Next Article in Journal
Two C23-Steroids and a New Isocoumarin Metabolite from Mangrove Sediment-Derived Fungus Penicillium sp. SCSIO 41429
Next Article in Special Issue
An Overview of Secondary Metabolites from Soft Corals of the Genus Capnella over the Five Decades: Chemical Structures, Pharmacological Activities, NMR Data, and Chemical Synthesis
Previous Article in Journal
Chrysomycin A Reshapes Metabolism and Increases Oxidative Stress to Hinder Glioblastoma Progression
Previous Article in Special Issue
Marine Prostanoids with Cytotoxic Activity from Octocoral Clavularia spp.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

New Meroterpenes from South China Sea Soft Coral Litophyton brassicum

1
Li Dak Sum Yip Yio Chin Kenneth Li Marine Biopharmaceutical Research Center, Ningbo University, Ningbo 315211, China
2
Institute of Drug Discovery Technology, Ningbo University, Ningbo 315211, China
3
Ningbo Institute of Marine Medicine, Peking University, Ningbo 315800, China
4
Xingzhi College, Zhejiang Normal University, Lanxi 321004, China
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2024, 22(9), 392; https://doi.org/10.3390/md22090392
Submission received: 31 July 2024 / Revised: 27 August 2024 / Accepted: 28 August 2024 / Published: 30 August 2024
(This article belongs to the Special Issue Bioactive Compounds from Soft Corals and Their Derived Microorganisms)

Abstract

:
A chemical investigation of the extracts from the soft coral Litophyton brassicum led to the isolation and identification of four new meroterpenes, brassihydroxybenzoquinone A and B (1 and 2) and brassinaphthoquinone A and B (3 and 4), along with two known related meroterpenes (5 and 6). Their structures were elucidated using high-resolution electrospray ionization mass spectrometry (HRESIMS), nuclear magnetic resonance (NMR) spectroscopy, and a comparison with the literature data. All compounds were evaluated for antibacterial activity against six pathogenic bacterial strains and for cytotoxic activity against three cancer cell lines. In the cytotoxic assay, all compounds were inactive at 10 μM against the A549, HeLa, and MDA-MB-231 cell lines. In the antibacterial assay, compounds 1 and 2 exhibited moderate inhibitory activity with minimum inhibitory concentrations (MIC) ranging from 8 to 64 μg/mL.

1. Introduction

The marine soft coral genus Litophyton belongs to the family Nephtheidae, order Alcyonacea, subclass Octocorallia. Currently, Litophyton comprises nearly 100 species, widely distributed throughout tropical and temperate waters such as the South China Sea, the Red Sea, and other regions of the Indo-Pacific Ocean [1]. Chemical investigations on Litophyton soft corals have revealed them to be prolific producers of bioactive secondary metabolites. Since the early 1970s, when novel cembrane diterpenes were first reported from L. viridis [2], numerous research groups worldwide have conducted chemical investigations on Litophyton, leading to significant discoveries. To date, nearly 200 secondary metabolites have been isolated and characterized from Litophyton corals over 50 years of research [1,3]. These compounds include sesquiterpenes, sesquiterpene dimers, diterpenes, norditerpenes, tetraterpenes, meroterpenes, steroids, ceramides, pyrimidines, and peptides [1]. A broad spectrum of pharmacological activities has been evaluated, including cytotoxic [4,5,6,7], antiviral [5], antibacterial [8], antifungal [9], antimalarial [10], anti-inflammatory [11], PTP1B inhibitory [12,13].
Meroterpenes, a class of compounds found in Litophyton, are relatively rare in the literature. For example, four meroterpenes isolated from the Red Sea soft coral Nephthea sp. were identified as potential SARS-CoV-2 main protease inhibitors [14]. Nine new meroditerpenoid-related metabolites were isolated from the Formosan soft coral Nephthea chabrolii, with proposed biosynthetic pathways [15]. Then, the next year, eight new meroditerpenoid-related metabolites were isolated from the organic extract of a Taiwanese soft coral Nephthea chabrolii [16]. It is noteworthy that the genus Nephthea was synonymized with Litophyton in 2016 due to their identical characteristics [1]. As part of our ongoing efforts to discover bioactive marine natural products from soft corals, specimens of Litophyton brassicum were collected from the South China Sea. Herein, we report the isolation and identification of six meroterpenes. The antibacterial activities of these compounds were evaluated against six pathogenic bacteria, and their antiproliferative effects were tested on three cancer cell lines.

2. Results

The acetone extract of soft coral Litophyton brassicum was subjected to repeated silica gel and reversed-phase silica gel column chromatography, followed by semipreparative HPLC, to afford four new meroterpenes (14) and two known compounds (5 and 6) (Figure 1). The known compounds were identified as chabrolohydroxybenzoquinone G [16] and chabrolohydroxybenzoquinone B [15].
Compound 1 was obtained as an optically active colorless oil. Its molecular formula was determined to be C26H38O4 based on HRESIMS analysis [M-H2O + H]+ m/z 397.2755, calcd. for C26H37O3, 397.2742), indicating eight degrees of unsaturation. In the 1H NMR spectrum, two aromatic signals resonating at δH 6.55 (1H, br. s) and 6.42 (1H, br. s) indicated the presence of a 1, 2, 4, 5-tetrasubstituted benzene system. Four olefinic protons δH 6.25 (1H, d, J = 9.8 Hz), 5.52 (1H, d, J = 9.8 Hz), 5.09 (1H, tq, J = 7.2, 1.4 Hz), and 5.05 (1H, tq, J = 7.4, 1.4 Hz), indicated the presence of regular prenyl moieties. The 1H, 13C NMR, DEPT, and HSQC spectra data confirmed the presence of 26 carbons, including a ketone carbonyl at δC 211.4, a benzene system, one disubstituted double bond [δC 122.6 (C-1), δC 129.9 (C-2)], two trisubstituted double bonds [δC 125.0 (C-6), δC 134.7 (C-7), δC 123.0 (C-14), and δC 132.8 (C-15)], an oxygenated quaternary carbon (δC 78.1), seven methylenes, and five methyls [δC 2.18 (C-7′), δH 1.60 (C-16), δH 1.66 (C-17), δH 1.55 (C-18), and δH 1.35 (C-19)]. The constitution of the side chain was elucidated initially by the 1H-1H COSY correlations of H-1/H-2, H2-4/H2-5/H-6, H2-8/H2-9/H2-10, and H2-12/H2-13/H-14 (Figure 2). The spectroscopic data were similar to those of the known compound chabrolohydroxybenzoquinones E [16], except for the missing of a methyl group on C-11, and the Δ10,11 double bond was reduced and oxidized to a keto group in 1. In addition, the HMBC correlations from H2-9, H2-10, H2-12, and H2-13 to C-11 implied the keto group was on C-11. The smaller coupling constant (J1,2 = 9.8 Hz) and the 1D NOE correlation (Figure 3) from H-6 to H2-8 suggested the Z configuration of the Δ1,2 and E configuration of the Δ6,7 double bonds, respectively. Comprehensive HMBC analysis allowed the complete assignment of the proton and carbon signals for 1 (Table 1 and Figure 2). As a result, the structure of 1 was elucidated as shown in Figure 1, named brassihydroxybenzoquinone A.
Compound 2 (brassihydroxybenzoquinone B) was also obtained as a colorless oil. The molecular formula of C27H40O4 derived by the HR-ESI-MS ion peak of [M-2H2O + H]+ at m/z 393.2798 (calcd. for C27H37O2, 393.2793) gave 8 degrees of unsaturation. The 1H and 13C NMR spectra of 2 (Table 1) were also similar to those of chabrolohydroxybenzoquinones E [16], except that the oxygen-bearing methylene [δC 60.5 (C-18)/δH 4.10] attached at C-11 in 2 rather than the methyl group in chabrolohydroxybenzoquinones E, which was proven by the HMBC correlations from H2-18 to C-10 (δC 128.8), C-11 (δC 138.4), and C-12 (δC 35.3). The configuration of the Δ1,2 double bond was also assigned as Z, based on the same method as for 1. The E geometry of Δ6,7 and Z geometry of Δ10,11 double bonds in 2 were supported by the 1D NOE enhancements of H-6/H2-8 and H-10/H2-12, respectively. We attempted to utilize the TDDFT-ECD calculation method to establish the absolute configuration of the C-3 position in compounds 1 and 2. However, the experimental ECD spectrum did not exhibit a clear Cotton effect (Figures S36 and S37), which is likely due to the C-3 position being located on the side chain, leading to its flexibility. As far as we know, determining the absolute configuration of C-3 in compounds 1 and 2 remains a challenging task.
Compound 3 (brassinaphthoquinone A) was a yellow oil. The molecular formula of C27H34O3, giving 11 degrees of unsaturation, was established by the HR-ESI-MS ion peak at m/z 389.2484 [M-H2O + H]+ (calcd. for C27H33O2, 389.2480). From the 1H NMR spectrum of 3, the resonances of three aromatic protons δH 7.96 (1H, d, J = 9.8 Hz), 7.90 (1H, d, J = 1.8 Hz), and 7.52 (1H, dd, J = 9.8, 1.8 Hz) indicated the presence of a 1, 2, 4-trisubstituted benzene system. One additional aromatic proton 6.81 (1H, br.s) was also observed. From the 1H and 13C NMR spectral data (Table 1), together with the HSQC data, 27 signals were assigned to two carbonyls, seven sp2 quaternary olefinic carbons, seven sp2 methine, an oxygenated sp3 methine [δC 66.1 (C-9)/δH 4.42], five sp3 methylene, and five methyls. The NMR data of 3 closely resembled those of chabrolonaphthoquinone A [15], a meroditerpenoid with a naphthoquinone moiety obtained from the Formosan soft coral Nephthea chabrolii. The differences were one more methyl group (C-18) on 3 rather than the carboxyl group on chabrolonaphthoquinone A, and C-9 was hydroxylated on 3. The different side chain was elucidated by the 1H-1H COSY correlations of H2-8/H-9/H-10 and the HMBC correlations from H-9 to C-7, C-10, and C-11. The key 1D NOE correlations (Figure 3) from H-6 to H2-8 and from H-10 to H2-12 established the E configuration of the Δ6,7 and Δ10,11 double bonds, respectively. Thus, the structure of 3 was elucidated, as shown in Figure 2.
Brassinaphthoquinone B (4) was also isolated as a yellow oil with a molecular formula of C27H34O3 on the basis of HR-ESI-MS ion peak at m/z 389.2471 [M-H2O + H]+ (calcd. for C27H33O2, 389.2480), which is the structural isomer of 3. The 1H and 13C NMR data (Table 1) of 4 were extremely similar to those of 3, except that the absence of hydroxyl group at C-9 in 4, and the oxygenated sp3 carbon [δC 60.5 (C-18)/δH 4.11] was a methylene in 4 rather than a methine in 3. It was deduced by analysis of the 1H-1H COSY correlations of H2-8/H2-9/H-10 and the HMBC correlations from H2-18 to C-10, C-11, and C-12. The 1D NOE enhancements between H-6/H2-8 and H-10/H2-12 indicated the E geometry of the Δ6,7 and Δ10,11 double bonds, respectively.
Compounds 16 were evaluated for cytotoxicity and antibacterial activities. In the cytotoxic assay, all compounds were inactive at 10 μM against the cell lines of A549, HeLa, and MDA-MB-231. An evaluation of the antibacterial activity against six pathogenic bacterial strains (Staphylococcus aureus, Escherichia coli, Pseudomonas aeruginosa, Bacillus subtilis, Vibrio parahaemolyticus, and Vibrio harveyi) showed that compounds 1 and 2 displayed moderate inhibitory activity (MIC 8–64 μg/mL) (Table 2). Interestingly, compound 2 may have potential as an antibiotic agent for controlling aquatic pathogens in the future.

3. Materials and Methods

3.1. General Chemical Experimental Procedures

NMR spectra were recorded on a Bruker AVANCE NEO 600 spectrometer (BrukerBiospin AG, Fällanden, Germany). 1H chemical shifts were referenced to the residual CDCl3 (7.26 ppm), and 13C chemical shifts were referenced to the CDCl3 (77.2 ppm) solvent peaks. High-resolution electrospray ionization mass spectra (HRESIMS) were performed on an ultra-high-performance liquid chromatograph (UPLC) and TIMS-QTOF high-resolution mass spectrometry (Waters, MA, USA). The purification was performed by reversed-phase high-performance liquid chromatography using a Shimadzu LC-20AT system (Shimadzu Corporation, Tokyo, Japan). The solvents used for HPLC were all Fisher HPLC grade. A Cosmosil C18-MS-II column (250 mm × 20.0 mm, id, 5 μm, Cosmosil, Nakalai Tesque Co., Ltd., Kyoto, Japan) was used for the preparative HPLC separation. Column chromatography was performed using silica gel (300–400 mesh, Qingdao Ocean Chemical Co., Ltd., Qingdao, China) and C18 reversed-phase silica gel (75 µm, Nakalai Tesque Co., Ltd., Kyoto, Japan).

3.2. Animal Material

Soft coral Litophyton brassicum was sampled off the coast of Xisha Islands, South China Sea, 12 m underwater, with a wet weight of 5.07 kg, and it was frozen immediately after collection. The specimens (XSSC201906) were deposited at the Li Dak Sum Yip Yio Chin Kenneth Li Marine Biopharmaceutical Research Center, Health Science Center, Ningbo University, China.

3.3. Extraction and Isolation

The soft coral samples were vacuum freeze-dried with a freeze-dryer, crushed in a pulverizer, and fully soaked in acetone at room temperature for 2 days each time, followed by ultrasonic extraction for 1 h, repeated soaking ultrasonic extraction for 4–5 times. The extract was filtered to remove the sample residue, and the extract was concentrated under reduced pressure. The extract was partitioned three times with Et2O and water (1:1, v:v), and the Et2O layer extract was concentrated under reduced pressure to obtain 80 g brown residue.
The 80 g of the extract was separated by gradient elution on a normal-phase silica gel column, yielding 11 fractions (FrA–FrK). Fr.G (1.8 g) was eluted with MeOH/H2O (75:15 to 100:0, v/v) on reversed-phase column chromatography to obtain five subfractions (Fr.G.1–Fr.G.5). Purification of Fr.G.3 by semi-preparative HPLC (MeCN/H2O, 72:18, 2 mL/min) gave compounds 1 (6.8 mg, tR = 62 min) and 3 (5.8 mg, tR = 71 min). Separation of Fr.H (1.54 g) on a reversed-phase column with MeOH/H2O (75:15~100:0, v/v) afforded seven subfractions (Fr.H.1~Fr.H.6). Fr.H.5 was purified by semipreparative reversed-phase HPLC (MeCN/H2O, 75: 15, 2 mL/min) to provide compounds 2 (5.7 mg, tR = 60 min) and 4 (7.8 mg, tR = 40 min). Separation of Fr.J (1.4334 g) on a reversed-phase column with MeOH/H2O (80:20~100:0, v/v) provided seven subfractions (Fr.J.1~Fr.J.6). Fr.J.6 was purified by semipreparative reversed-phase HPLC (MeCN/H2O, 80:20, 2 mL/min) to provide compounds 5 (6.7 mg, tR = 53 min) and 6 (8.8 mg, tR = 62 min).
Brassihydroxybenzoquinone A (1): colorless oil; {[α ] D 25 −17.40 (c 0.5, MeOH)}; UV (MeOH): 221 (3.38), 270 (2.51), 330 (2.57); IR (KBr) ν = 3400, 3100, 1766, 1651, 1261, 1079, 888, 630 cm−1; 1H and 13C NMR data, Table 1; HRESIMS m/z 397.2755 [M-H2O + H]+ (calcd. for C26H37O3, 397.2742).
Brassihydroxybenzoquinone B (2): colorless oil; {[α ] D 25 −18.67 (c 0.5, MeOH)}; UV (MeOH): 220 (3.27), 270 (2.48), 330 (2.55); IR (KBr) ν = 3574, 3154, 1436, 1261, 1037, 951, 710 cm−1; 1H and 13C NMR data, Table 1; HRESIMS m/z 393.2798 [M-2H2O + H]+ (calcd. for C27H37O2, 393.2793).
Brassinaphthoquinone A (3): yellow oil; {[α ] D 25 −23.47 (c 0.5, MeOH)}; UV (MeOH): 204 (3.39), 237 (3.23), 270 (3.05); IR (KBr) ν = 3582, 3088, 1765, 1692, 1450, 1139, 916, 749 cm−1; 1H and 13C NMR data, Table 1; HRESIMS m/z 389.2484 [M-H2O + H]+ (calcd. for C27H33O2, 389.2480).
Brassinaphthoquinone B (4): yellow oil; {[α ] D 25 −23.73 (c 0.5, MeOH)}; UV (MeOH): 204 (3.34), 245 (3.13), 270 (2.90); IR (KBr) ν = 3583, 2928, 1759, 1692, 1470, 1138, 915, 749 cm−1; 1H and 13C NMR data, Table 1; HRESIMS m/z 389.2480 [M-H2O + H]+ (calcd. for C27H33O2, 389.2481).

3.4. Antibacterial Assays

All isolated compounds were tested for antibacterial activities according to established methods [17]. Six bacterial strains were selected: S. aureus [CMCC (B) 26003], B. subtilis [CMCC (B) 63501], V. harveyi 1708B04 (accession number: MZ333451), V. parahaemolyticus (accession number: OL636376), P. aeruginosa [CMCC (B) 10104], and E. coli [CMCC (B) 44102], with penicillin G serving as the positive control. Compounds 16 were dissolved in DMSO and tested at concentrations of 128, 64, 32, 16, 8, 4, and 2 µg/mL. Briefly, bacteria were grown in Mueller–Hinton (MH) medium for 24 h at 28 °C with agitation (180 rpm), then diluted with sterile MH medium to match the 0.5 McFarland standard. One hundred microliters of each bacterial suspension and 100 μL of MH medium containing 0.002% 2,3,5-triphenyltetrazolium chloride, along with test or control compounds, were incubated. Inhibition data were recorded optically.

3.5. Cytotoxic Activity Assays

MDA-MB-231, HeLa, and A549 cells were cultured in DMEM (Gibco, Thermo Fisher Scientific, Inc., Waltham, MA, USA) supplemented with 10% FBS (Gibco, Thermo Fisher Scientific, Inc.). These cell lines were obtained from the Shanghai Cell Bank, Chinese Academy of Sciences. The cells were incubated at 37 °C in a humidified atmosphere containing 5% CO2. MTT assays were performed as described by Zhang et al. (2011) [18]. Briefly, cells were seeded into 96-well plates at a density of 5 × 103 cells/well and incubated for 12 h, followed by exposure to various test compounds at different concentrations for 48 h. Subsequently, the cells were stained with 20 μL of MTT solution (5 mg/mL) for 4 h. The medium and MTT solution were then removed, and 150 μL of DMSO was added to dissolve the formazan crystals. The plates were shaken at low speed for 10 min. Absorbance was measured at 490 nm using a microplate reader.

4. Conclusions

In summary, the chemical study of the soft coral Litophyton brassicum from the South China Sea led to the identification of four novel meroterpenes, brassihydroxybenzoquinone A and B and brassinaphthoquinone A and B, along with two known related meroterpenes. The structures of these compounds were elucidated using HRESIMS and NMR spectroscopy and corroborated with the existing literature. Compounds 1 and 2 exhibited moderate antibacterial activity, with minimum inhibitory concentrations (MIC) ranging from 8 to 64 μg/mL. The isolation of these novel terpenoids from marine soft corals underscores the rich chemical diversity of the Litophyton genus. These newly characterized compounds, with their unique structures, hold promise for developing innovative antimicrobial therapies, particularly amid escalating antibiotic resistance. Exploring their antimicrobial properties is, therefore, a critical endeavor for public health.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/md22090392/s1, Figures S1–S35: HRESIMS, IR, 1D and 2D NMR spectra of all new compounds 14; Figures S36 and S37: Experimental ECD spectra of compounds 1 and 2.

Author Contributions

X.C., J.Z., J.Y., and B.L. performed the extraction, isolation, and structure determination of the compounds and wrote the manuscript. H.O., H.H., and X.Y. performed the cytotoxic and antimicrobial bioassays. T.L., W.L., and S.H. collected the soft coral by SCUBA. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Fujian Provincial Key Laboratory of Innovative Drug Target Research (FJ-YW-2022KF01), the Li Dak Sum Yip Yio Chin Kenneth Li Marine Biopharmaceutical Development Fund of Ningbo University, the Ningbo Key Science and Technology Development Program (2021Z046), and the National 111 Project of China (D16013).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the Supplementary Materials file associated with this article.

Acknowledgments

We thank Zhenhua Long, Daning Li, and Da Huo of the Xisha Marine Science Comprehensive Experimental Station, South China Sea Institute of Oceanology, Chinese Academy of Sciences, for their assistance. We thank Xisha Marine Environment National Observation and Research Station. We would also like to thank the MS Center of the Institute of Drug Discovery Technology, Ningbo University.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yan, X.-Y.; Zhang, L.; Yang, Q.-B.; Ge, Z.-Y.; Liang, L.-F.; Guo, Y.-W. Genus Litophyton: A hidden treasure trove of structurally unique and diversely bioactive secondary metabolites. Mar. Drugs 2023, 21, 523. [Google Scholar] [CrossRef] [PubMed]
  2. Tursch, B. Chemical studies of marine invertebrates XIII: 2-Hydroxynephtenol, a novel cembrane diterpene from the soft coral Litophyton viridis. Bull. Soc. Chim. Belg. 1975, 84, 767–774. [Google Scholar] [CrossRef]
  3. Zong, Y.; Yang, J.-J.; Li, K.; Pei, Y.-F.; Hou, H.-Y.; Zhang, Y.; Wang, C.-L.; Li, P.L. Litoamentenes A-K, eleven undescribed cembranoids with cytotoxicity from the South China Sea soft coral Litophyton amentaceum. Phytochemistry 2024, 223, 114113. [Google Scholar] [CrossRef] [PubMed]
  4. Iwagawa, T.; Kusatsu, D.; Tsuha, K.; Hamada, T.; Okamura, H.; Furukawa, T.; Akiyama, S.-I.; Doe, M.; Morimoto, Y.; Iwase, F.; et al. Cytotoxic eunicellin-type diterpenes from the soft coral Litophyton viscudium [Erratum to document cited in CA156:358828]. Heterocycles 2012, 85, 2615. [Google Scholar] [CrossRef]
  5. Ellithey, M.S.; Lall, N.; Hussein, A.A.; Meyer, D. Cytotoxic, cytostatic and HIV-1 PR inhibitory activities of the soft coral Litophyton arboreum. Mar. Drugs 2013, 11, 4917–4936. [Google Scholar] [CrossRef] [PubMed]
  6. Zovko, K.M.; Ioannou, E.; Vagias, C.; Sawadogo, W.R.; Abdel-Razik, A.F.; Diederich, M.; Roussis, V. 4α-Methylated steroids with cytotoxic activity from the soft coral Litophyton mollis. Steroids 2016, 115, 130–135. [Google Scholar] [CrossRef] [PubMed]
  7. Mahmoud, A.H.; Zidan, S.A.H.; Samy, M.N.; Alian, A.; Abdelmohsen, U.R.; Fouad, M.A.; Kamel, M.S.; Matsunami, K. Cytotoxicity and chemical profiling of the Red Sea soft corals Litophyton arboreum. Nat. Prod. Res. 2022, 36, 4261–4265. [Google Scholar] [CrossRef] [PubMed]
  8. Abou, E.-K.L.T.; Hawas, U.W.; Al-Farawati, R.; El-Desouky, S.K. Sesquiterpenes from the Saudi Red Sea: Litophyton arboreum with their cytotoxic and antimicrobial activities. Z. Naturforsch. C 2018, 73, 9–14. [Google Scholar] [CrossRef] [PubMed]
  9. Hassan, N.H.; El-Hawary, S.S.; Emam, M.; Rabeh, M.A.; Abdelmohsen, U.R.; Selim, N.M. Potential inhibitors of CYP51 enzyme in dermatophytes by Red Sea soft coral Nephthea sp.: In silico and molecular networking studies. ACS Omega 2022, 7, 13808–13817. [Google Scholar] [CrossRef] [PubMed]
  10. Ahmed, M.M.A.; Ragab, E.A.; Zayed, A.; El-Ghaly, E.M.; Ismail, S.K.; Khan, S.I.; Ali, Z.; Chittiboyina, A.G.; Khan, I.A. Litoarbolide A: An undescribed sesquiterpenoid from the Red Sea soft coral Litophyton arboreum with an in vitro anti-malarial activity evaluation. Nat. Prod. Res. 2023, 37, 542–550. [Google Scholar] [CrossRef] [PubMed]
  11. Hsiao, T.-H.; Sung, C.-S.; Lan, Y.-H.; Wang, Y.-C.; Lu, M.-C.; Wen, Z.-H.; Wu, Y.-C.; Sung, P.-J. New anti-inflammatory cembranes from the cultured soft coral Nephthea columnaris. Mar. Drugs 2015, 13, 3443–3453. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, F.; Hua, Q.; Yao, L.-G.; Liang, L.-F.; Lou, Y.-X.; Lu, Y.-H.; An, F.-L.; Guo, Y.-W. One uncommon bis-sesquiterpenoid from Xisha soft coral Litophyton nigrum. Tetrahedron Lett. 2022, 88, 153571. [Google Scholar] [CrossRef]
  13. Yang, F.; Li, S.-W.; Zhang, J.; Liang, L.-F.; Lu, Y.-H.; Guo, Y.-W. Uncommon nornardosinane, seconeolemnane and related sesquiterpenoids from Xisha soft coral Litophyton nigrum. Bioorg. Chem. 2020, 96, 103636. [Google Scholar] [CrossRef] [PubMed]
  14. Abdelhafez, O.H.; Fahim, J.R.; Mustafa, M.; Aboulmagd, A.M. Abdelmohsen, Natural metabolites from the soft coral Nephthea sp. as potential SARS-CoV-2 main protease inhibitors. Nat. Prod. Res. 2022, 36, 2893–2896. [Google Scholar] [CrossRef] [PubMed]
  15. Sheu, J.H.; Su, J.H.; Sung, P.J.; Wang, G.H.; Dai, C.F. Novel meroditerpenoid-related metabolites from the formosan soft coral Nephthea chabrolii. J. Nat. Prod. 2004, 67, 2048–2052. [Google Scholar] [CrossRef] [PubMed]
  16. Su, J.-H.; Ahmed, A.F.; Sung, P.-J.; Wu, Y.-C.; Sheu, J.-H. Meroditerpenoids from a Formosan soft coral Nephthea chabrolii. J. Nat. Prod. 2005, 68, 1651–1655. [Google Scholar] [CrossRef] [PubMed]
  17. Shi, Y.; Pan, C.; Auckloo, B.N.; Chen, X.; Chen, C.-T.A.; Wang, K.; Wu, X.; Ye, Y.; Wu, B. Stress-driven discovery of a cryptic antibiotic produced by Streptomyces sp. WU20 from Kueishantao hydrothermal vent with an integrated metabolomics strategy. Appl. Microbiol. Biot. 2017, 101, 1395–1408. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, N.; Chen, Y.; Jiang, R.; Li, E.; Chen, X.; Xi, Z.; Guo, Y.; Liu, X.; Zhou, Y.; Che, Y. PARP and RIP 1 are required for autophagy induced by 11’-deoxyverticillin A, which precedes caspase-dependent apoptosis. Autophagy 2011, 7, 598–612. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of compounds 16.
Figure 1. Chemical structures of compounds 16.
Marinedrugs 22 00392 g001
Figure 2. 1H-1H COSY and key HMBC correlations of compounds 14.
Figure 2. 1H-1H COSY and key HMBC correlations of compounds 14.
Marinedrugs 22 00392 g002
Figure 3. Key 1D NOEs of compounds 14.
Figure 3. Key 1D NOEs of compounds 14.
Marinedrugs 22 00392 g003
Table 1. 1H NMR and 13CNMR data for compounds 1 to 4 at 600 MHz in CDCl3.
Table 1. 1H NMR and 13CNMR data for compounds 1 to 4 at 600 MHz in CDCl3.
Position1234
δH Mult (J, Hz)δC, TypeδH Mult (J, Hz)δC, TypeδH Mult (J, Hz)δC, TypeδH Mult (J, Hz)δC, Type
1′ 146.8, C 146.8, C 185.1, C 185.2, C
2′ 119.7, C 119.7, C 130.5, C 130.4, C
3′6.42, s112.6, CH6.42, s112.6, CH 132.3, C 132.0, C
4′ 147.6, C 147.6, C 186.1, C 186.1, C
5′ 124.7, C 124. 7, C 148.1, C 148.1, C
6′6.55, s118.3, CH6.55, s118.3, CH6.81, br.s135.9, CH6.81, q (1.6)135.9, CH
7′2.18, s16.1, CH32.17, s16.1, CH32.18, br.s16.4, CH32.18, d (1,6)16.6, CH3
16.25, d (9.8)122.6, CH6.25, d (9.8)122.6, CH7.96, d (7.9)126.5, CH7.96, d (7.9)126.4, CH
25.52, d (9.8)129.9, CH5.52, d (9.8)129.9, CH7.52, dd (7.9, 1.8)134.0, CH7.52, dd (7.9, 1.8)134.1, CH
3 78.1, C 78.1, C 148.8, C 149.1, C
41.63, m; 1.69, m41.1, CH21.68, m; 1.62, m41.0, CH22.80, t (7.6)36.1, CH22.77, t (7.6)36.3, CH2
52.10, m22.8, CH22.10, m22.8, CH22.40, q (7.4)29.5, CH22.35, q (7.5)29.4, CH2
65.09, tq (7.2, 1.4)125.0, CH5.10, tq (5.7, 1.4)124.86, CH5.26, m126.6, CH5.15, td (7.2,1.4)123.2, CH
7 134.7, C 134.9, C 133.6, C 136.5, C
81.93, t (7.2)39.1, CH21.98, t (7.5)40.0, CH22.13, m48.2, CH21.99, m40.0, CH2
91.64, m21.9, CH22.14, m; 2.10, m26.3, CH24.42, m66.1, CH2.14, m26.4, CH2
102.32, t (7.4)42.3, CH25.28, t (7.3)128.8, CH5.15, d (8.2)127.3, CH5.27, t (7.3)128.6 CH
11 211.4, C 138.4, C 138.5, C 138.6, C
122.40, t (7.4)43.0, CH22.11, m35.3, CH22.00, t (7.7)39.7, CH22.11, m35.4, CH2
132.23, q (7.4)22. 7, CH22.10, m27.2, CH22.07, m26.5, CH22.11, m27.2, CH2
145.05, tp (7.2, 1.4)123.0, CH5.10, tq (5.7, 1.4)124.3, CH5.07, tq (6.9, 1.4)124.1, CH5.10, tq (5.4, 1.2)124.3, CH
15 132. 8, C 131.9, C 131.8, C 131.2, C
161.60, s17.8, CH31.68, s25.9, CH31.59, s17.8, CH31.60, s17.9, CH3
171.66, s25.8, CH31.60, s17.9, CH31.67, s25.8, CH31.68, s25.9, CH3
181.55, s15.8, CH34.10, s60.5, CH21.66, s16.6, CH34.11, s60.5, CH2
191.35, s26.2, CH31.57, s16.1, CH31.59, s16.7, CH31.52, s16.2, CH3
20 1.35, s26.2, CH37.91, d (1.8)126.5, CH7.90, d (1.8)126.6, CH
Table 2. Inhibitory effects of 16 on six kinds of pathogenic bacteria.
Table 2. Inhibitory effects of 16 on six kinds of pathogenic bacteria.
MIC (µg/mL)
CompoundsS. AureusB. SubtilisV. HarveyiV. ParahaemolyticusE. coliP. Aeruginosa
13232>64326416
2>64816161616
3>64>64>64>64>64>64
4>64>64>64>64>64>64
5>6432>64643232
6>64>64>64>64>64>64
Penicillin a<0.5<0.5<0.5<0.5<0.5<0.5
a Positive control.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, X.; Zhang, J.; Yang, J.; Li, B.; Li, T.; Ouyang, H.; Lin, W.; Hu, H.; Yan, X.; He, S. New Meroterpenes from South China Sea Soft Coral Litophyton brassicum. Mar. Drugs 2024, 22, 392. https://doi.org/10.3390/md22090392

AMA Style

Chen X, Zhang J, Yang J, Li B, Li T, Ouyang H, Lin W, Hu H, Yan X, He S. New Meroterpenes from South China Sea Soft Coral Litophyton brassicum. Marine Drugs. 2024; 22(9):392. https://doi.org/10.3390/md22090392

Chicago/Turabian Style

Chen, Xiaoyi, Jiahui Zhang, Jiayu Yang, Bing Li, Te Li, Han Ouyang, Wenhan Lin, Hongyu Hu, Xia Yan, and Shan He. 2024. "New Meroterpenes from South China Sea Soft Coral Litophyton brassicum" Marine Drugs 22, no. 9: 392. https://doi.org/10.3390/md22090392

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