Natural nitric oxide (NO) inhibitors from Aristolochia mollissima

Zhen Dong a, Qiong Gua, Bao Chengb, Zhong-Bin Chenga, Gui-Hua Tanga, Zhang-Hua Suna, Jun-Sheng Zhanga, Jing-Mei Baoa and Sheng Yin*a
aGuangdong Provincial Key Laboratory of New Drug Design and Evaluation, School of Pharmaceutical Sciences, Sun Yat-sen University, Guangzhou, Guangdong 510006, P. R. China. E-mail: yinsh2@mail.sysu.edu.cn; Fax: +86-20-39943090; Tel: +86-20-39943090
bInstitute of Chinese Medical Sciences, Guangdong Pharmaceutical University, Guangzhou, Guangdong 510006, P. R. China

Received 2nd September 2014 , Accepted 16th October 2014

First published on 16th October 2014


Abstract

Six new sesquiterpenoids, aristomollins A–F (1–6), and 24 known analogues (7–30) were isolated from the leaves and stems of Aristolochia mollissima. Their structures were elucidated by spectroscopic analysis, and the absolute configurations of compounds 2–5 were determined by the chemical correlations and quantum chemical ECD calculations. Compound 1 represented an unprecedented 5,6-seco-4,5-cyclohumulane skeleton. All the compounds were examined for their inhibitory effects on the nitric oxide (NO) production induced by lipopolysaccharide (LPS) in BV-2 microglial cells, and compounds 4, 9, 28, and 30 exhibited pronounced inhibition of NO production with IC50 values in the range of 5.7–9.9 μM, being more active than the positive control, quercetin (IC50 = 15.7 μM).


Introduction

Neuroinflammation has been considered as one of the pathological factors in neurodegenerative diseases including Alzheimer's disease (AD), Parkinson's disease (PD), stroke, dementia, and amyotrophic lateral sclerosis (ALS).1,2 Activation of brain microglial cells and consequent overexpression of proinflammatory mediators such as nitric oxide (NO) are involved in the neuroinflammatory process. In addition, NO and superoxide lead to the formation of peroxynitrite, which results in numerous oxidation and potential destruction of host cellular constituents causing dysfunctional critical cellular processes, cell signaling pathway disruption, and brain cell death via cell apoptosis and necrosis.3,4 There are numerous evidences suggesting that suppression of proinflammatory mediators (such as NO) and further inhibition of the neuroinflammatory responses in microglia could attenuate the severity or delay the progress of these neurodegenerative disorder.5,6 Therefore, suppression of NO production in microglial cells might be an important and attractive therapeutic target for the treatment of neurodegenerative diseases.

Aristolochia mollissima Hance (Aristolochiaceae), a perennial shrub, is known as “Xun Gu Feng” in traditional Chinese medicine for its analgesic, anti-cancer, anti-rheumatic, and anti-inflammatory effects.7 Previous investigations on this plant revealed a number of sesquiterpenes, aristolochic acids, and aristolactams, some of which exhibited anti-inflammatory,8 antimicrobial,9 and analgesic activities.10 In our screening program aiming the discovery of natural NO inhibitors, the EtOAc fraction of the ethanolic extract of A. mollissima showed a certain inhibitory activity against the lipopolysaccharide (LPS)-induced NO production in BV-2 microglial cells. Subsequent chemical investigation led to the isolation of six new sesquiterpenoids (1–6), together with 24 known ones (7–30). Bioassay verified that compounds 3–5, 9, 17, and 28–30 were responsible for the NO inhibitory activities of the EtOAc fraction, with IC50 values ranging from 5.7 to 29.8 μM. Herein, details of the isolation, structural elucidation, and NO inhibitory activities of these compounds are described.

Results and discussion

The air-dried powder of the leaves and stems of A. mollissima was extracted with 95% EtOH at room temperature (rt) to give a crude extract, which was suspended in H2O and successively partitioned with EtOAc and n-BuOH. Various column chromatographic separations of the EtOAc extract afforded compounds 1–30 (Fig. 1).
image file: c4ra09612f-f1.tif
Fig. 1 Structures of compounds 1–30.

Compound 1, a colorless oil, had a molecular formula C15H20O, as determined by HRESIMS ion at m/z 239.1401 [M + Na]+ (calcd 239.1406). The IR absorption band at 1690 cm−1 indicated the presence of a carbonyl group. The 1H NMR spectrum showed two olefinic methyl singlets [δH 1.81 (3H, H3-14) and 1.44 (3H, H3-15)], a terminal double bond [δH 5.15 (1H, s, H-5a) and 4.87 (1H, s, H-5b)], a formyl proton [δH 9.51 (H-1)], three olefinic protons [δH 7.21 (s, H-3), 5.14 (dd, J = 7.6, 7.6 Hz, H-7), and 4.85 (dd, J = 7.4, 7.4 Hz, H-11)], and a series of aliphatic methylene multiplets. The 13C NMR spectrum, in combination with DEPT experiments, resolved 15 carbon resonances attributable to a highly conjugated aldehyde (δC 195.6), a terminal double bond (δC 112.0 and 146.4), three trisubstituted double bonds, two olefinic methyls, and four sp3 methylenes. As five of the six degrees of unsaturation were consumed by four double bonds and a carbonyl group, the remaining degree of unsaturation required the presence of an additional ring. In the 1H–1H COSY spectrum two structural fragments a (C-7–C-8–C-9) and b (C-11–C-12–C-13) were first established by the correlations observed (Fig. 2). The connectivities of the structural fragments a, b, the double bonds, the methyls, and the formyl group were achieved by analysis of the HMBC correlations (Fig. 2). In particular, HMBC correlations of H3-14/C-4, C-6, and C-7, H2-5/C-3, C-4, and C-6, and H-1/C-2, C-3, and C-13 incorporated Δ2, Δ4, and Δ6 between C-7 and C-13. Moreover, HMBC correlations from H3-15 to C-9, C-10, and C-11 further linked C-9 and C-11 via C-10 to afford an 11-membered macro ring. The geometry of Δ6 was assigned as Z by NOESY correlation between H-7 and CH3-14, while the geometries of Δ2 and Δ10 were both assigned as E by NOESY correlations of H-1/H-3 and H2-12/CH3-15, respectively. Thus, the structure of 1 was established as depicted and given the trivial name aristomollin A. Compound 1 featured an unprecedented 5,6-seco-4,5-cyclohumulane skeleton biogenetically related to co-isolated compounds 7 and 8 (Fig. 1).


image file: c4ra09612f-f2.tif
Fig. 2 Selected 1H–1H COSY (image file: c4ra09612f-u1.tif) and HMBC (image file: c4ra09612f-u2.tif) correlations of 1–3.

image file: c4ra09612f-f3.tif
Fig. 3 Key NOE correlations (image file: c4ra09612f-u3.tif) of 2, 3, and 5.

Compound 2, a colorless oil, had a molecular formula C15H22O4, as established by HRESIMS and ESIMS. The IR spectrum exhibited absorption bands for OH (3349 cm−1) and lactone (1761 cm−1) functionalities. The 1H NMR spectrum showed two methyl singlets [δH 1.80 (3H, H3-13) and 1.33 (3H, H3-14)], a terminal double bond [δH 4.92 (2H, brs, H-12)], two protons bonded to carbons bearing heteroatoms [δH 5.35 (dd, J = 5.1, 2.9 Hz, H-6) and 4.03 (dd, J = 8.0, 6.6 Hz, H-1)], and a series of aliphatic methylene multiplets. The 13C NMR spectrum, in combination with DEPT experiments, resolved 15 carbon resonances attributable to one carbonyl, a terminal double bond, two sp3 quaternary carbons, four sp3 methines (two bearing heteroatoms) four sp3 methylenes, and two methyls. As two of the five degrees of unsaturation were accounted for a double bond and a carbonyl, the remaining three degrees of unsaturation required 2 to be tricyclic. The aforementioned information was in support of a eudesmane-type sesquiterpene with a lactone ring. Detailed 2D NMR analyses (1H–1H COSY, HSQC, and HMBC) permitted the establishment of the gross structure of 2 as depicted in Fig. 2. The relative configuration of 2 was determined by analysis of the NOESY data and pyridine-induced solvent shifts. The cis-fused A/B ring system was established by the strong NOESY correlation of H-5/CH3-14, which was supported by the diagnostic carbon chemical shift of CH3-14 at δC 25.0, as the CH3-14 in trans-eudesmanes usually resonated at around δC 14.0.11–15 The NOESY correlations of H-5/H-6 and H-9α, H-9α/CH3-13, H-1/H-8β, and CH3-14/H-2α and H-3α indicated that the H-1, H-6, and the isopropenyl group were co-facial and arbitrarily assigned in α-orientation (Fig. 3). As no convincing evidence was observed in the NOESY spectrum to assign the configuration of 4-OH, the 1H NMR data of 2 was measured in CDCl3 and C5D5N to obtain the pyridine-induced solvent shifts.16,17 The solvent shifts of H-6 (ΔδCDCl3 − C5D5N = −0.35) and H-5 (ΔδCDCl3 − C5D5N = −0.34), indicating that the 4-OH/H-6 were 1,3-diaxial-oriented while 4-OH/H-5 were co-facial. Thus, 4-OH was assigned in α-orientation. The absolute configuration (AC) of 2 was determined by comparing its experimental electronic circular dichroism (ECD) spectrum with those calculated by the time-dependent density functional theory (TDDFT). In Fig. 4, the experimental ECD spectrum of 2 showed first negative and second positive Cotton effects at 230 and 192 nm, respectively, which matched the calculated ECD curve for 2a, an isomer with a 1R, 4R, 5S, 6R, 7R, and 10S configuration, indicating that 2 possessed the same AC. Thus, compound 2 was assigned as depicted and named aristomollin B.


image file: c4ra09612f-f4.tif
Fig. 4 Experimental ECD spectra (190–400 nm) of 2 and TDDFT calculated ECD spectra for 2a (1R, 4R, 5S, 6R, 7R, 10S) and enantiomer of 2a.

Compound 3, a colorless oil, had a molecular formula C15H24O2, as established by HRESIMS and ESIMS. The 1D NMR data of 3 were similar to those of aristoyunolin G (9)18 except for the absence of signals for the formyl group and the presence of the signals for a hydroxymethyl group [δH 4.15 (1H, d, J = 14.1 Hz, H-14a) and 4.04 (1H, d, J = 14.1 Hz, H-14b); δC 67.6], indicating 3 was a formyl-reduced derivative of 9. This was supported by the HMBC correlations of H-14/C-3, C-4, and C-5, H-3/C-14, and H-5/C-14 (Fig. 2). The chemical transformation of 9 to 3 by NaBH4 reduction further secured the structure of 3. As the AC of 9 was assigned as 5R, 10R, and 12R, the AC of 3 was consequently determined as depicted. Compound 3 was given the trivial name aristomollin C.

Aristomollin D (4) was found to possess the molecular formula C17H24O3 on the basis of HRESIMS data. The 1H and 13C NMR spectra of 4 showed high similarity to those of 9 except for the presence of an additional acetyl group signals [δH 1.99 (3H, s); δC 21.3 and 170.5], which indicated that 4 was an acetylated derivative of 9. This was supported by the severely downfield-shifted H-12 signal in 4 with respect to that in 9 (δH 4.96 in 4; δH 3.82 in 9) and by the HMBC correlation from H-12 to the carbonyl group (δC 170.5). The AC of 4 was assigned to be the same as that of 9 based on the chemical transformation of 9 to 4 by acetylation.

The molecular formula of aristomollin E (5) was deduced as C15H24O2 by HRESIMS data. Its 1D NMR spectra bore a resemblance to those of aristoyunolin H (10)18 except for the absence of signals for the formyl group and the presence of a hydroxymethyl group [δH 4.13 (1H, d, J = 13.7 Hz, H-14a) and 4.06 (1H, d, J = 13.7 Hz, H-14b); δC 67.8], indicating 5 was a formyl-reduced derivative of 10. This was supported by the HMBC correlations of H-14/C-3, C-4, and C-5, H-3/C-14, and H-5/C-14. The AC of 5 was determined to be the same as that of 10 (5S, 10S, and 12R) on the basis of the chemical transformation of 10 to 5 by NaBH4 reduction.

Compound 6 had a quasimolecular ion peak [M + Na]+ at m/z 259.1663 in the HRESIMS, corresponding to the molecular formula C15H24O2. The IR absorption bands at 3426 and 1718 cm−1 showed the presence of the OH and carbonyl groups. The 1H NMR spectrum showed three methyl singlets [δH 1.11 (H3-15), 0.98 (H3-12) 0.95 (H3-13)], a formyl doublet [δH 9.54 (d, J = 3.0 Hz, H-14)], and a number of aliphatic protons. The 15 carbon resonances were classified by DEPT experiments as three methyls, four sp3 methylenes, five sp3 methines, two sp3 quaternary carbons, and a formyl group. The above-mentioned information was very similar to that of 22,19 an aromadendrane sesquiterpenoid co-isolated in the current study, except for the presence of a formyl group [δH 9.54; δC 203.1] and a sp3 methine (δC 60.3) in 6 instead of a tertiary methyl (δC 24.4) and an oxygenated quaternary carbon (δC 80.3) in 22, indicating that 6 was a 4-dehydroxyl-14-oxidation derivative of 22. The HMBC correlations from the formyl proton (H-14) to C-3, C-4, and C-5, from H-3 to C-14, and from H-5 to C-14 afforded the gross structure as depicted. The NOESY interactions of H-1 with H-4, H-6, H-9α, and CH3-15 indicated that these protons were co-facial and arbitrarily assigned in α-orientation. The large coupling constant between H-5 and H-6 (J = 9.6 Hz) indicated a trans-relationship of these protons,19 and therefore H-5 was assigned in β-orientation. The NOESY correlations of H-5/CH3-13 and H-6/H-7 indicated the cis-cyclopropane moiety was β-oriented. Thus, compound 6 was deduced as depicted and named as aristomollin F.

The known compounds madolin W (7),20 madolin H (8),21 aristoyunnolin G (9),18 aristoyunnolin H (10),18 aristoyunnolin E (11),22 madolin F (12),21 aristolactone (13),19 versicolactone B (14),23 madolin U (15),22 aristoyunnolin B (16),22 (+)-isobicyclogermacrenal (17),24 madolin K (18),19 madolin T (19),25 spathulenol (20),26 15-hydroxyspathulenol (21),27 aromadendrane-4β,10β-diol (22),19 (−)-alloaromadendrane-4β,10β-diol (23),28 versicolactone C (24),23 manshurolide (25),29 aristoyunnolin F (26),22 versicolactone D (27),30 aristophyllide A (28),31 aristophyllide B (29),31 and aristoloterpenate-I (30)32 were identified by comparison of their NMR data with those in the literature.

Compounds 1–30 were evaluated for their inhibitory effects on the NO production in LPS-induced BV-2 microglial cells using the Griess assay.5 Compounds 1, 2, 6–8, 10–16, and 18–27 were inactive (<50% inhibition at 50 μM), while compounds 3, 5, 17, and 29 showed moderate inhibitory activities with IC50 values ranging from 15.7–29.8 μM. Compounds 4, 9, 28, and 30 showed remarkable inhibitory activities with IC50 values of 9.0, 9.9, 5.7, and 8.7 μM, respectively, being more active than the positive control quercetin (IC50 = 15.7 μM), a well-known NO inhibitor (Table 3). The inhibitory curves of 4, 28, and quercetin were represented in Fig. 5. To investigate whether the inhibitory activities of the active compounds were generated from their cytotoxicity, the effects of compounds 3–5, 9, 17, and 28–30 on LPS-induced BV-2 microglial cell viability were measured using the MTT method. These eight compounds (up to 80 μM) did not show any significant cytotoxicity with LPS treatment for 24 h.


image file: c4ra09612f-f5.tif
Fig. 5 The inhibitory curves of compounds 4, 28, and quercetin (positive control) on LPS-induced NO Production in BV-2 Cells.

Conclusions

In summary, six new sesquiterpenoids and 24 known analogues were isolated from leaves and stems of A. mollissima. Their structures were elucidated by spectroscopic analysis, chemical correlations, and quantum chemical ECD calculations. Compound 1 represented an unprecedented 5,6-seco-4,5-cyclohumulane skeleton biogenetically related to the co-isolated cyclohumulane sesquiterpenoids, 7 and 8. All the compounds were examined for their inhibitory effects on the nitric oxide (NO) production induced by lipopolysaccharide (LPS) in BV-2 microglial cells, and compounds 4, 9, 28, and 30 exhibited pronounced inhibition on NO production with IC50 values in the range of 5.7–9.9 μM, being more active than the positive control, quercetin (IC50 = 15.7 μM). As the NO-suppression activity may attenuate the severity or delay the progress of neurodegenerative disorder, the evaluation of these active compounds on certain neurodegenerative diseases models, such as AD and PD, needs further exploration.

Experimental section

General experimental procedures

Optical rotations were measured on a Perkin-Elmer 341 polarimeter, and CD spectra were obtained on an Applied Photophysics Chirascan spectrometer. UV spectra were recorded on a Shimadzu UV-2450 spectrophotometer. IR spectra were determined on a Bruker Tensor 37 infrared spectrophotometer. NMR spectra were measured on a Bruker AM-400 spectrometer at 25 °C. ESIMS was measured on a Finnigan LCQ Deca instrument, and HRESIMS was performed on a Waters Micromass Q-TOF instrument. A Shimadzu LC-20 AT equipped with a SPD-M20A PDA detector was used for HPLC and a YMC-pack ODS-A column (250 × 10 mm, S-5 μm, 12 nm) were used for semi-preparative HPLC separation. Silica gel (300–400 mesh, Qingdao Haiyang Chemical Co., Ltd.), C18 reversed-phase silica gel (12 nm, S-50 μm, YMC Co., Ltd.), Sephadex LH-20 gel (Amersham Biosciences), and MCI gel (CHP20P, 75–150 μm, Mitsubishi Chemical Industries Ltd.) were used for column chromatography. All solvents used were of analytical grade (Guangzhou Chemical Reagents Company, Ltd.).

Plant material

Leaves and stems of A. mollissima were collected in March 2013 from Jiangxi Province, P. R. China, and were authenticated by Prof. You-Kai Xu of Xishuangbanna Tropical Botanical Garden, Chinese Academy of Sciences. A voucher specimen (accession number: XGF201303) has been deposited at the School of Pharmaceutical Sciences, Sun Yat-sen University.

Extraction and isolation

The air-dried powder of leaves and stems of A. mollissima (5.0 kg) was extracted with 95% EtOH (4 × 10 L) at rt to give a crude extract (382 g), which was suspended in H2O (1.5 L) and successively partitioned with EtOAc (3 × 1.5 L) and n-BuOH (3 × 1.5 L). The EtOAc extract (175 g) was subjected to MCI gel column chromatography (CC) eluted with a MeOH/H2O gradient (1[thin space (1/6-em)]:[thin space (1/6-em)]9 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0) to afford five fractions (IV). Fr. I (5.6 g) was chromatographed over a C18 reversed-phase (RP-C18) column eluted with MeOH/H2O (5[thin space (1/6-em)]:[thin space (1/6-em)]5 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0) to afford five fractions (Fr. Ia–Ie). Fr. Ia (1.1 g) was separated on silica gel CC (PE/EtOAc, 6[thin space (1/6-em)]:[thin space (1/6-em)]1) to give 14 (150 mg). Fr. Ie (1.6 g) was separated on silica gel CC (PE/acetone, 3[thin space (1/6-em)]:[thin space (1/6-em)]1), followed by a Sephadex LH-20 column using ethanol as eluent to give 15 (22 mg), 16 (30 mg), 24 (15 mg), and 2 (7 mg). Fr. II (10.2 g) was subjected to silica gel CC (PE/EtOAc, 8[thin space (1/6-em)]:[thin space (1/6-em)]1 → 1[thin space (1/6-em)]:[thin space (1/6-em)]2) to give four fractions (Fr. IIa–IId). Fr. IIb (3.4 g) was purified on silica gel CC (PE/EtOAc, 5[thin space (1/6-em)]:[thin space (1/6-em)]1) to obtain 7 (50 mg), 18 (42 mg), and 6 (18 mg). Fr. IIc (0.9 g) was applied to silica gel CC (CH2Cl2/acetone, 25[thin space (1/6-em)]:[thin space (1/6-em)]1 → 5[thin space (1/6-em)]:[thin space (1/6-em)]1) to yield 9 (104 mg) and 10 (4 mg). Fr. IId (0.6 g) was chromatographed over a C18 reversed-phase (RP-C18) column eluted with MeOH/H2O (6[thin space (1/6-em)]:[thin space (1/6-em)]4 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0) to afford 3 (7 mg) and 5 (15 mg). Fr. III (28 g) was subjected to silica gel CC (PE/EtOAc, 5[thin space (1/6-em)]:[thin space (1/6-em)]1 → 1[thin space (1/6-em)]:[thin space (1/6-em)]1) to give ten fractions (Fr. IIIa–IIIj). Fr. IIIc (1.9 g) was subjected to a RP-C18 silica gel CC (MeOH/H2O, 6[thin space (1/6-em)]:[thin space (1/6-em)]4 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0), followed by a silica gel CC (PE/acetone, 20[thin space (1/6-em)]:[thin space (1/6-em)]1 → 1[thin space (1/6-em)]:[thin space (1/6-em)]1) to afford 19 (50 mg), 20 (122 mg), and 25 (35 mg). Fr. IIIe (230 mg) was chromatographed over silica gel CC (CH2Cl2/MeOH, 200[thin space (1/6-em)]:[thin space (1/6-em)]1) to yield 12 (3 mg). Fr. IV (44 g) was chromatographed over an MCI gel column eluted with a gradient of MeOH/H2O (6[thin space (1/6-em)]:[thin space (1/6-em)]4 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0) to give eight fractions (Fr. IVa–IVh). Fr. IVa (2.3 g) was separated over RP-C18 CC using a gradient of MeOH/H2O (6[thin space (1/6-em)]:[thin space (1/6-em)]4 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0) to yield 17 (300 mg) and 1 (22 mg). Fr. IVb (8.8 g) was subjected successively to a silica gel CC (PE/EtOAc, 60[thin space (1/6-em)]:[thin space (1/6-em)]1 → 5[thin space (1/6-em)]:[thin space (1/6-em)]1), a RP-18 silica gel CC (MeOH/H2O, 7[thin space (1/6-em)]:[thin space (1/6-em)]3 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0), and a Sephadex LH-20 column (EtOH) to yield 8 (40 mg), 13 (1.2 g), 21 (15 mg), 11 (12 mg), and 22 (6 mg). Fr. IVc (2.7 g) was applied to silica gel CC (PE/EtOAc, 40[thin space (1/6-em)]:[thin space (1/6-em)]1 → 1[thin space (1/6-em)]:[thin space (1/6-em)]1) to give Fr. IVc1–IVc3. Fr. IVc1 (840 mg) was separated on a silica gel CC (PE/EtOAc, 50[thin space (1/6-em)]:[thin space (1/6-em)]1 → 30[thin space (1/6-em)]:[thin space (1/6-em)]1) to give 26 (7.1 mg) and 27 (15 mg). Further purification of Fr IVc2 (1.1 g) by silica gel CC (PE/CHCl3, 4[thin space (1/6-em)]:[thin space (1/6-em)]1) afforded 4 (10.2 mg), 28 (10 mg), and 23 (4 mg). Fr IVc3 (600 mg) was separated over RP-C18 CC using a gradient of MeOH/H2O (7[thin space (1/6-em)]:[thin space (1/6-em)]3 → 10[thin space (1/6-em)]:[thin space (1/6-em)]0) to yield 29 (5.3 mg) and 30 (5 mg). The purity of compounds 1–30 was greater than 95% as determined by 1H NMR spectra (ESI).
Aristomollin A (1). Colorless oil; UV (MeOH) λmax (log[thin space (1/6-em)]ε) 228 (3.99) nm; IR (KBr) νmax 2923, 2853, 1690, 1459, 1377, 1219, and 1125 cm−1; 1H and 13C NMR data, see Tables 1 and 2; positive ESIMS m/z 217.2 [M + H]+; HRESIMS m/z 239.1401 [M + Na]+ (calcd for C15H20ONa, 239.1406).
Table 1 1H NMR data of compounds 1–6a
No. 1b 2c 3b 4b 5b 6b
a Data were recorded at 400 MHz, chemical shifts are in ppm, coupling constant J is in Hz.b In CDCl3.c In C5D5N.
1 9.51, s 4.03, dd (8.0, 6.6) 5.84, dd (17.6, 10.8) 5.86, dd (17.6, 10.9) 5.86, dd (17.6, 10.7) 2.01, m
2a   α 1.91, m 5.03, d (17.6) 5.03, dd (17.6, 1.1) 4.99, d (17.6) 1.59, m
2b   β 2.21, m 4.97, d (10.8) 4.95, dd (10.9, 1.1) 4.93, d (10.7)  
3a 7.21, s α 2.49, m 5.33, s 6.20, s 5.32, s 1.81, m
3b   β 2.59, m 4.97, s 6.19, s 4.94, s  
4           2.55, ddd (15.6, 8.1, 3.0)
5a 5.15, s 2.55, d (5.1) 2.59, s 3.39, s 2.61, s 1.46, m
5b 4.87, s          
6   5.35, dd (5.1, 2.9)       0.51, dd (9.6, 9.6)
7 5.14, dd (7.6, 7.6) 2.35, m 5.61, brs 5.63, dd (3.2, 3.2) 5.55, brs 0.66, ddd (10.6, 9.6, 6.2)
2.23, m 1.53, m 2.09, m 2.07, m 2.06, m 1.85, m
  1.82, m       0.94, m
2.10, m 2.31, m 1.64, m 1.38, m 1.30, m 1.57, m
  1.10, m 1.38, m 1.34, m 1.62, m 1.73, m
11a 4.85, dd (7.4, 7.4)   2.13, m 2.07, m 2.12, m  
11b     1.87, dd (13.6, 10.0) 1.85, d (14.7) 2.01, m  
12a 2.28, m 4.92, brs 3.86, m 4.96, m 3.89, m 0.98, s
12b            
13a 2.17, m 1.80, s 1.16, d (6.1) 1.15, d (6.2) 1.13, d (5.9) 0.95, s
13b            
14a 1.81, s 1.33, s 4.15, d (14.1) 9.63, s 4.13, d (13.7) 9.54, d (3.0)
14b     4.04, d (14.1)   4.06, d (13.7)  
15 1.44, s   0.93, s 0.72, s 0.90, s 1.11, s
      12-OAc 1.99, s    


Table 2 13C NMR (100 MHz) data of compounds 1–6
No. 1a 2b 3a 4a 5a 6a
a In CDCl3.b In C5D5N.
1 195.6, CH 67.4, CH 146.8, CH 145.5, CH 146.8, CH 57.6, CH
2 144.3, C 26.2, CH2 111.4, CH2 111.7, CH2 111.2, CH2 26.6, CH2
3 154.4, CH 27.0, CH2 114.2, CH2 137.2, CH2 114.1, CH2 26.0, CH2
4 146.4, C 75.9, C 149.2, C 150.7, C 149.8, C 60.3, CH
5 112.0, CH2 52.6, CH 49.1, CH 43.3, CH 49.9, CH 39.4, CH
6 132.6, C 79.3, CH 135.3, C 133.2, C 135.7, C 31.8, CH
7 134.8, CH 44.0, CH 125.7, CH 126.1, CH 124.6, CH 27.1, CH
8 24.0, CH2 19.5, CH2 22.8, CH2 22.8, CH2 22.9, CH2 20.2, CH2
9 38.4, CH2 32.6,CH2 28.5, CH2 27.9, CH2 28.6, CH2 44.3, CH2
10 134.2, C 37.0, C 38.5, C 38.3, C 38.6, C 75.1, C
11 126.9, CH 146.2, C 46.1, CH2 42.1, CH2 46.8, CH2 19.7, C
12 24.3, CH2 111.4, CH2 64.8, CH 69.1, CH 67.0, CH 28.6, CH3
13 26.7, CH2 21.7, CH3 22.7, CH3 19.8, CH3 23.2, CH3 15.9, CH3
14 13.9, CH3 25.0, CH3 67.6, CH2 194.1, CH 67.8, CH2 203.1, CH
15 14.4, CH3 178.9, C 26.4, CH3 25.8, CH3 25.8, CH3 20.2, CH3
      –OAc 21.3, CH3    
        170.5, C    


Table 3 IC50 values of the active compounds on LPS-induced NO production in BV-2 cells
Compound IC50a (μM) Compound IC50a (μM)
a Values are represented as means ± SD based on three independent experiments.b Positive control.
3 25.0 ± 3.1 28 5.7 ± 1.5
4 9.0 ± 1.6 29 15.7 ± 2.5
5 29.8 ± 3.2 30 8.7 ± 2.2
9 9.9 ± 0.9 Quercetinb 15.7 ± 2.1
17 19.5 ± 2.7    


Aristomollin B (2). Colorless oil; [α]20D +77 (c 0.23, CHCl3); IR (KBr) νmax 3349, 2934, 2865, 1761, 1453, 1218, and 1044 cm−1; 1H and 13C NMR data, see Tables 1 and 2; positive ESIMS m/z 289.1 [M + Na]+; HRESIMS m/z 289.1418 [M + Na]+ (calcd for C15H22O4Na, 289.1416).
Aristomollin C (3). Colorless oil; [α]20D +210 (c 0.14, CHCl3); IR (KBr) νmax 3377, 2964, 2924, 2865, 1642, 1059, and 912 cm−1; 1H and 13C NMR data, see Tables 1 and 2; positive ESIMS m/z 219.2 [M − H2O + H]+; negative ESIMS m/z 235.1 [M − H]; HRESIMS m/z 259.1668 [M + Na]+ (calcd for C15H24O2Na, 259.1669).
Aristomollin D (4). Colorless oil; [α]20D +180 (c 0.28, CHCl3); IR (KBr) νmax 3495, 2925, 1735, 1694, 1456, 1370, and 1244 cm−1; 1H and 13C NMR data, see Tables 1 and 2; positive ESIMS m/z 217.2 [M − HOAc + H]+; HRESIMS m/z 299.1632 [M + Na]+ (calcd for C17H24O3Na, 299.1623).
Aristomollin E (5). Colorless oil; [α]20D −247 (c 0.22, CHCl3); IR (KBr) νmax 3335, 2963, 2924, 2865, 1641, 1060, and 911 cm−1; 1H and 13C NMR data, see Tables 1 and 2; positive ESIMS m/z 219.2 [M − H2O + H]+; negative ESIMS m/z 235.2 [M − H]; HRESIMS m/z 259.1669 [M + Na]+ (calcd for C15H24O2Na, 259.1669).
Aristomollin F (6). Colorless oil; [α]20D −26 (c 0.33, CHCl3); IR (KBr) νmax 3426, 2928, 2867, 2718, 1718, 1455, 1376, and 1119 cm−1; 1H and 13C NMR data, see Tables 1 and 2; positive ESIMS m/z 219.2 [M − H2O + H]+; negative ESIMS m/z 235.3 [M − H]; HRESIMS m/z 259.1663 [M + Na]+ (calcd for C15H24O2Na, 259.1669).

ECD calculation of 2a

In general, conformational analyses were carried out via Monte Carlo searching using molecular mechanism with MMFF94 force field in the SPARTAN 04 (ref. 33) software package. The results showed 4 lowest energy conformers for 2a whose relative energy within 2.0 kcal mol−1. Subsequently, the conformers were reoptimized using DFT at the B3LYP/6-31+G (d) level in gas phase in the GAUSSIAN 09 program.34 The B3LYP/6-31+G (d) harmonic vibrational frequencies were also calculated to confirm their stability. The energies, oscillator strengths, and rotational strengths (velocity) of the first 60 electronic excitations were calculated using the TDDFT methodology at the B3LYP/6-311++G (2d, 2p) level in vacuum. The ECD spectra were simulated by the overlapping Gaussian function (half the bandwidth at 1/e peak height, σ = 0.6 eV).35 For 2a, the first 5 electronic excitations were adopted. To get the overall spectra, the simulated spectra of the lowest energy conformers were averaged according to the Boltzmann distribution theory and their relative Gibbs free energy (ΔG) (more details see ESI). Theoretical ECD spectrum of the corresponding enantiomer of 2a was obtained by directly inverse the ECD spectra of 2a.

Chemical transformation of 9 to 3

NaBH4 (1 mg) was added to a stirred solution of 9 (2 mg) in MeOH (0.5 mL), and the reaction was stirred at rt for 15 min. The mixture was then purified on Sephadex LH-20 (EtOH) to afford 3 (1.4 mg). Compound 3 was identified by the 1H NMR spectrum, MS data, and specific rotation.

Chemical transformation of 9 to 4

Acetic anhydride (200 μL) was added to a stirred solution of 9 (2 mg) in freshly distilled pyridine (1 mL). The reaction was stirred at rt for 10 h and quenched by adding 0.4 mL of H2O. After removal of solvent under vacuum, the residue was purified on a flash silica gel column eluting with CHCl3 to afford 4 (1.9 mg), which was identified by the 1H NMR spectrum, MS data, and specific rotation.

Chemical transformation of 10 to 5

To a stirred solution of 10 (2 mg) in MeOH (0.5 mL) was added NaBH4 (1 mg). The mixture was stirred at rt for 15 min, and then was subjected to Sephadex LH-20 (EtOH) column to obtain 5 (1.5 mg). Compound 5 was identified by the 1H NMR spectrum, MS data, and specific rotation.

Cell culture and viability assay

BV-2 microglial cells were obtained from Southern Medical University (SMU) Cell Bank (Guangzhou, People's Republic of China). Cells were plated into a 96-well plate (2 × 104 cells per well). After 24 h, they were pretreated with samples for 30 min and stimulated with 1 μg mL−1 LPS for another 24 h. The cell viability of the cultured cells was assessed by MTT assay. Briefly, BV-2 cells were incubated with 200 μL MTT solution (0.5 mg mL−1 in medium) for 4 h at 37 °C, and then the supernatants were removed and residues were dissolved in 200 μL DMSO. The absorbance was detected at 570 nm using a microplate reader (Molecular Devices, USA) and analyzed using a SoftMax Pro 5 software (Molecular Devices, USA).

Measurement of NO production

The NO concentration was measured by the Griess reaction. Briefly, BV-2 cells were treated with LPS (1.0 μg mL−1) and compounds for 24 h. After that, 100 μL of culture supernatant was allowed to react with 100 μL of Griess reagent (1% sulfanilamide, 0.1% N-1-naphthylethylenediamine dihydrochloride in 5% phosphoric acid) for 10 min at rt in the dark. Then, the optical density (100 μL per well) was measured at 540 nm using a microplate reader (Molecular Devices, USA). Sodium nitrite was used as a standard to calculate the nitrite concentration. Inhibition (%) = (1 − (ALPS+sample − Auntreated)/(ALPS − Auntreated)) × 100. The experiments were performed in triplicates, and the data are expressed as the mean ± standard deviation (SD) values. Quercetin was used as a positive control.

Acknowledgements

The authors thank the National Natural Science Foundation of China (no. 81102339) and the Opening Project of Guangdong Provincial Key Laboratory of New Drug Design and Evaluation (no. 2011A060901014) for providing financial support to this work.

Notes and references

  1. D. Luo, T. C. T. Or, C. L. H. Yang and A. S. Y. Lau, ACS Chem. Neurosci., 2014, 5, 855–866 CrossRef CAS PubMed.
  2. S. D. Skaper, P. Giusti and L. Facci, FASEB J., 2012, 26, 3103–3117 CrossRef CAS PubMed.
  3. P. Wang and J. L. Zweier, J. Biol. Chem., 1996, 271, 29223–29230 CrossRef CAS PubMed.
  4. R. Radi, A. Cassina and R. Hodara, Biol. Chem., 2002, 383, 401–409 CrossRef CAS PubMed.
  5. J. Li, K. W. Zeng, S. P. Shi, Y. Jiang and P. F. Tu, Fitoterapia, 2012, 83, 896–900 CrossRef CAS PubMed.
  6. B. Liu and J. S. Hong, J. Pharmacol. Exp. Ther., 2003, 304, 1–7 CrossRef CAS PubMed.
  7. T. S. Wu, A. G. Damu, C. R. Su and P. C. Kuo, Nat. Prod. Rep., 2004, 21, 594–624 RSC.
  8. G. X. Li, D. H.Wang and Q. Liu, China J. Chin. Mater. Med., 1985, 10, 39–41 CAS.
  9. J. Q. Yu, Z. X. Liao, X. Q. Cai, J. C. Lei and G. L. Zou, Environ. Toxicol. Pharmacol., 2007, 23, 162–167 CrossRef CAS PubMed.
  10. Y. M. Chen, L. N. Wang, C. Cui, J. Y. Ma, W. W. Wang, Z. X. Li, J. Gong and S. F. Ni, Anhui Nongye Kexue, 2013, 41, 4322–4323 CAS.
  11. Z. H. Cheng, T. Wu, S. W. A. Bligh, A. Bashall and B. Y. Yu, J. Nat. Prod., 2004, 67, 1761–1763 CrossRef CAS PubMed.
  12. A. Garcia and G. Delgado, J. Nat. Prod., 2006, 69, 1618–1621 CrossRef CAS PubMed.
  13. Z. Sun, B. Chen, S. Zhang and C. Hu, J. Nat. Prod., 2004, 67, 1975–1979 CrossRef CAS PubMed.
  14. N. Li, J. J. Chen and J. Zhou, J. Asian Nat. Prod. Res., 2005, 7, 279–282 CrossRef CAS PubMed.
  15. J. Kitajima, K. Kimizuka and Y. Tanaka, Chem. Pharm. Bull., 2000, 48, 77–80 CrossRef CAS.
  16. P. V. Demarco, E. Farkas, D. Doddrell, B. L. Mylari and E. Wenkert, J. Am. Chem. Soc., 1968, 90, 5480–5486 CrossRef CAS.
  17. B. N. Su, R. Misico, E. J. Park, B. D. Santarsiero, A. D. Mesecar, H. H. S. Fong, J. M. Pezzuto and A. D. Kinghorn, Tetrahedron, 2002, 58, 3453–3466 CrossRef CAS.
  18. R. B. Wu, Z. B. Cheng, Q. H. Han, T. T. Lin, J. W. Zhou, G. H. Tang and S. Yin, Chirality, 2014, 26, 189–193 CrossRef CAS PubMed.
  19. T. S. Wu, Y. Y. Chan and Y. L. Leu, Chem. Pharm. Bull., 2000, 48, 357–361 CrossRef CAS.
  20. T. S. Wu, Y. Y. Chan and Y. L. Leu, J. Nat. Prod., 2001, 64, 71–74 CrossRef CAS PubMed.
  21. Y. Y. Chan, Y. L. Leu and T. S. Wu, Tetrahedron Lett., 1998, 39, 8145–8148 CrossRef CAS.
  22. Z. B. Cheng, W. W. Shao, Y. N. Liu, Q. Liao, T. T. Lin, X. Y. Shen and S. Yin, J. Nat. Prod., 2013, 76, 664–671 CrossRef CAS PubMed.
  23. J. Zhang and L. X. He, Acta Pharm. Sin., 1986, 21, 273–278 CAS.
  24. G. Ruecker, R. Mayer, H. Wiedenfeld, B. S. Chung and A. Guellmann, Phytochemistry, 1987, 26, 1529–1530 CrossRef CAS.
  25. T. S. Wu, Y. Y. Chan and Y. L. Leu, J. Chin. Chem. Soc., 2000, 47, 957–960 CAS.
  26. C. Y. Ragasa, J. Ganzon, J. Hofilena, B. Tamboong and J. A. Rideout, Chem. Pharm. Bull., 2003, 51, 1208–1210 CrossRef CAS.
  27. C. Gaspar-Marques, M. F. Simoes and B. Rodriguez, J. Nat. Prod., 2004, 67, 614–621 CrossRef CAS PubMed.
  28. Z. H. Sun, C. Q. Hu and J. Y. Wang, Chin. Chem. Lett., 2007, 18, 1379–1382 CrossRef CAS PubMed.
  29. G. Ruecker, C. W. Ming, R. Mayer, G. Will and A. Guellmann, Phytochemistry, 1990, 29, 983–985 CrossRef CAS.
  30. H. Z. Xue, J. Zhang, L. X. He, C. H. He, Q. T. Zheng and R. Feng, Acta Pharm. Sin., 1989, 24, 917–922 CAS.
  31. T. S. Wu, Y. Y. Chan, Y. L. Leu, P. L. Wu, C. Y. Li and Y. Mori, J. Nat. Prod., 1999, 62, 348–351 CrossRef CAS PubMed.
  32. T. S. Wu, Y. Y. Chan, Y. L. Leu and Z. T. Chen, J. Nat. Prod., 1999, 62, 415–418 CrossRef CAS PubMed.
  33. Spartan 04, Wavefunction Inc., Irvine, CA Search PubMed.
  34. Gaussian 09, revision C.01, Gaussian, Inc., Wallingford, CT, 2009, full list of authors can be found in the ESI.
  35. P. J. Stephens and N. Harada, Chirality, 2010, 22, 229–233 CAS.

Footnotes

Electronic supplementary information (ESI) available: 1D and 2D NMR spectra of 1–6, 1H NMR spectra of known compounds (7–30). Detail information for ECD calculation. See DOI: 10.1039/c4ra09612f
These authors have contributed equally to this work.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.