Next Article in Journal
Cyclic [Cu-biRadical]2 Secondary Building Unit in 2p-3d and 2p-3d-4f Complexes: Crystal Structure and Magnetic Properties
Next Article in Special Issue
Yield and Composition of the Essential Oil of the Opopanax Genus in Turkey
Previous Article in Journal
Evaluation of the Chemical Composition of Selected Varieties of L. caerulea var. kamtschatica and L. caerulea var. emphyllocalyx
Previous Article in Special Issue
The Phenolic Compounds’ Role in Beer from Various Adjuncts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two Onnamide Analogs from the Marine Sponge Theonella conica: Evaluation of Geometric Effects in the Polyene Systems on Biological Activity

1
Department of Chemistry and Biochemistry, Graduate School of Advanced Science and Engineering, Waseda University, 3-4-1 Okubo, Shinjuku-ku, Tokyo 169-8555, Japan
2
Cell Biology Center, Institute of Innovative Research, Tokyo Institute of Technology, 4259 Nagatsuta, Midori-ku, Yokohama 226-8501, Japan
3
Research Institute for Science and Engineering, Waseda University, 3-4-1 Okubo, Shinjuku-ku, Tokyo 169-8555, Japan
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2524; https://doi.org/10.3390/molecules28062524
Submission received: 27 February 2023 / Revised: 6 March 2023 / Accepted: 7 March 2023 / Published: 9 March 2023

Abstract

:
Two previously unreported onnamide analogs, 2Z- and 6Z-onnamides A (1 and 2), were isolated from the marine sponge Theonella conica collected at Amami-Oshima Is., Kagoshima Prefecture, Japan. Structures of compounds 1 and 2 were elucidated by spectral analysis. Structure–activity relationships (SARs) for effects on histone modifications and cytotoxicity against HeLa and P388 cells were characterized. The geometry in the polyene systems of onnamides affected the histone modification levels and cytotoxicity.

1. Introduction

Marine invertebrates are a rich source of compounds with unique structures and biological activities. Marine invertebrates are composed of the phyla Porifera, Cnidaria, Mollusca, Echinodermata, Chordata and so forth. Marine sponges of the genus Theonella are well known for their rich secondary metabolites, including nonribosomal peptides, polyketides and terpenoids [1,2,3,4,5,6]. Onnamides or theopederins share a common core skeleton produced via polyketide and nonribosomal peptide biosynthetic pathways [7,8,9], and form a group with distinct structures and potent cytotoxicity against cancer cell lines. In this study, the isolation and structure elucidation of two unreported onnamide analogs, 2Z- and 6Z-onnamides A (1 and 2), are described, as well as known analogs 36.
Histone modifications play a crucial role in the epigenetic control of gene expression [10,11,12], and perturbations in this gene-switching system are related to chronic diseases such as cancer [13]. From this viewpoint, we have developed an in vitro cell-based assay system for evaluating the effects of compounds on multiple histone modifications in parallel [14]. Among 3750 extracts of marine organisms tested [14], the hydrophobic extract prepared from a marine sponge T. conica collected at Amami-Oshima Is., Kagoshima Prefecture, Japan, markedly enhanced the levels of trimethylated histone H3 lysine 27 (H3K27me3) and reduced the level of acetylated H4 lysine 5 (H4K5ac). Bioassay-guided isolation allowed us to identify onnamide analogs as active components in this marine sponge. Structure–activity relationships (SARs) for the activity of controlling histone modifications as well as for cytotoxicity were examined using six analogs (16). As the result, we observed that some of the histone modification changes occurred at the lower concentration than IC50 values for cytotoxicity. This finding imply different modes of action for these two biological activities that, however, were not fully supported because of the different sensitivity of HeLa cells to the compounds in the respective assays.
Therefore, our conclusion was that the activity in control of histone modification and cytotoxicity changes depending on the geometric isomerism in the side chains, but the different modes of action between these biological activities are not confirmed. Identification of the target gene expressions controlled by the histone modifications affected by onnamides should provide new insights into the underlying mechanisms of action for onnamides. A detailed investigation is currently underway.

2. Results and Discussion

A frozen specimen of T. conica (1020 g wet wt.) collected at Amami-Oshima Is., Kagoshima pref., Japan was extracted with CH3OH. The combined CH3OH extract was evaporated in vacuo and subjected to solvent partitioning [15], followed by ODS and SiO2 column chromatography. The obtained cytotoxic fractions were separated by repetitive reversed-phase HPLC on an ODS column to yield a mixture of 2Z- and 6Z-onnamides A (1, 2), onnamide A (3) [2], 4Z-onnamide A (4) [4], dihydroonnamide A (5) [3] and onnamide B (6) [3]. Final purification of the mixture of 1 and 2 was performed by recycling reversed-phase HPLC on a PHENYL-HEXYL column to afford 1 (3.5 mg) and 2 (1.5 mg) (3.4 × 10−3% and 1.5 × 10−3% yield based on wet weights). All structures of 16 were determined by 1H NMR and MS analyses (Table S1, Figures S1–S17).
The molecular formula of 2Z-onnamide A (1, C39H63N5O12) was determined to be the same as that of 3 by positive-mode high-resolution electrospray ionization mass spectrometry (HRESIMS) analysis (m/z 794.4536 [M + H]+, calcd for C39H64N5O12 794.4546, Δ − 1.3 ppm). The 1H NMR spectrum of 1 measured in CD3OD (Table 1) shared characteristic signals to that of 3, i.e., two O-methyls at CH3-30 and -32 (δH 3.23, 3.56), two doublet methyls of CH3-27 and -28 (δH 1.18, 0.97), two singlet methyls at CH3-33 and -34 (δH 0.86, 1.00), an exomethylene unit at CH2-29 (δH 4.80, 4.64), acetal methylene bridging C-16 and C-18 (O-CH2-O, δH 4.79, 5.21) and a hemiacetal methine proton (N-CH-O, δH 5.80). Additionally, oxymethine protons at 11, 13, 15, 16, 17, 21 and 26 (δH 3.88, 4.24, 3.99, 4.17, 3.66, 3.47 and 3.65, respectively) were observed, suggesting that the common tricyclic core structure (C-13 to C-34) in 3 was also preserved in 1. A spin system from H-2′ (δH 4.35) to H-5′ was suggested to be a part of arginine residue which was confirmed by the HMBC cross-peaks among H-2′/C-1 and H-5′/C-7′.
The distinct difference between 1 and 3 was observed in the side chain, as observed by the downfield-shifted signals for the polyene protons (H-2 to H-7). The overlapping signals of H-2 and H-3 were assigned based on the coupling constants of J2,3 (11.3 Hz), J4,5 (15.0 Hz) and J6,7 (15.0 Hz), which were obtained by processing data using a modified apodization function [16] (Figure 1). The geometry of the double bond between H-2 and H-3 in 1 was deduced as cis, confirming 2Z-onnamide A as 1.
The molecular formula of 6Z-onnamide A (2, C39H63N5O12) was also determined to be the same as that of 3 by the positive-mode HRESIMS (m/z 794.4530 [M + H]+, calcd for C39H64N5O12 794.4546, Δ − 2.0 ppm). The 1H NMR spectrum of 2 measured in CD3OD was superimposable with that of 3 (Table 2), except for the cis geometry in Δ6,7 (Figure 2), which indicated that 2 is 6Z-onnamide A.
The stereochemistry of 2Z- and 6Z-onnamides A (1 and 2) was assigned to the 11R, 13R, 15S, 16R, 17S, 18S, 21S, 22R, 25R, 26R and 2′S configuration, which is the same as onnamide A (3) and accounts for the biosynthetic pathway [8]. Analysis of the 1H NMR spectrum of 3 in CD3OD revealed that 3 undergoes photoisomerization by light irradiation to produce 1, 2 and 4Z-onnamideA (4) (Table S2). Calyculin and marinomycin analogs have been reported to undergo photoisomerization at their tetraene moieties [17,18]. Thus, we deduced that 1 and 2 are likewise artificially photoisomerized at the polyene moiety of 3. However, it remains unclear whether the geometric isomerism occurs in living T. conica by sunlight or in the laboratory under artificial light. Metabolomic analysis using freshly collected T. conica should provide an answer to this issue.
Compounds 16 showed significant cytotoxicity against HeLa cells with IC50 values of 38–540 nM (Table 3). SARs study using HeLa cells revealed that compound 5 with the reduced C21–C22 single bond is as potent as onnamide A (3), whereas 6 with a shorter side chain showed weaker activity (×1/8.2). The positions of the cis-trans isomerism in the side chains of onnamides also affects cytotoxicity: compound 4 was 1.7-fold more potent than 3, whereas 1 and 2 were 2.5-fold less cytotoxic than 3 (Figure 3).
Subsequently, the effects on histone modifications by onnamide A (3) were investigated using 16 monoclonal antibodies specific to each histone modification (Figure S18). The results revealed that 3 altered 13 histone modifications at concentrations of 70 and 140 nM, that are higher than the IC50 value (66 nM). Compound 3 enhanced the levels of trimethylated histone H3 lysine 4, 27 and 36 (H3K4me3, H3K27me3 and H3K36me3) and reduced the level of acetylated H4 lysine 5 (H4K5ac) at the lowest concentration (35 nM, Figure 4). All four histone modifications are related to cytotoxicity [19,20,21,22,23,24]. As H4K5ac is associated with DNA replication in the cell cycle, its decrease is consistent with the cell cycle arrest induced by 3 [14].
We also compared the effects of analogs 16 on histone modifications of H3K4me3, H3K27me3, H3K36me3 and H4K5ac at the same concentrations (35, 70 and 140 nM, Figure S19). Compound 6 administered at 70 nM (about 8 times less than the IC50 value of 540 nM) induced changes in the levels of the four histone modifications. However, enhanced cytotoxicity was also observed for 6 in this system (Figure S20), suggesting that the effects on histone modifications may be as the result of the cytotoxic effects by 6.
Onnamide A (3) and anisomycin [25,26] were reported to inhibit protein synthesis and to elicit ribotoxic stress response (RSR) [27]. RSR is induced in response to ribosomal impairment in the mitogen-activated protein kinase (MAPK)-mediated inflammatory signaling cascade. It includes activation of stress-activated protein kinases (SAPKs), such as p38 and c-Jun N-terminal kinase (JNK), and eventually causes cell death [28,29]. Moreover, activation of SAPKs by anisomycin was reported to be independent of protein synthesis inhibition [25], implying the possible explanation for the difference in cytotoxicity by 3 and anisomycin.
Anisomycin and onnamide A (3) bind to different sites on the ribosome (anisomycin binds to the A site of the ribosome, whereas 3 binds to the E site) [30,31]. The different signaling pathways caused by different binding sites likely explain the weaker cytotoxicity of anisomycin [27]. We had expected that the effects on histone modification could reflect the difference in cytotoxicity, but the similar effects of anisomycin on histone modifications to those of 3 were observed (Figure S21).
In this study, we could confirm that both cytotoxicity and control of histone modifications change depend on the geometric isomerism in the side chains. Evaluating the effects on the control of histone modifications can be effective way to distinguish modes of action by two different types of cytotoxic compounds, although in this case, the largely overlapping mechanisms by anisomycin and onnamides in cytotoxicity hampered us from clearly distinguishing them.

3. Materials and Methods

3.1. General Experimental Procedures

NMR spectra were recorded on an Avance (400 MHz) spectrometer (Bruker Corporation, Billerica, MA, USA). 1H and 13C NMR chemical shifts were referenced to the solvent peaks, δH 3.31 and δC 49.15 for CD3OD (FUJIFILM Wako Pure Chemical Corporation, Osaka, Japan). HRESI-MS spectra were measured on a Triple TOF 4600 (AB Sciex Pte. Ltd., Tokyo, Japan) in the positive mode. Optical rotation was determined on a DIP-1000 digital polarimeter (JASCO Corporation, Tokyo, Japan) in CH3OH. UV spectrum was recorded using a UV-1800 spectrophotometer (Shimadzu Corporation, Kyoto, Japan). IR spectrum was measured on a JIR-WINSPEC50 spectrometer(JEOL Ltd., Tokyo, Japan). Fluorescent images were obtained with an IX70 microscope equipped by DP72 (Olympus Corporation, Tokyo, Japan).

3.2. Biological Material

The same T. conica specimens as previous work [32,33] were used in this study. T. conica was collected by hand using SCUBA, Amami-Oshima Is., Kagoshima Prefecture, Japan (N 28° 06.82′, E 129° 21.09′) in June 2007. The sample was immediately frozen and kept at −25 °C until extraction.

3.3. Isolation

The frozen sponge specimen (1020 g wet wt.) was extracted with CH3OH (1 L × 5), and the combined extract was evaporated in vacuo. The concentrated extract was suspended in H2O and extracted with CHCl3, then n-C4H9OH. The CHCl3 and n- C4H9OH layers were combined and subjected to the Kupchan procedure [15] yielding n-hexane, CHCl3 and aqueous CH3OH layers [32,33]. CHCl3 layer was concentrated to dryness and then separated by ODS flash chromatography (CH3OH/H2O = 5:5, 7:3, CH3CN/H2O = 7:3, 85:15, CH3OH, CHCl3/CH3OH/H2O = 6:4:1) to yield 6 fractions. The fraction eluting with CH3OH/H2O (7:3), which affected the histone modifications in HeLa cells, was separated by silica gel column chromatography (CHCl3, CH3Cl3/CH3OH = 19:1, 9:1, CHCl3/CH3OH/H2O = 8:2:0.1, 7:3:0.5, 6:4:1, 5:5:2). The fr.8 was further separated by reversed-phase HPLC (COSMOSIL 5C18-AR-II) using with CH3OH/H2O (6:4) to yield 12 fractions. Of these, 8.5 mg of onnamide B (6, tR: 24.6 min), 44.2 mg of onnamide A (3, tR: 37.0 min) and 4.3 mg of dihydroonnamide A (5, tR: 51.6 min) were isolated. The crude fr. 8 was separated by 2-step reversed-phase HPLC (COSMOSIL 5C18-AR-II, CH3OH/H2O = 32.5:67.5), followed by final purification with recycle system (phenomenex PHENYL-HEXYL, CH3CN/H2O = 4:6), to yield 3.5 mg of 2Z-onnamide (1, 11 cycles, tR: 155 min, 3.4 × 10−3% yield based on wet weights) and 1.5 mg of 6Z-onnamide A (2, 11 cycles, tR: 150 min, 1.5×10−3% yield based on wet weights). The crude Fr. 10 was purified by recycling reversed-phase HPLC (phenomenex PHENYL-HEXYL, CH3CN/H2O = 4:6) to afford 2.5 mg of 4Z-onnamides A (4, 6 cycles, tR: 87.0 min).
2Z-onnamides A (1): yellow amorphous solid; [α]D22.8 + 48.5° (c 0.2, CH3OH); UV (CH3OH) λmax (logε) 299.4 (4.58) nm; IR (KBr film) νmax 3363, 2936, 1652, 1635, 1576, 1558, 1531, 1397, 1093, 1032 cm−1; ESIMS m/z 794.4539 [M + H]+ (calcd for C39H64N5O12 794.4546, Δ − 1.3 ppm); 1H and 13C NMR data, see Table 1.
6Z-onnamides A (2): yellow amorphous solid; [α]D23.0 +46.0° (c 0.2, CH3OH); UV (CH3OH) λmax (logε) 300.4 (4.73) nm; IR (KBr film) νmax 3417, 2925, 1644, 1583, 1537, 1403, 1316, 1092, 1032 cm−1; ESIMS m/z 794.4530 [M + H]+ (calcd for C39H64N5O12 794.4546, Δ − 2.0 ppm); 1H and 13C NMR data, see Table 2.
Onnamide A (3): yellow amorphous solid; HRESIMS m/z 794.4542 [M + H]+ (calcd for C39H64N5O12 794.4546, Δ − 0.5 ppm); 1H NMR data, see Table S1.
4Z-onnamides A (4): yellow amorphous solid; HRESIMS m/z 794.4552 [M + H]+ (calcd for C39H64N5O12 794.4546, Δ + 0.8 ppm); 1H NMR data, see Table S1.
Dihydroonnamide A (5): yellow amorphous solid; HRESIMS m/z 796.4692 [M + H]+ (calcd for C39H66N5O12 796.4702, Δ − 1.3 ppm); 1HNMR data, see Table S1.
Onnamide B (6): yellow amorphous solid; HRESIMS m/z 768.4391 [M + H]+ (calcd for C37H62N5O12 768.4389, Δ + 0.2 ppm); 1H NMR data, see Table S1.

3.4. Cell Culture

HeLa human cervical cancer cells were cultured at 37 °C under an atmosphere of 5% CO2 in Dulbecco’s modified Eagle’s medium (DMEM, Low Glucose, FUJIFILM Wako Pure Chemical Corporation), containing 10% of fetal bovine serum (FBS, Biowest, Nuaillé, France), 2 µg/mL of gentamicin reagent solution AND 10 µg/mL of antibiotic-antimycotic. P388 murine leukemia cells were propagated and maintained at 37 °C under an atmosphere of 5% CO2 in Roswell Park Memorial Institute medium (RPMI, FUJIFILM Wako Pure Chemical Corporation), containing HRDS solution (2, 2′-dithiobisethanol), and kanamycin sulfate. The J1 mouse embryonic stem cells (ESCs) were obtained from the American Type Culture Collection (ATCC, Manassas, VA, USA) and maintained on 0.1% gelatin-coated dishes with mitomycin C-treated mouse embryonic fibroblasts (MEFs, Kitayama Labes, Nagano, Japan) in the medium of DMEM supplemented with 15% FBS, 1% L-glutamine (Gibco, Thermo Fisher Scientific, Inc., Waltham, MA, USA), 1% non-essential amino acids (Gibco), 1% penicillin-streptomycin (P/S, Gibco), 0.18% 2-mercaptoethanol (Gibco) and 1000 U/mL leukemia inhibitory factor (LIF, Merck Chemicals GmbH, Darmstadt, Germany).

3.5. Cytotoxic Test

Cytotoxicity test was conducted as previously reported [34]. Briefly, HeLa cells in DMEM or P388 cells in RPMI (cell concentration, 10,000 cells/mL, 200 µL) were added to each well of 96-well microplates and kept in the incubator at 37 °C under an atmosphere of 5% CO2. After 24 h, samples in DMSO with various concentrations of onnamides (16) were added to each well. After 72 h cultivation, to each well was added 50 µL of 3-(4,5-dimethyl-2-thiazoyl)-2,5-diphenyl-2H tetrazolium bromide (MTT) saline solution (1 mg/mL, FUJIFILM Wako Pure Chemical Corporation) and they were then kept in the incubator at 37 °C under an atmosphere of 5% CO2. After 4 h, medium was removed by aspiration and 150 µL of DMSO was added to each well to lyse cells. Concentration of the reduced MTT was quantified, measuring the absorbance at 650 nm to estimate IC50 values.

3.6. Histone Modification Assay

Assay of histone modification levels was performed by an immunofluorescence using a previously reported method [14], with some modifications. Briefly, HeLa cells were incubated under the medium containing the sample for 20 h and then immunostained. Cells were fixed with 4% paraformaldehyde in PBS for 10 min, 1% Triton X-100 in PBS for 20 min, and blocked in Blocking One-P (Nacalai Tesque Inc., Kyoto, Japan) for 20 min and then incubated in Alexa Fluor 488 (Thermo Fisher Scientific, Inc., Waltham, MA, USA) or Cy3 (Thermo Fisher Scientific, Inc.)-labeled antibodies against each histone modification (1:1000, Monoclonal Antibody Institute, Nagano, Japan) for 2 h with Hoechst 33342 (1:2000, Dojindo Laboratories Co., Ltd., Kumamoto, Japan). Fluorescent images were obtained with a microscope. The relative fluorescence intensity was digitalized using CellProfilerTM software 3.0.0 [35] against those of the control wells. Anisomycin was purchased from FUJIFILM Wako Pure Chemical Corporation.

4. Conclusions

In conclusion, new onnamide analogs 2Z- and 6Z-onnamides A (1 and 2) and four known onnamides (36) were isolated from the marine sponge T. conica, and their structures were elucidated by MS and NMR spectral analyses. The combined effects on histone modifications and cytotoxicities by these compounds revealed that their modes of action follow those of anisomycin. Identifying modulations in the expression patterns of target genes caused by onnamides via changes in histone modifications should provide new insights into the underlying mechanisms of onnamide activity. A detailed investigation of these mechanisms is currently underway.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28062524/s1, Table S1: 1H NMR spectral data [δH mult. (J in Hz)] for onnamides (36) in CD3OD (400 MHz); Table S2: The integrated value of 1H NMR signal derived from compounds 1 to 4 when 3 was placed in an NMR tube and each time passed; Figures S1–S17: MS and NMR spectra of compounds 16; Figures S18–S21: the results of histone modification assay.

Author Contributions

Conceptualization: F.N. and Y.N.; investigation: F.N.; writing—original draft: F.N. and Y.N.; writing—review and editing: H.K. and N.F.; supervision: Y.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by research performed under a Waseda University Grant for Special Research Projects and the Japan Society for the Promotion of Science (JSPS) to 21H02073 (Y.N.), 25560408 (Y.N. and H.K.) and 21K20577 (F.N).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data from the present study are available in the article and Supplementary Materials.

Acknowledgments

We gratefully thank K. Hori, M. Hirayama and the crews of R/V Toyoshiomaru of Hiroshima University for supporting the collection of marine sponges. We thank Edanz “https://jp.edanz.com/ac (accessed on 10 February 2023)” for editing a draft of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Bewley, C.A.; Faulkner, D.J. Lithistid Sponges: Star Performers or Hosts to the Stars. Angew. Chem. Int. Ed. 1998, 37, 2162–2178. [Google Scholar] [CrossRef]
  2. Sakemi, S.; Ichiba, T.; Kohmoto, S.; Saucy, G.; Higa, T. Isolation and Structure Elucidation of Onnamide A, a New Bioactive Metabolite of a Marine Sponge, Theonella sp. J. Am. Chem. Soc. 1988, 110, 4851–4853. [Google Scholar] [CrossRef]
  3. Matsunaga, S.; Fusetani, N.; Nakao, Y. Eight New Cytotoxic Metabolites Closely Related to Onnamide A from Two Marine Sponges of the Genus Theonella . Tetrahedron 1992, 48, 8369–8376. [Google Scholar] [CrossRef]
  4. Kobayashi, J.; Itagaki, F.; Shigemori, H.; Sasaki, T. Three New Onnamide Congeners from the Okinawan Marine Sponge Theonella sp. J. Nat. Prod. 1993, 56, 976–981. [Google Scholar] [CrossRef]
  5. Fusetani, N.; Sugawara, T.; Matsunaga, S. Bioactive Marine Metabolites. 41. Theopederins A-E, Potent Antitumor Metabolites from a Marine Sponge, Theonella sp. J. Org. Chem. 1992, 57, 3828–3832. [Google Scholar] [CrossRef]
  6. Tsukamoto, S.; Matsunaga, S.; Fusetani, N.; Toh-e, A. Theopederins F-J: Five New Antifungal and Cytotoxic Metabolites from the Marine Sponge, Theonella Swinhoei. Tetrahedron 1999, 55, 13697–13702. [Google Scholar] [CrossRef]
  7. Cardani, C.; Fuganti, C.; Ghiringhelli, D.; Grasselli, P.; Pavan, M.; Valcurone, M.D. The Biosynthesis of Pederin. Tetrahedron Lett. 1973, 14, 2815–2818. [Google Scholar] [CrossRef]
  8. Piel, J.; Hui, D.; Wen, G.; Butzke, D.; Platzer, M.; Fusetani, N.; Matsunaga, S. Antitumor Polyketide Biosynthesis by an Uncultivated Bacterial Symbiont of the Marine Sponge Theonella Swinhoei. Proc. Natl. Acad. Sci. USA 2004, 101, 16222–16227. [Google Scholar] [CrossRef] [Green Version]
  9. Mosey, R.A.; Floreancig, P.E. Isolation, Biological Activity, Synthesis, and Medicinal Chemistry of the Pederin/Mycalamide Family of Natural Products. Nat. Prod. Rep. 2012, 29, 980. [Google Scholar] [CrossRef]
  10. Kouzarides, T. Chromatin Modifications and Their Function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [Green Version]
  11. Bannister, A.J.; Kouzarides, T. Regulation of Chromatin by Histone Modifications. Cell Res. 2011, 21, 381–395. [Google Scholar] [CrossRef] [PubMed]
  12. Pokholok, D.K.; Harbison, C.T.; Levine, S.; Cole, M.; Hannett, N.M.; Lee, T.I.; Bell, G.W.; Walker, K.; Rolfe, P.A.; Herbolsheimer, E.; et al. Genome-Wide Map of Nucleosome Acetylation and Methylation in Yeast. Cell 2005, 122, 517–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Portela, A.; Esteller, M. Epigenetic Modifications and Human Disease. Nat. Biotechnol. 2010, 28, 1057–1068. [Google Scholar] [CrossRef] [PubMed]
  14. Hayashi-Takanaka, Y.; Kina, Y.; Nakamura, F.; Becking, L.E.; Nakao, Y.; Nagase, T.; Nozaki, N.; Kimura, H. Histone Modification Dynamics as Revealed by a Multicolor Immunofluorescence-Based Single-Cell Analysis. J. Cell. Sci. 2020, 133, jcs.243444. [Google Scholar] [CrossRef] [PubMed]
  15. Kupchan, S.M.; Britton, R.W.; Ziegler, M.F.; Sigel, C.W. Bruceantin, a New Potent Antileukemic Simaroubolide from Brucea Antidysenterica. J. Org. Chem. 1973, 38, 178–179. [Google Scholar] [CrossRef]
  16. Lindon, J.C.; Ferrige, A.G. Digitisation and Data Processing in Fourier Transform NMR. Prog. Nucl. Magn. Reson. Spectrosc. 1980, 14, 27–66. [Google Scholar] [CrossRef]
  17. Matsunaga, S.; Fujiki, H.; Sakata, D.; Fusetani, N. Calyculins E, F, G, and H, Additional Inhibitors of Protein Phosphatases 1 and 2a, from the Marine Sponge Discodermia Calyx. Tetrahedron 1991, 47, 2999–3006. [Google Scholar] [CrossRef]
  18. Nicolaou, K.C.; Nold, A.L.; Milburn, R.R.; Schindler, C.S.; Cole, K.P.; Yamaguchi, J. Total Synthesis of Marinomycins A–C and of Their Monomeric Counterparts Monomarinomycin A and Iso-Monomarinomycin A. J. Am. Chem. Soc. 2007, 129, 1760–1768. [Google Scholar] [CrossRef]
  19. Cao, R.; Wang, L.; Wang, H.; Xia, L.; Erdjument-Bromage, H.; Tempst, P.; Jones, R.S.; Zhang, Y. Role of Histone H3 Lysine 27 Methylation in Polycomb-Group Silencing. Science 2002, 298, 1039–1043. [Google Scholar] [CrossRef] [Green Version]
  20. Kim, K.H.; Roberts, C.W.M. Targeting EZH2 in Cancer. Nat. Med. 2016, 22, 128–134. [Google Scholar] [CrossRef] [Green Version]
  21. Sobel, R.E.; Cook, R.G.; Perry, C.A.; Annunziato, A.T.; Allis, C.D. Conservation of Deposition-Related Acetylation Sites in Newly Synthesized Histones H3 and H4. Proc. Natl. Acad. Sci. USA 1995, 92, 1237–1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Kaneko, S.; Li, G.; Son, J.; Xu, C.-F.; Margueron, R.; Neubert, T.A.; Reinberg, D. Phosphorylation of the PRC2 Component Ezh2 Is Cell Cycle-Regulated and up-Regulates Its Binding to NcRNA. Genes Dev. 2010, 24, 2615–2620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Xie, P.; Tian, C.; An, L.; Nie, J.; Lu, K.; Xing, G.; Zhang, L.; He, F. Histone Methyltransferase Protein SETD2 Interacts with P53 and Selectively Regulates Its Downstream Genes. Cell Signal. 2008, 20, 1671–1678. [Google Scholar] [CrossRef] [PubMed]
  24. Guerra-Calderas, L.; González-Barrios, R.; Herrera, L.A.; Cantú de León, D.; Soto-Reyes, E. The Role of the Histone Demethylase KDM4A in Cancer. Cancer Genet. 2015, 208, 215–224. [Google Scholar] [CrossRef]
  25. Cano, E.; Hazzalin, C.A.; Mahadevan, L.C. Anisomycin-Activated Protein Kinases P45 and P55 but Not Mitogen-Activated Protein Kinases ERK-1 and -2 Are Implicated in the Induction of c-Fos and c-Jun. Mol. Cell. Biol. 1994, 14, 7352–7362. [Google Scholar]
  26. Zinck, R.; Cahill, M.A.; Kracht, M.; Sachsenmaier, C.; Hipskind, R.A.; Nordheim, A. Protein Synthesis Inhibitors Reveal Differential Regulation of Mitogen-Activated Protein Kinase and Stress-Activated Protein Kinase Pathways That Converge on Elk-1. Mol. Cell. Biol. 1995, 15, 4930–4938. [Google Scholar] [CrossRef] [Green Version]
  27. Lee, K.-H.; Nishimura, S.; Matsunaga, S.; Fusetani, N.; Horinouchi, S.; Yoshida, M. Inhibition of Protein Synthesis and Activation of Stress-Activated Protein Kinases by Onnamide A and Theopederin B, Antitumor Marine Natural Products. Cancer Sci. 2005, 96, 357–364. [Google Scholar] [CrossRef]
  28. Iordanov, M.S.; Pribnow, D.; Magun, J.L.; Dinh, T.H.; Pearson, J.A.; Chen, S.L.; Magun, B.E. Ribotoxic Stress Response: Activation of the Stress-Activated Protein Kinase JNK1 by Inhibitors of the Peptidyl Transferase Reaction and by Sequence-Specific RNA Damage to the Alpha-Sarcin/Ricin Loop in the 28S RRNA. Mol. Cell. Biol. 1997, 17, 3373–3381. [Google Scholar] [CrossRef] [Green Version]
  29. Vind, A.C.; Genzor, A.V.; Bekker-Jensen, S. Ribosomal Stress-Surveillance: Three Pathways Is a Magic Number. Nucleic Acids Res. 2020, 48, 10648–10661. [Google Scholar] [CrossRef]
  30. Garreau de Loubresse, N.; Prokhorova, I.; Holtkamp, W.; Rodnina, M.V.; Yusupova, G.; Yusupov, M. Structural Basis for the Inhibition of the Eukaryotic Ribosome. Nature 2014, 513, 517–522. [Google Scholar] [CrossRef] [Green Version]
  31. Dang, Y.; Schneider-Poetsch, T.; Eyler, D.E.; Jewett, J.C.; Bhat, S.; Rawal, V.H.; Green, R.; Liu, J.O. Inhibition of Eukaryotic Translation Elongation by the Antitumor Natural Product Mycalamide B. RNA 2011, 17, 1578–1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Sepe, V.; Ummarino, R.; D’Auria, M.V.; Taglialatela-Scafati, O.; Marino, S.D.; D’Amore, C.; Renga, B.; Chini, M.G.; Bifulco, G.; Nakao, Y.; et al. Preliminary Structure-Activity Relationship on Theonellasterol, a New Chemotype of FXR Antagonist, from the Marine Sponge Theonella Swinhoei. Mar. Drugs 2012, 10, 2448–2466. [Google Scholar] [CrossRef] [PubMed]
  33. Sepe, V.; D’Amore, C.; Ummarino, R.; Renga, B.; D’Auria, M.V.; Novellino, E.; Sinisi, A.; Taglialatela-Scafati, O.; Nakao, Y.; Limongelli, V.; et al. Insights on Pregnane-X-Receptor Modulation. Natural and Semisynthetic Steroids from Theonella Marine Sponges. Eur. J. Med. Chem. 2014, 73, 126–134. [Google Scholar] [CrossRef] [PubMed]
  34. Nakamura, F.; Maejima, H.; Kawamura, M.; Arai, D.; Okino, T.; Zhao, M.; Ye, T.; Lee, J.; Chang, Y.-T.; Fusetani, N.; et al. Kakeromamide A, a New Cyclic Pentapeptide Inducing Astrocyte Differentiation Isolated from the Marine Cyanobacterium Moorea Bouillonii. Bioorg. Med. Chem. Lett. 2018, 28, 2206–2209. [Google Scholar] [CrossRef]
  35. Carpenter, A.E.; Jones, T.R.; Lamprecht, M.R.; Clarke, C.; Kang, I.H.; Friman, O.; Guertin, D.A.; Chang, J.H.; Lindquist, R.A.; Moffat, J.; et al. CellProfiler: Image Analysis Software for Identifying and Quantifying Cell Phenotypes. Genome Biol. 2006, 7, R100. [Google Scholar] [CrossRef] [Green Version]
Figure 1. COSY, key HMBC cross-peaks and coupling constants of 2Z-onnamide A (1).
Figure 1. COSY, key HMBC cross-peaks and coupling constants of 2Z-onnamide A (1).
Molecules 28 02524 g001
Figure 2. COSY, key HMBC cross-peaks and coupling constants of 6Z-onnamide A (2).
Figure 2. COSY, key HMBC cross-peaks and coupling constants of 6Z-onnamide A (2).
Molecules 28 02524 g002
Figure 3. SARs of compounds 16 on the cytotoxicity against HeLa cells.
Figure 3. SARs of compounds 16 on the cytotoxicity against HeLa cells.
Molecules 28 02524 g003
Figure 4. Effects of onnamide A (3) on histone modifications. Quantification of the histone modification levels of H3K4me3, H3K27me3, H3K36me3 and H4K5ac after cultivation in medium containing samples for 20 h (samples: 35, 70 and 140 nM of 1, Ct.: DMSO, n = 3, mean ± S.D. ***: p < 0.001, **: p < 0.01, Dunnett test).
Figure 4. Effects of onnamide A (3) on histone modifications. Quantification of the histone modification levels of H3K4me3, H3K27me3, H3K36me3 and H4K5ac after cultivation in medium containing samples for 20 h (samples: 35, 70 and 140 nM of 1, Ct.: DMSO, n = 3, mean ± S.D. ***: p < 0.001, **: p < 0.01, Dunnett test).
Molecules 28 02524 g004
Table 1. NMR spectral data for 2Z-onnamide (1) in CD3OD (400/150 MHz).
Table 1. NMR spectral data for 2Z-onnamide (1) in CD3OD (400/150 MHz).
PositionδCδH Mult. (J in Hz)COSYHMBC
1168.5
2120.85.74 d (11.6)H-3C-1, C-4
3141.76.46 t (11.6)H-2, H-4C-1, C-5
4128.47.45 dd (15.0, 11.6)H-3, H-5
5142.46.43 dd (15.0, 10.7)H-4, H-6
6131.96.24 dd (15.0, 10.7)H-5, H-7
7140.55.95 dt (15.0, 6.9)H-6, H-8C-5
834.12.13 m, 2.23 mH-7, H-9
931.61.30 m, 1.47 mH-8, H-10
1037.01.29 m,1.43 mH-11, H-9
1171.23.65 mH-10, H-12
1237.51.54 mH-11, H-13
1378.83.47 dd (8.9, 3.4)H-12
1442.5
1580.73.66 d (9.9)H-16C-32
1675.84.17 dd (9.9, 6.6)H-15, H-17C-15, C-17, C-18, C-31
1771.03.99 dd (9.4, 6.6)H-16
1875.05.80 d (9.4)H-17C-20, C-31
20174.6
2174.14.24 s C-20, C-23, C-24
22101.5
2334.92.41 d (14.3), 2.32 d (14.3) C-22, C-24, C-25, C-29
24148.4
2543.22.20 mH-26, H-28C-24
2671.03.88 qd (6.5, 2.5)H-25, H-27C-25
2718.31.18 d (6.5)H-26, H-28C-25,26
2812.50.97 d (7.0)H-25, H-27C-24, C-25, C-26
29110.24.80 brs, 4.64 brs C-23, 25
3048.83.23 s C-22
3187.85.21d (6.8), 4.79 d (6.8) C-18
3262.13.56 s C-15
3314.30.86 s C-13, C-14, C-15, C-34
3423.71.00 s C-13, C-14, C-15, C-33
1′178.6
2′55.24.35 dd (7.3, 5.2)H-3′C-1, C-1′, C-3′
3′31.61.89 m, 1.74 mH-2′, H-4′C-1′
4′26.31.65 mH-3′, H-5′C-3′, C-5′
5′42.33.19 m, 3.26 mH-4′C-7′
7′158.7
Table 2. NMR spectral data for 6Z-onnamide (2) in CD3OD (400/150 MHz).
Table 2. NMR spectral data for 6Z-onnamide (2) in CD3OD (400/150 MHz).
PositionδCδH Mult. (J in Hz)COSYHMBC
1168.5
2120.86.11 d (15.0)H-3C-1, C-4
3141.77.25 dd (15.0, 11.2)H-2, H-4C-5
4128.46.35 dd (14.7, 11.2)H-3, H-5
5142.46.94 dd (14.7, 11.5)H-4, H-6
6131.96.13 t (11.5)H-5, H-7
7140.55.67 dt (11.5, 7.7)H-6, H-8
834.12.30 mH-7, H-9
931.61.50 m, 1.58 mH-8, H-10
1037.01.32 mH-11, H-9
1171.23.65 mH-10, H-12
1237.51.54 mH-11, H-13C-17
1378.83.49 mH-12
1442.5
1580.73.66 d (9.8)H-16C-32
1675.84.17 dd (9.8, 6.6)H-15, H-17C-15, C-17, C-18, C-31
1771.03.98 dd (9.1, 6.6)H-16
1875.05.83 d (9.1)H-17C-20
20174.6
2174.14.24 s C-20, 23, 24
22101.5
2334.92.41 d (14.3), 2.32 d (14.3) C-22, 24, 25, 29
24148.4
2543.22.20 mH-26, H-28
2671.03.87 qd (6.5, 2.6)H-25, H-27
2718.31.17 d (6.5)H-26C-25,26
2812.50.97 d (7.1)H-25C-24, 25, 26
29110.24.79 s, 4.64 brs C-23, 25
3048.83.24 s C-22
3187.85.23d (6.9), 4.80 d (6.9) C-18
3262.13.56 s C-15
3314.30.86 s C-13, 14, 15, 34
3423.71.00 s C-13, 14, 15, 33
1′178.6
2′55.24.38 dd (7.3, 5.2)H-3′C-1′
3′31.61.89 m, 1.75 mH-2′, H-4′
4′26.31.64 mH-3′, H-5′C-2′, 3′
5′42.33.18 m, 3.22 mH-4′C-7′
7′158.7
Table 3. Cytotoxicity against HeLa and P388 cells by compounds 16.
Table 3. Cytotoxicity against HeLa and P388 cells by compounds 16.
Cell LineIC50 (µM)
123456
HeLa0.170.150.0660.0380.0570.54
P3881.84.80.620.310.575.2
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

Nakamura, F.; Kimura, H.; Fusetani, N.; Nakao, Y. Two Onnamide Analogs from the Marine Sponge Theonella conica: Evaluation of Geometric Effects in the Polyene Systems on Biological Activity. Molecules 2023, 28, 2524. https://doi.org/10.3390/molecules28062524

AMA Style

Nakamura F, Kimura H, Fusetani N, Nakao Y. Two Onnamide Analogs from the Marine Sponge Theonella conica: Evaluation of Geometric Effects in the Polyene Systems on Biological Activity. Molecules. 2023; 28(6):2524. https://doi.org/10.3390/molecules28062524

Chicago/Turabian Style

Nakamura, Fumiaki, Hiroshi Kimura, Nobuhiro Fusetani, and Yoichi Nakao. 2023. "Two Onnamide Analogs from the Marine Sponge Theonella conica: Evaluation of Geometric Effects in the Polyene Systems on Biological Activity" Molecules 28, no. 6: 2524. https://doi.org/10.3390/molecules28062524

Article Metrics

Back to TopTop