11 - Co-Author - Picraviane A and B - Nortriterpenes With Limonoid-Like Skeletons Containing A Heptanolide E-Ring System From Picramnia Glazioviana

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Phytochemistry 163 (2019) 38–45

Contents lists available at ScienceDirect

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

Picraviane A and B: Nortriterpenes with limonoid-like skeletons containing a T


heptanolide E-ring system from Picramnia glazioviana
Leila Gimenesa,b, Joao Marcos Batista Juniora,c, Fernando Martins dos Santos Juniora,d,
Matheus da Silva Souzae, Liany Luna-Dulceyf, Javier Esteves Ellenae, Marcia Regina Cominettif,
Maria Fatima das Graças Fernandes da Silvaa, Paulo Cezar Vieiraa,g, Joao Batista Fernandesa,1,
Dan Staerkb,∗,1
a
Department of Chemistry, Federal University of Sao Carlos (UFSCar), 13565-905, Sao Carlos, SP, Brazil
b
Department of Drug Design and Pharmacology, University of Copenhagen, DK-2100, Copenhagen, Denmark
c
Institute of Science and Technology, Federal University of Sao Paulo, 12231-280, Sao Jose dos Campos, SP, Brazil
d
Department of Organic Chemistry, Chemistry Institute, Fluminense Federal University (UFF), Niterói, RJ, Brazil
e
São Carlos Institute of Physics, University of São Paulo, 13566-590, São Carlos, SP, Brazil
f
Department of Gerontology, Federal University of Sao Carlos, 13565-905, Sao Carlos, SP, Brazil
g
School of Pharmaceutical Sciences of Ribeirao Preto, University of Sao Paulo, 14040-903, Ribeirao Preto, SP, Brazil

ARTICLE INFO ABSTRACT

Keywords: Two highly oxygenated nortriterpenes, picraviane A and B, were isolated from the ethanolic extract of Picramnia
Picramnia glazioviana Engl. (Picramniaceae) glazioviana Engl. The structures were determined by analysis of HRMS and 2D NMR spectroscopic data. Single-
Nortriterpenes crystal X-ray diffraction data was also obtained for picraviane B. The absolute configuration of both compounds
Picravianes A and B were assigned by comparison of experimental and calculated vibrational and electronic circular dichroism
Vibrational circular dichroism
spectroscopy, respectively. Picraviane A showed a moderate cytotoxic activity against the triple negative MDA-
Electronic circular dichroism
Triple negative breast cancer
MB-231 breast cancer cell line. These compounds represent an undescribed class of natural products with li-
monoid-like skeletons containing a heptanolide as the E-ring.

1. Introduction their continued importance as research subject within chemical, bio-


logical, and medicinal research (Cragg and Newman, 2005; Pérez et al.,
The biosynthetic machinery in plants produces a plethora of spe- 2014; Yuan et al., 2016). The isolation of natural products is therefore,
cialized metabolites, and natural products are characterized by complex besides contributing to the understanding of natural selection and
chemical scaffolds that are highly oxygenated, contains multiple ring coevolution in plants, an important source of bioactive leads for drug
systems, and have specific stereochemical configurations. Thus, nature discovery and development.
has provided us with a large pool of specialized metabolites with unique The genus Picramnia, historically placed into the Simaroubaceae
and specific properties that have driven the pharmaceutical industry's family, was together with Alvaradoa, and recently Nothotalisia, moved
discoveries over the past century (Harvey et al., 2015; Kingston, 2011; into the Picramniaceae family due to the absence of quassinoids, which
Newman and Cragg, 2016). Thus, according to Newman and Cragg are characteristic taxonomical markers of Simaroubaceae (Diaz et al.,
(2016), 49 percent of all small-molecule drug entities approved for 2004; Fernando and Quinn, 1995; Jiwajinda et al., 2001; Thomas,
cancer treatment from the 1940s to 2014 were natural products or 2011). Picramnia comprises approximately 40 species distributed ex-
natural products derivatives. However, nature as a source of drug-like clusively in America tropics, i.e., from United States, Mexico and Cen-
chemicals is still largely unexplored, and from an estimated 250,000 to tral America to Colombia, Venezuela, Peru, Brazil, Paraguay, and Ar-
500,000 species of higher plants, only 5–15% have already been in- gentina in South America (Fernando and Quinn, 1995; Pirani and
vestigated phytochemically. In addition, the increasing number of un- Devecchi, 2016). Some Picramnia species have traditionally been used
described chemical plant constituents reported every year emphasize as vermicides (Solis et al., 1995) and for treatment of skin diseases

Corresponding author. Department of Drug Design and Pharmacology, Faculty of Health and Medical Sciences, University of Copenhagen, Universitetsparken 2,

DK-2100, Copenhagen, Denmark.


E-mail address: ds@sund.ku.dk (D. Staerk).
1
These senior authors contributed equally to this work.

https://doi.org/10.1016/j.phytochem.2019.03.024
Received 10 January 2019; Received in revised form 25 March 2019; Accepted 26 March 2019
0031-9422/ © 2019 Elsevier Ltd. All rights reserved.
L. Gimenes, et al. Phytochemistry 163 (2019) 38–45

(Balderrama et al., 2001; Robledo et al., 2015). Earlier phytochemical signals resembled the 1H NMR signals observed for the A and A′ rings of
investigations of Picramnia species showed the presence of anthraqui- the well-known limonoid limonin (Tavares et al., 2016), and this sub-
nones, anthrones, oxanthrones glycosides, and triterpenoids (Diaz et al., structure was further supported by HMBC correlations from H-1 to C-10
2004; Jacobs, 2003; Rodríguez-Gamboa et al., 2001, 1999), but no at δ 45.4, from H-1, H-2, and H-25 to the carbonyl group C-3 at δ 169.7,
undescribed natural products have been reported from Picramnia spe- and from H-23 and H-24 to each other's carbon atoms as well as to C-4
cies in recent years (Martínez et al., 2013; Robledo et al., 2015). at δ 80.8 and C-5 at δ 53.7.
Triple-negative breast cancer (TNBC) is a highly aggressive subtype The B, C and D-rings were subsequently identified based on com-
of breast cancer, with a low survival rate for patients (Zhang et al., bined use of COSY and HMBC spectra. Thus, the COSY spectrum
2012). This cancer form is characterized by the lack of expression of showed three spin systems corresponding to the C-5 – C-6 – C-7, the C-9
estrogen receptors, progesterone receptors, and human epidermal – C-11 – C-12, and the C-15 – C-16 backbones. These spin systems were
growth factor receptor-2 (Gluz et al., 2009). The treatment against this assembled based on HMBC correlations from H-7, H-9 and H-15β to the
subtype of breast cancer is performed through conventional che- quaternary C-8 (δ 43.4), from H-7, H-12 and H-15β to the quaternary C-
motherapy - but the treatment is causing several adverse reactions. 14 (δ 41.0), from H-11, H-12 and H-15α to the olefinic C-13 (δ 138.9),
Thus, there is an urgent need for discovery of natural product-based from H-12 and H-16 to C-18 (δ 118.2), and from H-15α and H-16 to C-
drug leads against TNBC. 17 (δ 149.0) – as can be seen from the HMBC correlations provided in
As part of our phytochemical studies of Picramnia species, and Fig. 1B. HMBC correlations from H-26 and H-27 to C-8 and C-14, re-
contributions to its chemosystematical classification, two undescribed spectively, supported the position of these methyl groups, and the up-
nortriterpenes, picraviane A (1) and B (2), with an undescribed skeleton field shift of H-7 and HMBC correlations from H-7 and H-2ʺ/H-6ʺ to the
were isolated from Picramnia glazioviana Engl. (Picramniaceae). In this carbonyl group at δ 165.0 (C-7ʺ) showed the presence of a benzoyl
paper, the isolation and full structure elucidation of compounds 1 and 2 group attached to C-7. The highly substituted heptanolide ring system
by 2D NMR spectroscopy, single-crystal X-ray crystallography, elec- (E-ring), was established based on HMBC correlations from H-19 to the
tronic circular dichroism and vibrational circular dichroism are de- C-17 – C-18 double bond, to the quaternary C-20 (δ 44.0) and the two
scribed, and a possible biogenetic pathway is proposed. The cytotoxic methyl groups attached to C-20 (δ 24.7 and 28.1, C-28 and C-29 re-
activity of 1 and 2 were also evaluated against the triple-negative breast spectively), and to C-2' (δ 169.9) of the acetyl group. Furthermore, the
cancer (TNBC) cell line MDA-MB-231. H-21 methylene group was identified as intervening C-20 and C-22
based on HMBC correlation to these two carbon atoms (Fig. 1). The
complete assignments of all 1H and 13C NMR resonances are given in
2. Results and discussion
Table 1, and all 2D correlations as well as 2D NMR spectra are given in
Supplementary data Table S1 and Figs. S1–S17, respectively.
Crude ethanol extract of P. glazioviana was redissolved in methanol-
The relative stereochemical configuration at C-1, C-5, C-7, C-8, C-9, C-
water (1:3) and successively extracted with hexane, dichloromethane
10, and C-14 of 1 was established using the NOESY spectrum. Thus,
and ethyl acetate. The dichloromethane phase was subjected to re-
NOESY cross peaks between H-1 and H-5, H-9 and H-23 as well as be-
versed-phase HPLC, leading to isolation of two undescribed nor-
tween H-27 and H-9, H-15α, H-16, and H2''/H-6'', confirmed these hy-
triterpenes, picraviane A (1) and B (2) containing a heptanolide E-ring
drogens to be on the same side of the molecule, i.e., tentatively assigned as
(Fig. 1A). The structure elucidation of 1 and 2 are described below.
α. Likewise, NOESY cross peaks between H-26 and H-6β, H-7, H-11β, H-
15β and H-25 confirmed these hydrogens to be on the other side of the
2.1. Structure elucidation of picravianes A (1) and B (2) molecule, i.e. tentatively assigned as β. These NOEs established the re-
lative stereochemistry of 1 as 1S*,5R*,7R*,8R*,9S*,10R*,14R*. However,
26
The dextrorotatory nortriterpene 1 ([α] D : +43.53 (c. 0.76, the flexibility of the E-ring did not allow unambiguous conclusions of the
MeOH)) was assigned the molecular formula C38H46O9 as determined configuration at C-19 based on the observed NOEs. Thus, density func-
by HRESIMS ([M+H]+ 647.3210, calcd. 647.3220, ΔM + 0.7 ppm), tional theory calculations of IR and VCD spectra were performed for the
indicating an index of hydrogen deficiency of 16. The structure was lowest-energy conformations of (1S,5R,7R,8R,9S,10R,14R,19S)-1 (Fig. 2A)
established based on spectroscopic data from 1D 1H NMR, 13C NMR, and its 19-epimer (1S,5R,7R,8R,9S,10R,14R,19R)-1 (Fig. 2B).
and DEPT135 as well as 2D COSY, HSQC, HMBC, H2BC, and NOESY The calculated [wB97XD/PCM(CHCl3)/6-311G**] IR and VCD
experiments. spectra of these two epimers (solid red lines) - as well as VCD of their
The 1H NMR spectrum of 1 displayed two methyl groups at δ 1.11 respective enantiomers (dotted lines) - are shown in Fig. 3A and Fig. 3B,
and 1.15 (3H, s, H-24 and H-23, respectively), an oxymethine at δ 4.21 respectively. The spatial arrangement of the A′- and E-rings combined
(1H, dd, J = 3.7 and 1.6 Hz, H-1), a methine at δ 2.48 (1H, dd, J = 13.8 with the opposite configuration for the chiral center at C-19 resulted in
and 2.0 Hz, H-5), and two diastereotopic proton pairs corresponding to some VCD bands with reversed signs, especially fundamentals 132 and
a methylene (δ 2.64, 1H, dd, J = 16.8 and 1.6 Hz, H-2α and δ 2.95, 1H, 171, which could be safely used to discriminate the two diastereomers.
dd, J = 16.8 and 3.7 Hz, H-2β) and an oxymethylene (δ 4.59, 1H, d, These bands arise mainly from the CeO stretching vibration of the
J = 13.3 Hz, H-25α and δ 4.63, 1H, d, J = 13.3 Hz, H-25β). These acetyl group at C-19 coupled with CH3 deformation modes of ring E

Fig. 1. A: Structure of picraviane A (1) and picraviane B (2). B: Key COSY (green) and HMBC (blue arrows pointing from H to C) correlations observed for 1. (For
interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

39
L. Gimenes, et al. Phytochemistry 163 (2019) 38–45

Table 1 and calculated IR and VCD data for compound (+)-1 (Fig. 3A, solid red
1
H NMR (600 MHz) and13C NMR (150 MHz) data of compounds 1 and 2 in lines), established the absolute stereochemical configuration of 1 as
chloroform-d. 1S,5R,7R,8R,9S,10R,14R,19S. In order to further support this assign-
Position 1 2 ment, a quantitative analysis was performed using the confidence level
algorithm (Debie et al., 2011) by assessing the neighbourhood simila-
δCa δH (J in Hz) a
δC δH (J in Hz) rities between the experimental spectrum and the computed spectra for
both epimers at C-19. The results indicated similarity values (Σ) for the
1 80.6 4.21 dd (3.7, 1.6) 80.3 4.05 dd (3.7, 1.6)
2 35.5 β: 2.95 dd (16.8, 3.7) 35.9 β: 2.93 dd (16.7, 3.7) IR spectra of 96.6 for 19S and 95.2 for 19R. The similarity measure-
α: 2.64 dd (16.8, 1.6) α: 2.61 dd (16.7, 1.5) ments of the VCD spectra confirmed compound (+)-1 as
3 169.7 170.2 1S,5R,7R,8R,9S,10R,14R,19S (ΣVCD = 74.5 for 19S and ΣVCD = 32.4 for
4 80.8 81.1
19R).
5 53.7 2.48 dd (14.4, 2.0) 60.7 1.71 br d (11.7, 1.8) 26
6 25.5 β: 1.97 td (14.4, 2.0) 18.3 1.62 m The structure of the dextrorotatory nortriterpene 2 ([α] D : +63.93
α: 1.72 dt (14.4, 2.6) (c. 0.41, MeOH)) was elucidated on the basis of spectroscopic data from
7 74.7 5.36 t (2.6) 33.6 1.64 m
1D 1H NMR, 13C NMR, and DEPT135 as well as 2D COSY, HSQC,
8 43.4 39.5
9 42.4 2.82 dd (10.3, 7.8) 44.7 2.14 dd (11.4, 5.8) HMBC, H2BC, and NOESY experiments. Compound 2 had a molecular
10 45.4 45.3 formula of C31H42O7 as determined by HRESIMS ([M
11 26.1 β: 2.55 ddd (19.4, 10.3, 4.1) 26.2 β: 2.40 m +H]+ = 527.2998, calcd. 527.3008, ΔM + 1.0 ppm) indicating an
α: 2.43 ddd (19.4, 7.8, 4.1) α: 2.19 m index of hydrogen deficiency of 11. The 13C NMR spectrum was iden-
12 116.8 5.77 br t (4.1) 118.4 5.95 br t (3.8)
13 138.9 136.9
tical to that observed for 1, except for the absence of signals for the
14 41.0 40.7 benzoyl group as well as the disappearance of the oxymethine observed
15 25.8 β: 1.87 m 26.8 β: 1.72 m for C-7 in 1 and the appearance of an additional methylene signal for C-
α: 1.38 dt (12.8, 3.8) α: 1.45 dt (12.8, 3.6) 7 in 2. This indicated that 2 has the same structure as 1, except for the
16 24.9 2.23 m 25.4 2.31 m
absence of the benzoyl group, which is in agreement with the 2J and 3J
17 149.0 148.8
18 118.2 119.2 HMBC correlations observed from H-5 at δ 1.71 to C-6 at δ 18.3 and C-7
19 73.3 5.28 s 73.6 5.36 s at δ 33.6.
20 44.0 44.2 A single-crystal X-ray diffraction study confirmed the molecular
21 43.8 β: 2.15 d (12.0) 43.8 β: 2.21 d (12.1) structure of 2 and allowed us to establish the same relative configura-
α: 2.31 d (12.0) α: 2.42 d (12.1)
22 169.1 169.3
tion as that observed for 1. Picraviane B crystallized in the non-cen-
23 30.6 1.15 s 31.0 1.28 s trosymmetric monoclinic space group P 21. The asymmetric unit com-
24 21.6 1.11 s 21.6 1.14 s prises one molecule of picraviane B and one water molecule from the
25 65.9 β: 4.63 d (13.3) 66.1 β: 4.60 d (13.1) crystallization medium (Fig. 4). A water molecule mediated the junc-
α: 4.59 d (13.3) α: 4.52 d (13.1)
tion of two picraviane B molecules through classical OeH•••O hydrogen
26 17.6 0.88 s 17.3 0.80 s
27 24.8 1.23 s 21.0 1.13 s
bonds [OW–HWA•••O1: 2.924(9) Å, ∠ = 172.9(2)° and OW–HWB•••O2:
28 24.7 1.14 s 24.7 1.15 s 2.930(9) Å, ∠ = 166.2(2)°]. As a consequence, two adjacent chains are
29 28.1 0.93 s 28.0 0.93 s placed together (Fig. 5A). These chains grow along the a-axis and are
1ʹ 20.8 2.01 s 20.8 2.04 s responsible to give rise to a C22 (8) graph set highlighted as an orange
2ʹ 169.9 170.0 dashed line in Fig. 5B.
1ʺ 130.1
2ʺ/6ʺ 129.6 8.02 XX′-system
Molecules are inserted on both sides of this chain forming a plane.
3ʺ/5ʺ 128.7 7.46 AA′-system As expected from the planar conformation of the raw material, a
4ʺ 133.6 7.59 tt (8.4, 1.2) layered structure could be observed. The dominant driving force in the
7ʺ 165.0 3D network assemblies are the non-classical CeH•••O hydrogen bonds
b between nearby picraviane B molecules functional groups [C2eH2A
Apparent splittings, s = singlet, d = doublet, t = triplet, br = broad.
a •••O6: 3.282(8) Å, ∠ = 146.1(2)°; C16eH16A•••O5: 3.327(1) Å,
Referenced relative to chloroform signal at δ 77.0 and 7.26 for13C and 1H,
∠ = 139.1(2)° and C21eH21A•••O5: 3.563(9) Å, ∠ = 168.8(2)°] and
respectively.
π•••π interactions connecting the aromatic centers of picraviane B from
adjacent planes, forming supramolecular wave-like layers (Fig. 5C).
(fundamental 132), as well as from CH2 twisting of ring A′ and the in-
This structure is in agreement with analysis of 2D NMR experiments,
plane CeH bending of H-19 coupled with the CH2 wagging of ring E
and all 2D correlations as well as 2D NMR spectra are given in
(fundamental 171). The excellent correlation between experimental

Fig. 2. Structures of the lowest-energy conformers of (+)-picraviane A (1) calculated for (1S,5R,7R,8R,9S,10R,14R,19S)-1 (A) and its 19-epimer
(1S,5R,7R,8R,9S,10R,14R,19R)-1 (B) at the wB97XD/PCM(CHCl3)/6-311G** level.

40
L. Gimenes, et al. Phytochemistry 163 (2019) 38–45

Fig. 3. Comparison of the observed IR and


VCD spectra of (+)-picraviane A (1) with
the calculated [wB97XD/PCM(CHCl3)/6-
311G**] IR and VCD spectra. A: The
(1S,5R,7R,8R,9S,10R,14R,19S)-diaster-
eomer (solid lines) and the corresponding
enantiomer (1R,5S,7S,8S,9R,10S,14S,19R)
(dotted line). B: The (1S,5R,7R,8R,9S,
10R,14R,19R)-diastereomer (solid lines)
and the corresponding enantiomer
(1R,5S,7S,8S,9R,10S,14S,19S) (dotted
lines). Numbers represent selected vibra-
tional modes.

Supplementary data Table S2 and Figs. S19–S34, respectively. The


complete assignment of all 1H and 13C resonances are given in Table 1.
Crystal data, structure refinement and crystallographic parameters are
shown in Supplementary data Table S3.
The relative stereochemical configuration of 2 was unambiguously
established to be 1S*,5R*,8R*,9S*,10R*,14S*,19S* by combined use of
2D NOESY and X-ray crystallography. The absolute stereochemical
configuration was determined by comparing experimental UV and ECD
data acquired in methanol with calculated UV and ECD spectra using
time-dependent density functional theory for the lowest-energy con-
formation of (1S,5R,8R,9S,10R,14S,19S)-2. The good agreement be-
tween experimental and calculated data, i.e., positive and negative
Cotton effects at 250 and 220 nm, respectively (Fig. 6), confirmed this
as the absolute stereochemical configuration of (+)-2. The ECD data
calculated for the 19R epimer of compound 2 is presented in the
Supplementary data Fig. S40. It is noteworthy that compound 2, despite
Fig. 4. ORTEP type diagram of the asymmetric unit of the monohydrate of 2. having the same spatial arrangement of atoms as that in compound 1,
Ellipsoids are drawn as 50% probability levels. has a change in the stereodescriptor for C-14 due to removal of the

Fig. 5. A: One water molecule mediates the junction of two


picraviane B (2) molecules giving rise to parallel chains. B:
The 1D chains grow along the a-axis in which multiple
water and picraviane B molecules are forming an ac-plane
via OeH•••O hydrogen bonds. C: 3D arrangement of crys-
talline picraviane B monohydrate showing the supramole-
cular wave-like layers.

41
L. Gimenes, et al. Phytochemistry 163 (2019) 38–45

Fig. 6. A: Comparison of the observed UV and ECD spectra of (+)-picraviane B (2) with the calculated [CAM-B3LYP/PCM(MeOH)/TZVP//B3LYP/6-31G*] UV and
ECD spectra of the (1S,5R,8R,9S,10R,14S,19S) stereo isomer. B: Structure of lowest-energy conformer identified for (1S,5R,8R,9S,10R,14S,19S).

benzoyl group at C-7 in compound 2. (intermediate D) results in E with the 29 carbon atoms as seen for
nortriterpenes. An allylic oxidation of E would lead to the formation of
intermediate F that would convert to G by a Baeyer-Villiger oxidation,
2.2. Biogenesis which presents the main skeleton of both 1 and 2. Finally, picraviane A
(1) and B (2) can be formed by further oxidations at C-19 and C-7,
Although the presence of the A and A′ rings in 1 and 2 are common which keeps the same stereochemistry at C-5, C-9, C-25, C-26 and C-27
in some types of limonoids in Rutaceae family, this is the first work as their suggested precursor. HRESIMS spectra of intermediates D, E
reporting the presence of these rings in nortriterpenes, which suggests and F are shown in Supplementary data Figs S37–S39.
that similar enzymes as those in the Sapindales order can be involved in
the formation of these compounds. Besides, limonoids and quassinoids
are biosynthetically related to euphol/tirucallol type precursors after 2.3. Cytotoxic and morphology evaluations
losing 4 and 10 carbons from the side-chain, respectively (Dewick,
2002). However, the formation of Picraviane A and B is very well Terpenoids are well-known for their cytotoxic activity (Chudzik
supported by the precursor oleanolic acid, a triterpene already isolated et al., 2015; Thoppil and Bishayee, 2011), and compounds 1 and 2 were
from some Picramnia species (Balderrama et al., 2001; Martínez et al., therefore evaluated for their cytotoxic activity against the MDA-MB-
2013), and also present in the crude ethanol extract of P. glazioviana 231 breast cancer cell line. Picraviane A exhibited moderate cytotoxic
(C30H47O3, [M-H]- = 455.3518, calcd. 435.3531, ΔM = + 2.8 ppm, activity with IC50 of 72.62 ± 0.18 μM, while picraviane B did not show
Supplementary data, Fig. S36) suggesting a plausible biogenetic significant activity within the tested range (IC50 > 100 μM). These
pathway for 1 and 2 as illustrated in Fig. 7. results indicate the benzoate unit present in 1 may be essential for the
Initially, the A and A′ rings can be formed by oxidation of C-1 and C- observed activity.
25 of the shown intermediate A, followed by two intramolecular cy- The moderate cytotoxicity of 1 on MDA-MB-231 cells is comparable
clization reactions and consequently loss of two water molecules to to the IC50 value reported for the natural triterpenoid 3-epimaslinic acid
form C containing the limonoid-type A and A′ rings. Allyllic oxidation isolated from Dillenia suffruticosa (Tor et al., 2015). These results em-
of C followed by decarboxylation and dehydration in ring D phasize the continued need for a deeper understanding of the effect of

Fig. 7. Biogenetic pathway proposed for 1 and 2.

42
L. Gimenes, et al. Phytochemistry 163 (2019) 38–45

Fig. 8. Effects of picraviane A (1) on cell


morphology of the triple negative MDA-MB-
231 breast cancer cell line. A: Cellular
morphology of MDA-MB-231 after 2, 4, 24
and 48 h of treatment with indicated con-
centrations of 1 [100 × amplification, scale
bar = 100 μm]. B: MDA-MB-231 cells
treated for 48 h with 1 at the three highest
concentration resulting in a decrease in cell
density compared to untreated cells (con-
trol) [200 × amplification, scale
bar = 50 μm].

inhibitors of cell lines from the aggressive triple negative breast cancer. compounds and negative ionization mode for the crude ethanol extract.
Hence, the effects on cell morphology of picraviane A (1) were eval- HPLC analyses were performed on a Shimadzu LC-10AD system con-
uated. As shown in Fig. 8A, incubation of compound 1 induced mor- sisting of a LC- 10AD pump (Kyoto, Japan), a SPD 10AD UV–Vis de-
phological changes on the tumor breast cells that are better visualized tector working with two wavelengths, and a SCL 10A VP system con-
after 24 h of treatment at the concentrations of 80 and 100 μM, where troller. This equipment was connected to a CBM – 10A and a Shimadzu
cells developed circular shapes accompanied by loss of cell density (red Class – VP software was used for data acquisition.
arrows). Moreover, in Fig. 8B after 48 h of treatment, compound 1 in- Optical rotations were measured at 589 nm using a
duced large morphological changes at 60, 80 and 100 μM, causing a Bellingham + Stanley ADP 410 polarimeter (Fischer Scientific,
total lack of adherence and appearance of round cells and cellular Hampton, NH, USA). IR and VCD spectra of compound 1 were recorded
debris, indicating cell death. with a Single-PEM ChiralIR-2X FT-VCD spectrometer (BioTools, Inc.,
Jupiter, FL, USA) using a resolution of 4 cm−1 and a collection time of
3. Conclusions 4 h. The optimum retardation of the ZnSe photoelastic modulator
(PEM) was set at 1400 cm−1. Minor instrumental baseline offsets were
In conclusion, the isolation of an undescribed class of nortriterpenes eliminated from the final VCD spectrum by subtracting the VCD spec-
with a limonoid-like skeleton containing a heptanolide E ring supports trum of compound 1 from that obtained for the solvent under the same
the classification of Picramnia sp. into the Picramniaceae family. In conditions. The IR and VCD spectra were recorded in a BaF2 cell with
addition to shear light on the chemodiversity in Picramnia spp., these 100 μm path length using CDCl3 as solvent. The sample was prepared
findings hold promises for identifying further nortriterpenes with po- using 5 mg of compound 1 dissolved in 120 μL of solvent. The IR
tential as lead compounds against triple negative breast cancer. spectrum of compound 2 was recorded on a Perkin Elmer Spectrum One
FT-IR spectrometer (Perkin Elmer, Waltham, MA, USA) using MeOH as
4. Material and methods solvent. UV and ECD spectra of 2 were recorded with a Jasco J-815
spectrometer (Jasco, Tokyo, Japan) in the 190–350 nm region using the
4.1. General experimental procedures following parameters: bandwidth 1 nm; response 1 s; scanning speed
100 nm/min; 3 accumulations; room temperature; sample in methanol
NMR spectra were acquired using a Bruker Avance III 600 MHz solution; 0.1 cm cell path length; concentration 0.32 mg/mL.
spectrometer equipped with a Bruker SampleJet sample changer and a
cryogenically cooled gradient inverse triple-resonance 1.7-mm TCI 4.2. Plant material and extraction
probe-head (Bruker Biospin, Rheinstetten, Germany) and on a 600 MHz
Bruker Avance III spectrometer equipped with a cryogenically cooled 5- Leaves of Picramnia glazioviana Engl. (Picramniaceae) were col-
mm DCH probe optimized for 13C and 1H, and a 400 MHz Bruker lected in Piuma and Marataizes, Espirito Santo, 3 km from Morro do
AvanceTM III NanoBay 9,4 T spectrometer. CDCl3 (δH = 7.26 ppm, Aghá, Rodovia ES-60, 20°52′S, 40°46′W, Atlantic Rainforest, Brazil in
δC = 77.0 ppm) was used as reference signal. Mass spectra were re- November 1988 and were identified by Dr. Jose Rubens Pirani. A
corded on a micrOTOF-Q II mass spectrometer (Bruker Daltonik GmbH, voucher specimen has been deposited at the Botanical Department, Sao
Bremen, Germany) equipped with an electrospray ionization (ESI) in- Paulo University, Sao Paulo, Brazil (Herbal SPF, collection number J. R.
terface. Mass spectra were acquired in positive ionization mode for pure Pirani et al., 2468). The leaves of P. glazioviana were dried in a

43
L. Gimenes, et al. Phytochemistry 163 (2019) 38–45

circulation oven at 40 °C and subsequently milled to yield 513 g of dry incorporated in HyperChem 8.0.10 software package. The puckering of
powdered plant material. This material was extracted with 2 l of rings A and E were also probed manually. A single conformer with
ethanol, filtered, and concentrated on a rotary evaporator under re- relative energy within 6 kcal mol−1 was selected and further geometry
duced pressure to yield 41.1 g of crude extract. optimized at the B3LYP/6-31G* level. This optimized conformer was
used for UV and ECD spectral calculations. TDDFT was employed to
4.3. Isolation and characterization data calculate the excitation energy (in nm) and rotatory strength R in the
dipole velocity (Rvel in cgs units: 10−40 esu2 cm2) form, at the CAM-
The crude ethanol extract was submitted to liquid-liquid parti- B3LYP level, combined with TZVP basis sets. The calculated rotatory
tioning by dissolving the extract in MeOH:H2O (1:3) and subsequently strengths from the first 30 singlet → singlet electronic transitions were
extracting sequentially with hexane, dichloromethane and ethyl acetate simulated into an ECD curve using Gaussian bands with a bandwidth of
(3 × 400 mL, for each solvent) - and then dried under reduced pressure. σ 0.25 eV. Cartesian coordinates of lowest-energy conformers are given
This procedure resulted in 1.53 g of dichloromethane extract. The di- in the last section named ‘CARTESIAN COORDINATES’ in
chloromethane extract was separated using semi-preparative reversed- Supplementary data.
phase HPLC on a Phenomenex Luna C18 column, 10 μm, 10 × 250 mm
(Phenomenex, Torrance, CA, USA) using isocratic elution with 4.5. Single crystal X-ray diffraction of picraviane B (2)
ACN:H2O (40:60) at a flow rate of 7 mL/min. Four injections of 100 μL
each of a solution of 370 mg/mL and collection based on chromato- During the extraction of the compound clear light colorless crystal
grams at 254 and 220 nm afford 24.4 mg of 1 and 8.2 mg of 2. plates were obtained by slow evaporation of a water:methanol solution
26 (8:2 v/v) at 5 °C. The crystallographic data for picraviane B (2) were
Picraviane A (1): white wax, [α] D : +43.53 (c. 0.76, MeOH). UV
collected at 293 K on an Bruker Enraf-Nonius diffractometer (Bruker
obtained from HPLC-PDA-HRESIMS analysis: λmax: 198, 232 and
Karlsruhe, Germany) with CCD detector system equipped with a Mo
255 nm. IR: (Fig. 2). 13C and 1H NMR data in Table 1. HRESIMS m/z [M
source (λ = 0.71073 Å). Data integration and cell determination were
+H]+ = 647.3210 (calculated for C38H47O9, 647.3220,
performed using the software Collect and Scalepack (Otwinowski and
ΔM = +0.7 ppm), [M + H-CH3CO2H]+ = 587.3003 (calculated for
Minor, 1997). Data reduction was carried out using the software Denzo-
C36H43O7, 587.3008, ΔM = +0.1 ppm).
26 SMN and Scalepack (Otwinowski and Minor, 1997). Using Olex2
Picraviane B (2): white wax, [α] D : +63.93 (c. 0.41, MeOH). UV (Dolomanov et al., 2009), the structure was solved (Olex2.solve struc-
obtained from HPLC-PDA-HRESIMS analysis: λmax: 229 and 252 nm. IR ture solution program, Bourhis et al., 2015) using Charge Flipping and
νmax: 2969, 1749, 1379, 1231, 1065 cm−1. 13C and 1H NMR data in the model obtained was refined by full–matrix least squares on F2 with
Table 1. HRESIMS m/z [M+H]+ = 527.2998 (calculated for C31H43O7, the SHELXL refinement package using Least Squares minimization
527.3008, ΔM = +1.0 ppm), [M + H-CH3CO2H]+ = 467.2791 (cal- (Sheldrick, 2015). Non-hydrogen atoms were refined with anisotropic
culated for C29H39O5, 467.2797, ΔM = +0.2 ppm). displacement parameters. The positions of hydrogen atoms were cal-
culated and were included in the structure factor calculations. All the
4.4. Computational methods, VCD and ECD spectra hydrogen atoms were placed in calculated positions and refined with
fixed individual displacement parameters [Uiso(H) = 1.2Ueq or
All density functional theory (DFT) and time-dependent DFT 1.5Ueq] according to the riding model. Molecular representations, ta-
(TDDFT) calculations were carried out at 298 K either in the gas phase, bles and pictures were generated by ORTEP-3 (Farrugia, 2012), Mer-
in chloroform or methanol solutions using the polarizable continuum cury 3.10.2 (Macrae et al., 2006) and Olex2 softwares. The Crystal-
model (PCM) in its integral equation formalism version (IEFPCM) in- lographic Information File was deposited in the Cambridge Structural
corporated in Gaussian 09 software. Calculations for compound 1 were Database (Groom et al., 2016) under the code CCDC 1862634. Copies of
performed for the arbitrarily chosen (1S,5R,7R,8R,9S,10R,14R,19S) the data can be obtained, free of charge, via www.ccdc.cam.ac.uk.
and (1S,5R,7R,8R,9S,10R,14R,19R) epimers. All conformational sear-
ches were carried out at the molecular mechanics level of theory with 4.6. Cytotoxicity assay
the Monte Carlo algorithm employing the MM + force field in-
corporated in the HyperChem 8.0.10 software package. The puckering MDA-MB-231 cells were obtained from Rio de Janeiro Cell Bank,
of rings A and E were also probed manually. For both epimers, three maintained in Dulbecco's modified eagle medium supplemented with
conformers with relative energy within 6 kcal mol−1 of the lowest- 10% fetal bovine serum and appropriate antibiotics. Cells were main-
energy conformer were selected and further geometry optimized at the tained at 37 °C in an incubator with 95% O2 and 5% CO2. The effects of
B3LYP/6-31G* level. In both cases, a single conformer was identified picraviane A (1) and B (2) on the viability of MDA-MB-231 was de-
with relative energy < 2.2 kcal mol−1, corresponding to more than termined by colorimetric assays (Luna-Dulcey et al., 2018; Mosmann,
97% of the total Boltzmann distribution. The lowest-energy conformer 1983) using 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bro-
for each configuration was selected for IR and VCD spectral calculations mide (MTT). MDA-MB-231 cells were seeded (1 × 104 cells/100 μL)
at the B3LYP/6-31G* level. These conformers were also further geo- into 96-well plates and incubated until 80% confluence. Subsequently,
metry optimized and had their IR and VCD spectra calculated at the cells were exposed to increasing concentrations (3.12–100 μM) of
wB97XD/PCM(CHCl3)/6-311G** level. IR and VCD spectra were cre- compounds. Cells were incubated for 48 h under the same conditions as
ated using dipole and rotational strengths from Gaussian, which were described above. Cytotoxicity assay was performed in comparison to
calculated at the same level used during geometry optimization steps the wells where the vehicle solvent (1% DMSO) was added instead of
and converted into molar absorptivities (M−1 cm−1). Each spectrum the tested compounds. Cells were treated with MTT (1 mg/mL) for 4 h.
was plotted as a sum of Lorentzian bands with half-widths at half- Medium was removed from the wells and the formed crystals were
maximum of 6 cm−1. The calculated wavenumbers were multiplied dissolved in DMSO. Absorbances were measured using a Labtech LT-
with a scaling factor of 0.97 and the final spectra were plotted using 4000 ELISA plate reader (Labtech, Heathfield, UK) at a wavelength of
Origin 8 software. 540 nm.
Calculations for compound 2 were performed for the arbitrarily
chosen (1S,5R,8R,9S,10R,14S,19S) configuration, based on the relative 4.7. Cell morphology of picraviane A (1)
configuration assigned by X-ray crystallography. The conformational
search was carried out at the molecular mechanics level of theory with MDA-MB-231 cells (1 × 105/well) were seeded in a 12 well-plate
the Monte Carlo algorithm employing the MM + force field and after 24 h and exposed to increasing concentrations of the

44
L. Gimenes, et al. Phytochemistry 163 (2019) 38–45

compound (25–100 μM) for additional 48 h. Cells were examined using Harvey, A.L., Edrada-Ebel, R., Quinn, R.J., 2015. The re-emergence of natural products
a Nikon T5100 inverted optical microscope (Nikon, Tokyo, Japan). for drug discovery in the genomics era. Nat. Rev. Drug Discov. 14, 111–129. https://
doi.org/10.1038/nrd4510.
Photos were taken at 2, 4, 24 and 48 h, and the efficiency of treatment Jacobs, H., 2003. Comparative phytochemistry of Picramnia and Alvaradoa, genera of the
was compared to control cells (with no treatment). newly established family Picramniaceae. Biochem. Syst. Ecol. 31, 773–783. https://
doi.org/10.1016/S0305-1978(02)00268-5.
Jiwajinda, S., Santisopasri, V., Murakami, A., Hirai, N., Ohigashi, H., 2001. Quassinoids
Conflicts of interest from Eurycoma longifolia as plant growth inhibitors. Phytochemistry 58, 959–962.
https://doi.org/10.1016/S0031-9422(01)00333-8.
The authors declare no conflict of interest. Kingston, D.G.I., 2011. Modern natural products drug discovery and its relevance to
biodiversity conservation. J. Nat. Prod. 74, 496–511. https://doi.org/10.1021/
np100550t.
Acknowledgments Luna-Dulcey, L., Tomasin, R., Naves, M.A., Da Silva, J.A., Cominetti, M.R., 2018.
Autophagy-dependent apoptosis is triggered by a semi-synthetic [6]-gingerol ana-
logue in triple negative breast cancer cells. OncoTarget 9, 30787–30804. https://doi.
L. G. would like to thank São Paulo Research Foundation (FAPESP)
org/10.18632/oncotarget.25704.
for the doctoral scholarships in Brazil and abroad (2013/20458-1 and Macrae, C.F., Edgington, P.R., McCabe, P., Pidcock, E., Shields, G.P., Taylor, R., Towler,
2016/18024-1). The authors would like to thank FAPESP (Grants # M., Van De Streek, J., 2006. Mercury: Visualization and analysis of crystal structures.
2012/25299-6, 2014/25222-9, 2015/07089-2 and 2016/23794-0) the J. Appl. Crystallogr. 39, 453–457. https://doi.org/10.1107/S002188980600731X.
Martínez, M.L., von Poser, G., Henriques, A., Gattuso, M., Rossini, C., 2013.
National Council for Scientific and Technological Development (CNPq, Simaroubaceae and Picramniaceae as potential sources of botanical pesticides. Ind.
J.E. grant #305190/2017-2), Coordination for the Improvement of Crops Prod. 44, 600–602. https://doi.org/10.1016/j.indcrop.2012.09.015.
Higher Education Personnel (CAPES, Finance Code 001) and the Centre Mosmann, T., 1983. Rapid colorimetric assay for cellular growth and survival:
Application to proliferation and cytotoxicity assays. J. Immunol. Methods 65, 55–63.
for Scientific Computing (NCC/GridUNESP). The 600 MHz NMR was https://doi.org/10.1016/0022-1759(83)90303-4.
acquired through a grant from ´Apotekerfonden af 1991′, The Carlsberg Newman, D.J., Cragg, G.M., 2016. Natural products as sources of new drugs from 1981 to
Foundation, and the Danish Agency for Science, Technology and 2014. J. Nat. Prod. 79, 629–661. https://doi.org/10.1021/acs.jnatprod.5b01055.
Otwinowski, Z., Minor, W., 1997. Macromolecular crystallography Part A. Methods
Innovation via the National Research Infrastructure funds. Enzymol. 276, 307–326. https://doi.org/10.1016/S0076-6879(97)76066-X.
Pérez, L.B., Still, P.C., Naman, C.B., Ren, Y., Pan, L., Chai, H.B., de Blanco, E.J.C., Ninh,
Appendix A. Supplementary data T.N., Van Thanh, B., Swanson, S.M., Soejarto, D.D., Kinghorn, A.D., 2014.
Investigation of Vietnamese plants for potential anticancer agents. Phytochemistry
Rev. 13, 727–739. https://doi.org/10.1007/s11101-014-9335-7.
Supplementary data to this article can be found online at https:// Pirani, J.R., Devecchi, M.F., 2016. Flora das cangas da Serra dos Carajás, Pará, Brasil:
doi.org/10.1016/j.phytochem.2019.03.024. Picramniaceae. Rodriguesia 67, 1447–1449. https://doi.org/10.1590/2175-
7860201667546.
Robledo, S.M., Cardona, W., Ligardo, K., Henao, J., Arbeláez, N., Montoya, A., Alzate, F.,
References Pérez, J.M., Arango, V., Vélez, I.D., Sáez, J., 2015. Antileishmanial effect of 5,3'-
hydroxy-7,4'-dimethoxyflavanone of Picramnia gracilis Tul. (Picramniaceae) fruit: in
Balderrama, L., Braca, A., Garcia, E., Melgarejo, M., Pizza, C., De Tommasi, N., 2001. vitro and in vivo studies. Adv. Pharmacol. Sci. 2015, 978379. https://doi.org/10.
Triterpenes and anthraquinones from Picramnia sellowii Planchon in Hook 1155/2015/978379.
(Simaroubaceae). Biochem. Syst. Ecol. 29, 331–333. https://doi.org/10.1016/S0305- Rodríguez-Gamboa, T., Fernandes, J.B., Fo, E.R., Da Silva, M.F.D.G., Vieira, P.C., Castro,
1978(00)00054-5. O., 1999. Two anthrones and one oxanthrone from Picramnia teapensis.
Bourhis, L.J., Dolomanov, O.V., Gildea, R.J., Howard, J.A.K., Puschmann, H., 2015. The Phytochemistry 51, 583–586. https://doi.org/10.1016/S0031-9422(99)00023-0.
anatomy of a comprehensive constrained, restrained refinement program for the Rodríguez-Gamboa, T., Fernandes, J.B., Rodrigues Filho, E., Da Silva, M.F., Vieira, P.C.,
modern computing environment - Olex2 dissected. Acta Crystallogr. Sect. A Found Barrios, Ch., M., Castro-Castillo, O., Victor, S.R., Pagnocca, F.C., Bueno, O.C.,
Crystallogr. 71, 59–75. https://doi.org/10.1107/S2053273314022207. Hebling, M.J.A., 2001. Triterpene benzoates from the bark of Picramnia teapensis
Chudzik, M., Korzonek-Szlacheta, I., Król, W., 2015. Triterpenes as potentially cytotoxic (Simaroubaceae). J. Braz. Chem. Soc. 12, 386–390. https://doi.org/10.1590/S0103-
compounds. Molecules 20, 1610–1625. https://doi.org/10.3390/ 50532001000300010.
molecules20011610. Sheldrick, G.M., 2015. SHELXT - integrated space-group and crystal-structure determi-
Cragg, G.M., Newman, D.J., 2005. Biodiversity: a continuing source of novel drug leads. nation. Acta Crystallogr. Sect. A Found. Crystallogr. 71, 3–8. https://doi.org/10.
Pure Appl. Chem. 77, 7–24. https://doi.org/10.1351/pac200577010007. 1107/S2053273314026370.
Debie, E., De Gussem, E., Dukor, R.K., Herrebout, W., Nafie, L.A., Bultinck, P., 2011. A Solis, P.N., Ravelo, A.G., Gonzalez, A.G., Gupta, M.P., Phillipson, J.D., 1995. Bioactive
confidence level algorithm for the determination of absolute configuration using anthraquinone glycosides from Picramnia antidesma ssp. fessonia. vol. 38. pp.
vibrational circular dichroism or Raman optical activity. ChemPhysChem 12, 477–480. https://doi.org/10.1016/0031-9422(94)00598-N.
1542–1549. https://doi.org/10.1002/cphc.201100050. Tavares, L.C., Fernandes, T.S., Ilha, V., Neto, A.T., Dos Santos, E.W., Burrow, R.A., Duarte,
Dewick, P.M., 2002. Medicinal Natural Products. A Biosynthetic Approach, 2 ed. John F.A., Flores, E.M.M., Silva, U.F., Mostardeiro, M.A., Morel, A.F., 2016. Limonin de-
Wiley & Sons Ltd, University of Nottingham, UK. rivatives: Synthesis using methodology in solution and heterogeneous medium and
Diaz, F., Chai, H.B., Mi, Q., Su, B.N., Vigo, J.S., Graham, J.G., Cabieses, F., Farnsworth, evaluation of the antimicrobial activity. J. Braz. Chem. Soc. 27, 161–178. https://doi.
N.R., Cordell, G.A., Pezzuto, J.M., Swanson, S.M., Kinghorn, A.D., 2004. Anthrone org/10.5935/0103-5053.20150266.
and oxanthrone C-glycosides from Picramnia latifolia collected in Peru. J. Nat. Prod. Thomas, W.W., 2011. Nothotalisia, a new genus of Picramniaceae from tropical America.
67, 352–356. https://doi.org/10.1021/np030479j. Brittonia 63, 51–61. https://doi.org/10.1007/s12228-010-9130-8.
Dolomanov, O.V., Bourhis, L.J., Gildea, R.J., Howard, J.A.K., Puschmann, H., 2009. Thoppil, R.J., Bishayee, A., 2011. Terpenoids as potential chemopreventive and ther-
OLEX2: a complete structure solution, refinement and analysis program. J. Appl. apeutic agents in liver cancer. World J. Hepatol. 3, 228–249. https://doi.org/10.
Crystallogr. 42, 339–341. https://doi.org/10.1107/S0021889808042726. 4254/wjh.v3.i9.228.
Farrugia, L.J., 2012. WinGX and ORTEP for Windows: an update. J. Appl. Crystallogr. 45, Tor, Y.S., Yazan, L.S., Foo, J.B., Wibowo, A., Ismail, N., Cheah, Y.K., Abdullah, R., Ismail,
849–854. https://doi.org/10.1107/S0021889812029111. M., Ismail, I.S., Yeap, S.K., 2015. Induction of apoptosis in MCF-7 cells via oxidative
Fernando, E.S., Quinn, C.J., 1995. Picramniaceae, a new family, and a recircumscription stress generation, mitochondria-dependent and caspase-independent pathway by
of Simaroubaceae. Taxon 44, 177–181. http://www.jstor.org/stable/1222440. ethyl acetate extract of Dillenia suffruticosa and its chemical profile. PLoS One 10,
Gluz, O., Liedtke, C., Gottschalk, N., Pusztai, L., Nitz, U., Harbeck, N., 2009. Triple-ne- 1–25. https://doi.org/10.1371/journal.pone.0127441.
gative breast cancer - current status and future directions. Ann. Oncol. 20, Yuan, H., Ma, Q., Ye, L., Piao, G., 2016. The traditional medicine and modern medicine
1913–1927. https://doi.org/10.1093/annonc/mdp492. from natural products. Molecules 21. https://doi.org/10.3390/molecules21050559.
Groom, C.R., Bruno, I.J., Lightfoot, M.P., Ward, S.C., 2016. The Cambridge structural Zhang, L., Hao, C., Dong, G., Tong, Z., 2012. Analysis of clinical features and outcome of
database. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 72, 171–179. 356 triple-negative breast cancer patients in China. Breast Care 7, 13–17. https://doi.
https://doi.org/10.1107/S2052520616003954. org/10.1159/000336539.

45

You might also like