Skip to content
Publicly Available Published by De Gruyter September 26, 2016

Unusual pyranosyl cembranoid diterpene from Sarcophyton trocheliophorum

  • Mohamed Shaaban EMAIL logo , Mohamed A. Ghani and Khaled A. Shaaban

Abstract

9-Hydroxy-10,11-dehydro-sarcotrocheliol (1), a new pyrane-based cembranoid diterpene, has been isolated along with three other known compounds, namely, sarcotrocheliol acetate (2), (+)-sarcophytol A (3), and (−)-sarcophytonin A (4), from the organic extract of the soft coral Sarcophyton trocheliophorum. The chemical structures of compounds 1–4 were determined on the basis of their 1D and 2D NMR [1H, 13C, 1H–1H correlation spectroscopy, heteronuclear multiple-quantum coherence, heteronuclear multiple-bond correlation, and nuclear Overhauser effect (NOE)] and mass spectra [electron ionization (EI), electrospray ionization, and high resolution mass spectrometry (HRMS)], and by comparison with related structures. The compounds 1–4 showed neither antimicrobial activity against 11 diverse tested microorganisms, nor cytotoxicity against brine shrimp, whereas the soft coral extract showed low cytotoxicity with a mortality rate of 1.7%.

1 Introduction

Soft corals belonging to the genus Sarcophyton are characterized by the production of cembranoid diterpenes, furanoditerpenes in which the isopropyl group may be functionalized as a γ-lactone [1], [2], [3], [4], along with norditerpenes and sesquiterpenes [5], [6]. The Sarcophyton genus is a rich source of secondary metabolites which are used as chemical defenses to deter predators [7]. Diverse biological activities were reported for these metabolites, including cytotoxicity [8], [9], [10], [11], [12], antimicrobial activity [13], and inhibition of the lipopolysaccharide-induced production of the proinflammatory cytokine tumor necrosis factor-α [14]. Among these metabolites are cembranoids, representing a major class of diterpenoids, which contain a 14-membered ring structure, decorated with a number of double bonds and methyl groups. They are prone to several enzymatic processes, including oxidation, oxidative rearrangements, photochemical ring contraction, and transannular cyclizations. Accordingly, such metabolites have the capability to be modified into remarkable marine metabolites leading to furan, furano-butenolide, oxirane, pyrane, and other ring-based cembranoids. The interesting published biological applications of cembranoids, for example, in the anticancer, antimicrobial, and anti-inflammatory field [7], [15], [16], encouraged us to further study the chemical constituents of the soft coral Sarcophyton trocheliophorum as possible leads for the discovery of new therapeutic agents.

In continuation with our research program on the discovery of bioactive and/or new metabolites from marine organisms, we are following up on our previous reports of novel bioactive metabolites from the shallow water Red Sea soft coral S. trocheliophorum (Family Alcyoniidae, Order Alcyonacea, subclass Octocorallia, class Anthozoa, phylum coelenterata) at the Hurghada coasts, Egypt [17], [18]. Herein we describe the isolation and structural determination of one new pyrane-based cembranoid, 9-hydroxy-10,11-dehydro-sarcotrocheliol (1) along with the recently reported sarcotrocheliol acetate (2) [19], [20], and two other known compounds, (+)-sarcophytol A (3) and (−)-sarcophytonin A (4). The chemical structures of compounds 1–4 were identified using NMR spectroscopy [1H, 13C, 1H–1H correlation spectroscopy (COSY), heteronuclear multiple-quantum coherence, heteronuclear multiple-bond correlation (HMBC), nuclear Overhauser effect (NOE)] and mass spectrometry [electron ionization (EI), chemical ionization (CI), electrospray ionization (ESI), ESI-high resolution mass spectrometry (HRMS)], and by comparison with related structures. The antimicrobial and cytotoxic activities of compounds 1–4 in comparison with the crude extract were also investigated.

2 Results and discussions

The physicochemical properties of the new compound 9-hydroxy-10,11-dehydro-sarcotrocheliol (1) are listed in Table 1. The chemical structures of sarcotrocheliol acetate (2), (+)-sarcophytol A (3), and (−)-sarcophytonin A (4) were fully assigned using 1D and 2D NMR (Figs. 1 and 2; Tables 2 and 3), mass spectrometry, and by comparison with the literature data [19], [20], [21], [22], [23].

Table 1:

Physicochemical properties of 9-hydroxy-10,11-dehydro-sarcotrocheliol (1).

AppearanceColorless oil
Rf0.25 (CH2Cl2/3% MeOH)
Anisaldehyde/H2SO4 spraying reagentPink turned later to brown
SolubilitySoluble in DMSO, MeOH, EtOH EtOAc, and CH2Cl2. Insoluble in hexane and water
Molecular formulaC20H34O3
MS ((+)-EI): m/z (%)322 (4) [M]+•, 304 (16) [M–H2O]+•, 286 (6), 261 (8), 243 (5), 217 (6), 206 (14), 194 (18), 177 (20), 159 (21), 136 (44), 121 (26), 107 (24), 93 (39), 81 (28), 69 (16), 55 (24), 43 (100)
HRMS ((+)-ESI)
 Found345.24000 [M+Na]+, 667.49000 [2M+Na]+
 Calcd.345.24000 [M+Na]+, 667.49078 [2M+Na]+
[α]D20+24° (c=0.19, CHCl3)
Fig. 1: Chemical structures of compounds 1–4.
Fig. 1:

Chemical structures of compounds 1–4.

Fig. 2: COSY (), selected HMBC (), and NOESY/NOE () correlations in 1–2.
Fig. 2:

COSY (

), selected HMBC (
), and NOESY/NOE (
) correlations in 1–2.

Table 2:

13C and 1H NMR of 9-hydroxy-10,11-dehydro-sarcotrocheliol (1) in CDCl3.

Position9-Hydroxy-10,11-dehydro-sarcotrocheliol (1)
δC (125 MHz)δH (300 MHz)
171.84.62 (dd, 15.4, 4.6)
245.41.34 (m)
318.61.62 (m), 1.36 (m)
434.02.26 (m)
575.0
675.43.46 (dd, 11.4, 1.7)
722.61.72 (m), 1.30 (m)
839.91.84 (m), 1.53 (m)
973.4
10135.55. 41 (d, 15.7)
11127.85.86 (m)
1242.32.76 (t, 11.2), 2.54 (dd, 11.5, 4.2)
13140.9
14124.05.28 (d, 17.1)
1529.11.20 (m)
1620.40.76 (d, 6.4)
1720.60.88 (d, 6.5)
1823.71.03 (s)
1929.31.31 (s)
2017.61.83 (s)

Multiplicities and J in Hz are given in parentheses. Chemical shifts δ are given in ppm.

Table 3:

13C and 1H NMR of sarcotrocheliol acetate (2) in CDCl3.

PositionSarcotrocheliol acetate (2)2 (lit. [19])
δC (125 MHz)δH (300 MHz)δC (150 MHz)δH (600 MHz)
171.24.55 (dd, 10.5, 4.6)71.44.52 (dd, 10.8, 5.4)
246.41.30 (m)46.51.26 (m)
318.91.54 (m)19.01.28 (m), 1.77 (m)
434.21.80 (m), 1.30 (m)34.3a1.51 (m), 1.23 (m)
573.573.7
673.45.40 (d, 9.7)73.45.37 (d, 10.7)
6-OCOCH3170.9171.0
6-OCOCH321.22.08 (s)21.32.06 (s)
728.91.69 (m), 1.54 (m)29.01.66 (m), 1.50 (m)
834.32.00 (m), 1.65 (m)34.41.97 (m), 1.63 (m)
9134.9135.0
10124.45. 09 (dd, 10.8, 5.0)124.55. 05 (dd, 10.2, 4.8)
1125.22.12 (m), 2.33 (m)25.32.07 (m), 2.30 (m)
1239.72.18 (m)39.82.15 (m), 2.10 (m)
13138.9139.0
14125.25.54 (d, 10.7)125.35.50 (d, 10.8)
1528.91.20 (m)29.01.17 (m)
1620.20.72 (d, 6.2)20.30.69 (d, 6.6)
1720.60.86 (d, 6.2)20.70.82 (d, 6.6)
1825.31.07 (s)25.41.04 (s)
1916.91.59 (s)17.01.56 (s)
2014.91.66 (s)15.01.62 (s)

Multiplicities and J in Hz are given in parentheses in comparison with the literature data. Chemical shifts δ are given in ppm.

aWrong atom numbering assignment in the literature; it has been corrected herein by comparison with our isolated sample of 2.

2.1 9-Hydroxy-10,11-dehydro-sarcotrocheliol

Compound 1 was obtained as optically active colorless oil using various chromatographic techniques. On thin-layer chromatography (TLC), compound 1 displayed a pink color reaction, which later turned to brown with the anisaldehyde/sulfuric acid developmental stain. The molecular weight of 1 was determined as 322 Dalton by EI-MS. The molecular formula of 1 was deduced as C20H34O3 by HRMS ((+)-ESI) (Table 1), 1H and 13C NMR, indicating the presence of four double-bond equivalents. The 1H and 13C NMR spectroscopic data (Table 2) of compound 1 showed resonances due to 20 carbons categorized by a heteronuclear single-quantum coherence experiment into five methyl, five methylene, seven methine (among them three sp2 and two sp3 oxy methines), and three quaternary carbons. Signals attributed to four olefinic carbons (δC=135.5 [δH=5.41], δC=127.8 [δH=5.86], δC=124.0 [δH=5.28], δC=140.9 ppm), accounting for two unsaturated double bonds, and implying that compound 1 contains bicyclic skeleton.

In the 1H, 13C NMR, and heteronuclear multiple-quantum coherence spectra, compound 1 featured resonances corresponding to five methyl groups: three tertiary methyl groups [C-18 (δH=1.03, δC=23.7 ppm), C-19 (δH=1.31, δC=29.3 ppm), C-20 (δH=1.83, δC=17.6 ppm)], and isopropyl doublet methyl signals [C-16 (δH=0.76, δC=20.4 ppm), C-17 (δH=0.88, δC=20.6 ppm)]; the in-between isopropyl methine carbon signal (C-15) was deduced at δC=29.1 (δH=1.20) ppm. The NMR spectra revealed further two doublet of doublet oxygenated methine functions resonating at δC=71.8 (δH=4.62, C-1) ppm and δC=75.4 (δH=3.46, C-6) ppm, and two quaternary oxygenated carbons at δC=75.0 (C-5) and δC=73.4 (C-9), along with a non-oxygenated methine signal appeared at δC=45.4 (δH=1.34, C-2) ppm (Table 2).

The planar constitution of 1 was finally confirmed, as depicted in Fig. 2, based on 1H–1H COSY and HMBC experiments. An extensive study of the 1H–1H COSY correlations established two proton sequences: the coupling between the olefinic proton shown at δH=5.28 (H-14) and the oxymethine proton at δH=4.62 ppm (H-1), as well as this proton (H-1) and the methine proton at δH=1.34 ppm (H-2), which in turn is correlated with both the isopropyl methine proton at δH=1.20 ppm (H-15) and the methylene protons found at δH=1.62, 1.36 ppm (H2-3); signals of the last methylene protons displayed a COSY correlation to CH2-4 (δH=2.26 ppm), confirming the connectivity of H-14/H-1/H-2/H-15 and H-2/H2-3/H2-4 (Fig. 2).

The HMBC couplings shown from H-1 to C-3, C-5, C-13 established the presence of a pyrane ring. The 2J HMBC correlation from CH3-18 to C-5 confirmed their direct attachment. The last methyl group (CH3-18) showed two further HMBC correlations to C-4 (δC=34.0 ppm), and C-6 (δC=75.4 ppm), proving the direct connectivity of C-4/C-5/C-6. The isopropyl group connectivity to the C-2 position was confirmed by the 3J HMBC correlations from both isopropyl methyl signals (H3-16, H3-17) toward C-2 (δC=45.4 ppm). The second sequence of proton patterns were identified through the shown HMBC correlations between the olefinic methyl H3-20 (δH=1.83 ppm) and C-13 (δC=140.9 ppm, 2J), CH-14 (δC=124.0 ppm, 3J), and CH2-12 (δC=42.3 ppm, 3J), proving the direct attachment of the corresponding 2-methyl-butylene partial structure to the pyrane ring (C14–C1). Based on the 1H–1H COSY data, the two olefinic protons at δH=5.86 ppm (CH-11) and δH=5.41 ppm (CH-10, d, J=15.7 Hz) are vicinal, forming one transoid olefinic double bond [–CH=CH–]. The last double bond (H-10/H-11) showed a direct attachment to CH2-12 (δH=2.76, 2.54 ppm) indicated by COSY correlation with H-11 along with HMBC correlations from H-10 (3J) and H-11 (2J) toward C-12. Proton signals of the last double bond (H-10/H-11) exhibited further two HMBC correlations versus the sp3 quaternary-oxygenated carbon C-9 (δC=73.4 ppm), to which the singlet tertiary methyl CH3-19 (δH=1.31, δC=29.3 ppm) is attached. The proton signals of CH3-19 displayed further HMBC correlations to H-10 and CH2-8, in which CH2-8 (δH=1.84, 1.53 ppm) proton signals displayed a COSY correlation to CH2-7 (δH=1.72, 1.30 ppm). CH2-7 showed a COSY correlation with the oxymethine CH-6 (δH=3.46 ppm). The last methylene protons (CH2-7) exhibited two further HMBC correlations at C-8 and C-6, confirming their location in between C-8 and C-6. Accordingly, connections between different parts of 1 were proven. The geometries of the double bonds at C-10, C-11, and C-14 are trans owing to the values of the coupling constant of H-10 (J=15.7 Hz), and chemical shifts of allylic methylene groups δC>30 ppm (C-12 and C-8 have δC=42.3 and 39.9 ppm, respectively) [19], [24]. The remaining 2D NMR correlations are in full agreement with structure 1 (Figs. 1 and 2). The relative configuration of 1 was proposed on the basis of the NOE cross peaks (Figs. 2 and 3) and by comparison with NOE spectra of our isolated 2, which has been recently reported from the Red Sea soft coral S. trocheliophorum as well [19]. The NOE observed between H-1/CH3-18, H-1/CH-14, H-1/CH3-20, and H-1/CH-2 suggested the cofacial orientation of H-1, CH3-18, CH3-20 (β), and H-2. In addition, the NOE correlation between CH3-18/H-6 and H-6/CH3-19 suggested their cofacial orientations as depicted in Fig. 2. Thus, the cumulative analysis of ESI-HRMS, 1H, 13C, COSY, heteronuclear single-quantum coherence, and HMBC spectroscopic data (Table 2; Fig. 2) established compound 1 as new pyrane-based cembranoid diterpene, which was thereby designated as 9-hydroxy-10,11-dehydro-sarcotrocheliol.

Fig. 3: Selected three-dimensional NOESY correlations of 9-hydroxy-10,11-dehydro-sarcotrocheliol (1).
Fig. 3:

Selected three-dimensional NOESY correlations of 9-hydroxy-10,11-dehydro-sarcotrocheliol (1).

2.2 Sarcotrocheliol (2), sarcophytol A (3), and sarcophytonin A (4)

Compounds 2–4 were obtained as colorless oils using a series of chromatographic techniques, and displayed physicochemical properties common to diterpenes (see the Experimental section). The molecular formulas of compounds 2–4 were established as C22H36O3, C20H32O, and C20H30O, respectively, by ESI-HRMS, 1H, and 13C NMR. Based on the full 1D and 2D NMR spectroscopic data (Figs. 1 and 2; Tables 3 and 4) along with HRMS analysis, structures of compounds 2–4 have been confirmed as sarcotrocheliol acetate (2), (+)-sarcophytol A (3), and (−)-sarcophytonin A (4, also known as deoxysarcophytoxide). Sarcotrocheliol acetate (2) has been recently reported from Red Sea soft coral S. trocheliophorum, and to have significant antibacterial activities against Staphylococcus aureus, Acinetobacter spp., and methicillin-resistant Staphylococcus aureus (MRSA) with minimum inhibitory concentration (MIC) range 1.53–4.34 mm [19]. The low value of the reported optical activity of compound 2 (lit. [19]: [α]D20=+8.4 (c=0.1, CHCl3)) compared with our isolated one ([α]D20=+74 (c=0.15, CHCl3)) is an indication that the reported 2 is mostly containing some impurities.

Table 4:

13C and 1H NMR of (+)-sarcophytol A (3) and (−)-sarcophytonin A (4) in CDCl3.

Position(+)-Sarcophytol A (3)(−)-Sarcophytonin A (4)
δC (125 MHz)δH (300 MHz)δC (150 MHz)δH (600 MHz)
1146.8126.9
269.84.99 (m)83.85.52 (m)
344.52.22 (m), 2.40 (m)125.65.08 (brd, 10.0)
4131.2140.0
5125.14.99 (m)39.12.12 (m), 2.25 (m)
624.52.10 (m)24.72.30 (m), 2.00 (m)
738.82.20 (m), 2.00 (m)125.34.86 (brd, 8.7)
8134.5132.9
9124.44.99 (m)40.22.06 (m), 1.84 (m)
1025.62.10 (m)23.42.00 (m)
1139.72.20 (m), 2.00 (m)123.75.00 (t, 7.8)
12135.7135.1
13121.15.99 (d, 11.5)37.11.86 (m)
14120.36.14 (d, 11.5)25.82.34 (m), 1.74 (m)
1527.12.58 (m)133.7
1624.41.04 (d, 6.9)78.24.48 (brs)
1725.41.11 (d, 6.8)10.11.64 (s)
1818.21.59 (s)14.81.69 (s)
1915.51.46 (s)15.01.58 (s)
2016.41.74 (d, 1.0)15.51.60 (d, 1.0)

Multiplicities and J in Hz are given in parentheses. Chemical shifts δ are given in ppm.

2.3 Biological activity

The crude extract of S. trocheliophorum and isolated compounds 1–4 are biologically inactive against Bacillus subtilis, S. aureus, Streptomyces viridochromogenes (Tü 57), Escherichia coli, Candida albicans, Mucor miehei, Chlorella vulgaris, Chlorella sorokiniana, Scenedesmus subspicatus, Rhizoctonia solani, and Pythium ultimum at 40 μg per disk. Compounds 1–4 and the crude extract were further examined for their cytotoxic activity against brine shrimp at a concentration 10 μg mL−1 (24 h). However, compounds 1–4 demonstrated no cytotoxicity, whereas the crude extract of S. trocheliophorum exhibited low cytotoxicity with a mortality rate of 1.7% [17], [18].

3 Experimental section

Optical rotations were measured by a polarimeter (Perkin-Elmer, model 343; Waltham, MA, USA). The NMR spectra were measured on Varian (Palo Alto, CA, USA) Unity 300 (300.145 MHz) and Varian Inova 500 (125.7 MHz) spectrometers. EI-MS spectra were recorded on a Finnigan MAT 95 spectrometer (70 eV) with perfluorkerosine as reference substance for EI-HRMS. ESI-HRMS were recorded on an Apex IV 7 Tesla Fourier-Transform Ion Cyclotron Resonance Mass Spectrometer (Bruker Daltonics, Billerica, MA, USA). Flash chromatography was carried out on silica gel (230–400 mesh). Rf values were measured on Polygram SIL G/UV254 TLC cards (Macherey-Nagel). Size exclusion chromatography was done on Sephadex LH-20 (Lipophilic Sephadex, Amersham Biosciences Ltd; purchased from Sigma-Aldrich Chemie, Steinheim, Germany).

3.1 Extraction, isolation, and purification

Details of extraction and chromatographic purification of the soft coral S. trocheliophorum were reported recently [17], [18], in which 10 fractions were obtained. Fraction FIII (7.2 g) was subjected to size exclusion fractionation on a Sephadex LH-20 column [dichloromethane (DCM)/40% MeOH] afforded two sub-fractions FIIIA (3.2 g) and FIIIB (2.8 g). Sub-fraction IIIA was deduced as a mixture of fatty acid and has been discarded. Sub-fraction FIIIB showed several UV-absorbing bands on TLC, which stained as violet on spraying with anisaldehyde/sulfuric acid and heating, indicating diterpene components. As a result, sub-fraction FIIIB was applied to silica gel column chromatography (eluted with cyclohexane-DCM gradient), followed by preparative thin-layer chromatography (cyclohexane-DCM, 2:1) and Sephadex LH-20 (DCM/40% MeOH), and afforded 9-hydroxy-10,11-dehydro-sarcotrocheliol (1, 169.5 mg), sarcotrocheliol acetate (2, 119.1 mg), sarcophytol A (3, 25.3 mg), and sarcophytonin A (4, 142.2 mg) as colorless oils. The remaining fractions and sub-fractions obtained from the soft coral S. trocheliophorum extract lacked the diterpenes analogs based on TLC and high-performance liquid chromatography analyses.

3.1.1 Sarcotrocheliol acetate (2)

Colorless oil, UV-absorbing (254 nm), and stained pink on spraying with anisaldehyde/sulfuric acid and changed later to brown. Rf=0.40 (CH2Cl2/3% MeOH). – [α]D20=+74° (c=0.15, CHCl3). – 1H (300 MHz, CDCl3) and 13C NMR (75 MHz, CDCl3): see Table 3. – MS ((+)-EI): m/z (%)=348.3 (8) [M]+•, 305.2 (4), 288.3 (32), 245.3 (7), 193.1 (9), 177.2 (100), 161.1 (24), 136.1 (26), 121.1 (34), 111.1 (28), 93.1 (70), 81.1 (44), 55.1 (32), 43.0 (100), 41.0 (28). – HRMS ((+)-ESI): m/z=349.27388 (calc. 349.27372 for C22H37O3, [M+H]+).

3.1.2 (+)-Sarcophytol A (3)

Colorless oil, UV-absorbing (254 nm), and stained pink on spraying with anisaldehyde/sulfuric acid and changed later to brown. Rf=0.62 (CH2Cl2/3% MeOH). – [α]D20=+30° (c=0.1, CHCl3). – 1H (300 MHz, CDCl3) and 13C NMR (125 MHz, CDCl3): see Table 4. – MS ((+)-EI): m/z (%)=288 (22) [M]+•, 273 (5) [M–CH3]+•, 189 (4), 161 (6), 147 (8), 138 (16), 137 (100), 109 (56), 93 (20), 81 (28), 69 (24), 68 (18), 41 (20). – HRMS ((+)-ESI): m/z=289.25266 (calc. 289.25259 for C20H33O, [M+H]+).

3.1.3 (−)-Sarcophytonin A (4)

Colorless oil, UV-absorbing (254 nm), and stained pink on spraying with anisaldehyde/sulfuric acid and later changed to brown. Rf=0.67 (CH2Cl2/3% MeOH). – [α]D20=37° (c=0.32, MeOH) (lit. [21]: [α]D20=92° (c=2.3 in CHCl3)). – 1H (300 MHz, CDCl3) and 13C NMR (125 MHz, CDCl3): see Table 4. – MS ((+)-EI): m/z (%)=286 (22) [M]+•, 203 (4), 175 (12), 163 (14), 149 (22), 135 (39), 121 (16), 107 (12), 93 (20), 81 (22), 41 (100). – HRMS ((+)-ESI): m/z=287.23703 (calc. 287.23695 for C20H31O, [M+H]+).

3.2 Antimicrobial activity and brine shrimp microwell cytotoxic assays

Details of our general antimicrobial activity testing were reported recently [17], [18]. The cytotoxicity assay was performed according to Takahashi et al. [25] and Sajid et al. [26] screening.

4 Supplementary information

NMR and HRMS spectra and other supplementary data associated with this article are given as Supplementary Information available online (http://dx.doi.org/10.1515/znb-2016-0144).

Acknowledgments

The authors are deeply thankful to Prof. H. Laatsch, Institute of Organic and Biomolecular Chemistry, Göttingen, for his support and lab facilities. They thank Dr. H. Frauendorf and R. Machinek for MS and NMR measurements. They would like to acknowledge F. Lissy for biological activity tests and A. Kohl for technical assistance. They also thank Dr. Eric E. Nybo (Department of Pharmaceutical Sciences, Ferris State University, Big Rapids, MI, USA) for reading and editing the manuscript. M. Shaaban thanks the German Academic Exchange Service (DAAD) for a short-term grant.

References

[1] J. Bernstein, U. Schmeuli, E. Zadock, Y. Kashman, I. Neeman, Tetrahedron1974, 30, 2817.10.1016/S0040-4020(01)97451-4Search in Google Scholar

[2] Y. Kashman, E. Zadock, I. Neeman, Tetrahedron1974, 30, 3615.10.1016/S0040-4020(01)97044-9Search in Google Scholar

[3] B. Tursch, Pure Appl. Chem. 1976, 48, 1.10.1351/pac197648010001Search in Google Scholar

[4] B. F. Bowden, J. C. Coll, S. J. Mitchell, G. J. Stoke, Aust. J. Chem. 1979, 32, 653.10.1071/CH9790653Search in Google Scholar

[5] D. J. Faulkner, Nat. Prod. Rep. 1984, 1, 551.10.1039/np9840100551Search in Google Scholar

[6] D. J. Faulkner, Nat. Prod. Rep. 1995, 12, 223.10.1039/np9951200223Search in Google Scholar

[7] J. C. Coll, Chem. Rev.1992, 92, 613.10.1021/cr00012a006Search in Google Scholar

[8] A. J. Weinheimer, J. A. Matson, M. B. Hoassain, D. van der Helm, Tetrahedron Lett.1977, 18, 2923.10.1016/S0040-4039(01)83115-4Search in Google Scholar

[9] J.-H. Sheu, A. F. Ahmed, R.-T. Shiue, C.-F. Dai, Y.-H. Kuo, J. Nat. Prod. 2002, 65, 1904.10.1021/np020280rSearch in Google Scholar PubMed

[10] C.-Y. Duh, S.-K. Wang, H.-K. Tseng, J.-H. Sheu, M. Y. Chiang, J. Nat. Prod. 1998, 61, 844.10.1021/np980021vSearch in Google Scholar PubMed

[11] C.-Y. Duh, S.-K. Wang, M.-C. Chia, M. Y. Chiang, Tetrahedron Lett.1999, 40, 6033.10.1016/S0040-4039(99)01194-6Search in Google Scholar

[12] K. Yamada, T. Ujiie, K. Yoshida, T. Miyamoto, R.Higuchi, Tetrahedron1997, 53, 4569.10.1016/S0040-4020(97)00169-5Search in Google Scholar

[13] T. L. Aceret, J. C. Coll, Y. Uchio, P. W. Sammarco, Comp. Biochem. Physiol., Part C: Pharmacol., Toxicol. Endocrin. 1998, 120, 121.10.1016/S0742-8413(98)00032-2Search in Google Scholar

[14] H. Takaki, R. Koganemaru, Y. Iwkawa, R. Higuchi, T. Miyamoto, Biol. Pharm. Bull. 2003, 26, 380.10.1248/bpb.26.380Search in Google Scholar PubMed

[15] Y. Li, G. Pattenden, Nat. Prod. Rep.2011, 28, 1269.10.1039/c1np00023cSearch in Google Scholar PubMed

[16] H. N. Kamel, M. Slattery, Pharm. Biol.2005, 43, 253.10.1080/13880200590928852Search in Google Scholar

[17] M. Shaaban, M. A. Ghani, K. A. Shaaban, Z. Naturforsch.2013, 68b, 939.10.5560/znb.2013-3131Search in Google Scholar

[18] K. A. Shaaban, M. A. Ghani, M. Shaaban, Br. J. Pharm. Res.2015, 5, 192.10.9734/BJPR/2015/14757Search in Google Scholar

[19] K. O. Al-Footy, W. M. Alarif, F. Asiri, M. M. Aly, S. N. Ayyad, Med. Chem. Res.2015, 24, 505.10.1007/s00044-014-1147-1Search in Google Scholar

[20] A. Abdel-Lateff, W. M. Alarif, S. N. Ayyad, S. S. Al-Lihaibi, S. A. Basaif, Nat. Prod. Res. 2015, 29, 24.10.1080/14786419.2014.952637Search in Google Scholar PubMed

[21] M. Kobayashi, T. Nakagawa, H. Mitsuhashi, Chem. Pharm. Bull. 1979, 27, 2382.10.1248/cpb.27.2382Search in Google Scholar

[22] K. Nichitani, K. Harada, N. Sano, K. Sato, K. Yamakawa, Chem. Pharm. Bull.1991, 39, 2514.10.1248/cpb.39.2514Search in Google Scholar

[23] H. Miyaoka, S. Taira, H. Mitome, K. Iguchi, K. Matsumoto, C. Yokoo, Y. Yamada. Chem. Lett.1996, 25, 239.10.1246/cl.1996.239Search in Google Scholar

[24] B.-N. Su, Y. Takaishi, J. Nat. Prod. 1999, 62, 1325.10.1021/np990145nSearch in Google Scholar PubMed

[25] A. Takahashi, S. Kurasawa, D. Ikeda, Y. Okami, T. Takeuchi, J. Antibiot. 1989, 32, 1556.10.7164/antibiotics.42.1556Search in Google Scholar PubMed

[26] I. Sajid, C. B. F. Fondja Yao, K. A. Shaaban, S. Hasnain, H. Laatsch, World J. Microbiol. Biotechnol.2009, 25, 601.10.1007/s11274-008-9928-7Search in Google Scholar


Supplemental Material:

The online version of this article (DOI: 10.1515/znb-2016-0144) offers supplementary material, available to authorized users.


Received: 2016-6-28
Accepted: 2016-7-13
Published Online: 2016-9-26
Published in Print: 2016-12-1

©2016 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 30.5.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2016-0144/html
Scroll to top button