Anda di halaman 1dari 5

Phytochemistry 87 (2013) 6064

Contents lists available at SciVerse ScienceDirect

Phytochemistry
journal homepage: www.elsevier.com/locate/phytochem

Strigone, isolation and identication as a natural strigolactone from Houttuynia


cordata
Takaya Kisugi a, Xiaonan Xie a, Hyun Il Kim a, Kaori Yoneyama a, Aika Sado b, Kohki Akiyama b,
Hideo Hayashi b, Kenichi Uchida c, Takao Yokota c, Takahito Nomura a, Koichi Yoneyama a,
a
b
c

Weed Science Center, Utsunomiya University, 350 Mine-machi, Utsunomiya 321-8505, Japan
Graduate School of Life and Environmental Sciences, Osaka Prefecture University, 1-1 Gakuen-cho, Naka-ku, Sakai, Osaka 599-8531, Japan
Department of Biosciences, Teikyo University, 1-1 Toyosatodai, Utsunomiya 320-8551, Japan

a r t i c l e

i n f o

Article history:
Received 5 October 2012
Received in revised form 8 November 2012
Available online 3 January 2013
Keywords:
Houttuynia cordata
Saururaceae
Strigol
Strigolactone
Strigone
Sorgomol
5-Deoxystrigol
Orobanche minor
Phelipanche ramosa
Striga hermonthica
Orobanchaceae

a b s t r a c t
(+)-Strigone was described earlier in a paper on isolation of strigol and then recently examined for hyphal
branching activity in arbuscular mycorrhizal fungi as a strigolactone. Herein, it was isolated from root
exudates of Houttuynia cordata, and its structure was conrmed by direct comparison with synthetic
standards in LCMS/MS, GCMS, and 1H and 13C NMR analyses. The stereochemistry of strigone was
determined by comparing the CD spectra and RRt in chiral LCMS/MS with those of synthetic (+)-strigone
and ()-strigone. Four stereoisomers of strigone exhibited clearly different levels of stimulation activity
on the seeds of three root parasitic plants, Orobanche minor, Phelipanche ramosa, and Striga hermonthica.
(+)-Strigone was a highly potent germination stimulant on S. hermonthica and also on P. ramosa, but less
active than ent-20 -epi-strigone on O. minor. In addition to strigone, H. cordata was found to produce strigol, sorgomol, and 5-deoxystrigol, indicating that this plant produces mainly strigol-type strigolactones
derived from 5-deoxystrigol.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
Strigolactones (SLs) are carotenoid-derived plant secondary
metabolites which in the rhizosphere induce seed germination of
root parasitic plants, witchweeds (Striga spp.) and broomrapes
(Orobanche and Phelipanche spp.), and promote root colonization
by arbuscular mycorrhizal (AM) fungi (Akiyama et al., 2005; Bouwmeester et al., 2007; Xie et al., 2010). In planta, SLs function as
a novel class of hormones which are involved in the regulation of
shoot (Gomez-Roldan et al., 2008; Umehara et al., 2008) and root
architecture (Koltai, 2011; Ruyter-Spira et al., 2011), and various
developmental processes through cross-talk with other hormones.
SLs are distributed widely in the plant kingdom (Xie et al., 2010),
and even mosses and Charales produce and utilize them presumably in regulation and promotion of rhizoid elongation (Delaux
et al., 2012; Proust et al., 2011).
More than 15 SLs have been puried and characterized from
root exudates of various plant species (Xie et al., 2010) since strigol
was rst isolated from cotton root exudates as a germination stim Corresponding author. Tel.: +81 286495152.
E-mail address: yoneyama@cc.utsunomiya-u.ac.jp (K. Yoneyama).
0031-9422/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.phytochem.2012.11.013

ulant for S. lutea (Cook et al., 1966, 1972). All natural SLs characterized so far have a tricyclic lactone (ABC part) connected to a
butenolide group (D-ring) via an enol ether bridge (Xie et al.,
2010). They have different substituents on the A/B ring but the
same CD moiety which has been suggested to be essential for biological activity (Zwanenburg et al., 2009). Natural SLs can be divided into two groups, strigol-type and orobanchol-type SLs
carrying the a-oriented and b-oriented C ring, respectively (Yoneyama et al., 2012). The two simplest SLs, 5-deoxystrigol and its isomer ent-20 -epi-5-deoxystrigol appear to be the common precursors
for the other SLs, the former for strigol-type SLs and the latter for
orobanchol-type SLs. The biosynthesis of these two types of SLs
seems to be regulated somewhat independently in plant species
like Chinese milk vetch (Yoneyama et al., 2012).
In addition to these known SLs, detection of novel SLs including
methoxy-5-deoxystrigol isomers from rice (Jamil et al., 2011), SL1
and SL2 from maize (Jamil et al., 2012), and didehyro-orobanchol
(or strigol) isomers from red clover (Sato et al., 2003), tobacco
(Xie et al., 2007), and tomato (Lpez-Rez et al., 2008) have been
reported. Working with some of these novel SLs in our laboratory
was limited due to lack of enough pure samples for structural elucidation. In earlier studies, Ma et al. (2004) reported germination

T. Kisugi et al. / Phytochemistry 87 (2013) 6064

61

stimulating activities of aqueous and methanol extracts from various Chinese medicinal herbs on the seeds of witchweed, Striga
hermonthica. In fact, the root culture of one of these medicinal
herbs, Menispermum dauricum, produces strigol (Yasuda et al.,
2003). By further analysis of the aqueous and methanol extracts
of Houttuynia cordata, that has been reported to be active on S. hermonthica seed germination (Ma et al., 2004, 2005), it was established that root exudates from this plant grown hydroponically
elicited the seed germination of Orobanche minor. Herein, are described the isolation and identication of (+)-strigone (1) (Cook
et al., 1972; Akiyama et al., 2010) as a natural SL from root exudates of H. cordata, and the signicantly different germination
stimulation activities of its stereoisomers on the seeds of three root
parasites, O. minor, Phelipanche ramosa, and S. hermonthica (Fig. 1).

2. Results and discussion


2.1. Isolation and identication of (+)-strigone (1)
H. cordata was grown hydroponically and root exudates were
collected in a manner similar to that described previously (Yoneyama et al., 2012). The root exudates were partitioned with EtOAc
and 0.2 M K2HPO4 solution to obtain a neutral EtOAc fraction. This
was subjected to silica gel column chromatography with stepwise
elution using n-hexane and EtOAc. The EtOAcn-hexane (1:1, v/v)
fraction, which showed germination stimulating activity on O. minor, was further puried using reversed-phase HPLC to obtain pure
1.
The molecular formula of 1 was established to be C19H20O6 by
HRESITOFMS. The molecular ion at m/z 344 [M]+ in GCMS
and sodium adduct ion at m/z 367 [M+Na]+ in LCMS supported
this molecular formula.
The 1H and 13C NMR spectra of 1 were similar to those of 5deoxystrigol (4) except for the absence of two methylene protons.
Correlations among four methylene protons in the A-ring moiety
were observed in 1H1H COSY spectra, and these methylene protons appeared downeld compared to those of 4. Analysis of the
MS and NMR spectroscopic data indicated the presence of a ketone
at C-5 or C-7 in 1. So far, 7-oxoorobanchol and its acetate were reported as natural SLs having a ketone (Xie et al., 2009). In their 1H
NMR spectra, in particular chemical sifts of methylene protons in
the A-ring were slightly different from those of 1. By contrast,
the 1H NMR spectra of strigone and its epimer (Akiyama et al.,
2010; Cook et al., 1972) were very similar to that of 1. The fragmentation pattern of 1 in GCMS spectrum was also in good agree-

10
8a

8b

O
3a

4a

6'
2'

5'

2.2. Germination stimulating activities of strigone stereoisomers on O.


minor, P. ramosa, and S. hermonthica
O

OH

4'

strigone (1)

HO

7'

strigol (2)

O
O

sorgomol (3)

ment with those of authentic ()-strigone and its epimer (Fig. 2a).
The RRt of 1 was identical to those of ()-strigone in both GCMS
and LCMS/MS analyses (Fig. 2b). Thus, 1 was determined to be
strigone.
To elucidate the stereochemistry of 1, (+)- and ()-strigones
were synthesized from optically pure strigol enantiomers (see Section 4) and the CD spectra of each were compared with that of 1.
The spectral pattern of (+)-strigone was identical to that of 1, and
chiral LCMS/MS analyses also supported this assignment. Consequently, 1 was determined to be (+)-strigone (1).

C
O

3'

Fig. 2. GCMS analysis of 1. (a) EIMS spectrum of 1. (b) Total ion chromatograms
of authentic strigones (upper) and 1 (lower).

5-deoxystrigol (4)

Fig. 1. Chemical structures of SLs produced by H. cordata.

The germination stimulating activities of strigone stereoisomers


on O. minor, P. ramosa, and S. hermonthica seeds were shown in
Fig. 3. Towards O. minor, ent-20 -epi-1, with an a-oriented C ring
and a C-20 (R) conguration as in orobanchol, had the highest activity. The activity of 1, having a b-oriented C ring and a C-20 (R)
conguration as in strigol, was lower than that of ent-20 -epi-1
and ent-1, and 20 -epi-1 was the least active among them. These
results indicate that in the four stereoisomers of strigone the
a-orientation of the C ring has a greater inuence than that of
the C-20 conguration on germination stimulation of O. minor
seeds. By contrast, 1 exhibited the highest activity and ent-20 -epi1 was the second active on seed germination of P. ramosa, indicating that the C-20 (R) conguration is far more inuential in this case.
This is also true with S. hermonthica, where 1 was an exceptionally

62

T. Kisugi et al. / Phytochemistry 87 (2013) 6064

4. Experimental

O. minor

Germination (%)

100
80
60

4.1. General procedures

ent-1
2'-epi-1

ent-2'-epi-1
40
20
0
0.01

0.1

10

100

10

100

Concentration (nM)

P. ramosa
Germination (%)

100
80
60
40
20
0
0.01

0.1

Concentration (nM)

NMR spectra were obtained with either a JEOL JNM-AL400 NMR


or a JEOL JMN-ECA-500 spectrometer in CDCl3. The resonances of
residual CDCl3 at dH 7.26 and dC 77.0 ppm were used for internal references for 1H and 13C NMR spectra, respectively. CD spectra were
recorded with either a JASCO J-720W or a JASCO J-820 spectropolarimeter. Optical rotation was measured with a JASCO P-2100 polarimeter. Mass spectra were recorded on a Shimadzu GCMS-QP2010
Plus instrument in the direct injection mode. EI/GCMS spectra were
obtained with a JEOL JMS-Q1000GC/K9 on a DB-5 (J&W Scientic,
Agilent) capillary column (5 m  0.25 lm) using a He carrier gas
(3 ml min1). GCMS was operated according to the method reported earlier (Yokota et al., 1998) with some minor modications.
Column temperature was kept at 130 C for the rst 1.5 min, and elevated to 270 C by a 6 C min1 gradient. High-resolution mass spectrum was obtained with an Agilent 6250 Q-TOF mass spectrometer
equipped with an ESI source. LCMS analyses were performed by
using a Quattro LC tandem MS instrument from Micromass (Manchester, U.K.) under the following conditions: L-column2 ODS column (CERI, Japan, 50  2.1 mm, 2 lm) applying a (MeOH:H2O)
gradient system initially 30:70, 45:55 at 3 min, 50:50 at 8 min,
70:30 at 12 min, 100:0 at 15 min with a ow rate of 0.2 ml min1.

S. hermonthica

Germination (%)

100

4.2. Plant material and collection of root exudates

80
60
40
20
0
0.01

0.1

10

100

Concentration (nM)
Fig. 3. Germination stimulating activities of strigone stereoisomers toward O.
minor, P. ramosa and S. hermonthica seeds. Data are presented as means se (n = 3).

potent germination stimulant among the four stereoisomers. The


results with P. ramosa and S. hermonthica were in good agreement
with those reported by Thuring et al. (1997) for stereochemical
structureactivity relationship of GR24. Importance of the a-orientation of the C ring for germination stimulation on O. minor seems
to be similar to that reported for S. gesnerioides (Ueno et al., 2011).
In addition to 1, strigol (2), sorgomol (3), and 5-deoxystrigol (4)
were also isolated from the root exudates of H. cordata. Therefore,
this plant produces mainly strigol-type SLs derived from 4.

3. Concluding remarks
In the present study, strigone (1) was isolated as a natural SL.
Although 1 has the molecular formula same to those of putative
didehydro-orobanchol (strigol) isomers, 1 appeared to be different
from those detected in tobacco, red clover, and tomato by LCMS/
MS and GCMS analyses (data not shown). The observation that
highly potent germination stimulation activity of 1 on S. hermonthica while it is a rather weak stimulant on O. minor implies that different root parasitic plants may respond to an SL with different
sensitivities. Therefore, it would be expected to nd species-specic SLs (or specic mixtures of SLs) for root parasites with very
narrow host ranges such as O. hederae which parasitizes only ivy
(Fernndez-Aparicio et al., 2009).

Orobanche minor Sm. (clover broomrape) seeds were collected


from mature plants grown in Watarase basin of Tochigi Prefecture,
Japan. P. ramosa and S. hermonthica seeds were kindly provided by
Profs. Philippe Delavault (University of Nantes, France) and A. G. T.
Babiker (ARC, Sudan), respectively. H. cordata (dokudami) plants
were collected on the campus of Utsunomiya University in Japan.
The plants were washed and transferred to plastic containers
(53.5  33.5  14 cm, W  L  H) containing 20 l tap water. Seven
containers each containing about 50 seedlings were placed in a
growth room maintained at 1722 C under natural daylight conditions. After 1-week cultivation, the root exudates released into
tap water were adsorbed onto activated charcoal (2 g for 20 l,
Wako Pure Chemical Industries Ltd., Osaka, Japan, for chromatography) as reported previously (Akiyama et al., 2005). The tap water
growth media and activated charcoal were exchanged every 2
3 days. The exudates adsorbed to activated charcoal were eluted
with acetone. After evaporation of acetone in vacuo, the residue
was dissolved in 0.2 M K2HPO4 (50 ml, pH 8.3) and extracted with
EtOAc (3  50 ml). The EtOAc extracts were combined, dried
(MgSO4), and concentrated in vacuo.
4.3. Seed germination assay
Germination assays on O. minor, P. ramosa and S. hermonthica
were conducted as reported previously (Yoneyama et al., 1998,
2007). Seeds were conditioned at 23 C for 1 week (O. minor and P.
ramosa) or at 30 C for 2 weeks (S. hermonthica). Each test solution,
unless otherwise mentioned, contained less than 0.1% (v/v) CH3CN.
4.4. Purication and isolation of strigolactones
The crude EtOAc extract of root exudates (322.2 mg) was subjected to silica gel (Merck, 230400 mesh) CC with stepwise elution system of n-hexaneEtOAc (100:00:100) to obtain 11
fractions. Germination stimulating activities on O. minor were observed in Frs. 48 (n-hexaneEtOAc, 70:3030:70), and these active fractions except for Fr. 6 were found to contain known SLs

T. Kisugi et al. / Phytochemistry 87 (2013) 6064

(5-deoxystrigol (4) in Frs. 4 and 5, strigol (2) and sorgomol (3) in


Frs. 7 and 8) by LCMS/MS analysis. Fr. 6 (n-hexaneEtOAc,
50:50, 12.2 mg) was further puried by reversed-phase HPLC
(Inertsil ODS-3, 250  10 mm, 5 lm, GL Sciences, Japan; ow rate:
3.0 ml min1; UV detection: 240 nm; oven: 25 C) with CH3CN
H2O (45:55, v/v) to yield (+)-strigone (1) (0.16 mg) (RRt 32.0 min).
To isolate known SLs, Fr. 4 (n-hexaneEtOAc, 70:30, 1.0 mg) was
puried by reversed-phase HPLC (Inertsil ODS-3, 250  10 mm,
5 lm, GL Sciences Inc.; ow rate: 3.0 ml min1; UV detection
238 nm; oven: 30 C) with CH3CNH2O (55:45, v/v) to obtain (+)5-deoxystrigol (4) (0.2 mg, RRt 42.7 min). Frs. 7 and 8 (n-hexane
EtOAc, 40:6030:70) were combined and subjected to reversedphase HPLC (Inertsil ODS-3, 250  10 mm, 5 lm) with a gradient
of MeCNH2O (3050%) to give an active fraction. This fraction
(2.0 mg) was further puried with Develosil CN-HG-5 column
(Nomura Chemical Co., Ltd., 250  4.6 mm; CH3CNH2O (30:70,
v/v); ow rate: 1 ml min1; UV detection: 238 nm; oven: 30 C)
to obtain (+)-strigol (2) (1.0 mg, RRt 24.2 min) and sorgomol (3)
(0.2 mg, RRt 22.1 min).
4.5. Identication of known SLs
The structures and stereochemistries of known SLs were elucidated by GCMS analyses, chiral LCMS/MS analyses, and comparing CD and 1H NMR spectra with authentic compounds (data not
shown).
4.6. LCMS/MS analysis of stereoisomers by using chiral column
Puried SLs and authentic SLs were subjected to LCMS/MS
analyses using a chiral column (Daicel, Japan, CHIRALPAK AD-3R,
150  2.1 mm, 3 lm; eluent: MeOH; ow rate: 0.2 ml min1;
oven: 10 C for strigone isomers, 20 C for strigol (2) isomers,
30 C for 5-deoxystrigol (4) isomers). RRt (min) of authentic SLs
were; (+)-strigone, 10.04; ()-ent-strigone, 7.50; (+)-strigol, 3.49;
()-ent-strigol, 2.89; (+)-5-deoxystrigol, 5.47; (+)-20 -epi-5-deoxystrigol, 3.69; ()-ent-5-deoxystrigol, 4.01; ()-ent-20 -epi-5-deoxystrigol, 4.18. RRt of strigone (1), strigol (2), and 5-deoxystrigol (4)
isolated from H. cordata root exudates were identical to those of
synthetic standards of (+)-strigone (1), (+)-strigol (2), and (+)-5deoxystrigol (3), respectively.
4.7. (+)-Strigone (1)
CD (c 0.0016 CH3CN) kmax (De) nm: 269 (1.22), 246 (52.28),
227 (5.73), 217 (5.28). GCEIMS, m/z (rel. int): 344 [M]+ (1),
247 (4), 230 (5), 215 (1), 97 (100). HRESITOFMS m/z:
345.1342 [M+H]+ (calcd. for C19H21O6, m/z 345.1333). 1H NMR
(500 MHz, CDCl3) d: 1.29 (3H, s, CH3), 1.31 (3H, s, CH3), 1.841.95
(2H, m, H-7), 2.03 (3H, t, J = 1.5 Hz, H-70 ), 2.432.49 (1H, brdt,
J = 5.5, 17.6 Hz, H-6a), 2.542.61 (1H, ddd, J = 5.4, 10.9, 17.6 Hz,
H-6b), 2.642.69 (1H, dt, J = 2.6, 17.2 Hz, H-4a), 2.892.94 (1H,
ddd, J = 0.6, 9.1, 17.2 Hz, H-4b), 3.693.74 (1H, m, H-3a), 5.69
(1H, brddd, J = 0.6, 2.3, 8.1 Hz, H-8b), 6.13 (1H, quin, J = 1.4 Hz, H20 ), 6.94 (1H, quin, J = 1.6 Hz, H-30 ), 7.51 (1H, d, J = 2.6 Hz, H-60 ).
13
C NMR (125 MHz, CDCl3) d: 10.7, 26.0, 26.7, 33.4, 34.6, 35.1,
36.4, 38.9, 87.3, 100.6, 111.9, 135.9, 138.0, 140.8, 151.7 (carbonyl
carbons could not be observed).
4.8. Synthesis of (+)-strigone, ()-ent-strigone, (+)-20 -epi-strigone and
()-ent-20 -epi-strigone
()-Strigol (2) and ()-20 -epi-strigol were synthesized as reported previously (Reizelman et al., 2000). ()-Strigol was puried
on a semi-preparative Chiralpak AD-H HPLC column
(250  10 mm, Daicel, Osaka, Japan) employing isocratic elution

63

with isopropanoln-hexane (1:1, v/v) at a ow rate of


2.4 ml min1. Compounds eluted from the column were monitored
with a UV detector. (+)-Strigol and ()-ent-strigol eluted as a single
peak at 14.7 and 9.3 min, respectively, were collected. ()-20 -epiStrigol was puried as above but using EtOHn-hexane (1:1, v/v)
as an eluent. (+)-20 -epi-Strigol and ()-ent-20 -epi-strigol eluted as
a single peak at 11.9 and 13.5 min, respectively, were collected.
The enantiomeric purities of four stereoisomers were estimated
to be >99% e.e., respectively by analytical HPLC using a Chiralpak
AD-H column (250  4.6 mm, 5 lm, Daicel, Osaka, Japan). (+)-Strigol: [a]D25 + 237 (c 0.55, CHCl3) ([a]D25 + 271 (c 0.50, CHCl3),
Hirayama and Mori, 1999); ()-ent-strigol: [a]D25  236 (c 0.56,
CHCl3) ([a]D25  277 (c 0.37, CHCl3), Hirayama and Mori, 1999);
(+)-20 -epi-strigol: [a]D25 + 124 (c 0.44, CHCl3) ([a]D24 + 145 (c
1.24, CHCl3), Hirayama and Mori, 1999); ()-ent-20 -epi-strigol:
[a]D25  133 (c 0.43, CHCl3) ([a]D25  141 (c 1.38, CHCl3), Hirayama and Mori, 1999).
To a solution of (+)-strigol (2) (11 mg, 0.032 mmol) in dry CH2Cl2 was added pyridinium dichromate (PDC, 18 mg, 0.049 mmol).
The reaction mixture was stirred at room temperature for 7 h, diluted with Et2O, and ltered. The solvent was then evaporated
and the residue was applied to a semi-preparative Inertsil SIL100A HPLC column (250  10 mm, 5 lm, GL Sciences, Osaka, Japan) employing isocratic elution with EtOHn-hexane (1:4, v/v)
at a ow rate of 2.4 ml min1 to give (+)-strigone (1)
(6.4 mg, 0.019 mmol, 59%). ()-ent-Strigol, (+)-20 -epi-strigol and
()-ent-20 -epi-strigol were oxidized and puried as above to give
()-ent-strigone, (+)-20 -epi-strigone and ()-ent-20 -epi-strigone,
respectively, in a similar yield.

4.8.1. (+)-Strigone (1)


CD (c 0.0050, CH3CN) kmax (De) nm: 269 (1.37), 246 (23.9), 228
(4.66), 217 (4.30). [a]D25 + 225 (c 0.32, CHCl3). EI-MS m/z (rel.
int.): 344 [M]+ (4), 247 (7), 230 (9), 215 (3), 97 (100). 1H NMR
(400 MHz, CDCl3) d: 1.29 (3H, s, CH3), 1.31 (3H, s, CH3), 1.841.97
(2H, m, H-7), 2.03 (3H, t, J = 1.5 Hz, H-70 ), 2.46 (1H, dt, J = 17.8,
5.2 Hz, H-6a), 2.58 (1H, ddd, J = 5.6, 10.7, 17.4 Hz, H-6b), 2.66
(1H, dt, J = 17.3, 2.7 Hz, H-4a), 2.91 (1H, dd, J = 9.3, 16.8 Hz, H4b), 3.693.74 (1H, m, H-3a), 5.69 (1H, dd, J = 1.7, 8.0 Hz, H-8b),
6.14 (1H, t, J = 1.3 Hz, H-20 ), 6.95 (1H, t, J = 1.6 Hz, H-30 ), 7.51 (1H,
d, J = 2.7 Hz, H-60 ). 13C NMR (100 MHz, CDCl3) d: 10.7, 26.0, 26.7,
33.4, 34.6, 35.1, 36.4, 38.9, 87.3, 100.6, 112.0, 135.9, 137.9, 140.8,
151.7, 162.7, 170.2, 170.6, 198.3.

4.8.2. ()-ent-Strigone
CD (c 0.0050, CH3CN) kmax (De) nm: 269 (1.17), 246 (24.4), 227
(4.06), 216 (2.85). [a]D25  222 (c 0.35, CHCl3). 1H and 13C NMR
and mass data were the same as for (+)-strigone.

4.8.3. (+)-20 -epi-Strigone


CD (c 0.0050, CH3CN) kmax (De) nm: 250 (12.2), 227 (18.1).
[a]D25 + 83 (c 0.22, CHCl3). EI-MS m/z (rel. int.): 344 [M]+ (4),
247 (8), 230 (9), 215 (3), 97 (100). 1H NMR (400 MHz, CDCl3) d:
1.28 (3H, s, CH3), 1.31 (3H, s, CH3), 1.841.96 (2H, m, H-7), 2.04
(3H, t, J = 1.5 Hz, H-70 ), 2.45 (1H, dt, J = 17.6, 5.2 Hz, H-6a), 2.58
(1H, ddd, J = 5.7, 10.5, 17.3 Hz, H-6b), 2.65 (1H, dt, J = 17.3,
2.7 Hz, H-4a), 2.91 (1H, dd, J = 9.0, 17.3 Hz, H-4b), 3.693.74 (1H,
m, H-3a), 5.69 (1H, dd, J = 1.7, 8.3 Hz, H-8b), 6.16 (1H, t,
J = 1.3 Hz, H-20 ), 6.92 (1H, t, J = 1.6 Hz, H-30 ), 7.48 (1H, d,
J = 2.7 Hz, H-60 ). 13C NMR (100 MHz, CDCl3) d: 10.8, 26.1, 26.8,
33.4, 34.7, 35.1, 36.4, 39.0, 87.3, 100.1, 112.0, 136.3, 138.1, 140.8,
151.0, 162.5, 169.9, 170.7, 198.3.

64

T. Kisugi et al. / Phytochemistry 87 (2013) 6064

4.8.4. ()-ent-20 -epi-Strigone


CD (c 0.0050, CH3CN) kmax (De) nm: 251 (11.4), 226 (16.3).
[a]D25  73 (c 0.18, CHCl3); 1H and 13C NMR and mass data were
the same as for (+)-20 -epi-strigone.
Acknowledgments
This work was supported by the Program for Promotion of Basic
and Applied Research for Innovations in Bio-oriented Industry and
in part by JSPS KAKENHI Grant Number 23380061.
References
Akiyama, K., Matsuzaki, K., Hayashi, H., 2005. Plant sesquiterpenes induce hyphal
branching in arbuscular mycorrhizal fungi. Nature 435, 824827.
Akiyama, K., Ogasawara, S., Ito, S., Hayashi, H., 2010. Structural requirements of
strigolactones for hyphal branching in AM fungi. Plant Cell Physiol. 51, 1104
1117.
Bouwmeester, H.J., Roux, C., Lopez-Raez, J.A., Bcard, G., 2007. Rhizosphere
communication of plants, parasitic plants and AM fungi. Trends Plant Sci. 12,
224230.
Cook, C.E., Whichard, L.P., Turner, B., Wall, M.E., Egley, G.H., 1966. Germination of
witchweed (Striga lutea Lour.): isolation and properties of a potent stimulant.
Science 154, 11891190.
Cook, C.E., Whichard, L.P., Wall, M.E., Egley, G.H., Coggon, P., Luhan, P.A., McPhail,
A.T., 1972. Germination stimulants. II. The structure of strigol a potent seed
germination stimulant for witchweed (Striga lutea Lour.). J. Am. Chem. Soc. 94,
61986199.
Delaux, P.-M., Xie, X., Timme, R.E., Puech-Pages, V., Dunand, C., Lecompte, E.,
Delwiche, C.F., Yoneyama, K., Bcard, G., Sjalon-Delmas, N., 2012. Origin of
strigolactones in the green lineage. New Phytol. 194, 857871.
Fernndez-Aparicio, M., Flores, F., Rubiales, D., 2009. Recognition of root exudates
by seeds of broomrape (Orobanche and Phelipanche) species. Ann. Bot. 103, 423
431.
Gomez-Roldan, V., Fermas, S., Brewer, P.B., Puech-Pags, V., Dun, E.A., Pillot, J.-P.,
Letisse, F., Matusova, R., Danoun, S., Portais, J.-C., Bouwmeester, H., Bcard, G.,
Beveridge, C.A., Rameau, C., Rochange, S.F., 2008. Strigolactone inhibition of
shoot branching. Nature 455, 189194.
Hirayama, K., Mori, K., 1999. Plant bioregulators, 5. Synthesis of (+)-strigol and (+)orobanchol, the germination stimulants, and their stereoisomers by employing
lipase-catalyzed asymmetric acetylation as the key step. Eur. J. Org. Chem. 1999,
22112217.
Jamil, M., Rodenburg, J., Charnikhova, T., Bouwmeester, H.J., 2011. Pre-attachment
Striga hermonthica resistance of New Rice for Africa (NERICA) cultivars based on
low strigolactone production. New Phytol. 192, 964975.
Jamil, M., Kanampiu, F.K., Karaya, H., Charnikhova, T., Bouwmeester, H.J., 2012.
Striga hermonthica parasitism in maize in response to N and P fertilisers. Field
Crops Res. 134, 110.
Koltai, H., 2011. Strigolactones are regulators of root development. New Phytol. 190,
545549.
Lpez-Rez, J.A., Charnikhova, T., Gmez-Roldn, V., Matusova, R., Kohlen, W., De
Vos, R., Verstappen, F., Puech-Pages, V., Bcard, G., Mulder, P., Bouwmeester, H.,
2008. Tomato strigolactones are derived from carotenoids and their
biosynthesis is promoted by phosphate starvation. New Phytol. 178, 863874.
Ma, Y.Q., Cheng, J.M., Inanaga, S., Shui, J.F., 2004. Induction and inhibition of Striga
hermonthica (Del.) Benth. germination by extracts of traditional Chinese
medicinal herbs. Agron. J. 96, 13491356.

Ma, Y., Shui, J., Inanaga, S., Cheng, J., 2005. Stimulatory effects of Houttuynia cordata
Thunb, on seed germiation of Striga hermonthica (Del.) Benth. Allelopathy J. 15,
4956.
Proust, H., Hoffmann, B., Xie, X., Yoneyama, K., Schaefer, D.G., Yoneyama, K., Nogu,
F., Rameau, C., 2011. Strigolactones regulate protonema branching and act as a
quorum sensing-like signal in the moss Physcomitrella patens. Development
138, 15311539.
Reizelman, A., Scheren, M., Nefkens, G.H.L., Zwanenburg, B., 2000. Synthesis of all
eight stereoisomers of the germination stimulant strigol. Synthesis 13, 1944
1951.
Ruyter-Spira, C., Kohlen, W., Charnikhova, T., van Zeijl, A., van Bezouwen, L., de
Ruijter, N., Cardoso, C., Lopez-Raez, J.A., Matusova, R., Bours, R., Verstappen, F.,
Bouwmeester, H., 2011. Physiological effects of the synthetic strigolactone
analog GR24 on root system architecture in Arabidopsis: another belowground
role for strigolactones? Plant Physiol. 155, 721734.
Sato, D., Awad, A.A., Chae, S.H., Yokota, T., Sugimoto, Y., Takeuchi, Y., Yoneyama, K.,
2003. Analysis of strigolactones, germination stimulants for Striga and
Orobanche, by high-performance liquid chromatography/tandem mass. J.
Agric. Food Chem. 51, 11621168.
Thuring, J.W.J.F., Nefkens, G.H.L., Zwanenburg, B., 1997. Asymmetric synthesis of all
stereoisomers of the strigol analogue GR24. Dependence of absolute
conguration on stimulatory activity of Striga hermonthica and Orobanche
crenata seed germination. J. Agric. Food Chem. 45, 22782283.
Ueno, K., Fujiwara, M., Nomura, S., Mizutani, M., Sasaki, M., Takikawa, H., Sugimoto,
Y., 2011. Structural requirements of strigolactones for germination induction of
Striga gesnerioides seeds. J. Agric. Food Chem. 59, 92269231.
Umehara, M., Hanada, A., Yoshida, S., Akiyama, K., Arite, T., Takeda-Kamiya, N.,
Magome, H., Kamiya, Y., Shirasu, K., Yoneyama, K., Kyozuka, J., Yamaguchi, S.,
2008. Inhibition of shoot branching by new terpenoid plant hormones. Nature
455, 195200.
Xie, X., Kusumoto, D., Takeuchi, Y., Yoneyama, K., Yamada, Y., Yoneyama, K., 2007.
20 -Epi-orobanchol and solanacol, two unique strigolactones, germination
stimulants for root parasitic weeds, produced by tobacco. J. Agric. Food Chem.
55, 80678072.
Xie, X., Yoneyama, K., Kurita, J., Harada, Y., Yamada, Y., Takeuchi, Y., Yoneyama, K.,
2009. 7-Oxoorobanchyl acetate and 7-oxoorobanchol as germination
stimulants for root parasitic plants from ax (Linum usitatissimum). Biosci.
Biotechnol. Biochem. 73, 13671370.
Xie, X., Yoneyama, K., Yoneyama, K., 2010. The strigolactone story. Annu. Rev.
Phytopathol. 48, 93117.
Yasuda, N., Sugimoto, Y., Kato, M., Inanaga, S., Yoneyama, K., 2003. (+)-Strigol, a
witchweed seed germination stimulant, from Menispermum dauricum root
culture. Phytochemistry 62, 11151119.
Yokota, T., Sakai, H., Okuno, K., Yoneyama, K., Takeuchi, Y., 1998. Alectrol and
orobanchol, germination stimulants for Orobanche minor, Allelopathy Journal
from its host red clover. Phytochemistry 49, 19671973.
Yoneyama, K., Takeuchi, Y., Ogasawara, M., Konnai, M., Sugimoto, Y., Sassa, T., 1998.
Cotylenins and fusicoccins stimulate seed germination of Striga hermonthica
(Del.) Benth and Orobanche minor Smith. J. Agric. Food Chem. 46, 15831586.
Yoneyama, K., Yoneyama, K., Takeuchi, Y., Sekimoto, H., 2007. Phosphorus
deciency in red clover promotes exudation of orobanchol, the signal for
mycorrhizal symbionts and germination stimulant for root parasites. Planta
225, 10311038.
Yoneyama, K., Xie, X., Kim, H.I., Kisugi, T., Nomura, T., Sekimoto, H., Yokota, T.,
Yoneyama, K., 2012. How do nitrogen and phosphorus deciencies affect
strigolactone production and exudation? Planta 235, 11971207.
Zwanenburg, B., Mwakaboko, A.S., Reizelman, A., Anilkumar, G., Sethumadhavan, D.,
2009. Structure and function of natural and synthetic signalling molecules in
parasitic weed germination. Pest Manag. Sci. 65, 478491.

Anda mungkin juga menyukai