Next Article in Journal
Seasonal Variation of Mycosporine-Like Amino Acids in Three Subantarctic Red Seaweeds
Previous Article in Journal
The Protective Effect of Echinochrome A on Extracellular Matrix of Vocal Folds in Ovariectomized Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Probing Anti-Proliferative 24-Homoscalaranes from a Sponge Lendenfeldia sp.

1
Doctoral Degree Program in Marine Biotechnology, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
2
Doctoral Degree Program in Marine Biotechnology, Academia Sinica, Taipei 11529, Taiwan
3
National Museum of Marine Biology and Aquarium, Pingtung 94450, Taiwan
4
Research Center for Chinese Herbal Medicine, College of Human Ecology, Chang Gung University of Science and Technology, Taoyuan 33303, Taiwan
5
Department of Marine Biotechnology and Resources, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
6
Graduate Institute of Marine Biology, National Dong Hwa University, Pingtung 94450, Taiwan
7
Department of Marine Recreation, National Penghu University of Science and Technology, Penghu 88046, Taiwan
8
Institute of Chemistry, Academia Sinica, Taipei 11529, Taiwan
9
Chinese Medicine Research and Development Center, China Medical University Hospital, Taichung 40447, Taiwan
10
Graduate Institute of Natural Products, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2020, 18(2), 76; https://doi.org/10.3390/md18020076
Submission received: 30 December 2019 / Revised: 16 January 2020 / Accepted: 22 January 2020 / Published: 24 January 2020

Abstract

:
In the current study, an NMR spectroscopic pattern-based procedure for probing scalarane derivatives was performed and four new 24-homoscalaranes, lendenfeldaranes A–D (14), along with three known compounds, 12α-acetoxy-22-hydroxy-24-methyl-24-oxoscalar-16-en- 25-al (5), felixin F (6), and 24-methyl-12,24,25-trioxoscalar-16-en-22-oic acid (7) were isolated from the sponge Lendenfeldia sp. The structures of scalaranes 17 were elucidated on the basis of spectroscopic analysis. Scalaranes 17 were further evaluated for their cytotoxicity toward a series of human cancer cell lines and the results suggested that 5 and 7 dominated in the anti- proliferative activity of the extract. The 18-aldehyde functionality was found to play a key role in their activity.

Graphical Abstract

1. Introduction

Since the first scalarane-type derivative, scalarin, was originally isolated from Cocaspongia scalaris [1], more than three hundred scalarane sesterterpenoids have been obtained from cyanobacteria and marine organisms [2,3]. Compounds of this type demonstrate a wide spectrum of interesting biological properties, such as anti-inflammation [4], cytotoxicity [5,6,7], anti-feedant [8,9,10], anti-microbial activity [11,12], ichthyotoxicity [13], anti-tubercular activity [14], anti-HIV [15], and inhibition of the nuclear hormone receptor [16]. In order to seek novel anti-proliferative substances from marine organisms, a chemical and bioactive investigation was carried out on the organic extracts of a marine sponge identified as Lendenfeldia sp. (family—Thorectidae). The ethyl acetate (EtOAc) extract of Lendenfeldia sp. was found to exhibit anti-proliferative activity against human cancer cell lines, including human acute lymphoblastic leukemia (MOLT-4), human chronic myelogenous leukemia (K-562), human histiocytic lymphoma (U-937), and human T-cell lymphoblastic lymphoma (SUP-T1) with IC50 values < 0.625 μg/mL. The bioassay-guided isolation, combined with an NMR spectroscopic pattern-based procedure, was used to explore the anti- proliferative scalarane substances, and led to the isolation four new 24-homoscalaranes, lendenfeldaranes A–D (14), along with three known metabolites, 12α-acetoxy-22-hydroxy-24- methyl-24-oxoscalar-16-en-25-al (5) [17], felixin F (6) [18], and 24-methyl-12,24,25-trioxoscalar-16- en-22-oic acid (7) [17]. In the current study, the comprehensive workflow of isolation, structure elucidation and an anti-proliferative evaluation were implemented on scalaranes 17 (Figure 1).

2. Results and Discussion

Lendenfeldarane A (1) was obtained as an amorphous powder and assigned the molecular formula C28H42O6 (eight degrees of unsaturation) from its (+)-HRESIMS at m/z 497.28736 [M + Na]+ (calcd. for C28H42O6 + Na, 497.28712). The 1H NMR data of 1 (Table 1), showed six singlet methyls at δH 0.75, 0.86, 0.96, 1.16, 2.14, and 2.29, one olefinic proton at δH 6.90 (1H, br s), and one oxymethine proton at δH 4.77 (1H, br s). The diastereotopic geminal protons at δH 3.85 (1H, d, J = 11.6 Hz) and 4.02 (1H, d, J = 11.6 Hz) were assumed to be an oxymethylene group. Analyses of the 13C and distortionless enhancement by polarization transfer (DEPT) spectra of 1 (Table 1) revealed the existence of 28 carbon resonances, including six methyls, eight sp3 methylenes (including one oxymethylene), five sp3 methines (including one oxymethine), four sp3 quaternary carbons, one sp2 methine, and four sp2 quaternary carbons (including three carbonyls). Based on the 1H and 13C NMR spectra, 1 was found to possess an acetoxy group (δH 2.14, 3H, s; δC 170.4, C; 21.5, CH3). An additional unsaturated functionality was indicated by 13C resonances at δC 139.7 (CH-16) and 137.9 (C-17), suggesting the presence of a trisubstituted olefin. Thus, four degrees of unsaturation were accounted for, and the above NMR data—as well as the unassigned degrees of unsaturation of 1—implied a tetracyclic analogue.
The gross structure of 1 was further established from its 2D NMR spectra. From the coupling information in the COSY spectrum of 1 (Figure 2), it was possible to establish four partial structure units between H2-1/H2-2/H2-3, H-5/H2-6/H2-7, H-9/H2-11/H-12, and H-14/H2-15/H2-16. The heteronuclear multiple bond correlation (HMBC) spectrum connected these substructures by the connectivity between H-5/C-10; H-16/C-17; H3-20/C-3, C-4, C-5, C-19; H3-21/C-7, C-8, C-9, C-14; H2-22/C-1, C-9, C-10; H3-23/C-12, C-13, C-14, C-18; and H3-26/C-17, C-24, indicating a scalarane-type sesterterpenoid structure (Figure 2). Furthermore, the acetoxy and carboxylic acid groups positioned at C-12 and C-18 were determined by the HMBC, from H-12 to the acetate carbonyl at δC 170.4 and from H-18 to C-25 (δC 175.1), respectively.
The relative stereochemistry of 1 was elucidated by correlations in the NOESY experiment. Using the conventional method for analyzing the stereochemistry, the α- and β-configurations were assigned at H-5 and C-10-hydroxymethyl, respectively, to anchor the stereochemical analysis. In the NOESY spectrum of 1 (Figure 3), H-9 correlated with H-5, but not with H3-21 and H2-22, suggesting that these two protons (H-5 and H-9) were situated on the same face and were α-oriented, and that the Me-21 and C-10-hydroxymethyl groups were β-oriented at C-8 and C-10, respectively. H-14 exhibited correlations with H-9 and H-18, but not with H3-21 and H3-23, demonstrating that H-14 and H-18 were α-oriented. Additionally, the proton signal of a methyl group at δH 0.96 (H3-23) displayed a correlation with H-12 (δH 4.77), which indicated the β-orientations of Me-23 and H-12. The NOESY spectrum showed a correlation between H3-26 and H-16, revealing the E geometry of the C-16/17 carbon–carbon double bond. It was found that the NMR data of 1 were similar to those of a known scalarane analogue, 12α-acetoxy-22-hydroxy-24-methyl-24-oxoscalar-16-en-25-al (5), from an Australian sponge, Lendenfeldia sp. [17], except that the aldehyde group in 5 was replaced by a carboxylic acid group in 1. Based on the above findings, the structure of 1 was accordingly assigned, as shown in Figure 1, and named lendenfeldarane A (Supplementary Materials, Figures S1–S8).
Compound 2 (lendenfeldarane B) was obtained as an amorphous powder and its molecular formula was determined as C26H40O6, based on a sodiated adduct ion peak [M + Na]+ at m/z 471.27171 in (+)-HRESIMS (calcd. for C26H40O6 + Na, 471.27142). The 1H NMR data of 2 (Table 1) showed five singlet methyls at δH 0.76, 0.87, 1.30, 1.34, and 2.40 and one oxymethine proton at δH 3.53 (1H, ddd, J = 10.8, 10.8, 4.8 Hz). The diastereotopic geminal protons at δH 3.93 (1H, dd, J = 11.4, 1.2 Hz) and 4.08 (1H, d, J = 11.4 Hz) were assumed to be an oxygenated methylene group. The 13C and DEPT data of 2 suggested the presence of 26 carbons that were similar to those of a known scalarane, felixin F (6) [18], including a carboxylic carbon at δC 172.4, two ketone carbons at δC 212.6 and 221.9, an oxymethine carbon at δC 72.7, and an oxymethylene carbon at δC 62.7. Analysis of these NMR data suggested that compounds 2 and 6 are closely related, with the only difference being that the β- aldehyde group at C-18 in 6 was replaced by a β-carboxylic acid group in 2. Based on the analyses of the COSY, HMBC, and NOESY spectra, as well as the specific rotation data (2: [α] D 20 +49 (c 0.99, CHCl3), 6: [α] D 20 +55 (c 0.04, CHCl3); ref [18] 6: [α] D 25 +54 (c 0.4, CHCl3)), Compound 2 was finally assigned, as shown in Figure 1, and named as lendenfeldarane B (Supplementary Materials, Figures S9–S16).
The molecular formula of lendenfeldarane C (3) was determined as C26H40O4 from an [M + Na]+ ion at m/z 439.28188 (calcd. for C26H40O4 + Na, 439.28174) and NMR data (Table 2), indicating seven degrees of unsaturation. The 1H NMR data of 3 showed four singlet methyls at δH 0.78, 0.86, 1.08, 1.13, one doublet methyl at δH 1.37 (J = 6.5 Hz), and two oxymethine protons at δH 4.60 (1H, br s) and 4.79 (1H, q, J = 6.5 Hz). The diastereotopic geminal protons at δH 3.92 (1H, d, J = 11.5 Hz) and 4.05 (1H, d, J = 11.5 Hz) were assumed to be an oxymethylene group. Analyses of the 13C NMR and DEPT spectrum of 3 revealed the existence of 26 carbon resonances, including five methyls, nine sp3 methylenes (including one oxymethylene), five sp3 methines (including two oxymethines), four sp3 quaternary carbons, and three sp2 quaternary carbons (including one ester carbonyl). Based on the 13C spectrum, 3 was found to possess an ester carbonyl (δC 172.6) and an unsaturated degree was indicated by the 13C chemical shifts at δC 133.5 (C-18) and 165.2 (C-17), suggesting the presence of a tetrasubstituted olefin. Thus, the above NMR data and the remaining five unsaturated degrees of 3 required a pentacyclic analogue.
The gross structure of 3 was constructed from its 2D NMR spectra. From the COSY spectrum (Figure 2), five partial structure units between H2-1/H2-2/H2-3, H-5/H2-6/H2-7, H-9/H2-11/H-12, H-14/ H2-15/H2-16, and H-24/H3-26 were established. The HMBC spectrum connected these fractional structures by the key correlations between H-5/C-10; H2-15/C-17; H3-20/C-3, C-4, C-5, C-19; H3-21/ C-7, C-8, C-9, C-14; H2-22/C-1, C-9, C-10; and H3-23/C-12, C-13, C-14, C-18, indicating a scalarane skeleton. The COSY correlation between H3-26/H-24 and the HMBC from H-24 to C-17, C-18, and C-25 allowed the establishment of a 5-methyl-2(5H)-furanone. In the NOESY experiement of 3 (Figure 3), H3-23 correlated with H3-21 and H-12, indicating the β-orientation of Me-23 and H-12, respectively. The orientation of Me-26 was determined to be β-oriented, based on the comparison of the NMR chemical shifts of Me-26 (δH 1.37, 3H, d, J = 6.5 Hz; δC 18.5) in 3 with those of previous reported scalarane analogues, phyllactones A (δH 1.51, 3H, d, J = 6.5 Hz; δC 19.6) and B (δH 1.38, 3H, d, J = 6.5 Hz; δC 18.5) (Figure 4) [5]. Hence, the structure of 3 was determined to be a new sesterterpenoid and this metabolite was found to be the 12-epi-compound of a known 24- homoscalarane, 23-hydroxy-20-methylscalarolide [19,20], and should be named lendenfeldarane C (Supplementary Materials, Figures S17–S24).
Compound 4 (lendenfeldarane D) has a molecular formula of C30H44O6, as established by (+)- HRESIMS at m/z 523.30301 (calcd. for C30H44O6 + Na, 523.30307). The 1H and 13C NMR data indicated that 4 possessed a structural skeleton similar to that of 3 (Table 2). Comparison of the 1H and 13C NMR spectra of 4 with those of 3 revealed that the C-12 oxymethine resonance at δC 69.9 observed in 3 was moved to δC 73.8 in 4, and the C-22 oxymethene resonance at δC 63.0 observed in 3 was moved to δC 64.7 in 4. Similarly, the 1H NMR spectrum of 4 displayed two additional acetate methyl signals at δH 1.97 and 2.07 (both 3H × s), relative to 3. Therefore, the differences between compounds 3 and 4 are that the hydroxy groups at C-12 and C-22 in 3 were replaced by acetoxy groups in 4. The gross structure of 4 is supported by the HMBC and COSY correlations (Figure 2). The stereochemical configuration was identical to that of other scalarane sesterterpenes based on the NOESY cross- peaks at H-5/H-9, H-9/H-14, H3-20/H-22, H-22/H3-21, H3-21/H3-23, and H3-23/H-12 (Figure 3). Thus, the structure of 4 was determined and named as lendenfeldarane D (Supplementary Materials, Figures S25–S32).
Based on the cytotoxicity that was demonstrated by the EtOAc extract of Lendenfeldia sp., all of the isolates 17 were assessed for their cytotoxicity toward the cancer cell lines MOLT-4, K-562, U-937, and SUP-T1 (Table 3). Compound 5 showed the most potent cytotoxicity toward MOLT-4 cells (IC50 = 0.31 μM). Since both compounds 5 and 7 were the major components, it was suggested that the cytotoxicity of the extract from Lendenfeldia sp. was attributed to these two scalaranes and the aldehyde groups in 5 and 7 played a significant role in their cytotoxicity.

3. Material and Methods

3.1. General Experimental Procedures

The optical rotation values were measured using a Jasco P-1010 digital polarimeter (Jasco, Tokyo, Japan). The IR spectra were obtained with a Thermo Scientific Nicolet iS5 FT-IR spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA). The NMR spectra were recorded on a 600 or a 400 MHz Jeol ECZ NMR (Jeol, Tokyo, Japan) and a 500 MHz Varian Unity INOVA NMR spectrometer (Varian, Palo Alto, CA, USA), using the residual CHCl3 signals (δH 7.26 ppm) and CDCl3C 77.0 ppm) as the internal standards for 1H and 13C NMR, respectively. The coupling constants (J) are presented in Hz. ESIMS and HRESIMS were recorded using a Bruker 7 Tesla solariX FTMS system (Bruker, Bremen, Germany). The column chromatography was carried out with silica gel (230–400 mesh; Merck, Darmstadt, Germany). The TLC was performed on plates that were precoated with Kieselgel 60 F254 (0.25 mm thick, Merck, Darmstadt, Germany), then sprayed with 10% H2SO4 solution, followed by heating to visualize the spots. The normal-phase HPLC (NP-HPLC) was performed using a system comprising a pump (L-7110; Hitachi, Tokyo, Japan), an injection port (Rheodyne, 7725; Rohnert Park, CA, USA), and a semi-preparative normal-phase column (YMC- Pack SIL, 250 × 20 mm, 5 μm; Sigma-Aldrich, St. Louis, MO, USA). The reverse-phase HPLC (RP- HPLC) was performed using a system comprising a pump (L-2130; Hitachi), a photodiode array detector (L-2455; Hitachi), an injection port (Rheodyne; 7725), and a reverse-phase column (Luna 5 μm, C18(2) 100Å AXIA Packed, 250 × 21.2 mm; Phenomenex, Torrance, CA, USA).

3.2. Animal Material

The specimens of the marine sponge Lendenfeldia sp. were collected by hand, using self-contained underwater breathing apparatus (SCUBA), while diving off the coast of Southern Taiwan on 5 September 2012, and stored in a freezer until extraction. The sponge material was identified by Dr. Yusheng M. Huang, Department of Marine Recreation, National Penghu University of Science and Technology, Taiwan, by comparison—as described in a previous publication [21]. A voucher specimen (NMMBA-TWSP-12006) was deposited in the National Museum of Marine Biology and Aquarium, Pingtung, Taiwan.

3.3. Extraction and Isolation

The sliced bodies of Lendenfeldia sp. (wet weight 1.21 kg) were extracted with EtOAc. The EtOAc layer (5.09 g) was separated on silica gel and eluted using a mixture of n-hexane and EtOAc (stepwise, 100:1–pure EtOAc) to yield 11 fractions A–K. Fraction F was separated by NP-HPLC, using a mixture of n-hexane and EtOAc (3:1, flow rate: 3.0 mL/min) to afford seven fractions F1–F7. Fraction F3 was separated by RP-HPLC using a mixture of MeOH and H2O (7:3, flow rate: 5 mL/min) to afford 4 (1.2 mg). Fraction G was chromatographed on silica gel and eluted using n-hexane/ acetone (8:1—pure acetone) to afford eight fractions G1–G8. Fraction G3 was separated by NP-HPLC using a mixture of n-hexane and acetone (2.5:1, flow rate: 3.0 mL/min) to afford 10 fractions G3A– G3J, including Compound 7 (fraction G3C, 57.5 mg). Fraction G3D was separated by RP-HPLC using a mixture of MeOH and H2O (8:2, flow rate: 5 mL/min) to afford 3 (0.8 mg). Fraction H was separated on silica gel and eluted using a mixture of n-hexane and acetone (6:1–2:1) to obtain 14 fractions H1–H14. Fraction H5 was re-purified by NP-HPLC, using a mixture of n-hexane and acetone (5:1, flow rate: 3.0 mL/min) to afford 5 (49.2 mg) and 6 (2.0 mg), respectively. Fraction K was separated by NP-HPLC using a mixture of n-hexane and acetone (2:1, flow rate: 3.0 mL/min) to afford 1 (11.4 mg) and 2 (7.6 mg).
Lendenfeldarane A (1): amorphous powder; [α] D 20 +17 (c 0.99, CHCl3); IR (ATR) νmax 3552–2420 (broad), 3467, 1716, 1667 cm−1; 1H (CDCl3, 400 MHz) and 13C (CDCl3, 100 MHz) NMR data, see Table 1; ESIMS: m/z 497 [M + Na]+; HRESIMS: m/z 497.28736 (calcd. for C28H42O6 + Na, 497.28712).
Lendenfeldarane B (2): amorphous powder; [α] D 20 +49 (c 0.99, CHCl3); IR (ATR) νmax 3444–2309 (broad), 3399, 1740, 1730 cm−1; 1H (CDCl3, 600 MHz) and 13C (CDCl3, 150 MHz) NMR data, see Table 1; ESIMS: m/z 471 [M + Na]+; HRESIMS: m/z 471.27171 (calcd. for C26H40O6 + Na, 471.27142).
Lendenfeldarane C (3): amorphous powder; [α] D 20 +56 (c 0.04, CHCl3); IR (ATR) νmax 3436, 1731 cm−1; 1H (CDCl3, 500 MHz) and 13C (CDCl3, 125 MHz) NMR data, see Table 2; ESIMS: m/z 439 [M + Na]+; HRESIMS: m/z 439.28188 (calcd. for C26H40O4 + Na, 439.28174).
Lendenfeldarane D (4): amorphous powder; [α] D 20 +38 (c 0.05, CHCl3); IR (ATR) νmax 1738, 1672 cm−1; 1H (CDCl3, 600 MHz) and 13C (CDCl3, 150 MHz) NMR data, see Table 2; ESIMS: m/z 523 [M + Na]+; HRESIMS: m/z 523.30301 (calcd. for C30H44O6 + Na, 523.30307).
12α-Acetoxy-22-hydroxy-24-methyl-24-oxoscalar-16-en-25-al (5): amorphous powder; [α] D 20 +24 (c 2.46, CHCl3); IR (ATR) νmax 1732, 1702, 1662 cm−1; 1H (CDCl3, 400 MHz) and 13C (CDCl3, 100 MHz) NMR data were found to be in complete agreement with previous report [17]; ESIMS: m/z 481 [M + Na]+.
Felixin F (6): amorphous powder; [α] D 20 +55 (c 0.04, CHCl3) (ref. [18] [α] D 25 +54 (c 0.4, CHCl3)); IR (ATR) νmax 3430, 1701 cm1; 1H (CDCl3, 400 MHz) and 13C (CDCl3, 100 MHz) NMR data were found to be in complete agreement with previous report [18]; ESIMS: m/z 455 [M + Na]+.
24-Methyl-12,24,25-trioxoscalar-16-en-22-oic acid (7): amorphous powder; [α] D 20 +68 (c 0.04, CHCl3) (ref. [17] [α] D 21 +33.5 (c 1, CHCl3)); IR (ATR) νmax 3468–2388 (broad), 1732, 1702, 1662 cm1; 1H (CDCl3, 400 MHz) and 13C (CDCl3, 100 MHz) NMR data were found to be in complete agreement with previous report [17]; ESIMS: m/z 451 [M + Na]+.

3.4. MTT Cell Proliferative Assay

The anti-proliferative properties of the metabolites against a limited panel of human tumor cell lines, including MOLT-4, K-562, U-937, and SUP-T1, were assayed. The cell lines were purchased from the American Type Culture Collection (ATCC). The cells were seeding at 2 × 104 and were cultured in 96-well plates. The cytotoxic effect of the tested compounds was determined by the (3- (4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT)-cell proliferation assay (Sigma- M2128; Sigma-Aldrich, St. Louis, MO, USA) after 72 h. A total of 50 μL of MTT solution was added to each well for 1 h and an ELISA reader (Anthoslabtec Instrument, Salzburg, Austria) was used (OD = OD570 − OD620) for the IC50 value, calculated with CalcuSyn software.

4. Conclusions

The marine sponge belonging to the genus Lendenfeldia has proven to be a prolific producer of bioactive metabolites, especially sesterterpenoids with a scalarane skeleton. In the present study, seven 24-homoscalaranes were obtained from the Lendenfeldia sp. that was collected from the waters of Southern Taiwan, including four new 24-homoscalaranes, lendenfeldaranes A–D (14), along with three known analogues, 12α-acetoxy-22-hydroxy-24-methyl-24-oxo-scalar-16-en-25-al (5), felixin F (6), and 24-methyl-12,24,25-trioxoscalar-16-en-22-oic acid (7). The anti-cancer assessments indicated that 57 showed the most promising anti-proliferative activities against tumor cells. The structure-activity relationship (SAR) discussions also suggested the pivotal role of 18-aldehyde functionality in the activity against leukemia and lymphoma. Overall, these results can support the potential use of the marine sponge, genus Lendenfeldia, as a therapeutic agent in the treatment of cancer. We have therefore begun to culture this potentially useful sponge in tanks, using our highly developed aquaculture technology, for the extraction of natural products in order to establish a stable supply of bioactive materials, which will also protect the natural population and habitats from over-exploitation.

Supplementary Materials

The following are available online at https://www.mdpi.com/1660-3397/18/2/76/s1. HRESIMS, 1D, and 2D NMR spectra of compounds 14.

Author Contributions

B.-R.P., S.S.-F.Y., C.-Y.D., and P.-J.S. conceived and designed the experiments; B.-R.P., Y.-Y.C., J.-H.S., Y.M.H., and Y.-H.C. performed the sample collections, species identification, extraction, isolation, and structures determination; the pharmacological experiments were carried out by Y.-Y.C. and M.-C.L.; J.-H.S. and P.-J.S. contributed reagents and analysis tools; B.-R.P., K.-H.L. and P.-J.S. participated in data interpretation, wrote the manuscript and revised the paper. All authors have read and agreed to the published version of the manuscript

Acknowledgments

This research was supported by grants from the National Museum of Marine Biology and Aquarium; the National Dong Hwa University; and the Ministry of Science and Technology (Grant Nos: MOST 106-2320-B-291-001-MY3 and 107-2320-B-291-001-MY3), Taiwan, awarded to P.-J.S.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fattorusso, E.; Magno, S.; Santacroce, C.; Sica, D. Scalarin, a new pentacyclic C-25 terpenoid from the sponge Cacospongia scalaris. Tetrahedron 1972, 28, 5993–5997. [Google Scholar] [CrossRef]
  2. Mo, S.; Krunic, A.; Pegan, S.D.; Franzblau, S.G.; Orjala, J. An antimicrobial guanidine-bearing sesterterpene from the cultured cyanobacterium Scytonema sp. J. Nat. Prod. 2009, 72, 2043–2045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. González, M.A. Scalarane sesterterpenoids. Curr. Bioact. Compd. 2010, 6, 178–206. [Google Scholar] [CrossRef] [Green Version]
  4. Marshall, L.A.; Winkler, J.D.; Griswold, D.E.; Bolognese, B.; Roshak, A.; Sung, C.-M.; Webb, E.F.; Jacobs, R. Effects of scalaradial, a type II phospholipase A2 inhibitor, on human neutrophil arachidonic acid mobilization and lipid mediator formation. J. Pharmacol. Exp. Ther. 1994, 268, 709–717. [Google Scholar]
  5. Fu, X.; Zeng, L.M.; Su, J.Y.; Pais, M.; Potier, P. Scalarane-type bishomosesterterpenes from the sponge Phyllospongia foliascens. J. Nat. Prod. 1992, 55, 1607–1613. [Google Scholar] [CrossRef]
  6. Jiménez, J.I.; Yoshida, W.Y.; Scheuer, P.J.; Lobkovsky, E.; Clardy, J.; Kelly, M. Honulactones:  new bishomoscalarane sesterterpenes from the Indonesian sponge Strepsichordaia aliena. J. Org. Chem. 2000, 65, 6837–6840. [Google Scholar] [CrossRef]
  7. Chill, L.; Rudi, A.; Aknin, M.; Loya, S.; Hizi, A.; Kashman, Y. New sesterterpenes from Madagascan Lendenfeldia sponges. Tetrahedron 2004, 60, 10619–10626. [Google Scholar] [CrossRef]
  8. Cimino, G.; De Rosa, S.; De Stefano, S.; Sodano, G. The chemical defense of four Mediterranean nudibranchs. Comp. Biochem. Physiol. 1982, 73, 471–474. [Google Scholar] [CrossRef]
  9. Rogers, S.D.; Paul, V.J. Chemical defenses of three Glossodoris nudibranchs and their dietary Hyrtios sponges. Mar. Ecol. Prog. Ser. 1991, 77, 221–232. [Google Scholar] [CrossRef]
  10. Avila, C.; Paul, V.J. Chemical ecology of the nudibranch Glossodoris pallida: Is the location of diet-derived metabolites important for defense? Mar. Ecol. Prog. Ser. 1997, 150, 171–180. [Google Scholar] [CrossRef]
  11. Kamel, H.N.; Kim, Y.B.; Rimoldi, J.M.; Fronczek, F.R.; Ferreira, D.; Slattery, M. Scalarane sesterterpenoids: Semisynthesis and biological activity. J. Nat. Prod. 2009, 72, 1492–1496. [Google Scholar] [CrossRef]
  12. Bergquist, P.R.; Cambie, R.C.; Kernan, M.R. Scalarane sesterterpenes from Collospongia auris, a new thorectid sponge. Biochem. Syst. Ecol. 1990, 18, 349–357. [Google Scholar] [CrossRef]
  13. Gavagnin, M.; Mollo, E.; Docimo, T.; Guo, Y.-W.; Cimino, G. Scalarane metabolites of the nudibranch Glossodoris rufomarginata and its dietary sponge from the South China Sea. J. Nat. Prod. 2004, 67, 2104–2107. [Google Scholar] [CrossRef]
  14. Youssef, D.T.A.; Shaala, L.A.; Emara, S. Antimycobacterial scalarane-based sesterterpenes from the Red Sea sponge Hyrtios erecta. J. Nat. Prod. 2005, 68, 1782–1784. [Google Scholar] [CrossRef]
  15. Chang, L.C.; Otero-Quintero, S.; Nicholas, G.M.; Bewley, C.A. Phyllolactones A–E: New bishomoscalarane sesterterpenes from the marine sponge Phyllospongia lamellosa. Tetrahedron 2001, 57, 5731–5738. [Google Scholar] [CrossRef]
  16. Nam, S.-J.; Ko, H.; Shin, M.; Ham, J.; Chin, J.; Kim, Y.; Kim, H.; Shin, K.; Choi, H.; Kang, H. Farnesoid X- activated receptor antagonists from a marine sponge Spongia sp. Bioorg. Med. Chem. Lett. 2006, 16, 5398–5402. [Google Scholar] [CrossRef]
  17. Kazlauskas, R.; Murphy, P.T.; Wells, R.J. Five new C26 tetracyclic terpenes from a sponge (Lendenfeldia sp.). Aust. J. Chem. 1982, 35, 51–59. [Google Scholar] [CrossRef]
  18. Lai, Y.-Y.; Chen, L.-C.; Wu, C.-F.; Lu, M.-C.; Wen, Z.-H.; Wu, T.-Y.; Fang, L.-S.; Wang, L.-H.; Wu, Y.-C.; Sung, P.-J. New cytotoxic 24-homoscalarane sesterterpenoids from the sponge Ircinia felix. Int. J. Mol. Sci. 2015, 16, 21950–21958. [Google Scholar] [CrossRef]
  19. Hochlowski, J.E.; Faulkner, D.J.; Bass, L.S.; Clardy, J. Metabolites of the dorid nudibranch Chromodoris sedan. J. Org. Chem. 1983, 48, 1738–1740. [Google Scholar] [CrossRef]
  20. Harinantenaina, L.; Brodie, P.J.; Maharavo, J.; Bakary, G.; TenDyke, K.; Shen, Y.; Kingston, D.G.I. Antiproliferative homoscalarane sesterterpenes from two Madagascan sponges. Bioorg. Med. Chem. 2013, 21, 2912–2917. [Google Scholar] [CrossRef] [Green Version]
  21. Bergquist, P.R. A revision of the supraspecific classification of the orders Dictyoceratida, Dendroceratida, and Verongida (class Demospongiae). N. Z. J. Zool. 1980, 7, 443–503. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The structures of lendenfeldaranes A–D (14), 12α-acetoxy-22-hydroxy-24-methyl-24-oxo- scalar-16-en-25-al (5), felixin F (6), and 24-methyl-12,24,25-trioxoscalar-16-en-22-oic acid (7).
Figure 1. The structures of lendenfeldaranes A–D (14), 12α-acetoxy-22-hydroxy-24-methyl-24-oxo- scalar-16-en-25-al (5), felixin F (6), and 24-methyl-12,24,25-trioxoscalar-16-en-22-oic acid (7).
Marinedrugs 18 00076 g001
Figure 2. The key COSY correlations ( Marinedrugs 18 00076 i001) and heteronuclear multiple bond correlation (HMBC) ( Marinedrugs 18 00076 i002) of 14.
Figure 2. The key COSY correlations ( Marinedrugs 18 00076 i001) and heteronuclear multiple bond correlation (HMBC) ( Marinedrugs 18 00076 i002) of 14.
Marinedrugs 18 00076 g002
Figure 3. The selected NOESY correlations ( Marinedrugs 18 00076 i003) of 14.
Figure 3. The selected NOESY correlations ( Marinedrugs 18 00076 i003) of 14.
Marinedrugs 18 00076 g003
Figure 4. The 1H and 13C NMR chemical shifts in the methyl group in the α,β-unsaturated-γ-lactone moiety in phyllactones A and B and lendenfeldaranes C (3) and D (4).
Figure 4. The 1H and 13C NMR chemical shifts in the methyl group in the α,β-unsaturated-γ-lactone moiety in phyllactones A and B and lendenfeldaranes C (3) and D (4).
Marinedrugs 18 00076 g004
Table 1. The 1H and 13C NMR data for 24-homoscalaranes 1 and 2 (CDCl3).
Table 1. The 1H and 13C NMR data for 24-homoscalaranes 1 and 2 (CDCl3).
C/H12
δH (J in Hz) aδC Multiple bδH (J in Hz) cδC Multiple d
12.12 m; 0.48 ddd (12.8, 12.8, 3.2)34.4, CH22.13 m; 0.80 m33.8, CH2
21.50 m17.8, CH21.44 m; 1.65 m17.9, CH2
31.19 m; 1.43 m41.6, CH21.18 m; 1.43 m41.4, CH2
4 33.0, C 33.0, C
50.94 m56.8, CH0.95 br d (12.6)56.8, CH
61.43 m18.3, CH21.53 m18.2, CH2
71.03 m; 1.82 ddd (12.8, 3.2, 3.2)42.2, CH20.96 m; 1.91 ddd (13.2, 3.6, 3.6)42.1, CH2
8 37.8, C 38.7, C
91.50 m49.3, CH1.26 m62.8, CH
10 41.7, C 42.8, C
112.07 m24.9, CH22.62 dd (12.6, 1.8); 3.34 dd (14.4, 12.6)39.0, CH2
124.77 br s75.7, CH 221.9, C
13 38.9, C 52.8, C
141.31 m52.2, CH1.21 m58.0, CH
152.27 m23.3, CH21.65 m; 1.94 ddd (12.6, 4.2, 1.8)30.1, CH2
166.90 br s139.7, CH3.53 ddd (10.8, 10.8, 4.8)72.7, CH
17 137.9, C3.22 dd (12.0, 10.8)54.8, CH
183.79 br s48.0, CH3.18 d (12.0)51.3, CH
190.75 s21.9, CH30.76 s21.7, CH3
200.86 s33.8, CH30.87 s33.7, CH3
211.16 s16.1, CH31.30 s16.4, CH3
223.85 d (11.6); 4.02 d (11.6)62.9, CH23.93 dd (11.4, 1.2); 4.08 d (11.4)62.7, CH2
230.96 s15.5, CH31.34 s15.3, CH3
24 199.3, C 212.6, C
25 175.1, C 172.4, C
262.29 s25.3, CH32.40 s33.4, CH
OAc-12 170.4, C
2.14 s21.5, CH3
a 400 MHz. b 100 MHz. c 600 MHz. d 150 MHz.
Table 2. The 1H and 13C NMR data for 24-homoscalaranes 3 and 4 (CDCl3).
Table 2. The 1H and 13C NMR data for 24-homoscalaranes 3 and 4 (CDCl3).
C/H34
δH (J in Hz) aδC Multiple bδH (J in Hz) cδC Multiple d
12.16 m; 0.80 m34.0, CH22.01 m; 0.52 ddd (13.8, 13.8, 3.0)34.7, CH2
21.63 m18.4, CH21.63 m18.2, CH2
31.18 m; 1.43 m41.7, CH21.15 ddd (9.0, 9.0, 4.2); 1.43 m41.5, CH2
4 33.0, C 33.0, C
51.05 m56.8, CH1.01 dd (13.8, 3.6)57.1, CH
61.56 m; 1.91 m16.9, CH21.56 m; 1.93 m17.0, CH2
71.10 m; 1.89 m42.0, CH21.11 ddd (12.0, 12.0, 4.2); 1.92 m41.9, CH2
8 37.6, C 37.5, C
91.56 m52.3, CH1.26 m53.2, CH
10 41.8, C 40.2, C
111.89 m; 2.18 m27.1, CH21.99 m; 2.20 m23.3, CH2
124.60 br s69.9, CH5.54 t (3.0)73.8, CH
13 40.2, C 38.4, C
141.60 m50.0, CH1.55 m51.2, CH
152.18 m; 2.35 m24.1, CH22.23 m; 2.39 m24.0, CH2
161.46 m; 1.56 m18.0, CH21.60 m18.0, CH2
17 165.2, C 163.6, C
18 133.5, C 132.6, C
190.78 s21.8, CH30.83 s21.9, CH3
200.86 s33.9, CH30.89 s33.7, CH3
211.08 s16.3, CH30.98 s16.4, CH3
223.92 d (11.5); 4.05 d (11.5)63.0, CH24.15 dd (12.0, 1.2); 4.58 d (12.0)64.7, CH2
231.13 s21.7, CH31.19 s21.3, CH3
244.79 q (6.5)78.2, CH4.78 q (6.0)77.7, CH
25 172.6, C 171.1, C
261.37 d (6.5)18.5, CH31.36 d (6.0)18.6, CH3
OAc-12 169.9, C
1.97 s21.2, CH3
OAc-22 170.9, C
2.07 s21.2, CH3
a 500 MHz. b 125 MHz. c 600 MHz. d 150 MHz.
Table 3. The anti-proliferative effects of scalaranes 17.
Table 3. The anti-proliferative effects of scalaranes 17.
CompoundCell lines IC50 (μM)
MOLT-4K-562U-937SUP-T1
139.54NANA33.02
234.93NANANA
36.3111.695.749.00
429.83NANANA
50.313.042.355.90
65.679.716.9712.33
71.491.045.887.49
Doxorubicin a0.020.130.040.09
a Doxorubicin was used as a positive control; NA: not active at 20 μg/mL for 72 h.

Share and Cite

MDPI and ACS Style

Peng, B.-R.; Lai, K.-H.; Chen, Y.-Y.; Su, J.-H.; Huang, Y.M.; Chen, Y.-H.; Lu, M.-C.; Yu, S.S.-F.; Duh, C.-Y.; Sung, P.-J. Probing Anti-Proliferative 24-Homoscalaranes from a Sponge Lendenfeldia sp. Mar. Drugs 2020, 18, 76. https://doi.org/10.3390/md18020076

AMA Style

Peng B-R, Lai K-H, Chen Y-Y, Su J-H, Huang YM, Chen Y-H, Lu M-C, Yu SS-F, Duh C-Y, Sung P-J. Probing Anti-Proliferative 24-Homoscalaranes from a Sponge Lendenfeldia sp. Marine Drugs. 2020; 18(2):76. https://doi.org/10.3390/md18020076

Chicago/Turabian Style

Peng, Bo-Rong, Kuei-Hung Lai, You-Ying Chen, Jui-Hsin Su, Yusheng M. Huang, Yu-Hsin Chen, Mei-Chin Lu, Steve Sheng-Fa Yu, Chang-Yih Duh, and Ping-Jyun Sung. 2020. "Probing Anti-Proliferative 24-Homoscalaranes from a Sponge Lendenfeldia sp." Marine Drugs 18, no. 2: 76. https://doi.org/10.3390/md18020076

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop