Next Article in Journal
Anti-Inflammatory Cembrane-Type Diterpenoids and Prostaglandins from Soft Coral Lobophytum sarcophytoides
Previous Article in Journal
Characterization of an Alkaline GH49 Dextranase from Marine Bacterium Arthrobacter oxydans KQ11 and Its Application in the Preparation of Isomalto-Oligosaccharide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Diterpenes with a Fused 6-5-6-6 Ring System Isolated from the Marine Sponge-Derived Fungus Trichoderma harzianum

Department of Medicinal Molecular Chemistry, Osaka University of Pharmaceutical Sciences, 4-20-1, Nasahara, Takatsuki, Osaka 569-1094, Japan
*
Author to whom correspondence should be addressed.
Mar. Drugs 2019, 17(8), 480; https://doi.org/10.3390/md17080480
Submission received: 5 July 2019 / Revised: 7 August 2019 / Accepted: 15 August 2019 / Published: 19 August 2019

Abstract

:
New diterpenes, namely, trichodermanins F–H, with a fused 6-5-6-6 ring system were isolated from the fungus Trichoderma harzianum OUPS-111D-4 separated from a marine sponge Halichondria okadai. These chemical structures were elucidated by 1D and 2D NMR as well as high-resolution fast atom bombardment mass spectrometry (HRFABMS) spectral analyses. We established their absolute stereostructures by application of the modified Mosher’s method or circular dichroism (CD) spectroscopy. In addition, their cytotoxicities were assessed using several cancer cell lines, with 1 and 2 exhibiting modest activities.

1. Introduction

A number of secondary metabolites derived from marine microorganisms have diverse structures and exhibit unexpected biological activities [1,2,3,4]. This motivated us to target marine microorganisms as a source of seed compounds for antitumor chemotherapy agents, and we have reported many cytotoxic metabolites to date [5,6,7,8]. As part of this ongoing study, we examined the metabolites of the fungus Trichoderma harzianum separated from a piece of the marine sponge Halichondria okadai. We have already reported the isolation, structural determination, and cytotoxicity of the decalin derivatives tandyukisins A–F derived from this fungus [9,10,11] as well as those of trichodermanins C, D, and E (1), classified as diterpenes with a rare fused 6-5-6-6 ring system [12]. In the continuing search for cytotoxic metabolites from this fungal strain, we isolated three new compounds, namely, trichodermanins F–H (24). Except for our study, there are only two other reports of natural products consisting of this ring system, i.e., trichodermanin A [13] and wickerols A and B [14]. We have previously reported the first determination of the absolute configurations of trichodermanins C and D by application of the modified Mosher’s method [15]. In the present study, we established the absolute configuration of trichodermanin E (1), which was confirmed by circular dichroism (CD) spectroscopy. In addition, we also examined the absolute stereostructures of the newly isolated trichodermanins F–H (24). We herein describe the evaluation of the cytotoxic activities of 14, although they exhibited zero or weak activities.

2. Results and Discussion

T. harzianum, a fungus from H. okadai, was cultured at 27 °C for six weeks in a medium (70 L) containing 1% glucose, 1% malt extract, and 0.05% peptone in artificial seawater adjusted to pH 7.5. After incubation, the EtOAc extract of the culture filtrate was purified via cytotoxic assay-directed fractionation using a stepwise combination of silica gel column and Sephadex LH-20 chromatography, followed by reverse-phase HPLC, affording trichodermanins E (1) (0.93 mg, 0.011%), F (2) (3.7 mg, 0.043%), G (3) (4.6 mg, 0.053%), and H (4) (1.3 mg, 0.015%)
The molecular formula of trichodermanin E (1) was established as C20H34O3 based on the molecular ion m/z 345.2397 [M + Na]+ in high-resolution fast atom bombardment mass spectrometry (HRFABMS) (Figure S7). Absorption in the IR spectrum at 3383 cm−1 indicated the presence of hydroxy groups. The relative stereostructure of 1 has previously been reported using distortionless enhancement by polarization transfer (DEPT), heteronuclear multiple quantum coherence spectroscopy (HMQC), and NOESY experiments with the 1H and 13C NMR spectra of 1 (Figure 1, and Figures S1–S6, Table 1, and Table S1) [12]. However, the assignment of its absolute configuration was not performed. In the present study, the absolute stereostructure of 1 was elucidated by acylation with p-bromobenzoyl chloride to apply the dibenzoate rule [16]. In the CD spectrum of the dibenzoate 1a, the negative Cotton effect (Δε258–15) demonstrated that the screw sense of two benzoate chromophores at C-1 and C-2 was counterclockwise. Therefore, the absolute configurations for C-1 and C-2 were revealed as R and R, respectively (Figure 2) [16].
Trichodermanin F (2) was assigned the formula C20H34O2 by HRFABMS m/z 289.2534 [M − Na]+ (calcd for C20H34O3Na: 289.2532), which contained one less oxygen atom than that of 1 (Figure S14). Consideration of the 1H and 13C NMR spectra of 2 (Table 1 and Table S2, and Figures S8, S9 and S12), including 2D NMR spectra, suggested that compound 2 had one secondary methyl (C-17); four tertiary methyls (C-16, C-18, C-19, and C-20); six sp3-hybridized methylenes (C-1, C-7, C-9, C-10, C-13, and C-14); five sp3 methines (C-2, C-3, C-6, C-11, and C-12), including one oxygen-bearing sp3 methine (C-2); and four quaternary sp3 carbons (C-4, C-5, C-8, and C-15), including one oxygen-bearing quaternary sp3 carbon (C-15). 1H–1H correlation spectroscopy (COSY) revealed three partial structures (Figure 3 and Figure S10). In the HMBC spectrum (Figure 3 and Figure S13), the correlation from 15-methyl to C-11, C14, and C-15; from 8-methyl to C-7, C-8, C-9, and C-12; from germinal dimethyl (H-18 and H-19) to C-4, C-5, and C-6; from H-12 to C-3, C-4, and C-5; and from H-13 to C-12 showed three cyclohexane rings and a cyclopentane comprising a fused 6-5-6-6 ring system to give the planar structure of 2, as shown in Figure 3.
For the stereochemistry of 2, the relative configuration and conformation were investigated by NOESY experiments (Table S2 and Figure 4 and Figure S11). In the NOESY analysis of 2, correlations from H-2α to H-19; from H-3 to H-11, H-14β, and H-20; and from H-1β to H-20 indicated that cyclohexane ring A existed in the boat conformation with 3-CH3 in the α-orientation. In cyclohexane ring B, the above correlations and the NOESY cross peaks (H-7α/H-12 and H-18) demonstrated that ring B existed in the chair conformation and that 5-CH3 (C-18), H-7α, and H-12 were oriented in coaxial arrangements. The NOESY correlations between H-3, H-11, and H-14β indicated that cyclohexane ring C also existed in the chair conformation. In addition, the correlations (H-18/H-13α, H-11/H-20, H-14α/H-16, and H-12/H-16) revealed that H-11, H-13α, H-14β, and 15-CH3 (C-16) were oriented in coaxial arrangements and that the ring juncture between cyclohexane rings B and C was trans (Figure 4 and Figure S11, and Table S2).
A difference in the structural features of 2 and 1 was the absence of a secondary hydroxyl group at C-1. Therefore, we applied the modified Mosher’s method [15] to determine their absolute stereostructures. The 1H chemical shift differences between the (S)- and (R)-MTPA esters 2a and 2b showed positive on the right side and negative on the left side of the MTPA plane; therefore, we revealed an S configuration at C-2 (Figure 5 and Table S5).
Trichodermanin G (3) was assigned the formula C20H34O3 by HRFABMS m/z 345.2395 [M + Na]+ (calcd for C20H34O3Na: 345.2396), which was the same as that of 1 (Figure S21). The NMR spectral features (Table 1 and Table S3, and Figures S15, S16 and S19) resembled those of 1, except for the proton signals for H-2 (δH 1.45, m and δH 2.66, ddd), H-3 (δH 2.10, m), and H-10 (δH 4.38, ddd) and the carbon signals of C-1 (δC 72.6), C-2 (δC 41.8), C-3 (δC 26.7), C-9 (δC 52.2), C-10 (δC 72.9), and C-11 (δC 55.0) in 3. The 1H–1H COSY correlations (H-1/H-6, H-1/H-2, and H-2/H-3), and HMBC correlation from H-17 to C-2 indicated that the hydroxy methine at C-2 in 1 were replaced by a methylene in 3 (Table 1 and Table S3, and Figures S17 and S20). In addition, the 1H–1H COSY correlations (H-9/H-10, H-10/H-11, and H-11/H-12) and HMBC correlations (H-20/C-9, H-12/C-10, and H-16/C-11) demonstrated that the methylene at C-10 in 1 were replaced by a hydroxy methine in 3 (Table 1 and Table S3, and Figures S17 and S20). The HMBC correlations for the 6-5-6-6 skeleton were the same as those of 1 and 3 (Table S3 and Figure S20). The above evidence confirmed the planar structure of 3. The study for the stereochemistry of 3 is described later (together with that of 4).
Trichodermanin H (4) had the same molecular formula C20H34O3 by HRFABMS m/z 345.2415 [M + Na]+ (calcd for C20H34O3Na: 345.2396) as that of 1 and 3 (Figure S28). Comparing the NMR spectra of 4 with those of 3 (Table 1, Tables S3 and S4, and Figures S22, S23 and S26), we proposed that the differences in the NMR chemical shifts at H-3 (δH 1.88), H-10 (δH 1.59, 1.80), H-17 (δH 3.60, 3.95), C-2 (δC 36.9), C-3 (δC 34.0), C-9 (δC 43.5), C-10 (δC 21.6), C-11 (δC 44.1), and C-17 (δC 68.1) for 4 from those for 3 were caused by a change in the linkage position of the two hydroxy groups, i.e., 4 was a positional isomer of 3. The NMR spectral features (Table 1, Tables S1 and S4, and Figures S22, S23 and S26) resembled those of 1, except for the proton signals for H-2 (δH 2.74, ddd and δH 1.64, ddd) and H-17 (δH 3.60, dd and 3.95, dd) and the carbon signals of C-1 (δC 72.1) and C-17 (δC 68.1) in 4. These data and the 1H–1H COSY correlation between H-3 and H-17 and H-9 and H-10 indicated the planar structure of 2 (Figure S24). This structure was supported by the HMBC correlations observed being similar to those of 1, 2, and 3 (Table S4, and Figure S27). The most remarkable structural feature of 4 was that 3-CH3 (C-17) in the previous trichodermanins was replaced with a hydroxy methyl group.
The relative stereochemical configurations and conformations of 3 and 4 were investigated by NOESY experiments (Tables S3 and S4, Figure 6, and Figures S18 and S25). The key NOESY correlations for ring A (H-2α/H-19, H-1/H20, and H-3/H-20), ring B (H-7α/H-12, H-7α/H-18, and H-12/H-18), and ring C (H-3/H-11, H-11/H-14β, and H-13α/H-18) showed that the relative configurations and conformations in the 6-5-6-6 ring system of 3 and 4 were identical with those of the above metabolites (Tables S3 and S4, Figure 6, and Figures S18 and S25). The absolute configurations of 3 and 4 were not positively elucidated because of the small amount of material available. However, it is unreasonable to assume that only 3 and 4 derived from the same fungus are enantiomers of other trichodermanins. Therefore, we suggest the absolute stereostructures of 3 and 4 as shown in Figure 1 based on assumed biosynthetic relatedness. These stereostructural hypotheses will be confirmed using the modified Mosher’s method [15], the quantum chemical equivalent circulating density (ECD) calculation method [17], or synthetic research in the future.
As a primary screen for antitumor activity, the cancer cell growth inhibitory properties of trichodermanins E–H (14) were examined using murine P388 leukemia, human HL-60 leukemia, and murine L1210 leukemia cell lines. The results are shown in Table 2. Both 1 and 2 exhibited modest activities against the cell lines, although they were inferior to that of 5-fluorouracil (Table 2). However, 3 and 4 did not inhibit cell growth at all (Table 2). In the study of the structure–activity relationship inferred from the activities of 14 and previous trichodermanins E and D, the binding positions (C-1, C-2, C-6, C-10, and C-17) of the hydroxy groups did not contribute to the enhancement of activity. The number had some influence, i.e., the activities of diol derivatives as 2 and trichodermanins C were more potent than those of the triol derivatives as 1, 3, 4, and trichodermanin D. In particular, trichodermanins C with a carbonyl group at C-1 showed strong activity comparable to 5-fluorouracil. We speculate that the significant cell growth inhibition of trichodermanins C was caused by the change of conformation of A ring due to the carbonyl group at C-1. A comparison of cytotoxicity with normal cell and an elucidation of the detailed mechanism of activity will continue to be examined after supply of these fungal metabolites.

3. Materials and Methods

3.1. General Experimental Procedures

These are the same procedures as those in recent reports [7,8,9,10,11]. NMR spectra were recorded on an Agilent-NMR-vnmrs (Agilent Technologies, Santa Clara, CA, USA) 600 with tetramethylsilane (TMS) as an internal reference. FABMS was recorded using a JEOL JMS-7000 mass spectrometer (JEOL, Tokyo, Japan). IR spectra were recorded on a JASCO FT/IR-680 Plus (Tokyo, Japan). Optical rotations were measured using a JASCO DIP-1000 digital polarimeter (Tokyo, Japan). Silica gel 60 (230–400 mesh, Nacalai Tesque, Inc., Kyoto, Japan) was used for column chromatography with medium pressure. Octadecyl silica (ODS) HPLC was run on a JASCO PU-1586 (Tokyo, Japan) equipped with a differential refractometer RI-1531 (Tokyo, Japan) and Cosmosil packed column 5C18-MSII (25 cm × 20 mm i.d., Nacalai Tesque, Inc., Kyoto, Japan). Analytical thin-layer chromatography (TLC) was performed on precoated Merck aluminum sheets (DC-Alufolien Kieselgel 60 F254, 0.2 mm, Merck, Darmstadt, Germany) with the solvent system CH2Cl2/MeOH (19:1) (Nacalai Tesque, Inc., Kyoto, Japan), and compounds were viewed under a UV lamp (AS ONE Co., Ltd., Osaka, Japan) and sprayed with 10% H2SO4 (Nacalai Tesque, Inc., Kyoto, Japan), followed by heating.

3.2. Fungal Material

This study is a follow-up report for this fungal strain. As shown in the previous reports [9], the fungus Trichoderma harzianum was isolated from a piece of inner tissue of the marine sponge Halichondria okadai collected at Osaka Bay, Japan, in October 2008. The fungal strain was identified by Techno Suruga Laboratory Co., Ltd. The sponge was wiped with EtOH and a cutting applied to the surface of nutrient agar layered in a Petri dish. Serial transfers of one of the resulting colonies provided a pure strain of T. harzianum.

3.3. Culturing and Isolation of Metabolites

The EtOAc extract (8.6 g) of the culture filtrate, which was obtained as described in previous reports [9,10,11,12], was chromatographed on a silica gel column with a CHCl3/MeOH gradient as the eluent to afford Fr. 1 (5% MeOH in CHCl3 eluate, 65.6 mg) and Fr. 2 (5% MeOH in CHCl3 eluate, 119.7 mg). Fr. 1 was purified by ODS HPLC using MeOH/H2O (80:20) as the eluent to afford Fr. 3 (11.3 mg) and 2 (3.7 mg, tR 60 min). Fr. 3 was purified by HPLC using MeCN/H2O (40:60) as the eluent to afford 1 (0.93 mg, tR 27.5 min). Fr. 2 was purified by ODS HPLC using MeOH/H2O (80:20) as the eluent to afford Fr. 4 (9.9 mg). Fr. 4 was purified by HPLC using MeCN/H2O (30:70) as the eluent to afford 3 (0.8 mg, tR 58 min) and 4 (1.3 mg, tR 60 min)
Trichodermanins E (1): pale yellow oil; [α] D 22 + 188.0 (c 0.09, MeOH); IR (neat) νmax/cm−1: 3383. HRFABMS m/z 345.2397 [M + Na]+ (calcd for C20H34O3Na: 345.2396); NMR data, see Table 1 and Table S1 and the previous report [12].
Trichodermanins F (2): pale yellow oil; [α] D 22 + 22.4 (c 0.24, MeOH); IR (neat) νmax/cm−1: 3398. HRFABMS m/z 289.2534 [M − Na]+ (calcd for C20H34O3Na: 289.2532). NMR data, see Table 1 and Table S2.
Trichodermanins G (3): pale yellow oil; [α] D 22 + 8.0 (c 0.08, MeOH); IR (neat) νmax/cm−1: 3387. HRFABMS m/z 345.2395 [M + Na]+ (calcd for C20H34O3Na: 345.2396). NMR data, see Table 1 and Table S3.
Trichodermanin H (4): pale yellow oil; [α] D 22 + 52.0 (c 0.04, MeOH); IR (neat) νmax/cm−1: 3359. FABMS m/z (rel. int.): 345 ([M + Na]+, 37.4%) 305 (43.8%), 287 (54.5%), 115 (100%). HRFABMS m/z 345.2415 [M + Na]+ (calcd for C20H34O3Na: 345.2396). NMR data, see Table 1 and Table S4.

3.4. Formation of Dibenzoate of 1

To a solution of 1 (1.8 mg, 5.6 μmol) in abs. pyridine (0.3 mL), p-bromobenzoyl chloride (3.0 mg, 13.6 μmol) was added, and the reaction mixture was stirred at room temperature for 15 h. Water (1.0 mL) was added to the reaction mixture and extracted using CH2Cl2. The organic layer was evaporated under reduced pressure, and the residue was purified by HPLC using 100% MeOH as the eluent to afford dibenzoate 1a (2.2 mg, 58.8%) as a pale yellow oil.
Dibenzoate 1a: pale yellow oil; HRFABMS m/z 709.1138 [M + Na]+ (calcd for C34H40O579Br2Na: 709.1139). 1H NMR data (TMS, CDCl3) δ ppm: 1.04 (3H, s), 1.13, 1.21 (3H, d), 1.22 (3H, s), 1.23 (3H, s), 1.24 (3H, s), 1.26 (3H, s), 1.28, 1.47, 1.56, 1.58, 1.68, 1.82, 1.85, 1.88, 1.92, 2.07, 2.25, 5.50 (1H, d), 5.85 (1H, dd), 7.56 (Ar. H), 7.81 (Ar. H), 7.88 (Ar. H).

3.5. Formation of the (S)- and (R)-MTPA Esters of 2

To a solution of 2 (1.0 mg, 3.3 μmol) in abs. pyridine (0.3 mL), (R)-(–)-MTPA chloride (3.0 mg, 15.7 μmol) was added, and the reaction mixture was stirred at room temperature for 2 h. The reaction mixture was evaporated under reduced pressure, and the residue was purified by HPLC using MeOH–H2O (90:10) as the eluent to afford (S)-MTPA ester 2a (0.8 mg, 47.2%).
2 (1.2 mg, 3.9 μmol) and (S)-(+)-MTPA chloride (3.0 mg, 15.7 μmol) were treated with the same procedure to afford (R)-MTPA ester 2b (1.5 mg, 73.7%).
MTPA ester 2a: pale yellow oil; HRFABMS m/z 545.2860 [M + Na]+ (calcd for C30H41F3O4Na: 545.2855) (Figure S32). 1H NMR data and spectrum including 2D NMR spectra are listed in Table S5, and Figures S29–S31 of Supplementary Materials.
MTPA ester 2b: pale yellow oil; HRFABMS m/z 545.2849 [M + Na]+ (calcd for C30H41F3O4Na: 545.2855) (Figure S36). 1H NMR data and spectrum including 2D NMR spectra are listed in Table S5, and Figures S33–S35 of Supplementary Materials.

3.6. Assay for Cytotoxicity

Cytotoxic activities of 14 were examined by the same procedure to date [5,6,7,8,9,10,11,12,18], i.e., the 3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl- 2H-tetrazolium bromide (MTT) method. P388, HL-60, and L1210 cells were cultured in RPMI 1640 Medium (10% fetal calf serum) at 37 °C in 5% CO2. The test materials were dissolved in dimethyl sulfoxide (DMSO) to give a concentration of 10 mM, and the solution was diluted with the Essential Medium to yield concentrations of 200, 20, and 2 μM, respectively. Each solution was combined with each cell suspension (1 × 105 cells/mL) in the medium, respectively. After incubating at 37 °C for 72 h in 5% CO2, grown cells were labeled with 5 mg/mL MTT in phosphate-buffered saline (PBS), and the absorbance of formazan dissolved in 20% sodium dodecyl sulfate (SDS) in 0.1 N HCl was measured at 540 nm with a microplate reader (MTP-310, CORONA electric). Each absorbance value was expressed as percentage relative to that of the control cell suspension that was prepared without the test substance using the same procedure as that described above. All assays were performed three times, semilogarithmic plots were constructed from the averaged data, and the effective dose of the substance required to inhibit cell growth by 50% (IC50) was determined.

3.7. The Origin of the Cell Lines

P388 cell line was obtained from Dr. Numata. HL-60 cell line was obtained from Dr. Kawai. L1210 cell line was from Dr. Endo.

4. Conclusions

In this study, three new terpenes with a rare fused 6-5-6-6 ring system, namely, trichodermanins E–H (1–4), were isolated from the fungus T. harzianum separated from the marine sponge H. okadai. Spectral analyses and chemical transformations were utilized to elucidate the absolute stereostructures of 1 and 2. The relative stereostructures of 3 and 4 were assumed to be those of the same class of metabolites. Therefore, we deduced that their absolute configurations were also the same as those of other trichodermanins. In cytotoxic assays performed using three cancer cell lines, 1 and 2 exhibited moderate activity against the cell lines, although they were inferior to that of 5-fluorouracil.

Supplementary Materials

The following are available online at https://www.mdpi.com/1660-3397/17/8/480/s1, Table S1: Spectral data including 2D NMR data for 1, Table S2: Spectral data including 2D NMR data for 2, Table S3: Spectral data including 2D NMR data for 3, Table S4: Spectral data including 2D NMR data for 4, Table S5: Spectral data for 2a and 2b, Figure S1: 1H NMR spectrum of 1 in CDCl3, Figure S2: 13C NMR spectrum of 1 in CDCl3, Figure S3: 1H–1H COSY of 1, Figure S4: NOESY of 1, Figure S5: HMQC of 1, Figure S6: HMBC of 1, Figure S7: FABMS of 1, Figure S8: 1H NMR spectrum of 2 in CDCl3, Figure S9: 13C NMR spectrum of 2 in CDCl3, Figure S10: 1H–1H COSY of 2, Figure S11: NOESY of 2, Figure S12: HMQC of 2, Figure S13: HMBC of 2, Figure S14: FABMS of 2, Figure S15: 1H NMR spectrum of 3 in CDCl3, Figure S16: 13C NMR spectrum of 3 in CDCl3, Figure S17: 1H–1H COSY of 3, Figure S18: NOESY of 3, Figure S19: HMQC of 3, Figure S20: HMBC of 3, Figure S21: FABMS of 3, Figure S22: 1H NMR spectrum of 4 in CDCl3, Figure S23: 13C NMR spectrum of 4 in CDCl3, Figure S24: 1H–1H COSY of 4, Figure S25: NOESY of 4, Figure S26: HMQC of 4, Figure S27: HMBC of 4, Figure S28: FABMS of 4, Figure S29: 1H NMR spectrum of 2a in CDCl3, Figure S30: 1H–1H COSY of 2a, Figure S31: NOESY of 2a, Figure S32: FABMS of 2a, Figure S33: 1H NMR spectrum of 2b in CDCl3, Figure S34: 1H–1H COSY of 2b, Figure S35: NOESY of 2b, Figure S36: FABMS of 2b.

Author Contributions

T.Y., A.F., and T.K conceived and designed the experiments; T.Y. and A.F. performed the experiments; T.Y. analyzed the data; and T.Y. wrote the paper.

Funding

This research received no external funding.

Acknowledgments

We thank Endo (Kanazawa University) for supply of the cancer cells. We are grateful to M. Fujitake and K. Minoura of this university for MS and NMR measurements, respectively.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nicoletti, R.; Vinale, F. Bioactive Compounds from Marine-Derived Aspergillus, Penicillium, Talaromyces and Trichoderma Species. Mar. Drugs 2018, 16, 408. [Google Scholar] [CrossRef] [PubMed]
  2. Imhoff, J. Natural Products from Marine Fungi—Still an Underrepresented Resource. Mar. Drugs 2016, 14, 19. [Google Scholar] [CrossRef] [PubMed]
  3. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.G.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2017, 34, 235–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Blunt, J.W.; Carroll, A.R.; Copp, B.R.; Davis, R.A.; Keyzers, R.A.; Prinsep, M.R. Marine natural 369 products. Nat. Prod. Rep. 2018, 35, 8–53. [Google Scholar] [CrossRef] [PubMed]
  5. Yamada, T.; Iritani, M.; Ohishi, H.; Tanaka, K.; Minoura, K.; Doi, M.; Numata, A. Pericosines, antitumour metabolites from the sea hare-derived fungus Periconia byssoides. Structures and biological activities. Org. Biomol. Chem. 2007, 5, 3979–3986. [Google Scholar] [CrossRef] [PubMed]
  6. Yamada, T.; Kitada, H.; Kajimoto, T.; Numata, A.; Tanaka, R. The relationship between the CD Cotton effect and the absolute configuration of FD-838 and its seven stereoisomers. J. Org. Chem. 2010, 75, 4146–4153. [Google Scholar] [CrossRef] [PubMed]
  7. Yamada, T.; Kikuchi, T.; Tanaka, R. Altercrasin A, a novel decalin derivative with spirotetramic acid, produced by a sea urchin-derived Alternaria sp. Tetrahedron Lett. 2015, 56, 1229–1232. [Google Scholar] [CrossRef]
  8. Yamada, T.; Tanaka, A.; Nehira, T.; Nishii, T.; Kikuchi, T. Altercrasins A–E, decalin derivatives, from a sea-urchin-derived Alternaria sp.: Isolation and structural analysis including stereochemistry. Mar. Drugs 2019, 17, 218. [Google Scholar] [CrossRef] [PubMed]
  9. Yamada, T.; Mizutani, Y.; Umebayashi, Y.; Inno, N.; Kawashima, M.; Kikuchi, T.; Tanaka, R. A novel ketoaldehyde decalin derivative, produced by a marine sponge-derived Trichoderma harzianum. Tetrahedron Lett. 2014, 55, 662–664. [Google Scholar] [CrossRef]
  10. Yamada, T.; Umebayashi, Y.; Kawashima, M.; Sugiura, Y.; Kikuchi, T.; Tanaka, R. Determination of the chemical structures of tandyukisins B–D, isolated from a marine sponge-derived fungus. Mar. Drugs 2015, 13, 3231–3240. [Google Scholar] [CrossRef] [PubMed]
  11. Suzue, M.; Kikuchi, T.; Tanaka, R.; Yamada, T. Tandyukisins E and F, novel cytotoxic decalin derivatives isolated from a marine sponge-derived fungus. Tetrahedron Lett. 2016, 57, 5070–5073. [Google Scholar] [CrossRef]
  12. Yamada, T.; Suzue, M.; Arai, T.; Kikuchi, T.; Tanaka, R. Trichodermanins C–E, new diterpenes with a fused 6-5-6-6 ring system produced by a marine sponge-derived fungus. Mar. Drugs 2017, 15, 169. [Google Scholar] [CrossRef] [PubMed]
  13. Sun, P.X.; Zheng, C.J.; Li, W.C.; Huang, F.; Qin, L.P. Trichodermanin A, a novel diterpenoid from endophytic fungus culture. J. Nat. Med. 2011, 65, 381–384. [Google Scholar] [CrossRef] [PubMed]
  14. Yamamoto, T.; Izumi, N.; Ui, H.; Sueki, A.; Masuma, R.; Nonaka, K.; Hirose, T.; Sunazuka, T.; Nagai, T.; Yamada, H.; et al. Wickerols A and B: Novel anti-influenza virus diterpenes produced by Trichoderma atroviride FKI-3849. Tetrahedron 2012, 68, 9267–9271. [Google Scholar] [CrossRef]
  15. Ohtani, I.; Kusumi, T.; Kashman, Y.; Kakisawa, H. High-field FT NMR application of Mosher’s method. The absolute configurations of marine terpenoids. J. Am. Chem. Soc. 1991, 113, 4092–4096. [Google Scholar] [CrossRef]
  16. Harada, N.; Nakanishi, K. Circular Dichroic Spectroscopy–Exciton Coupling in Organic Stereochemistry; University Science Books: Sausalito, CA, USA, 1983. [Google Scholar]
  17. Sun, Z.H.; Gu, J.; Ye, W.; Wen, L.X.; Lin, Q.B.; Li, S.N.; Che, Y.C.; Li, H.H.; Zhang, W.M. Geospallins A–C: New thiodiketopiperazines with inhibitory activity against angiotensin-converting enzyme from a deep-Sea-derived Fungus Geosmithia pallida FS140. Mar. Drugs 2018, 16, 464. [Google Scholar] [CrossRef] [PubMed]
  18. Miki, I.; Ishihara, N.; Ohtoshi, M.; Kase, H. Simple colorimetric cell-cell adhesion assay using MTT-stained leukemia cells. J. Immiunol. Methods 1993, 164, 155–261. [Google Scholar] [CrossRef]
Figure 1. Structures of trichodermanins E–H (14).
Figure 1. Structures of trichodermanins E–H (14).
Marinedrugs 17 00480 g001
Figure 2. Circular dichroism (CD) spectrum of the dibenzoate of trichodermanin E (1a).
Figure 2. Circular dichroism (CD) spectrum of the dibenzoate of trichodermanin E (1a).
Marinedrugs 17 00480 g002
Figure 3. Key 1H–1H COSY and HMBC correlations for (1,2).
Figure 3. Key 1H–1H COSY and HMBC correlations for (1,2).
Marinedrugs 17 00480 g003
Figure 4. Key NOESY correlations for 2.
Figure 4. Key NOESY correlations for 2.
Marinedrugs 17 00480 g004
Figure 5. 1H chemical-shift differences between the (S)- and (R)-MTPA esters 2a and 2b.
Figure 5. 1H chemical-shift differences between the (S)- and (R)-MTPA esters 2a and 2b.
Marinedrugs 17 00480 g005
Figure 6. Key NOESY correlations for 3 and 4.
Figure 6. Key NOESY correlations for 3 and 4.
Marinedrugs 17 00480 g006
Table 1. NMR spectral data for 1–4 in CDCl3.
Table 1. NMR spectral data for 1–4 in CDCl3.
Position1234
δHaδCδHaδCδHaδCδHaδC
2.54ddd36.5 (t)
4.11d80.4 (d)1.68m 4.20dd72.6 (d)4.24dd72.1 (d)
3.88dd83.7 (d)4.30ddd74.2 (d)2.66ddd41.8 (t)2.74ddd36.9 (t)
1.45m 1.64ddd
31.88qd36.6 (d)1.98qd37.9 (d)2.10m26.7 (d)1.88dddd34.0 (d)
4 41.2 (s) 40.8 (s) 38.9 (s) 38.9 (s)
5 39.4 (s) 39.0 (s) 39.0 (s) 38.5 (s)
61.50dd53.2 (d)1.62m41.8 (d)1.41dd52.2 (d)1.45dd52.2 (d)
1.78dd40.9 (t)1.71dd42.5 (t)1.69dd41.3 (t)1.74dd40.9 (t)
1.70dd 1.56dd 1.65dd 1.68dd
8 39.6 (s) 39.7 (s) 38.8 (s) 39.2 (s)
1.03m43.5 (t)1.02m43.9 (t)1.48m52.2 (t)1.03m43.5 (t)
1.43m 1.43m 1.43m
10α1.59m21.6 (t)1.60m21.6 (t)4.38ddd72.9 (d)1.59m21.5 (t)
10β1.80m 1.81m 1.80m
111.81dd44.2 (d)1.85m44.4 (d)1.85dd55.0 (d)1.88ddd44.1 (d)
121.32d51.8 (d)1.30d51.9 (d)1.28d50.6 (d)1.28d52.0 (d)
13α1.23ddd26.3 (t)1.18ddd26.4 (t)1.19ddd25.9 (t)1.28m25.5 (t)
13β1.72ddd 1.68m 1.74ddd 1.52m
14α1.64ddd41.1 (t)1.62m41.2 (t)1.60ddd41.0 (t)1.67m40.87 (t)
14β1.46ddd 1.46ddd 1.50ddd 1.50m
15 73.6 (s) 73.8 (s) 73.1 (s) 73.6 (s)
161.18s20.5 (q)1.18s20.4 (q)1.22s21.5 (q)1.18s20.5 (q)
171.23d20.0 (q)1.19d20.4 (q)1.09d22.4 (q)3.60dd68.1 (t)
3.95dd
18ax0.99s25.7 (q)0.95s25.3 (q)0.97s25.8 (q)0.99s25.7 (q)
19eq1.04s25.2 (q)0.96s25.4 (q)1.14s24.9 (q)1.13s25.5 (q)
200.98s19.8 (q)1.13s20.3 (q)1.17s21.4 (q)0.92s19.5 (q)
a1H chemical shift values (δ ppm from SiMe4) followed by multiplicity.
Table 2. Cytotoxicity of metabolites 13 against cancer cell lines.
Table 2. Cytotoxicity of metabolites 13 against cancer cell lines.
CompoundsCell line P388Cell line HL-60Cell line L1210
IC50 (μM) aIC50 (μM) aIC50 (μM) a
152.1 ± 1.359.8 ± 2.2125.2 ± 4.3
258.9 ± 1.2 42.9 ± 3.041.5 ± 2.5
3>300>300>300
4>30085.3 ± 2.173.2 ± 2.2
DMSO (control)>300>300>300
5-fluorouracil b3.9 ± 0.63.7 ± 0.14.2 ± 0.4
a DMSO was used as vehicle. b Positive control.

Share and Cite

MDPI and ACS Style

Yamada, T.; Fujii, A.; Kikuchi, T. New Diterpenes with a Fused 6-5-6-6 Ring System Isolated from the Marine Sponge-Derived Fungus Trichoderma harzianum. Mar. Drugs 2019, 17, 480. https://doi.org/10.3390/md17080480

AMA Style

Yamada T, Fujii A, Kikuchi T. New Diterpenes with a Fused 6-5-6-6 Ring System Isolated from the Marine Sponge-Derived Fungus Trichoderma harzianum. Marine Drugs. 2019; 17(8):480. https://doi.org/10.3390/md17080480

Chicago/Turabian Style

Yamada, Takeshi, Ayano Fujii, and Takashi Kikuchi. 2019. "New Diterpenes with a Fused 6-5-6-6 Ring System Isolated from the Marine Sponge-Derived Fungus Trichoderma harzianum" Marine Drugs 17, no. 8: 480. https://doi.org/10.3390/md17080480

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