You are on page 1of 5

DOI: 10.1002/ajoc.

201402219

Cycloaddition

Communication

| Very Important Paper |

Enantioselective Cycloadditions of Vinyl Cyclopropanes and


Nitroolefins for Functionally and Optically Enriched
Nitrocyclopentanes
Wen-Ke Li, Ze-Shui Liu, Long He,* Tai-Ran Kang,* and Quan-Zhong Liu*[a]
Abstract: Palladium-catalyzed enantioselective cycloaddition of vinyl cyclopropane dicarbonitriles and nitroolefins
was achieved to generate structurally and optically enriched nitrocyclopentanes with three consecutive chiral
carbon centers in up to 96 % yield and with up to 92 % ee.
The two diastereoisomers produced in each reaction are
readily separable by simple chromatography. The present
method allows the preparation of both diastereoisomers
in optically enriched form.
Scheme 1. Formal (3+2) cycloadditions of nitroolefins and vinyl cyclopropanes.

The rapid and efficient assembly of optically enriched cyclopentanes with multiple chiral centers has received much attention from the organic and medicinal chemistry communities.[1]
A straightforward approach would be a transition-metal-catalyzed asymmetric formal (3+2)-cycloaddition between a 1,3dipole and an olefin; it would rapidly construct structurally
and optically enriched cyclopentanes. In this regard, allenoates
and trimethylenemethanes are precursors for versatile threecarbon fragments which are widely used in (3+2)-cycloadditions to form structurally enriched cyclopentanes.[2] Also, racemic malonate-derived donor-acceptor (D-A) cyclopropanes
have received much attention over the past decade for their
synthetic utility.[3] For example, Johnson and co-workers reported an elegant dynamic kinetic asymmetric (3+2) cycloaddition
of racemic cyclopropanes with aldehydes or imines.[4] Tang and
co-workers also described a highly enantioselective cyclopentannulation of indoles with donor-acceptor cyclopropanes
using copper catalysts.[5] Vinyl cyclopropanes are readily activated by transition-metal catalysts such as Pd,[6] Ir,[7] Rh,[8] Ru,[9]
Ni,[10] and Fe[11] to form the corresponding zwitterionic complex. The in situ-generated 1,3-dipoles are reactive in the
formal (3+2)-cycloaddition with electron-deficient olefins. In
this context, Goldberg and Stoltz reported a palladium-catalyzed diastereoselective (3+2) cycloaddition of vinyl cyclopro-

panes and b-2-dinitrostyrenes (Scheme 1).[12] After that, the Shi


and Trost groups reported chiral-palladium-catalyzed enantioselective (3+2)-cycloadditions of vinyl cyclopropanes and electron-deficient alkenes.[13] Nitroolefins are versatile dipolarophiles in (3+2)-cycloadditions that provide valuable nitrogencontaining products,[14] which can be easily converted into medicinally and synthetically important compounds via the Nef
reaction, reduction, and other transformations.[15] However, to
date, the enantioselective cycloaddition of vinyl cyclopropanes
and nitroolefins has been unsuccessful, and furthermore, the
enantioselective transformation of vinyl cyclopropanes is still illusive, with only few examples reported.[13, 16]
The enantioselective cycloaddition of vinyl cyclopropane derivatives and nitroolefins would allow the formation of optically active nitrocyclopentanes, a precursor of cyclopentyl amines,
which are well-represented in bioactive compounds[17] and are
also at the core of numerous bioactive natural products, such
as those shown in Figure 1.[18]

[a] W.-K. Li, Z.-S. Liu, Prof. Dr. L. He, Prof. Dr. T.-R. Kang, Prof. Dr. Q.-Z. Liu
Chemical Synthesis and Pollution Control Key Laboratory of
Sichuan Province and College of Chemistry and Chemical Engineering
China West Normal University
Nanchong 637000 (P. R. China)
E-mail: longhe@cwnu.edu.cn
quanzhongliu@cwnu.edu.cn
Supporting information for this article is available on the WWW under
http://dx.doi.org/10.1002/ajoc.201402219.
Asian J. Org. Chem. 2015, 4, 28 32

Figure 1. Formal (3+2) cycloadditions of nitroolefins and vinyl cyclopropanes.

28

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Communication
cyclopentane 3 a was isolated as the major isomer in 85 %
total yield and with 5:1 diastereoselectivity after stirring at
room temperature for 24 h (entry 1, Table 1). Encouraged by
this results, we investigated the reaction with several commercially available chiral phosphine ligands, as shown in Figure 2.
Although reactions with various chiral ligands surveyed afforded the expected cycloadducts in varying yields and selectivity (entries 211, Table 1), the steric hindrance of chiral ligands has a significant effect on the reactivity and enantioselectivity of the cycloaddition reaction. The ligands with bulky
substituents provided poorer results. For example, the reaction
with Me-Duphos (L2) afforded 3 a as a major isomer in 95 %
total yield although only low enantioselectivity (38 % ee,
entry 2). On the other hand, bulky ligand L3 afforded better selectivity (16:1 d.r. and 60 % ee) but only 78 % yield was obtained (entry 3). iPr-Duphos (L4) did not work at all (entry 4).
Other chiral diphenylphosphinobutanes were also investigated,
and (R,R)-2,3-diphenylphosphinobutane (L5) and DIOP (L6) gave
poor enantioselectivity (entries 5 and 6). The reactions with
BINAP and its derivatives (L7L9) delivered excellent yields and
moderate to good enantioselectivity (entries 79), but almost
the same diastereoselectivity (1.1:11.4:1). Among the ligands
investigated, (S)-MeO-BIPHEP (L10) gave the highest enantioselectivity (80 and 89 % ee for the two diastereomers 3 a and 3 a,
respectively), although the d.r. value remained 1.4:1. Fortunately, the two diastereoisomers could be easily separated by
simple chromatography. Chiral phophororamdite L11 was not
suitable and low ees were obtained. In terms of enantioselectivity and yield, (S)-MeO-BIPHEP (L10) was the best ligand.
The reaction was further optimized by investigating the effects of solvents, temperature, and the results are summarized
in Table 2. A solvent screening indicated that aprotic solvents
afforded the corresponding cycloadduct 3 a in good to excellent yields and with 5091 % ee. Because the two diastereoisomers can be readily separated
by simple chromatography, the
diastereomeric ratio was not
considered a selection criterion.
Among the solvents surveyed,
diethyl ether provided the same
diastereoselectivity as THF, but
THF afforded the best enantioselectivity (entries 13, Table 2).
Toluene was the best solvent,
which provided 3 as the major
isomer in 93 % total yield and
with 82 % and 91 % ee for the
two stereoisomers (entries 48).
Chlorinated solvents provided
low
diastereoselectivity,
although moderate enantioselectivity was achieved (entries 9
11). Polar aprotic solvents such
as DMSO and acetonitrile provided better diastereomeric ratios.
These results may be from the
interaction of the polar solvent

Table 1. The effects of ligands on the cycloaddition of 1 a and 2 a.[a]

Entry

Ligand

Yield [%]

d.r.
3 a:3 a[b]

ee [%]
3 a, 3 a[c]

1
2
3
4
5
6
7
8
9
10
11

L1
L2
L3
L4
L5
L6
L7
L8
L9
L10
L11

85
95
78
<5
74
90
87
97
97
92
85

5:1
7.7:1
16:1

6.8:1
1:1.2
1:1.1
1:1.4
1:1.1
1:1.4
1:0.95

38, 15
60, 25

11, 17
13, 22
75, 85
40, 43
79, 87
80, 89
35, 43

[a] The reaction was carried out at 0.1 mmol scale using the chiral ligand
(10 mol %) and [Pd(dba)2] as the catalyst in anhydrous THF. [b] d.r. values
were determined by 1H NMR spectroscopy of crude products. [c] The ees
were determined by HPLC after each diastereoisomer was obtained.

With this background, we commenced our investigation of


the reaction of b-nitrostyrene 1 a and vinyl cyclopropane 2 a
(Table 1) catalyzed by bis(dibenzylideneacetone)palladium(0)
([Pd(dba)2]) and achiral 1,2-bis(diphenylphosphino) benzene
(L1) as the ligand (Figure 2). To our delight, the expected nitro-

Figure 2. The ligands surveyed in this work.


Asian J. Org. Chem. 2015, 4, 28 32

www.AsianJOC.org

29

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Communication
Table 2. Optimization of the reaction.[a]

Table 3. Substrate scope.[a]

Entry

Solvent

Yield [%]

d.r.
3 a:3 a

ee [%]
3 a, 3 a

Entry

R1

1
2
3
4
5
6
7
8
9
10
11
12
13
14[b]
15[c]
16[d]
17[e]

THF
1,4-dioxane
diethyl ether
toluene
benzene
o-xylene
PhCF3
PhCN
DCM
CHCl3
DCE
DMSO
CH3CN
toluene
toluene
toluene
toluene

92
88
90
93
91
87
92
95
90
89
88
90
80
98
93
88
93

1:1.1
1:1.3
1:1.1
1:1.4
1:1.3
1:1.2
1:1
1:0.5
1:0.9
1:1
1:0.8
1:0.3
1:0.4
1:1.4
1:1.4
1:1
1:1.4

80,
79,
75,
82,
73,
70,
79,
53,
70,
65,
77,
70,
75,
87,
81,
39,
73,

1
2
3
4
5
6
7
8
9
10
11
12
13

CN
CN
CN
CN
CN
CN
CN
CN
CN
CN
CN
CN
CN

14
15
16

CN (1 a)
CN (1 a)
CO2Me
(1 b)
SO2Ph (1 c)

89
89
85
91
85
80
85
50
68
65
85
50
70
89
87
73
87

[a] The reaction was carried out at 0.1 mmol scale at room temperature
in the solvents indicated, in all cases, 2-vinylcyclopropane-1,1-dicarbonitrile (1 equiv.) was used and [Pd(dba)2] (5 mol %) and chiral ligand L10
(10 mol %) were used as the catalyst. The diastereoselectivity was determined by 1H NMR spectroscopy spectra of crude products and the ee
values of each diastereoisomer were determined by HPLC after purification by chromatography. [b] 10 mol % of [Pd(dba)2] and 10 mol % of L10
were used. [c] 2.5 mol % of [Pd(dba)2] and 5 mol % of L10 used. [d] Run at
0 8C. [e] Run at 50 8C. DCE = 1,2-dichloroethane; DCM = dichloromethane;
DMSO = dimethyl sulfoxide; THF = tetrahydrofuran.

17

www.AsianJOC.org

Product Yield
[%]

d.r.
3:3[b]

ee [%]
3, 3[c]

Ph (2 a)
4-FPh (2 b)
4-ClPh (2 c)
4-BrPh (2 d)
4-MeOPh (2 e)
4-MePh (2 f)
3-MeOPh (2 g)
3-BrPh (2 h)
3-ClPh (2 i)
2-MeOPh (2 j)
2-NO2Ph (2 k)
1-naphthyl (2 l)
2-thiophenyl
(2 m)
2-furyl (2 n)
tBu (2 o)
Ph (2 a)

3a
3b
3c
3d
3e
3f
3g
3h
3i
3j
3k
3l
3m

93
96
84
79
90
96
96
75
70
97
0
60
95

1:1.4
1:1.4
1:1.3
1:1.1
1:1.2
1:1.4
1:1.3
1:1.4
1:1.4
1:1.6

1:0.9
1:1.2

82,
80,
77,
60,
79,
89,
84,
69,
70,
60,

76,
77,

3n
3o
3p

89
56
49

1:1.2
1;1.5
1:0.92

77, 85
73, 89
17, 27

Ph (2 a)

3q

91
89
87
67
82
92
70
70
80
76
65
87

[a] The reaction was carried out at 0.1 mmol scale at room temperature
in toluene. In all cases, 2-vinylcyclopropane-1,1-dicarbonitrile (1 equiv.)
and nitroolefin (1.1 equiv.) was used and [Pd(dba)2] (5 mol %) and chiral
ligand L10 (10 mol %) were used as the catalyst. [b] The diastereoselectivity was determined by 1H NMR spectroscopy of crude products. [c] ee
values of each diastereoisomer were determined by HPLC after purification by chromatography.

enes were also appropriate substrates affording the cycloaddition products in good to high yields with moderate enantioselectivity (entries 79).
The reaction of a nitroolefin with an ortho-electron-donating
group also delivered good yields and moderate enantioselectivity (entry 10). However, a substrates containing a strong
electron-withdrawing was not reactive (entry 11). The presence
of heteroaromatic rings such as thiophenyl (entry 13) and furyl
(entry 14) is well tolerated under the present conditions, but
gave lower enantioselectivity compared with a phenyl-substituted nitroolefin. It should be noted that the reaction of an
alkyl-substituted nitroolefin 2 o provided good enantioselectivity, although only 56 % yield was obtained (entry 15).
Finally, several vinyl cyclopropane derivatives were also investigated. The yield and enantioselectivity of the reaction of
vinyl cyclopropane dicarboxylic acid methyl ester 1 b were not
satisfactory (entry 16). (2-vinylcyclopropane-1, 1-disulfonyl)dibenzene 1 c was not the suitable substrate (entry 17). In our investigation, the diastereoselectivity was disappointing; howev-

and the nitro functionality of nitroolefin 2 a (entries 12 and 13).


Lower ee values were obtained when the [Pd(dba)2] loading
was reduced to 2.5 mol % or increased to 10 mol %, with no
yield reduction (entries 14 and 15). Performing the reaction at
a lower temperature or increasing to 50 8C was unsuccessful
for achieving better results (entries 16 and 17).
With the optimal reaction conditions established, various
trans-b-nitrostyrene derivatives 2 were surveyed with vinyl cyclopropane dicarbonitrile 1 a. As summarized in Table 3,
among the nitroolefins that were screened, the yield and enantioselectivity depended greatly on the electronic nature of substituents attached to the aromatic ring of the nitroolefins. Various trans-b-nitrostyrenes bearing electron-donating and electron-withdrawing groups at the para position reacted smoothly to afford cycloaddition products in high yields and with
moderate to good enantioselectivity (entries 26). para-Methylsubstituted nitroolefin 2 f afforded the best enantioselectivity
(89 and 92 % ee, entry 6). meta-Substituted trans-b-nitrostyrAsian J. Org. Chem. 2015, 4, 28 32

(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)
(1 a)

R2

30

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Communication
2.78 ppm (m, 1 H). 13C NMR (100 MHz, CDCl3): d = 131.00, 130.29,
129.94, 129.71, 128.07, 122.41, 113.80, 113.29, 90.92, 58.62, 45.42,
41.89, 41.27 ppm. IR (KBr): g = 3438, 2922, 2255, 1639, 1457, 1371,
739, 695 cm1. 3 a: White solid, aD25:7 = 7.48 (c = 0.06, CH2Cl2),
m.p. 110 8C, 91 % ee determined by HPLC analysis (Chiralpak AD-H
column, 5 % IPA in n-hexane, flow rate: 1.2 mL min1, 210 nm); retention time: t (major) = 17.53 min, t (minor) = 15.02 min. 1H NMR
(400 MHz, CDCl3): d = 7.607.34 (m, 5 H), 5.96 (ddd, J = 17.0, 10.2,
8.0 Hz, 1 H), 5.405.33 (m, 2 H), 5.24 (dd, J = 11.0, 9.1 Hz, 1 H), 4.39
(d, J = 11.1 Hz, 1 H), 3.58 (dt, J = 17.7, 8.8 Hz, 1 H), 3.13 (dd, J = 14.4,
10.2 Hz, 1 H), 2.70 ppm (dd, J = 14.4, 7.7 Hz, 1 H). 13C NMR (100 MHz,
CDCl3): d = 133.53, 130.48, 129.77, 129.70, 127.98, 120.75, 114.38,
113.49, 90.44, 58.27, 46.77, 41.17, 39.09 ppm. IR (KBr): g = 3440,
2924, 2255, 1640, 1555, 1372, 714 cm1. HRMS (ESI): m/z: calcd for
C15H13N3O2 : 268.1080 [M+H]; found: 268.1071.

er, the two stereoisomers for each reaction can be easily separated by simple column chromatography. This allows the present method to be practical in organic synthesis.
The ee values of the cycloadduct were further increased by
simple recrystallization from a mixed solvent of n-hexane and
ethyl acetate. For example, after recrystallization, the optically
pure isomers 3 d and 3 d (> 99 % ee) were thus obtained as
single crystals and their absolute configurations were determined by X-ray diffraction (Figure 3 and see the Supporting Information).[19] The absolute configurations of other cycloadducts were tentatively assigned by comparison.

Acknowledgements
This work was financially supported by the National Natural
Science Foundation of China (NSFC) (21102116) and the Innovative Research Team in College of Sichuan Province (grant no.
14TD0016).
Keywords: cycloaddition nitroalkenes nitrocyclopentanes
palladium vinyl cyclopropane dicarbonitriles
[1] a) M. Lautens, W. Klute, W. Tam, Chem. Rev. 1996, 96, 49; b) T. Hudlicky,
J. D. Price, Chem. Rev. 1989, 89, 1467; c) C. E. Masse, J. S. Panek, Chem.
Rev. 1995, 95, 1293; d) G. Helmchen, M. Ernst, G. Paradies, Pure Appl.
Chem. 2004, 76, 495; e) R. D. Little in Comprehensive Organic Synthesis,
Vol. 5 (Eds.: B. M. Trost, I. Fleming), Pergamon, New York, 1991, pp. 239.
[2] For recent examples with allenoates as a three-carbon fragment, see
a) H. Xiao, Z. Chai, C. Zheng, Y. Yang, W. Liu, J. Zhang, G. Zhao, Angew.
Chem. Int. Ed. 2010, 49, 4467 4470; Angew. Chem. 2010, 122, 4569
4572; b) J. E. Wilson, G. C. Fu, Angew. Chem. Int. Ed. 2006, 45, 1426
1429; Angew. Chem. 2006, 118, 1454 1457; c) B. J. Cowen, S. J. Miller, J.
Am. Chem. Soc. 2007, 129, 10988; For recent examples utilizing trimethylenemethane as a three-carbon fragment, see d) B. M. Trost, N.
Cramer, S. M. Silverman, J. Am. Chem. Soc. 2007, 129, 12396; e) B. M.
Trost, J. P. Stambuli, S. M. Silverman, U. Schworer, J. Am. Chem. Soc.
2006, 128, 13328; f) H. M. L. Davies, B. Xiang, N. Kong, D. G. Stafford, J.
Am. Chem. Soc. 2001, 123, 7461.
[3] For D-A cyclopropane reviews, see: a) H.-U. Reissig, Top. Curr. Chem.
1988, 144, 73; b) H. N. C. Wong, M. Y. Hon, C. W. Tse, Y. C. Yip, J. Tanko, T.
Hudlicky, Chem. Rev. 1989, 89, 165; c) H.-U. Reissig, R. Zimmer, Chem.
Rev. 2003, 103, 1151; For Lewis catalyzed non asymmetric reaction of DA cyclopropanes, see: d) P. D. Pohlhaus, J. S. Johnson, J. Org. Chem.
2005, 70, 1057; e) M. Zhu, J. Liu, J. Yu, L. Chen, C. Zhang, L. Wang, Org.
Lett. 2014, 16, 1856.
[4] a) A. T. Parsons, J. S. Johnson, J. Am. Chem. Soc. 2009, 131, 3122; b) A. T.
Parsons, A. G. Smith, A. Neel, J. S. Johnson, J. Am. Chem. Soc. 2010, 132,
9688; c) P. D. Pohlhaus, S. D. Sanders, A. T. Parsons, W. Li, J. S. Johnson,
J. Am. Chem. Soc. 2008, 130, 8642.
[5] a) D. B. England, T. D. O. Kuss, R. G. Keddy, M. A. Kerr, J. Org. Chem. 2001,
66, 4704; b) H. Xiong, H. Xu, S. Liao, Z. Xie, Y. Tang, J. Am. Chem. Soc.
2013, 135, 7851.
[6] a) S. Sebelius, V. J. Olsson, O. A. Wallner, K. J. Szab, J. Am. Chem. Soc.
2006, 128, 8150; b) K. Burgess, J. Org. Chem. 1987, 52, 2046; c) I. Shimizu, Y. Ohashi, J. Tsuji, Tetrahedron Lett. 1985, 26, 3825.
[7] J. Moran, A. G. Smith, R. M. Carris, J. S. Johnson, M. Krische, J. Am. Chem.
Soc. 2011, 133, 18618.
[8] P. A. Wender, C. O. Husfeld, E. Langkopf, J. A. Love, J. Am. Chem. Soc.
1998, 120, 1940.
[9] B. M. Trost, F. D. Toste, H. Shen, J. Am. Chem. Soc. 2000, 122, 2379.

Figure 3. The absolute configuration of cycloadducts 3 d and 3 d.

In summary, we have described an enantioselective cycloaddition of vinyl cyclopropane dicarbonitriles and nitroolefins
that affords optically enriched nitrocyclopentane derivatives
with three consecutive chiral stereocenters in up to 96 % yield
and with up to 92 % ee. The work presents the first enantioselective (3+2) cycloaddition of vinyl cyclopropane and nitroolefins. The two diastereomeric isomers can be easily separated
by simple chromatography even if the diastereomeric ratio is
only 1:1.4, which allows the present method to be attractive in
organic synthesis.

Experimental Section
A mixture of nitroolefin 2 a (1.1 equiv.), vinyl cyclopropane 1 a
(1 equiv.), and (S)- MeO- BIPHEP (0.1 equiv., 10 mol %) and
[Pd(dba)2] (0.05 equiv., 5 mol %) in toluene (2 mL) was stirred under
an atmosphere of argon for 24 h at room temperature. Upon completion of the reaction, the solvent was removed under reduced
pressure, and the residue was purified by column chromatography
on silica gel using mixed solvent of EtOAc and hexanes as the
eluent to afford the optically enriched cycloadducts 3 a and 3 a.
3 a: White solid, a25:7
D = 42.18 (c = 0.04, CH2Cl2), m.p. 126 8C, 82 %
ee determined by HPLC analysis (Chiralpak AD-H column, 5 % isopsopyl alcohol (IPA) in n-hexane, flow rate: 1.2 mL min1, 210 nm);
retention time: t (major) = 13.02 min, t (minor) = 14.17 min. 1H NMR
(400 MHz, CDCl3): d = 7.517.43 (m, 5 H), 5.69 (ddd, J = 17.4, 10.1,
7.5 Hz, 1 H), 5.51 (dd, J = 9.5, 7.7 Hz, 1 H), 5.435.39 (m, 2 H), 4.60 (d,
J = 7.7 Hz, 1 H), 3.713.63 (m, 1 H), 2.952.90 (m, 1 H), 2.85
Asian J. Org. Chem. 2015, 4, 28 32

www.AsianJOC.org

31

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Communication
[10] a) Y. Sumida, H. Yorimitsu, K. Oshima, Org. Lett. 2008, 10, 4677; b) R. K.
Bowman, J. S. Johnson, Org. Lett. 2006, 8, 573.
[11] A. P. Dieskau, M. S. Holzwarth, B. Plietker, J. Am. Chem. Soc. 2012, 134,
5048.
[12] A. F. G. Goldberg, B. M. Stoltz, Org. Lett. 2011, 13, 4474.
[13] a) B. M. Trost, P. J. Morris, Angew. Chem. Int. Ed. 2011, 50, 6167; Angew.
Chem. 2011, 123, 6291; b) B. M. Trost, P. J. Morris, S. J. Sorague, J. Am.
Chem. Soc. 2012, 134, 17823; c) L. Mei, Y. Wei, Q. Xu, M. Shi, Organometallics 2012, 31, 7591.
[14] a) T. Arai, A. Mishiro, N. Yokoyama, K. Suzuki, H. Sato, J. Am. Chem. Soc.
2010, 132, 5338; b) M.-X. Xue, X. M. Zhang, L.-Z. Gong, Synlett 2008,
691; c) Y. K. Liu, H. Liu, W. Du, L. Yue, Y. C. Chen, Chem. Eur. J. 2008, 14,
9873; d) J. W. Xie, K. Yoshida, K. Takasu, Y. Takemoto, Tetrahedron Lett.
2008, 49, 6910.
[15] a) D. Amantini, F. Fringuelli, O. Piermatti, F. Pizzo, L. Vaccaro, J. Org.
Chem. 2003, 68, 9263; b) E. C. Taylor, B. Liu, J. Org. Chem. 2003, 68,
9938.

Asian J. Org. Chem. 2015, 4, 28 32

www.AsianJOC.org

[16] A. T. Parsons, M. M. J. Campbell, J. S. Johnson, Org. Lett. 2008, 10, 2541.


[17] S. Yusuf, K. K. Teo, J. Pogue, L. Dyal, I. Copland, H. Schumacher, B. Ingelheim, G. Dagenais, P. Sleight, C. N. Anderson, N. Engl. J. Med. 2008, 358,
1547.
[18] a) D. E. Brodersen, W. M. Clemons, A. P. Carter, R. J. Morgan-Warren, B. T.
Wimberly, V. Ramakrishnan, Cell 2000, 103, 1143; b) J. T. Malinowski, R. J.
Sharpe, J. S. Johnson, Science 2013, 340, 180.
[19] CCDC 1007484 (3 d) and CCDC 1007485 (3 d) contain the supplementary crystallographic data for this paper. These data can be obtained free
of charge from The Cambridge Crystallographic Data Centre via
www.ccdc.cam.ac.uk/data_request/cif.

Received: October 16, 2014


Published online on November 13, 2014

32

 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like