分享

1

 昵称12061175 2013-05-10
The natural alkaloid isoanabasine: synthesis from
2,30-bipyridine, efficient resolution with BINOL, and assignment
of absolute configuration by Moshers method
Chuan-Qing Kang, Yan-Qin Cheng, Hai-Quan Guo, Xue-Peng Qiu and Lian-Xun Gao*
The State Key Laboratory of Polymer Physics and Chemistry, Changchun Institute of Applied Chemistry,
Chinese Academy of Sciences, Graduate School of Chinese Academy of Sciences, 5625 Renmin Street, Changchun 130022, PR China
Received 18 April 2005; accepted 10 May 2005
Abstract—A highly efficient and practical resolution of racemic 1-benzylisoanabasine, which was synthesized by reduction of the
benzyl salt of 2,30-bipyridine, has been achieved through molecular complexation with (R)-BINOL or (S)-BINOL to afford pure
enantiomers (100% ee). The two enantiomers of the natural alkaloid isoanabasine have been obtained by debenzylation of the corresponding
enantiomeric 1-benzylisoanabasine. Using Moshers method by NMR techniques, the absolute configuration of ()-isoanabasine
has been assigned as the (R)-configuration for the first time. Moreover, an unexpected rotamer ratio of Moshers amide
was observed. The syn-form of two rotamers of (R)-MTPA-(R)-isoanabasine was predominant over the anti-form.
 2005 Elsevier Ltd. All rights reserved.
1. Introduction
The rarely investigated minor natural alkaloid isoanabasine
1 (Fig. 1) has been discovered only in low quantities
in the desert plant Anabasis aphylla L., which can be
found in the northwest of China and Central Asia.1 Recently,
much attention has been paid to the synthesis
and biological activities of its skeleton analogues including
natural and synthetic alkaloids, such as anabasine,
anatabine, cytisine, and some other pyridyl piperidines.
2,3 These analogues displayed potential as therapeutic
agents for peripheral and central nervous
system diseases and other disorders because of their
affinity activities to neuronal nicotinic acetylcholine
receptors (nAChRs) but with lower toxicities than the
well-known alkaloid nicotine.3 Experience has shown
that important biological activities of chiral natural
products are related with their absolute configuration,
so that methods providing single enantiomers were
of utmost importance. However, to the best of our
knowledge, the enantiomers of 1 had never gained as
much attention, no report has ever clarified the absolute
configuration or focused on the preparation of its enantiomer.
Limited preliminary studies related to 1 were
conducted only on reactions of the pyridine or the piperidine
ring during the 1970s in the former Soviet Union.4
Herein, we report our approach to the two enantiomers
of 1 via synthesis and resolution, plus assignment of
absolute configuration of 1 by Moshers method.
2. Results and discussion
2.1. Synthesis of racemic alkaloids
Syntheses of substituted piperidines have been reported
using a variety of methods including stereoselective
N N
R
1 R = H
2 R = Bn
0957-4166/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.tetasy.2005.05.012
* Corresponding author. Tel.: +86 431 526 2259; fax: +86 431 568
5653; e-mail: lxgao@ciac.jl.cn
OH
OH OH
RO CO2H OH
RO CO2H
5 R = H
6 R = p-Toluoyl
(R)-7 (S)-7
Figure 1. Chiral reagents bearing double acidic groups for resolution.
Tetrahedron: Asymmetry 16 (2005) 2141–2147
Tetrahedron:
Asymmetry
approaches in the development of new drugs.5 However,
those methods involving enantioselective reaction usually
lead to a complex mixture of products, low yield,
or quantitative consumption of the chiral auxiliary
meaning that they are not practical. All reported preparations
of 1 were based on the reduction of 2,30-bipyridine
3 in two approaches. One involved the direct use
of a large amount of reductive metals in low to moderate
yields.6a,b The other method was the oxidation of
2,30-bipyridine prior to catalytic hydrogenation with
Pd/C.6c,d In view of the advantages of a pre-constructed
heterocyclic skeleton upon reduction of bipyridines to
pyridyl piperidines, we have utilized an improved reductive
approach to 1 from 3.
Formation of the piperidine ring of 1 was accomplished
on a large scale by selective reduction of 30-pyridyl
through benzylation on less hindered nitrogen atom of
3 (Scheme 1). Benzyl salt 4, obtained by heating a solution
of 3 and benzyl chloride in acetonitrile, was elaborated
to 2 via hydrogenation in the presence of
triethylamine catalyzed with PtO2 or Pd/C. Debenzylation
of 2 was performed by substituent exchange reaction
on nitrogen with benzyl chloroformate and then
hydrolysis with 37% HCl to afford racemic 1 in total
67% yield.
2.2. Enantiomerically pure alkaloids from resolution
Molecular complexation, based on molecular recognition
between host and guest molecules directed by steric
complementary action and specific intermolecular
forces (such as hydrogen bonding and second-order
interaction), has proven to be an effective method for
the resolution of organic molecules.7 Optically active
1,10-bi-2-naphthol 7 (BINOL) has shown excellent performances
in this field. A few successful examples on
resolution of alkaloids by forming host–guest complex
with resolving reagent bearing double carboxylic or
phenolic groups have been published.8 The selectively
formed diastereomeric salt in well-matched chirality, if
possible, would display marked differences in physical
or chemical properties such as solubility and stability
compared with the alkali guest and acidic host themselves
or the chirality-mismatched diastereomeric salt.
Three readily available chiral reagents bearing double
acidic groups, l-tartaric acid 5, l-O,O0-ditoluoyl tartaric
acid 6 and (R)-7 (Fig. 1), were selected as candidates to
screen out better resolving process. As shown in Table 1,
fortunately, (R)-7 could form crystals with one enantiomer
of 2 in ethanol in a molar ratio of 1:2 [(R)-7:rac-2].
The solvent was then changed with methanol to give a
similar result. Acetonitrile and toluene were not suitable
for the resolution. Varying the ratio of the resolving reagent
and racemic alkaloid to 1:1 did not afford better
results. Hence, we selected (R)-7 and (S)-7 as resolving
reagents to optimize the resolution process.
The typical resolution process employed is as follows.
A solution of (R)-7 (0.5 equiv) and rac-2 (1.0 equiv) in
ethanol was stirred under reflux for 0.5 h. The crystals
formed during cooling to room temperature were collected
by filtration and purified by recrystallization.
The resulting crystals were characterized as a 1:1 molecular
complex of 2 and (R)-7 by 1H NMR and elemental
analysis.9 Decomposition of the resolving complex with
15% NaOH afforded the ()-enantiomer of 2 (100%
ee,10 79% yield based on half of the starting rac-2),
which displayed ?a20
D ? 65.4 (c 2.0, ethanol). All mother
liquids from resolution and recrystallization, enriched
in the (+)-enantiomer of 2, were combined and washed
with 15% NaOH to remove (R)-7. The resulting residues
were applied to form molecular complex with (S)-7
(0.5 equiv) to afford (+)-2 (100% ee,10 83% yield based
on half of the starting rac-2), which displayed ?a20
D ?
t65.2 (c 2.0, ethanol). Both (R)- and (S)-7 were recovered
in >80% yield by acidification of the corresponding
alkali aqueous solution, respectively (Scheme 2).
Unfortunately, no crystals of the molecular complex of
(R)-7 and ()-2 suitable for X-ray crystallography structure
analysis have been obtained until now. However, by
1H NMR (Fig. 2), we have observed that the hydroxy
N N
3
N N
4 Bn
Cl
N N
Bn
N NH
BnCl, MeCN
reflux, 4 hr
94%
H2, Et3N, EtOH
PtO2 at 1 atm
or Pd/C at 10 atm
95%
1) CbzCl, toluene
reflux, 15 hr
2) 37% HCl
reflux, 6 hr
76%
2
1
Scheme 1.
Table 1. Screening of resolving reagents by formation of precipitatesa
Alkaloids Resolving reagents
L-5 L-6 (R)-7
1   
2   +
a Those formed precipitates after mixing were marked with +,
otherwise with .
rac-2 + (R)-7
(1.0 equiv : 0.5 equiv)
(R)-7 (-)-2
(-)-2
100% ee
(R)-7 (-)-2
(+)-2
(+)-2, (-)-2
recycled
(R)-7
(S)-7
(0.5 equiv)
(+)-2 (S)-7 (+)-2
100% ee
recycled
(S)-7
(mother liquor)
Scheme 2. Schematic procedure of the resolution.
2142 C.-Q. Kang et al. / Tetrahedron: Asymmetry 16 (2005) 2141–2147
protons in the molecular complex (R)-7?()-2 have
shown a downfield shift from d 5.02 [in (R)-7] to d
6.80 ppm, while the protons on benzyl methylene
showed a double doublet peak with a large coupling constant
(J = 60.0 and 13.2 Hz). The 1:1 mixture of (S)-7
and ()-2 in the same concentration11 to the molecular
complex (R)-7?()-2 exhibited great differences and that
the hydroxy protons also showed a downfield shift from
d 5.02 to d 6.50 ppm and the protons on benzyl methylene
showed double doublet with small split (J = 18.0
and 13.2 Hz).12 The larger split in 1H NMR spectra of
the molecular complex (R)-7?()-2 demonstrated the
stronger hindered rotation of benzyl around its C–N
bond, which should come from the stronger binding between
(R)-7 and ()-2. These phenomena suggested that
the hydrogen bonding present in both diastereomeric
mixtures, and the stronger interactions in (R)-7?()-2
originated from the matching of chirality of the host
and guest in terms of chiral recognition, which was a
fundamental driver to the stereoselectivity of (R)-7 to
()-2 or (S)-7 to (+)-2 during resolution.
Based on the debenzylation procedure described in the
last section, both enantiomers of 2 were converted to
corresponding pure enantiomers of 1, which displayed
a specific rotation of 15.2 (c 1.0, ethanol) with the same
sign to their benzyl precursors.
2.3. Assignment of absolute configuration of ()-1
After pioneering work in the 1970s, Moshers method
has been one of the most useful techniques amongst
those employed to determine the absolute configuration
of organic compounds.13 Hoye and Renner extended
and modified Moshers method to cyclic secondary
amines, which have a stereogenic center on the ring.14
It has been proven that Hoyes method was the most efficient
technique for the assignment of the absolute
configuration of substituted cycloamines15a–f even
though the inconveniences presented in MTPA amide
synthesis and NMR spectra analysis might limit its
application.15g,h
In principle, the most stable conformation of the
Mosher amides derived from the secondary amines
(such as those derived from piperidine) has the trifluoromethyl
group coplanar with the carbonyl group
syn-periplanar and with the nitrogen anti-periplanar,
the phenyl group on MTPA moiety of the amides
shielded one of four possible quadrants of the molecule
depending on the specific rotamer and absolute stereochemistry.
Furthermore, it was reasonable to assume
that the larger pyridyl group at the 3-position in the
equatorial position was favorable over that at the axial
position. For Moshers amide isomers 8 and 9 (Scheme
3), one rotamer of 8 in which the axial protons shielded
most at the 2- and 6-positions, should take the same
rotation conformation around the amide bond to that
of 9 in which most-shielded axial protons at 3- and 5-
positions and vice versa.16
Based on the above principles and in light of the 1H
NMR, 13C NMR, 1H–1H COSY, and 1H–13C HMQC,
the chemical shifts of the rotamers were assigned and
we easily knew that the minor of 8 and the major of 9
were a pair of rotamers, which had taken the syn-form
while another pair were anti-rotamers (Table 2). Hence,
the molar ratio17 of syn-8 to anti-8 was 1:3.1 similar to
what has generally been reported14,15a–e and that of
syn-9 to anti-9 was unexpectedly 2.8:1, which had been
presented only once in the literature.15f From the differences
of chemical shifts (DdS–R), the axial 20-proton in
the syn-8 and the axial 30-proton in syn-9 should take
syn-form with phenyl on corresponding MTPA moiety
based on the carbonyl plane. Thus, the absolute stereochemistry
on stereogenic center of ()-1 was deduced to
be an (R)-configuration. The rendered 3D representations
have clearly exhibited the absolute configuration
Figure 2. Partial 1H NMR spectra of diastereomeric 1:1 mixtures of
()-2 with (R)-7 and (S)-7.
(-)-1
(R)-MTPACl
(S)-MTPACl
N
N O
CF3
Ph
OMe
N
N O
CF3
OMe
Ph
anti-8
anti-9
O
Cl
CF3
Ph
OMe
O
Cl
CF3
OMe
Ph
N
N
O
Ph CF3
OMe
syn-9
N
N
O
MeO CF3
Ph
syn-8
+
+
1
3 2
4 6
5
Scheme 3. Conversion of ()-1 to corresponding Mosher amides.16
Table 2. Assignment of chemical shifts (ppm) of protons on piperidinyl
ring of Mosher amides
Ha syn-8
(minor)
anti-8
(major)
syn-9
(major)
anti-9
(minor)
DdS–R
syn anti
2a 2.41 2.96 3.17 2.99 0.76 0.03
2e 4.10 4.83 4.08 4.80 0.02 0.05
3 2.86 2.78 1.67 3.05 1.19 0.24
4a 1.70 1.75 1.87 1.76 0.18 0.01
4e 2.05 1.84 1.76 2.02 0.29 0.18
5a 1.63 0.48 1.50 1.55 0.13 1.07
5e 1.92 1.18 1.79 1.55 0.13 0.37
6a 2.57 2.91 2.62 2.30 0.05 0.61
6e 4.85 4.01 4.78 3.84 0.07 0.17
a Protons at axial position were marked with a and those at equatorial
position marked with e.
C.-Q. Kang et al. / Tetrahedron: Asymmetry 16 (2005) 2141–2147 2143
and the shielding circumstance in Mosher amides (Fig.
3).
3. Conclusions
In conclusion, we have synthesized the natural alkaloid
isoanabasine 1 via a facile and practical route in which
the intermediate 1-benzylisoanabasine 2 was resolved
to both enantiomers in 100% ee efficiently by forming
a molecular complex with (R)- and (S)-1,10-bi-2-naphthol
7. The preliminary investigation to the chiral recognition
involving in resolution, studied by 1H NMR,
revealed that the matching of chirality between (R)-7
and (R)-()-2 [or (S)-7 and (S)-(+)-2] strengthened the
hydrogen bonding of the host–guest complex. Enantiomerically
pure 1 was obtained by debenzylation of corresponding
enantiomeric 2. The absolute configuration
of ()-1 was assigned as an (R)-configuration using
Moshers method by NMR techniques. During the
deduction, it was observed from unexpected rotamer
ratio of Mosher amide derived from (R)-()-1 and (S)-
MTPACl that the syn-rotamer predominated over the
anti-one.
4. Experiments
Generally, chiral HPLC measurements were carried out
on a Shimadzu Class-VP workstation through CHIRALCEL
 OD-H column (Daicel Chemical Industries,
Ltd) with a UV detector at 254 nm. Optical rotations
were measured on a Perkin Elmer 341LC polarimeter.
ESI-MS measurements were conducted with a LCQ
instrument. Unless otherwise noted, 1H spectra were recorded
on a Bruker 600 MHz spectrometer and the proton
chemical shifts referenced to TMS as internal
standard at 0.00 ppm. 13C NMR chemical shifts were reported
in parts per million relative to the solvent CDCl3
at 77.0 ppm or DMSO-d6 at 39.5 ppm. Melting points
were measured with a hot-stage microscope XT-4. Elemental
analysis was carried out with a VarioEL instrument.
TLC was performed on aluminum TLC-layers
Silica gel GF-254. Detection was done by UV light
(254 and 365 nm). All materials were available commercially
without further purification. All products were
dried in vacuum prior to further reaction and analysis.
4.1. Preparation of 10-benzyl-2,30-bipyridinium chloride 4
A stirring solution of 2,30-bipyridine 3 (15.60 g,
0.10 mol) and benzyl chloride (13.8 mL, 0.12 mol) in
acetonitrile (100 mL) was heated to reflux for 4 h. After
being cooled to room temperature, the precipitates were
collected by filtration, washed with acetonitrile (30 mL),
and dried in vacuum to give 26.56 g of 10-benzyl-2,30-
bipyridinium chloride 4 as buff granule (94% yield).
Mp 78–80 C. 1H NMR (600 MHz, DMSO-d6): d
10.15 (s, 1H), 9.38 (d, 1H, J = 6.0 Hz), 9.27 (d, 1H,
J = 8.4 Hz), 9.79 (d, 1H, J = 4.2 Hz), 8.39 (d, 1H,
J = 7.8 Hz), 8.30 (dd, 1H, J = 7.8, 6.0 Hz), 8.08 (dt,
1H, J = 7.8, 1.8 Hz), 7.68 (d, 2H, J = 6.6 Hz), 7.59 (dd,
1H, J = 7.2, 4.8 Hz), 7.44 (m, 3H), 6.10 (s, 2H). 13C
NMR (150 MHz): d 150.1, 149.6, 144.4, 143.0, 142.5,
138.4, 138.0, 134.3, 129.2, 129.1, 128.9, 128.5, 125.1,
122.0, 63.2. ESI-MS (m/z): [MI]+ 247.0.
4.2. Preparation of racemic 1-benzylisoanabasine [i.e., 1-
benzyl-3-(pyridin-2-yl)piperidine] rac-2
A suspension of 4 (11.28 g, 0.040 mol), triethylamine
(6.2 mL, 0.044 mol), and 10% Pd/C (0.6 g) in ethanol
(100 mL) was stirred in an autoclave under hydrogen
atmosphere at 10 atm for 6–8 h at room temperature.
Upon deflation of the hydrogen, the insoluble species
were removed by filtration through Celite. After
evaporation of the ethanol, the residues were dissolved
in dichloromethane (100 mL) and washed with 10%
NaOH (30 mL · 2) and water. The organic layer was
dried over magnesium sulfate. The solvent was removed
by evaporation to give 9.57 g of 1-benzylisoanabasine 2
as colorless oil (95% yield). 1H NMR (600 MHz,
CDCl3): 8.52 (d, 1H, J = 4.2 Hz), 7.57 (dt, 1H, J = 7.8,
1.8 Hz), 7.35 (d, 2H, J = 7.2 Hz), 7.30 (t, 2H, J = 7.2
Hz), 7.24 (t, 1H, J = 7.2 Hz), 7.16 (d, 1H, J = 7.8 Hz),
7.09 (dd, 1H, J = 7.8, 4.8 Hz), 3.59 (s, 2H), 3.07
(m, 2H), 2.94 (d, 1H, J = 11.4 Hz), 2.28 (t, 1H, J = 10.8
Hz), 2.07 (m, 1H), 1.97 (dt, 1H, J = 12.6, 1.8 Hz),
1.78 (m, 2H), 1.61 (m, 1H). 13C NMR (150 MHz):
d 163.6, 149.1, 138.1, 136.3, 129.2, 128.2, 127.0, 122.0,
121.3, 63.4, 59.0, 53.6, 44.6, 30.5, 25.3. ESI-MS (m/z):
[M+H]+ 253.1.
An alternative hydrogenation using PtO2 as catalyst was
carried out at ambient pressure in a similar procedure.
4.3. Preparation of racemic isoanabasine [i.e., 3-(pyridin-
2-yl)piperidine] rac-1
A stirred solution of 2 (2.52 g, 0.010 mol) and benzyl
chloroformate (1.7 mL, 0.012 mol) in toluene (30 mL)
was heated to reflux for 10 h under a nitrogen atmosphere.
After being cooled to room temperature, the
Figure 3. 3D representations of rotamers of Mosher amides 8 and 9
from minimizing energy.
2144 C.-Q. Kang et al. / Tetrahedron: Asymmetry 16 (2005) 2141–2147
reaction mixture was filtered through silica gel (200
mesh) and eluted with ethyl acetate (100 mL) to remove
tar and decolorizing. The filtrate was then washed with
saturated NaHCO3 (40 mL · 2) and water. The solvent
was removed by evaporation and the residues dissolved
in a mixture of acetic acid (20 mL) and 37% HCl
(10 mL). The acidic solution was stirred and heated to
reflux for 6 h and the acids then distilled off. The
residues were suspended in 10% NaOH (30 mL) and
extracted with dichloromethane (50 mL · 2). The organic
layer was dried over magnesium sulfate and evaporated
to give 1.23 g of rac-1 as viscous oil (76% yield).
1H NMR (600 MHz, CDCl3): d 8.49 (d, 1H, J = 4.8
Hz), 7.56 (dt, 1H, J = 7.8, 1.8 Hz), 7.12 (d, 1H, J =
7.8 Hz), 7.07 (dd, 1H, J = 7.2, 4.8 Hz), 3.23 (m, 1H), 3.15
(s, 1H), 3.08 (d, 1H, J = 12.8 Hz), 2.86 (m, 2H), 2.66
(dt, 1H, J = 12.2, 2.4 Hz), 2.00 (dd, 1H, J = 12.6,
3.0 Hz), 1.76 (m, 2H), 1.60 (m, 1H). 13C NMR (150
MHz): d 163.4, 149.0, 136.3, 121.6, 121.3, 51.7, 46.1,
45.2, 30.6, 26.1. ESI-MS (m/z): [M+H]+ 163.1.
4.4. Resolution of racemic 1-benzylisoanabasine to enantiomers
(R)-()-2 and (S)-(+)-2
A solution of rac-2 (5.56 g, 0.022 mol) and (R)-1,10-bi-2-
naphthol (R)-7 (3.15 g, 0.011 mol) in ethanol (25 mL)
was stirred at 70 C for 0.5 h and cooled naturally to
room temperature. After 1 h, the precipitates were
collected by filtration and purified by recrystallization
in ethanol (15 mL). The resultant molecular complex
was suspended in 15% NaOH (12 mL) and extracted
with dichloromethane (30 mL · 3). The extract was
washed with saturated NaHCO3, dried over magnesium
sulfate, and evaporated to dryness to give 2.20 g of
(R)-()-2 (79% yield). All filtrates from the separation
of the crystals and recrystallization were combined and
evaporated to remove ethanol. The residues were
dissolved in dichloromethane (60 mL) and washed with
15% NaOH (15 mL · 2) to remove (R)-7. All aqueous
solutions containing sodium salt of (R)-7 were combined
and acidified with 10% HCl to form precipitates, which
were collected by filtration, washed with water, and
recrystallized in toluene to give 2.57 g of (R)-7 (82%
yield). The organic phase enriched with (S)-(+)-2 was
then dried with magnesium sulfate and evaporated to
give an oil, which was complexed with (S)-7 (3.15 g,
0.011 mol) in ethanol to give 2.31 g of (S)-(+)-2
(83% yield) in similar procedure as described above.
(S)-7 was recovered in 80% yield and the terminal
residual oil, usually >0% ee, could be used for resolution
again. (R)-(+)-2: ?a20
D ? 65.4 (c 2.0, ethanol), 100% ee
[tR(R) = 5.5 min]. (S)-()-2: ?a20
D ? t65.2 (c 2.0,
ethanol), 100% ee [tR(S) = 6.6 min].9,10
4.5. (R)-()-Isoanabasine (R)-()-1
Prepared from (R)-()-2 in a similar procedure to rac-1
from rac-2. ?a20
D ? 15.2 (c 1.0, ethanol).
4.6. (S)-(+)-Isoanabasine (S)-(+)-1
Prepared from (S)-(+)-2 in a similar procedure to rac-1
from rac-2. ?a20
D ? t15.2 (c 1.0, ethanol).
4.7. Preparation of Moshers amide (3R,20S)-1-(2-methoxy-
2-trifluoromethyl-phenylaceto)-3-(pyridin-2-yl)piperidine
anti- and syn-8
To a solution of (R)-()-1 (21.0 mg, 0.12 mmol) and triethylamine
(18.4 lL, 0.13 mmol) in dichloromethane
(2.0 mL) was added dropwise a solution of (R)-()-2-
methoxy-2-trifluoromethyl-phenylacetyl chloride [(R)-
MTPACl, 24.5 lL, 0.13 mmol] in dichloromethane
(1 mL) at 0 C. The reaction mixture was stirred for
another 4 h at room temperature and diluted with
dichloromethane (5 mL) followed by quenching with
water (3 mL). The organic layer was separated and
washed with water (3 mL) again, dried over magnesium
sulfate, and evaporated to dryness. The residues were
purified with silica-gel chromatography column to give
40.0 mg of rotameric mixture of 8 as oil (86% yield),
which showed a single Rf value on TLC. The molar ratio
of syn-8 to anti-8 was 1:3.1 characterized by 1H NMR.
anti-8: 1H NMR (600 MHz, CDCl3): d 8.53 (d, 1H,
J = 4.5 Hz), 7.62 (d, 2H, J = 7.4 Hz), 7.61 (m, 1H),
7.41 (m, 2H), 7.39 (d, 1H, J = 7.3 Hz), 7.18 (d, 1H,
J = 7.8 Hz), 7.13 (dd, 1H, J = 7.3, 5.0 Hz), 4.83 (d, 1H,
J = 12.9 Hz), 4.01 (d, 1H, J = 13.5 Hz), 3.68 (s, 3H),
2.96 (t, 1H, J = 12.4 Hz), 2.91 (dt, 1H, J = 13.6,
2.8 Hz), 2.78 (tt, 1H, J = 11.4, 3.7 Hz), 1.84 (d, 1H,
J = 12.3 Hz), 1.75 (dt, 1H, J = 12.5, 4.0 Hz), 1.18 (dt,
1H, J = 13.6, 3.1 Hz), 0.48 (m, 1H). 13C NMR
(150 MHz): d 163.7, 161.6, 149.1, 136.5, 134.1, 129.1,
128.1, 126.5, 123.6 (q, J = 288 Hz), 122.0, 121.7, 84.7
(q, J = 24.9 Hz), 55.3, 47.2, 45.1, 43.6, 30.3, 23.5. syn-
8: 1H NMR (600 MHz, CDCl3): d 8.43 (d, 1H,
J = 4.1 Hz), 7.55 (dt, 1H, J = 7.7, 1.5 Hz), 7.44 (dd,
2H, J = 8.0, 3.2 Hz), 7.30 (m, 3H), 7.08 (dd, 1H,
J = 7.3, 5.0 Hz), 6.97 (d, 1H, J = 7.8 Hz), 4.85 (overlap
with a peak of anti-8, 1H), 4.10 (dd, 1H, J = 13.5,
1.7 Hz), 3.85 (s, 3H), 2.86 (tt, 1H, J = 11.8, 3.7 Hz),
2.57 (dt, 1H, J = 13.0, 2.7 Hz), 2.41 (t, 1H,
J = 12.5 Hz), 2.05 (d, 1H, J = 13.4 Hz), 1.92 (dt, 1H,
J = 13.3, 2.5 Hz), 1.70 (m, 1H), 1.63 (m, 1H). 13C
NMR (150 MHz): d 164.0, 161.3, 149.2, 136.5, 133.8,
129.0, 128.2, 126.3, 123.5 (q, J = 288 Hz), 121.8, 121.4,
84.9 (q, J = 24.9 Hz), 56.2, 50.5, 44.8, 42.9, 30.6, 25.2.
ESI-MS (m/z): [M+H]+ 378.3.
4.8. Preparation of Moshers amide (3R,20R)-1-(2-methoxy-
2-trifluoromethyl-phenylaceto)-3-(pyridin-2-yl)piperidine
anti- and syn-9
The synthesis of 9 was performed by amidation of (R)-
()-1 with (S)-MTPACl to give 41.8 mg of rotameric
mixture (90% yield) in similar procedure to that of 8.
The molar ratio of syn-9 to anti-9 was 2.8:1. syn-9: 1H
NMR (600 MHz, CDCl3): d 8.54 (d, 1H, J = 4.5 Hz),
7.58 (d, 2H, J = 7.6 Hz), 7.53 (dt, 1H, J = 7.7, 1.6 Hz),
7.43 (m, 3H), 7.16 (m, 1H), 6.21 (d, 1H, J = 7.8 Hz),
4.78 (dt, 1H, J = 13.0, 1.9 Hz), 4.08 (dt, 1H, J = 13.0,
1.9 Hz), 3.62 (s, 3H), 3.17 (dd, 1H, J = 11.7, 13.5 Hz),
2.62 (dt, 1H, J = 13.0, 2.4 Hz), 1.87 (m, 1H), 1.78 (m,
2H), 1.67 (tt, 1H, J = 11.9, 3.7 Hz), 1.50 (m, 1H). 13C
NMR (150 MHz): d 163.5, 160.3, 147.8, 137.8, 134.5,
129.1, 128.3, 126.8, 123.6 (q, J = 288 Hz), 123.3, 122.3,
84.5 (q, J = 24.9 Hz), 55.4, 49.3, 43.0, 42.8, 29.7, 24.7.
C.-Q. Kang et al. / Tetrahedron: Asymmetry 16 (2005) 2141–2147 2145
anti-9: 1H NMR (600 MHz, CDCl3): d 8.64 (d, 1H,
J = 4.5 Hz), 7.76 (dt, 1H, J = 7.7, 1.6 Hz), 7.47 (m,
2H), 7.36 (m, 3H), 7.29 (m, 2H), 4.80 (overlap with a
peak of syn-9, 1H), 3.84 (d, 1H, J = 13.4 Hz), 3.78 (s,
3H), 3.05 (m, 1H), 2.99 (m, 1H), 2.30 (dt, 1H,
J = 12.0, 4.5 Hz), 2.02 (d, 1H, J = 13.4 Hz), 1.76 (overlap
with a peak of syn-9, 1H), 1.55 (m, 2H). 13C NMR
(150 MHz): d 164.2, 160.8, 147.9, 138.3, 133.9, 129.1,
128.2, 126.4, 123.7 (q, J = 288 Hz), 122.6, 122.5, 84.9
(q, J = 24.9 Hz), 56.1, 46.7, 45.9, 43.5, 30.2, 25.0. ESIMS
(m/z): [M+H]+ 378.3.
Acknowledgments
We appreciate greatly Dr. Zheng Bian and Dr. Xiaoli
Bais valuable suggestions and corrections to this paper.
We also express our thanks to Mr. Yijie Wu and Ms.
Fengying Jing for their help on NMR data acquirement.
References
1. Duke, J. A. Handbook of Phytochemical Constituents of
GRAS Herbs and Other Economic Plants; CRC Press:
Boca Raton, FL, 1992, cited in online database (http://
www.ars-grin.gov:8080): USDA, ARS, National Genetic
Resources Program, Phytochemical and Ethnobotanical
Databases, National Germplasm Resources Laboratory,
Beltsville, Maryland, 2005.
2. Selected synthesis published recently: (a) Danieli, B.;
Lesma, G.; Passarella, D.; Sacchetti, A.; Silvani, A.;
Virdis, A. Org. Lett. 2004, 6, 493–496; (b) Amat, M.;
Canto′ , M.; Llor, N.; Bosch, J. Chem. Commun. 2002, 526–
527; (c) Stehl, A.; Canto, M.; Schulz, K. Tetrahedron 2002,
58, 1343–1354; (d) Romanelli, M. N.; Manetti, D.;
Scapecchi, S.; Borea, P. A.; Dei, S.; Bartolini, A.;
Ghelardini, C.; Gualtieri, F.; Guandalini, L.; Varani, K.
J. Med. Chem. 2001, 44, 3946–3955; (e) Felpin, F. X.;
Girard, S.; Vo-Thanh, G.; Robins, R. J.; Villie′ras, J.;
Lebreton, J. J. Org. Chem. 2001, 66, 6305–6312; Synlett
2000, 1646–1648; (f) O Neill, B. T.; Yohannes, D.;
Bundesmann, M. W.; Arnold, E. P. Org. Lett. 2000, 2,
4201–4204; (g) Andre′s, J. M.; Herra′iz-Sierra, I.; Pedrosa,
R.; Pe′rez-Encabo, A. Eur. J. Org. Chem. 2000, 1719–1726;
(h) Balasubramanian, T.; Hassner, A. Tetrahedron: Asymmetry
1998, 9, 2201–2205; (i) Weymann, M.; Pfrengle, W.
Synthesis 1997, 1151–1160; (j) Yang, C. M.; Tanner, D. D.
Can. J. Chem. 1997, 95, 616–620; (k) Wanner, M. J.;
Koomen, G.-J. J. Org. Chem. 1996, 61, 5581–5586; (l)
Genisson, Y.; Mehmandoust, M.; Marazano, C. Heterocycles
1994, 39, 811–818.
3. For biological activities research, recent reviews: (a)
Toma, L.; Barlocco, D.; Gelain, A. Expert Opin. Ther.
Patents 2004, 14, 1029–1040; (b) Romanelli, M. N.;
Gualtieri, F. Med. Res. Rev. 2003, 23, 393–426; (c)
Bunnelle, W. H.; Decker, M. W. Expert Opin. Ther.
Patents 2003, 13, 1003–1021.
4. (a) Yunusov, T. K.; Akbarov, K.; Otroshchenko, O. S.
Uzb. Khim. Zh. 1980, 77–90; (b) Sadikov, A. S.;
Otroshchenko, O. S.; Yunusov, T. K. Pol. J. Chem.
1979, 53, 367–377; (c) Kurbatov, Y. V.; Zalyalieva, S. V.
Khim. Geterotsikl. Soedin. 1977, 1535–1537.
5. For reviews, see: (a) Buffat, M. G. P. Tetrahedron 2004, 60,
1701–1729; (b) Laschat, S.; Dickner, T. Synthesis 2000,
1781–1813; (c) Bailey, P. D.; Millwood, P. A.; Smith, P. D.
J. Chem. Soc., Chem. Commun. 1998, 633–640; (d) Baliah,
V.; Jeyraman, R.; Chandrasekaran, L. Chem. Rev. 1983,
83, 379–423.
6. (a) Lunn, G. J. Org. Chem. 1992, 57, 6317–6320; (b)
Smith, C. R. J. Am. Chem. Soc. 1931, 53, 277–283; (c)
Plaquevent, J.-C.; Chichaoui, I. Bull. Soc. Chim. Fr. 1996,
133, 369–380; (d) Plaquevent, J.-C.; Chichaoui, I. Tetrahedron
Lett. 1993, 34, 5287–5288.
7. (a) Cheng, X.; Hou, G. H.; Xie, J. H.; Zhou, Q. L. Org.
Lett. 2004, 6, 2381–2383; (b) Liao, J.; Sun, X. X.; Cui, X.;
Yu, K. B.; Zhu, J. Chem. Eur. J. 2003, 9, 2611–2615; (c)
Kaupp, G. Angew. Chem., Int. Ed. Engl. 1994, 33, 728–
729; (d) Toda, F.; Tohi, Y. J. Chem. Soc., Chem. Commun.
1993, 1238–1240; (e) Toda, F. Top. Curr. Chem. 1987, 140,
43–69.
8. (a) Bortolini, O.; Fogagnolo, M.; Fantin, G.; Medici, A.
Chem. Lett. 2003, 32, 206–207; (b) Nishikawa, K.;
Tsukada, H.; Abe, S.; Kishimoto, M.; Yasuoka, N.
Chirality 1999, 11, 166–171; (c) Toda, F.; Shinyama, T.
J. Chem. Soc., Perkin Trans. 1 1997, 1759–1762; (d)
Zaderenko, P.; Lopez, P.; Ballesteros, P.; Takumi, H.;
Toda, F. Tetrahedron: Asymmetry 1995, 6, 381–384; (e)
Hatano, K.; Takeda, T.; Saito, R. J. Chem. Soc., Perkin
Trans. 2 1994, 579–584.
9. The integrals of all proton peaks in 1H NMR spectra and
element contents were matched with the formula of
molecular complex (R)-7?()-2. 1H NMR (53.8 mg in
0.5 mL of CDCl3): d 8.37 (d, 1H, J = 4.2 Hz), 7.84 (d, 2H,
J = 9.0 Hz), 7.81 (d, 2H, J = 8.4 Hz), 7.52 (dt, 1H,
J = Hz), 7.29 (d, 4H, J = 9.0 Hz), 7.22 (m, 9H), 7.06 (m,
2H), 3.36 (dd, 2H, J = 60.0, 13.2 Hz), 3.02 (m, 1H), 2.97
(d, 1H, J = 11.4 Hz), 2.81 (d, 1H, J = 10.8 Hz), 2.13 (t, 1H,
10.8 Hz), 1.95 (dt, 1H, J = 11.4, 3.0 Hz), 1.86 (d, 1H,
J = 11.4 Hz), 1.68 (m, 2H), 1.47 (qd, 1H, J = 12.0, 7.5 Hz).
Elemental analysis result: calculated for C37H34N2O2:
C, 82.50; H, 6.36; N, 5.20. Found: C, 82.41; H, 6.45; N,
5.19.
10. Enantiomeric excess analysis was performed with nhexane/
isopropanol/diethylamine (in volume ratio of 70/
30/0.1) as mobile phases. The retention time of ()-2 was
5.5 min, while that of (+)-2 was 6.6 min under a flow
velocity of 1.0 mL/min.
11. The 1H NMR spectra of the molecular complex of (R)-7
and ()-2 were measured in concentration of 53.8 mg of
the molecular complex in 0.6 mL of CDCl3, and that of
(S)-7 and ()-2 in concentration of 25.2 mg of ()-2 and
28.6 mg of (S)-7 in 0.6 mL of CDCl3.
12. Besides protons on benzyl methylene, few of other protons
also showed a few differences in chemical shifts between
two diastereomeric complexes.
13. (a) Sullivan, G. R.; Dale, J. A.; Mosher, H. S. J. Org.
Chem. 1973, 38, 2143–2147; (b) Dale, J. A.; Dull, D. L.;
Mosher, H. S. J. Org. Chem. 1969, 34, 2543–2549; Review:
(c) Seco, J. M.; Quin?oa′ , E.; Riguera, R. Chem. Rev. 2004,
104, 17–117, and references cited therein.
14. (a) Hoye, T. R.; Renner, M. K. J. Org. Chem. 1996, 61,
8480–8495; (b) Hoye, T. R.; Renner, M. K. J. Org. Chem.
1996, 61, 2056–2064.
15. (a) Kanger, T.; Kriis, K.; Pehk, T.; Mu¨u¨ risepp, A.-M.;
Lopp, M. Tetrahedron: Asymmetry 2002, 13, 857–865; (b)
Davis, H. M. L.; Hodges, L. M. J. Org. Chem. 2002, 67,
5683–5689; (c) Itoh, T.; Nagata, K.; Miyazaki, M.;
Kameoka, K.; Ohsawa, A. Tetrahedron 2001, 57, 8827–
8839; (d) Beshore, D. C.; Dinsmore, C. J. Tetrahedron
Lett. 2000, 41, 8735–8739; (e) Dinsmore, C. J.; Zartman,
C. B. Tetrahedron Lett. 2000, 41, 6309–6312; (f) Davies, H.
M. L.; Matasi, J. J.; Hodges, L. M.; Huby, N. J. S.;
Thornley, C.; Kong, N.; Houser, J. H. J. Org. Chem. 1997,
62, 1095–1105; (g) Huang, X.; Fujioka, N.; Pescitelli, G.;
Koehn, F. E.; Williamson, R. T.; Nakanishi, K.; Berova,
2146 C.-Q. Kang et al. / Tetrahedron: Asymmetry 16 (2005) 2141–2147
N. J. Am. Chem. Soc. 2002, 124, 10320–10335; (h)
Chalard, P.; Bertrand, M.; Canet, I.; The′ry, V.; Remuson,
R.; Jeminet, G. Org. Lett. 2000, 2, 2431–2434.
16. Both amides were composed of two inseparable rotamers
from hindered rotation around amido C–N bond. The
rotamers showed two different groups of peaks and one
predominated over another.
17. The rotamer ratios of Mosher amides measured by NMR
after purification were consistent with those before
purification.
C.-Q. Kang et al. / Tetrahedron: Asymmetry 16 (2005) 2141–2147 2147

    本站是提供个人知识管理的网络存储空间,所有内容均由用户发布,不代表本站观点。请注意甄别内容中的联系方式、诱导购买等信息,谨防诈骗。如发现有害或侵权内容,请点击一键举报。
    转藏 分享 献花(0

    0条评论

    发表

    请遵守用户 评论公约

    类似文章 更多