 3. Proposed Mechanism for [3+3] Cycloaddition of Azomethine Ylides and 2-Indolylnitroethylenes In conclusion, we have presented the first example of chemo- and stereoselective [3+3] cycloaddition of azomethine ylides with 2-indolylnitroethylenes employing the Cu(I)/PhPhosferrox complex as the catalyst. A wide range of highly substituted tetrahydro-γ-carboline derivatives were synthesized in moderate to high yields (up to 92%) with moderate to high chemoselectivities (up to 94:6 c.r.) and excellent level of stereoselectivities (up to >98:2 d.r., >99% ee in most cases). This reaction provides an efficient protocol for constructing structurally diverse chiral tetrahydro-γ-carboline derivatives. Further application of this new [3+3] cycloaddition strategy to other reaction system and asymmetric synthesis of bioactive molecules is underway.

ASSOCIATED CONTENT Supporting Information Experimental details, characterization of new compounds, crystallographic data, NMR and HPLC spectra. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author *Email: [email protected]

Notes The authors declare no competing financial interest.

Page 4 of 6

This work is supported by the National Natural Science Foundation of China (No. 21372074), the Shanghai Committee of Science and Technology (No. 14431902500, 14PJD013), and Shanghai Yangfan Program (No. 14YF1404600).

REFERENCES (1) (a) Khorana, N.; Smith, C.; Herrick-Davis, K.; Purohit, A.; Teitler, M.; Grella, B.; Dukat, M.; Glennon, R. A. J. Med. Chem. 2003, 46, 3930−3937; (b) Kawasaki, T.; Higuchi, K. Nat. Prod. Rep. 2005, 22, 761−793; (c) Alekseyev, R. S.; Kurkin, A. V.; Yurovskaya, M. A. Chem. Heterocycl. Comp. 2011, 46, 1169−1198; (d) Henderson, A. J.; Deering, D.; Grabowski, J. F.; Hadden, M.; Jiang, X.; Khmelnitsky, Y.; Luche, M.; Surman, M. D.; Cheetham, S.; Vickers, S.; Viggers, J.; Guzzo, P. R. Bioorg. Med. Chem. Lett. 2010, 20, 7024−7028. (2) (a) Garnier, J.; Mahuteau, J.; Plat, M.; Merienne, C. Phytochemistry 1989, 28, 308−309; (b) Itoh, A.; Tanahashi, T.; Nagakura, N.; Nishi, T. Phytochemistry 2003, 62, 359−369; (c) Limbach, M.; Surayakanta, D.; Janssen, A.; Es-Sayed, M.; Magull, J. M.; de Meijere, A. Eur. J. Org. Chem. 2005, 610−617; (d) Trujillo, J. I. Meyers, M. J.; Anderson, D. R.; Hedge, S.; Mahoney, M. W.; Vernier, W. F.; Buchler, I. P.; Wu, K. K.; Yang, S.; Hartmann, S. J.; Reitz, D. B. Bioorg. Med. Chem. Lett. 2007, 17, 4657−4663. (3) (a) Harbert, C. A.; Plattner, J. J.; Welch, W. M. J. Med. Chem. 1980, 23, 635−643; (b) Orme, M. W.; Sawyer, J. S.; Schultze, L. M. WO 2002000657 A2, 2002; (c) Wu, Y.-T.; Hsu, Y.-L.; Wu, P.-L. Heterocycles 2008, 75, 1191−1197; (d) Bridoux, A.; Millet, R.; Pommery, J.; Pommery, N.; Henichart, J.-P. Bioorg. Med. Chem. 2010, 18, 3910−3924; (e) Ivachtchenko, A. V.; Frolov, E. B.; Mitkin, O. D.; Tkachenko, S. E.; Okun, I. M.; Khvat, A. V. Bioorg. Med. Chem. Lett. 2010, 20, 78−82; (f) Fretz, H.; Valdenaire, A.; Pothier, J.; Hilpert, K.; Gnerre, C.; Peter, O.; Leroy, X.; Riederer, M. A. J. Med. Chem. 2013, 56, 4899−4911; (g) Wang, S.; Wang, Y.; Liu, W.; Liu, N.; Zhang, Y.; Dong, G.; Liu, Y.; Li, Z.; He, X.; Miao, Z.; Yao, J.; Li, J.; Zhang, W.; Sheng, C. ACS Med. Chem. Lett. 2014, 5, 506−511; (h) Zhang, D.-B.; Yu, D.-G.; Sun, M.; Zhu, X.-X.; Yao, X.-J.; Zhou, S.Y.; Chen, J.-J.; Gao, K. J. Nat. Prod. 2015, 78, 1253−1261. (4) For selected examples: (a) Taylor, M. S.; Jacobsen, E. N. J. Am. Chem. Soc. 2004, 126, 10558−10559; (b) Seayad, J.; Seayad, A. M.; List, B. J. Am. Chem. Soc. 2006, 128, 1086−1087; (c) Wanner, M. J.; van der Haas, R. N. S.; de Cuba, K. R.; van Maarseveen, J. H.; Hiemstra, H. Angew. Chem. Int. Ed. 2007, 46, 7485−7487; (d) Li, C.; Xiao, J. J. Am. Chem. Soc. 2008, 130, 13208−13209; (e) BouHamdan, F. R.; Leighton, J. L. Angew. Chem. Int. Ed. 2009, 48, 2403−2406; (f) Muratore, M. E.; Holloway, C. A.; Pilling, A. W.; Ian Storer, R.; Trevitt, G.; Dixon, D. J. J. Am. Chem. Soc. 2009, 131, 10796−10797; (g) Lalonde, M. P.; McGowan, M. A.; Rajapaksa, N. S.; Jacobsen, E. N. J. Am. Chem. Soc. 2013, 135, 1891−1894; (h) Yin, Q.; Wang, S.-G.; You, S.-L. Org. Lett. 2013, 15, 2688−2691; (i) Ghislieri, D.; Houghton, D.; Green, A. P.; Willies, S. C.; Turner, N. J. ACS Catal. 2013, 3, 2869−2872; (j) Liu, X.; Meng, Z.; Li, C.; Lou, H.; Liu, L. Angew. Chem. Int. Ed. 2015, 54, 6012−6015. (5) (a) Vecchietti, V.; Clarke, G. D.; Colle, R.; Giardina, G.; Petrone, G.; Sbacchi, M. J. Med. Chem. 1991, 34, 2624−2633; (b) Bandini, M.; Melloni, A.; Piccinelli, F.; Sinisi, R.; Tommasi, S.; UmaniRonchi, A. J. Am. Chem. Soc. 2006, 128, 1424−1425; (c) Lee, Y.; Klausen, R. S.; Jacobsen, E. N. Org. Lett. 2011, 13, 5564−5567; d) Sheng, Y.-F.; Li, G.-Q.; Kang, Q.; Zhang, A.-J.; You, S.-L. Chem. Eur. J. 2009, 15, 3351−3354; (e) Fandrick, D. R.; Johnson, C. S.; Fandrick, K. R.; Reeves, J. T.; Tan, Z.; Lee, H.; Song, J. J.; Yee, N. K.; Senanayake, C. H. Org. Lett. 2010, 12, 748−751. (6) (a) 1,3-Dipolar Cycloaddition Chemistry; Padwa, A., Ed.; Wiley: New York, 1984; Vol. 1, pp 653−732. (b) Synthetic Applications of 1,3-Dipolar Cycloaddition Chemistry Towards Heterocycles and Natural Products; Padwa, A., Pearson, W. H., Eds.; Wiley: New York, 2003; Vol. 59, pp 1−942.

ACKNOWLEDGMENT ACS Paragon Plus Environment

Page 5 of 6

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 Catalysis

(7) For recent reviews: (a) Adrio, J.; Carretero, J. C. Chem. Commun. 2011, 47, 6784−6794; (b) Zhang, W. Chem. Lett. 2013, 42, 676−681; (c) Adrio, J.; Carretero, J. C. Chem. Commun. 2014, 50, 12434−12446; (d) Nájera, C.; Sansano, J. M. J. Organomet. Chem. 2014, 771, 78−92; (e) Narayan, R.; Potowski, M.; Jia, Z.-J.; Antonchick, A. P.; Waldmann, H. Acc. Chem. Res. 2014, 47, 1296−1310; (f) Hashimoto, T.; Maruoka, K. Chem. Rev. 2015, 115, 5366−5412. (8) (a) Wang, M.; Wang, Z.; Shi, Y.-H.; Shi, X.-X.; Fossey, J. S.; Deng, W.-P. Angew. Chem. Int. Ed. 2011, 50, 4897−4900; (b) Wang, Z.; Luo, S.; Zhang, S.; Yang, W.-L.; Liu, Y.-Z.; Li, H.; Luo, X.; Deng, W.-P. Chem. - Eur. J. 2013, 19, 6739−6745; (c) Yang, W.-L.; Liu, Y.-Z.; Luo, S.; Yu, X.; Fossey, J. S.; Deng, W.-P. Chem. Commun. 2015, 51, 9212−9215; (d) Wang, Z.; Yu, X.; Tian, B.-X.; Payne, D.-T.; Yang, W.-L.; Liu, Y.-Z.; Fossey, J. S.; Deng, W.-P. Chem. Eur. J. 2015, 21, 10457−10465; (e) Yang, W.-L.; Tang, F.-F.; He, F.S.; Li, C.-Y.; Yu, X.; Deng, W.-P. Org. Lett. 2015, 17, 4822−4825; (f) He, F.-S.; Zhu, H.; Wang, Z.; Gao, M.; Yu, X.; Deng, W.-P. Org. Lett. 2015, 17, 4988−4991. (9) Some pioneering works in [3+3] cycloaddition of azomethine ylides: (a) Tong, M.-C.; Chen, X.; Tao, H.-Y.; Wang, C.-J. Angew. Chem. Int. Ed. 2013, 52, 12377−12380; (b) Guo, H.; Liu, H.; Zhu, F.L.; Na, R.; Jiang, H.; Wu, Y.; Zhang, L.; Li, Z.; Yu, H.; Wang, B.; Xiao, Y.; Hu, X.-P.; Wang, M. Angew. Chem. Int. Ed. 2013, 52, 12641−12645; (c) Huang, J.; Luo, S.; Gong, L. Acta Chim. Sin. 2013, 71, 879−883; (d) Shi, F.; Zhu, R.-Y.; Dai, W.; Wang, C.-S.; Tu, S.-J. Chem. - Eur. J. 2014, 20, 2597−2604; (e) Yuan, C.; Liu, H.; Gao, Z.; Zhou, L.; Feng, Y.; Xiao, Y.; Guo, H. Org. Lett. 2015, 17, 26−29; (f) Wang, C.-S.; Zhu, R.-Y.; Zhang, Y.-C.; Shi, F. Chem. Commun. 2015, 51, 11798−11801; (g) Wu, Y.; Qiao, G.; Liu, H.; Zhang, L.; Sun, Z.; Xiao, Y.; Guo, H. RSC Adv. 2015, 5, 84290–84294. [6+3] cycloaddition of azomethine ylides: (h) Potowski, M.; Bauer, J. O.; Strohmann, C.; Antonchick, A. P.; Waldmann, H. Angew. Chem. Int. Ed. 2012, 51, 9512–9516; (i) He, Z.-L.; Teng, H.-L.; Wang, C.-J. Angew. Chem. Int. Ed. 2013, 52, 2934–2938; (j) Liu, H.; Wu, Y.; Zhao, Y.; Li, Z.; Zhang, L.; Yang, W.; Jiang, H.; Jing, C.; Yu, H.; Wang, B.; Xiao, Y.; Guo, H. J. Am. Chem. Soc. 2014, 136, 2625−2629; (k) Teng, H.-L.; Yao, L.; Wang, C.-J. J. Am. Chem. Soc. 2014, 136, 4075−4080; (l) Li, Q.-H.; Wei, L.; Wang, C.-J. J. Am. Chem. Soc. 2014, 136, 8685−8692; (m) He, Z.-L.; Sheong, F. K.; Li, Q.-H.; Lin, Z.; Wang, C.-J. Org. Lett. 2015, 17, 1365−1368. (10) (a) Ni, Q.; Zhang, H.; Grossmann, A.; Loh, C. C. J.; Merkens, C.; Enders, D. Angew. Chem. Int. Ed. 2013, 52, 13562−13566; (b) Talukdar, R.; Tiwari, D. P.; Saha, A.; Ghorai, M. K. Org. Lett. 2014, 16, 3954−3957. (11) Selected examples for [3+2] cycloaddition of azomethine ylides to nitroalkenes: (a) Cabrera, S.; Arrayás, R. G.; Carretero, J. C. J. Am. Chem. Soc. 2005, 127, 16394−16395; (b) Yan, X.-X.; Peng, Q.; Zhang, Y.; Zhang, K.; Hong, W.; Hou, X.-L.; Wu, Y.-D. Angew. Chem. Int. Ed. 2006, 45, 1979−1983; (c) Arai, T.; Mishiro, A.; Yokoyama, N.; Suzuki, K.; Sato, H. J. Am. Chem. Soc. 2010, 132, 5338−5339; (d) Arai, T.; Yokoyama, N.; Mishiro, A.; Sato, H. Angew. Chem. Int. Ed. 2010, 49, 7895−7898; (e) Kim, H. Y.; Li, J.Y.; Kim, S.; Oh, K. J. Am. Chem. Soc. 2011, 133, 20750−20753; (f) Conde, E.; Bello, D.; Cózar, A. D.; Sánchez, M.; Vázquez, M. A.; Cossío, F. P. Chem. Sci. 2012, 3, 1486−1491; (g) Narayan, R.; Bauer, J. O.; Strohmann ,C.; Antonchick, A. P.; Waldman, H. Angew. Chem. Int. Ed. 2013, 52, 12892−12896; (h) Castelló, L. M.; Nájera, C.; Sansano, J. M.; Larrañaga, O.; Cózar, A. D.; Cossío, F. P. Org. Lett. 2013, 15, 2902−2905; (i) Bai, X.-F.; Song, T.; Xu, Z.; Xia, C.-G.; Huang, W.-S.; Xu, L.-W. Angew. Chem. Int. Ed. 2015, 54, 5255−5259; (j) Arai, T.; Tokumitsu, C.; Miyazaki, T.; Kuwano, S.; Awata, A. Org. Biomol. Chem. 2016, 14, 1831–1839. (12) (a) Ayerbe, M.; Arrieta, A.; Cossío, F. P.; Linden, A. J. Org. Chem. 1998, 63, 1795−1805; (b) Vivanco, S.; Lecea, B.; Arrieta, A.; Prieto, P.; Morao, I.; Linden, A.; Cossío, F. P. J. Am. Chem. Soc. 2000, 122, 6078−6092; (c) Li, Q.; Ding, C.-H.; Hou, X.-L.; Dai, L.-X.

Org. Lett. 2010, 12, 1080−1083; (d) Imae, K.; Konno, T.; Ogata, K.; Fukuzawa, S. Org. Lett. 2012, 14, 4410−4413; (e) Conde, E.; Rivilla, I.; Larumbe, A.; Cossío, F. P. J. Org. Chem. 2015, 80, 11755−11767. (13) (a) Esquivias, J.; Arrayás, R. G.; Carretero, J. C. Angew. Chem. Int. Ed. 2006, 45, 629−633; (b) Jia, Y.-X.; Xie, J.-H.; Duan, H.-F.; Wang, L.-X.; Zhou, Q.-L. Org. Lett. 2006, 8, 1621−1624.

ACS Paragon Plus Environment

ACS Catalysis

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 6 of 6