You are on page 1of 4

Letter

pubs.acs.org/OrgLett

Functionalized Tetrahydropyridines by Enantioselective Phosphine-


Catalyzed Aza-[4 + 2] Cycloaddition of N‑Sulfonyl-1-aza-1,3-dienes
with Vinyl Ketones
Huamin Wang, Wei Zhou, Mengna Tao, Anjing Hu, and Junliang Zhang*
Shanghai Key Laboratory of Green Chemistry and Chemical Processes, School of Chemistry and Molecular Engineering, East China
Normal University, 3663 North Zhongshan Road, Shanghai 200062, P. R. China
*
S Supporting Information

ABSTRACT: The development of electron-demand disfavored [4 + 2] cycloaddition of two electron-deficient reacting partners
poses a considerable challenge. An enantioselective aza-[4 + 2] cycloaddition of electron-deficient N-sulfonyl-1-aza-1,3-dienes is
possible with vinyl ketones via phosphine catalysis, which provides facile access to a wide range of enantioenriched
trifluoromethylated tetrahydropyridines in up to 97% yield with 97% ee and >20:1 dr. The mechanistic study indicated that this
cycloaddition proceeds via a tandem intermolecular aza-Rauhut−Currier/intramolecular aza-Michael addition reaction.

T etrahydropyridines are an important class of six-


membered nitrogen-containing heterocycles which are
widely found in a variety of natural products and biologically
with alkyl vinyl ketones have been reported by Loh and
Zhong,6a Shi,6b,c and Wu6d by utilizing the chiral phosphine
catalysts derived from natural or unnatural amino acids.
active compounds (Figure 1).1 Meanwhile, tetrahydropyridines However, the development of more efficient new chiral
phosphine catalyst to further expand the range of N-sulfonyl-
1-aza-1,3-butadienes is still desirable.
Meanwhile, the introduction of a CF3 group at a specific site
to a bioactive molecule is considered as a more and more useful
strategy to improve its physicochemical, biological, and
pharmaceutical properties;7 for example, a trifluoromethylated
piperidine was introduced to the janus kinase inhibitors, which
observed a significant enhancement of its pharmaceutical
activity (Figure 1).8 However, the asymmetric synthesis of
trifluoromethylated heterocycles is largely lagging due to the
limited availability of trifluoromethylated substrates and robust
chiral catalysts. Recently, we developed several new phosphine
Figure 1. Bioactive molecules having tetrahydropyridine and catalysts, such as Xiao-Phos, Wei-Phos, and Peng-Phos, which
piperidine cores. have shown good performance in enantioselective Rauhut−
Currier reactions and allylation reactions.9 With these chiral
phosphine catalysts in hand, we wished to examine their
are important organic synthons which can be readily reduced to
performance in the enantioselective aza-[4 + 2] cycladdition.
piperidines.2 Therefore, tremendous efforts for the synthesis of
Herein, we wish to report the first phosphine catalyzed
tetrahydropyridines, especially in an asymmetric manner, have
been undertaken, among which an organocatalytic [4 + 2] asymmetric aza-[4 + 2] cycloadditions of 1-aza-1,3-butadienes
cycloaddition reaction3 has been widely recognized as one of derived from β-fluoroalkylated-α,β-enones with vinyl ketones,
powerful approaches. However, compared to the well- which provide a facile access to valuable chiral fluoroalkylated
established Brønsted acids4a−e and proline4f−h-catalyzed tetrahydropyridines (Scheme 1). Besides β-fluoroalkylated α,β-
enantioselective aza-[4 + 2] cycloaddition reactions, the enone derived imines, the chalcone-derived imines could also
development of other organocatalysts such as chiral phosphine5 furnish the desired products with satisfactory diastereo- and
has been less explored and unbalanced. To date, only a few
examples of asymmetric phosphine-catalyzed intermolecular Received: February 17, 2017
aza-[4 + 2] cycloadditions of N-sulfonyl-1-aza-1,3-butadienes

© XXXX American Chemical Society A DOI: 10.1021/acs.orglett.7b00489


Org. Lett. XXXX, XXX, XXX−XXX
Organic Letters Letter

Scheme 1. Enantioselective Phosphine-Catalyzed Aza [4 + 2] Table 1. Variation of Aza Diene and Aryl Vinyl Ketone
Cycloaddition Components

entry 1, R1/Rf Ar yielda (%) eeb (%)


enantioselectivity. Furthermore, the mechanistic study indi-
cated that this cycloaddition proceeds via a tandem 1 1a, Ph/CF3 2a, Ph 3aa, 92 95
intermolecular aza-Rauhut−Currier/intramolecular aza-Michael 2c 1b, 4-MeC6H4/CF3 2a 3ba, 91 95
addition reaction. 3 1c, 4-MeOC6H4/CF3 2a 3ca, 84 94
Aza diene 1a and vinyl ketone 2a were selected as model 4 1d, 4-PhC6H4/CF3 2a 3da, 94 95
substrates for screening chiral phosphine catalysts (Figure 2). 5 1e, 4-FC6H4/CF3 2a 3ea, 83 91
6 1f, 4-ClC6H4/CF3 2a 3fa, 88 94
7 1g, 4-BrC6H4/CF3 2a 3ga, 90 96
8 1h, 4-IC6H4/CF3 2a 3ha, 92 96
9 1i, 4-NCC6H4/CF3 2a 3ia, 82 92
10 1j, 4-MeO2CC6H4/CF3 2a 3ja, 88 91
11 1k, 4-MeO2SC6H4/CF3 2a 3ka, 62 95
12 1l, 2-FC6H4/CF3 2a 3la, 97 91
13 1m, 2-O2NC6H4/CF3 2a 3ma, 96 85
14 1n, 3-ClC6H4/CF3 2a 3na, 90 95
Figure 2. Phosphine catalysts employed in this study.
15 1o, 3-BrC6H4/CF3 2a 3oa, 92 94
16 1p, 2-naphthyl/CF3 2a 3pa, 92 97
17d 1q, 2-furyl/CF3 2a 3qa, 72 92
To our delight, chiral sulfinamide phosphines (P1 and P2)
18 1r, Ph/CF3 2a 3ra, 93 94
could catalyze the reaction, but low yields and ee values were
19 1s, Ph/CF3 2a 3sa, 79 94
obtained (Table S1, entries 1 and 2). Further screening showed
20 1t, Ph/C2F5 2a 3ta, 65 92
that the amide moiety plays a significant effect on the
21e 1v, cyclopropyl/CF3 2a 3va, 70 82
enantioselectivity; the acetyl- (P3) and p-toluoyl- derived
22 1c 2b, 4-MeC6H4 3cb, 84 92
(P4) amide phosphines delivered relatively lower ee’s (Table
23 1c 2c, 4-MeOC6H4 3cc, 90 92
S1, entries 3 and 4), while phosphine P5 bearing a more acidic
24 1c 2d, 4-PhC6H4 3cd, 90 93
amide moiety derived from 3,5-ditrifluoromethylbenzonic acid
25 1c 2e, 4-FC6H4 3ce, 75 91
turned out to be a promising catalyst, furnishing the desired [4
26 1c 2f, 4-ClC6H4 3cf, 85 90
+ 2] cycloaddition product in 77% yield with 90% ee (Table S1,
27 1c 2g, 4-BrC6H4 3cg, 84 92
entry 5). The introduction of different aryl groups at the ortho-
28 1c 2h, 4-O2NC6H4 3ch, 60 90
position of the phenyl ring, structured as P6 and P7, resulted in
29d 1c 2i, 4-NCC6H4 3ci, 60 92
elevated enantioselectivity and diastereoselectivity (Table S1,
30 1c 2j, 4-CF3C6H4 3cj, 74 90
entries 6 and 7). With the use of the best catalyst P7 as the
31 1c 2k, 2-naphthyl 3ck, 95 90
catalyst, various solvents were then screened, and acetone was
32d 1c 2l, 2-furyl 3cl, 89 93
found to be the best solvent with regard to activity (Table S1, a
entries 8−12). Further increase of the catalyst loading to 7.5 Yield of isolated trans isomer, dr >20:1; determined by 19F NMR
analysis of the crude reaction mixture. bDetermined by HPLC analysis.
mol % led to a slightly higher yield (Table S1, entry 13). A c
10 mol % of P7 was used. d10 mol % of P7 was used at 0 °C. e10 mol
slightly higher ee was observed when the reaction was run at 0 % of P7 was used at −20 °C.
°C but led to lower yield (Table S1, entry 14).
With the optimal reaction conditions in hand, we then
turned to explore the generality of this cycloaddition process entry 21). Moreover, we examined the scope of the
with a variety of 1-aza-1,3-dienes with ketone 2a (Table 1). cycloaddition reaction of 1c with a series of aryl vinyl ketones
Regardless of their electronic nature and the substitution 2. Aryl vinyl ketones with electron-donating groups such as 4-
pattern of the aryl ring, high yields and ee values were furnished Me (2b), 4-OMe (2c), and 4-Ph (2d) on the phenyl ring
(Table 1, entries 1−15). The 2-naphthyl group and heterocyclic exhibited higher enantioselectivity (Table 1, entries 22−24).
ring were also applicable, furnishing the corresponding Halogenated substrates 2e−g were also well tolerated, affording
products 3pa and 3qa in 72−92% yields with 92−97% ee’s the corresponding products 3ce−cg in 75−85% yields with
(Table 1, entries 16 and 17). When the substituent R2 on N- 90−92% ee (Table 1, entries 25−27). Vinyl ketones 2h−j with
sulfonylimine 1 was changed, other groups, including p- electron-withdrawing aromatic rings also successfully afforded
methoxyphenyl and phenyl, were all well tolerated in the the corresponding products 3ch−cj with good to excellent
reactions (Table 1, entries 18 and 19). Changing the fluoroalkyl enantioselectivity (Table 1, entries 28−30). 2-Naphthyl- and 2-
group from a CF3 group to the C2F5 group also worked well to furyl-substituted vinyl ketones afforded the desired products in
give the corresponding product 3ta with 92% ee in 65% yield 90% and 93% ee, respectively (Table 1, entries 31 and 32).
(Table 1, entry 20). Notably, alkyl-substituted β-trifluoroalkyl Next, we turned our attention to the challenge of
α,β-enone derived imine also worked well giving rise to the enantioselective [4 + 2] cycloaddition of alkyl vinyl ketones
fused heterocyclic 3va in acceptable yield with 82% ee (Table 1, with β-fluoroalkylated enone derived imines. For the alkyl vinyl
B DOI: 10.1021/acs.orglett.7b00489
Org. Lett. XXXX, XXX, XXX−XXX
Organic Letters Letter

ketones, catalyst P6 was identified as the best catalyst, and the Scheme 2. Gram-Scale Synthesis of 3a and Further Synthetic
substrate scope was then examined (Table S2). In general, the Transformationsa
reaction worked quite well, delivering the corresponding
products with excellent ee values and high dr values. Similar
to the reaction with aryl vinyl ketone, in general, β-
fluoroalkylated enone derived imines with electronically and
sterically different aromatic groups were also examined, and the
reaction could proceed smoothly to afford the desired products
(3bn−fo) with high stereoselectivity (Table S2, entries 1−13).
The absolute configuration of 3dn was established by single-
crystal X-ray diffraction analysis10 (see the SI).
The enantioselective cycloaddition between chalcone-derived
imines (4) and methyl vinyl ketone (2n) poses considerable
challenges due to their relatively lower reactivity.7a,7d Inspired a
Gram-scale reaction and synthetic applications. Reaction conditions:
by above results and with diverse catalysts in hand, we made (a) P7 (5 mol %), 1a (2.8 mmol), 2a (1.5 equiv), acetone (0.1 M), rt,
our efforts to this cycloaddition. Gratifyingly, as shown in Table 88% yield, 94% ee, dr > 20:1; (b) NaBH4, 0 °C, 90% yield, 90% ee, dr
2, the corresponding cycloadducts 5an−fn could be obtained in > 20:1; (c) RuCl3, NaIO4, 0 °C, 75% yield, 92% ee, dr > 20:1; (d)
TFA/thioanisole, rt, 70% yield, 93% ee, dr > 20:1; (e) BF3·Et2O,
Table 2. [4 + 2] Cycloaddition of Methyl Vinyl Ketones with Et3SiH, 70 °C, 80% yield, 94% ee, dr > 20:1; (f) NaCNBH3, MeOH/
Chalcone-Derived Iminesa AcOH, rt, 94% ee, dr >20:1.

idine 6d was also obtained in good yield by chemoselective


reduction the olefin moiety of 3aa without any reduction of the
carbonyl group.
In summary, we have developed an enantioselective
entry R1 dr yieldb (%) eec (%) phosphine-catalyzed aza-[4 + 2] cycloaddition of N-sulfonyl-
1 4a, Ph 13:1 5an, 70 90
1-aza-1,3-dienes with vinyl ketones, which provides a facile
2 4b, 4-MeoC6H4 12:1 5bn, 65 90
access to various functional trifluoromethylated tetrahydropyr-
3 4c, 4-FC6H4 13:1 5cn, 60 90
idines in moderate to good yields with high diastereo- and
4 4d, 4-ClC6H4 10:1 5dn, 60 90 enantioselectivities. Mechanistic study11 indicated that this aza-
5 4e, 4-BrC6H4 13:1 5en, 63 92 [4 + 2] cycloaddition proceeded via a tandem aza-Rauhut−
6 4f, 4-CF3C6H4 12:1 5fn, 67 90 Currier/intramolecular aza-Michael addition reaction. Further-
7 4g, perfluorophenyl >20:1 5gn, 75 92 more, the phosphonium enolate intermediate12 plays a dual
a role, acting as a key intermediate for the first aza-Rauhut−
Unless otherwise stated, reactions were performed by using 0.1 mmol
of 4, 0.3 mmol of 2n, 10 mol % of P2, and 30 mol % of 2-chlorophenol
Currier reaction and a real catalyst for the second aza-Michael
in 1.0 mL of CHCl3 at 0 °C for 72 h. bYield of isolated cycloadduct. addition step. Further efforts toward the application of these
c
Determined by HPLC analysis. novel multifunctional phosphine catalysts for other asymmetric
transformations are currently underway in our group and will
be reported in due course.


60−70% yields with 90−92% ee’s regardless of the substituent,
from electron-donating group MeO to weak electron-with-
ASSOCIATED CONTENT
drawing groups such as F, Cl, and Br and even strong electron-
withdrawing group CF3 with the chiral sulfinamide phosphine * Supporting Information
S

P2 as the catalyst, indicating that this type of cycloaddition is The Supporting Information is available free of charge on the
sensitive to the structure of the substrate and an appropriate ACS Publications website at DOI: 10.1021/acs.or-
catalyst must be correctly chosen. In addition, pentafluoro- glett.7b00489.
substituted N-tosyl ketimine gave the product 5gn with 75% Experimental procedures; spectroscopic data for the
yield and 92% ee (Table 2, entry 7). substrates and products (PDF)
In order to display the potential synthetic applications of this X-ray data for (−)-3dn (CIF)


methodology, a gram-scale reaction of 1a and 2a was also
carried out to furnish 1.2 g of the desired product 3aa in 88% AUTHOR INFORMATION
yield and 94% ee with the use of 5 mol % of P7 as the catalyst
Corresponding Author
(Scheme 2). The selective reduction of carbonyl group
provided effective access to valuable chiral alcohol 6a in 90% *E-mail: jlzhang@chem.ecnu.edu.cn.
yield as a single diastereomer (>20:1 dr) and 90% ee. The ORCID
diastereoselective dihydroxylation of the olefin moiety with Junliang Zhang: 0000-0002-4636-2846
RuCl3/NaIO4 delivered a functionalized piperidine 6b in 75%
yield with 92% ee. The deprotection of tosyl group was realized Notes
The authors declare no competing financial interest.


by treatment of 3aa with TFA/thioanisole, and the imine
product 6c was obtained as single diastereomer in 70% yield
with 93% ee and trifluoromethylated piperidine 7c was ACKNOWLEDGMENTS
furnished in 90% yield with 94% ee by subsequent reduction We are grateful to the 973 Program (2015CB856600), the
with sodium cyanoborohydride. The enantioenriched piper- National Natural Science Foundation of China (21425205,
C DOI: 10.1021/acs.orglett.7b00489
Org. Lett. XXXX, XXX, XXX−XXX
Organic Letters Letter

21672067), and the Changjiang Scholars and Innovative (6) (a) Shi, Z.; Yu, P.; Loh, T.-P.; Zhong, G. Angew. Chem., Int. Ed.
Research Team in University (PCSIRT) for financial support. 2012, 51, 7825. (b) Zhang, X.; Chen, G.; Dong, X.; Wei, Y.; Shi, M.


Adv. Synth. Catal. 2013, 355, 3351. (c) Zhang, X. N.; Dong, X.; Wei,
REFERENCES Y.; Shi, M. Tetrahedron 2014, 70, 2838. (d) Wang, G.; Rexiti, R.; Sha,
F.; Wu, X. Tetrahedron 2015, 71, 4255.
(1) (a) Nantermet, P. G.; Barrow, J. C.; Selnick, H. G.; Homnick, C. (7) For leading references, see: (a) Fujita, T.; Iwasa, J.; Hansch, C. J.
F.; Freidinger, R. M.; Chang, R. S. L.; O’Malley, S. S.; Reiss, D. R.; Am. Chem. Soc. 1964, 86, 5175. (b) Leo, A.; Hansch, C.; Elkins, D.
Broten, T. P.; Ransom, R. W.; Pettibone, D. J.; Olah, T.; Forray, C. Chem. Rev. 1971, 71, 525. (c) Katritzky, A. R.; Lobanov, V. S.;
Bioorg. Med. Chem. Lett. 2000, 10, 1625. (b) Michael, J. P. Nat. Prod. Karelson, M. Chem. Soc. Rev. 1995, 24, 279. (d) Huang, Y. Y.; Yang, X.;
Rep. 2004, 21, 625. (c) Maison, W. Pipecolic Acid Derivatives. In Chen, Z.; Verpoort, F.; Shibata, N. Chem. - Eur. J. 2015, 21, 8664.
Highlights in Bioorganic Chemistry; Wiley-VCH: New York, 2004; p 18. (8) For janus kinase inhibitors, see: (a) Childers, M. L.; Fuller, P.;
(d) Kawagishi, F.; Toma, T.; Inui, T.; Yokoshima, S.; Fukuyama, T. J. Guerin, D.; Katz, J. D.; Pu, Q.; Scott, M. E.; Thompson, C. F.;
Am. Chem. Soc. 2013, 135, 13684. (e) Lei, B. L.; Zhang, Q. S.; Yu, W. Martinez, M.; Falcone, D.; Torres, L.; Deng, Y.; Kuruklasuriya, R.;
H.; Ding, Q. P.; Ding, C. H.; Hou, X. L. Org. Lett. 2014, 16, 1944. Zeng, H.; Bai, Y.; Kong, N.; Liu, Y.; Zheng, Z. WO Patent
(f) Aridoss, G.; Amirthaganesan, S.; Jeong, Y. T. Bioorg. Med. Chem. 2014146491, A1, 2014. (b) Dinsmore, C.; Fuller, P.; Guerin, D.;
Lett. 2010, 20, 2242. Katz, J. D.; Thompson, C. F.; Falcone, D.; Deng, W.; Torres, L.; Zeng,
(2) (a) Rubiralta, M.; Giralt, E.; Diez, A. Piperidine: Struture, H.; Bai, Y.; Fu, J.; Kong, N.; Liu, Y.; Zheng, Z. WO Patent
Preparation, Reactivity, and Synthetic Applications of Piperidine and Its 2014146493, A1, 2014.
Derivatives; Elsevier: New York, 1991. (b) Hwang, Y. C.; Fowler, F. W. (9) (a) Su, X.; Zhou, W.; Li, Y.; Zhang, J. Angew. Chem., Int. Ed. 2015,
J. Org. Chem. 1985, 50, 2719. (c) Michael, J. P. Nat. Prod. Rep. 2004, 54, 6874. (b) Zhou, W.; Su, X.; Tao, M.; Zhu, C.; Zhao, Q.; Zhang, J.
21, 625. (d) Sahn, J. J.; Su, J. Y.; Martin, S. F. Org. Lett. 2011, 13, 2590. Angew. Chem., Int. Ed. 2015, 54, 14853. (c) Zhou, W.; Chen, P.; Tao,
(e) Sahn, J. J.; Martin, S. F. Tetrahedron Lett. 2011, 52, 6855. (f) Jian, M.; Su, X.; Zhao, Q.; Zhang, J. Chem. Commun. 2016, 52, 7612.
T.; Sun, L.; Ye, S. Chem. Commun. 2012, 48, 10907. (d) Chen, P.; Yue, Z.; Zhang, J.; Lv, X.; Wang, L.; Zhang, J. Angew.
(3) For a review of [4 + 2] cycloadditions, see: (a) Klier, L.; Tur, F.; Chem., Int. Ed. 2016, 55, 13316. (e) Zhou, W.; Gao, L.; Tao, M.; Su,
Poulsen, P. H.; Jørgensen, K. A. Chem. Soc. Rev. 2017, 46, 1080. X.; Zhao, Q.; Zhang, J. Huaxue Xuebao 2016, 74, 800.
(b) Jeon, B.-S.; Wang, S.-A.; Ruszczycky, M. W.; Liu, H.-W. Chem. Rev. (10) The X-ray crystal structure information is available at the
2016, DOI: 10.1021/acs.chemrev.6b00578. (c) Masson, G.; Lalli, C.; Cambridge Crystallographic Data Centre (CCDC) under deposition
Benohoud, M.; Dagousset, G. Chem. Soc. Rev. 2013, 42, 902. no. CCDC 1500207 ((−)-3dn).
(4) For selected examples of chiral Brønsted acids catalyed by aza-[4 (11) Details of experiments and results are summarized in the
+ 2] cycloaddition, see: (a) Dagousset, G.; Retailleau, P.; Masson, G.; Supporting Information.
Zhu, J. Chem. - Eur. J. 2012, 18, 5869. (b) Dagousset, G.; Zhu, J.; (12) Some examples for phosphonium enolate as Brønsted base:
Masson, G. J. Am. Chem. Soc. 2011, 133, 14804. (c) Xu, A. H.; Zuend, (a) Wang, H.-Y.; Zhang, K.; Zheng, C.-W.; Chai, Z.; Cao, D.-D.;
S. J.; Woll, M. G.; Tao, Y.; Jacobsen, E. N. Science 2010, 327, 986. Zhang, J.-X.; Zhao, G. Angew. Chem., Int. Ed. 2015, 54, 1775. (b) Lou,
(d) Liu, H.; Dagousset, G.; Masson, G.; Retailleau, P.; Zhu, J. J. Am. Y.-P.; Zheng, C.-W.; Pan, R.-M.; Jin, Q.-W.; Zhao, G.; Li, Z. Org. Lett.
Chem. Soc. 2009, 131, 4598. (e) Akiyama, T.; Morita, H.; Fuchibe, K. J. 2015, 17, 688. (c) Wang, H.-Y.; Zheng, C.-W.; Chai, Z.; Zhang, J.-X.;
Am. Chem. Soc. 2006, 128, 13070. For recent examples of proline- Zhao, G. Nat. Commun. 2016, 7, 12720.
catalyed aza-[4 + 2] cycloaddition, see: (f) Han, B.; He, Z.-Q.; Li, J.-L.;
Li, R.; Jiang, K.; Liu, T.-Y.; Chen, Y.-C. Angew. Chem., Int. Ed. 2009,
48, 5474. (g) Sundén, H.; Ibrahem, I.; Eriksson, L.; Córdova, A. Angew.
Chem., Int. Ed. 2005, 44, 4877. (h) Li, J.-L.; Liu, T.-Y.; Chen, Y.-C. Acc.
Chem. Res. 2012, 45, 1491.
(5) For reviews related to phosphines catalysis, see: (a) Ye, L.-W.;
Zhou, J.; Tang, Y. Chem. Soc. Rev. 2008, 37, 1140. (b) Cowen, B. J.;
Miller, S. J. Chem. Soc. Rev. 2009, 38, 3102. (c) Wei, Y.; Shi, M. Acc.
Chem. Res. 2010, 43, 1005. (d) Wang, S.-X.; Han, X.; Zhong, F.; Wang,
Y.; Lu, Y. Synlett 2011, 2011, 2766. (e) Zhao, Q.-Y.; Lian, Z.; Wei, Y.;
Shi, M. Chem. Commun. 2012, 48, 1724. (f) Wei, Y.; Shi, M. Chem.
Rev. 2013, 113, 6659. (g) Fan, Y.; Kwon, O. Chem. Commun. 2013, 49,
11588. (h) Xie, P.; Huang, Y. Eur. J. Org. Chem. 2013, 2013, 6213.
(i) Wang, Z.; Xu, X.; Kwon, O. Chem. Soc. Rev. 2014, 43, 2927.
(j) Voituriez, A.; Marinetti, A.; Gicquel, M. Synlett 2015, 26, 142.
(k) Xie, P.; Huang, Y. Org. Biomol. Chem. 2015, 13, 8578. (l) Li, W.;
Zhang, J. Chem. Soc. Rev. 2016, 45, 1657. (m) Wang, T.; Han, X.;
Zhong, F.; Yao, W.; Lu, Y. Acc. Chem. Res. 2016, 49, 1369. For
selected examples since 2015 because of space limitations, see: (n) Gu,
Y.; Hu, P.; Ni, C.; Tong, X. J. Am. Chem. Soc. 2015, 137, 6400.
(o) Zhang, L.; Liu, H.; Qiao, G.; Liu, Y.; Xiao, Y.; Guo, H.; Hou, Z. J.
Am. Chem. Soc. 2015, 137, 4316. (p) Lee, S.; Fujiwara, Y.; Nishiguchi,
A.; Kalek, M.; Fu, G. J. Am. Chem. Soc. 2015, 137, 4587. (q) Yao, W.;
Dou, X.; Lu, Y. J. Am. Chem. Soc. 2015, 137, 54. (r) Dong, X.; Liang,
L.; Li, E.; Huang, Y. Angew. Chem., Int. Ed. 2015, 54, 1621. (s) Ziegler,
D.; Fu, G. J. Am. Chem. Soc. 2016, 138, 12069. (t) Wang, T.; Yu, Z.;
Hoon, D.; Phee, C.; Lan, Y.; Lu, Y. J. Am. Chem. Soc. 2016, 138, 265.
(u) Li, E.; Jin, H.; Jia, P.; Dong, X.; Huang, Y. Angew. Chem., Int. Ed.
2016, 55, 11591. (v) Sankar, M. G.; Castro, M. G.; Golz, C.;
Strohmann, C.; Kumar, K. Angew. Chem., Int. Ed. 2016, 55, 9709.
(w) Han, X.; Chan, W.-L.; Yao, W.; Wang, Y.; Lu, Y. Angew. Chem., Int.
Ed. 2016, 55, 6492. (x) Cai, L.; Zhang, K.; Kwon, O. J. Am. Chem. Soc.
2016, 138, 3298. (y) Satpathi, B.; Ramasastry, S. S. V. Angew. Chem.,
Int. Ed. 2016, 55, 1777.

D DOI: 10.1021/acs.orglett.7b00489
Org. Lett. XXXX, XXX, XXX−XXX

You might also like