Next Article in Journal
Copper(I) Complexes with Terphenyl-Substituted NPN Ligands Bearing Pyridyl Groups: Synthesis, Characterization, and Catalytic Studies in the S-Arylation of Thiols
Previous Article in Journal
Effect of Contrasting Redox Potential Evolutions and Cap Management Techniques on the Chemical Composition of Red Wine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Amesilide, a New Bicyclic Polyketide from the Marine Fungus Amesia nigricolor MUT6601

by
Giang Nam Pham
1,†,
Matteo Florio Furno
2,
Juan A. Garcia-Sanchez
3,
Patrick Munro
3,
Fatouma Mohamed Abdoul-Latif
4,
Laurent Boyer
3,
Giovanna Cristina Varese
2 and
Mohamed Mehiri
1,*
1
Marine Natural Products Team, Institut de Chimie de Nice, Université Côte d’Azur, CNRS UMR 7272, 06108 Nice, France
2
Department of Life Sciences and Systems Biology, University of Torino, Mycotheca Universitatis Taurinensis (MUT), Viale Mattioli 25, 10125 Torino, Italy
3
INSERM U1065, Centre Méditerranéen de Médecine Moléculaire (C3M), Bâtiment Universitaire ARCHIMED, 151 Route de Saint Antoine de Ginestière BP, 23194 Nice, France
4
Medicinal Research Institute, Center for Studies and Research of Djibouti, IRM-CERD, Route de l’Aéroport, Haramous, Djibouti City P.O. Box 486, Djibouti
*
Author to whom correspondence should be addressed.
Present address: Faculty of Pharmacy, PHENIKAA University, Yen Nghia, Ha Dong, Hanoi 12116, Vietnam.
Molecules 2025, 30(15), 3169; https://doi.org/10.3390/molecules30153169
Submission received: 8 December 2024 / Revised: 25 May 2025 / Accepted: 28 May 2025 / Published: 29 July 2025
(This article belongs to the Section Natural Products Chemistry)

Abstract

A new bicyclic polyketide, amesilide (1), along with the previously reported metabolites, chamisides A (2), B (3), and E (4), chaetoconvosins B (5) and C (6), and chaetochromins A (7) and B (8), were isolated from the marine fungus Amesia nigricolor MUT6601. The structures of the compounds were determined by extensive spectrometric (HRMS) and spectroscopic (1D and 2D NMR) analyses, as well as specific rotation. Absolute configurations of the stereogenic centers of amesilide (1) were determined by a comparison of its experimental circular dichroism (CD) spectrum with its time-dependent density functional theory (TD-DFT) electronic circular dichroism (ECD) spectra. Among them, chaetochromins A (7) and B (8) showed strong antibacterial activity against Staphylococcus aureus S25 (MBC values of 12.50 µM and MIC values of 6.25 µM) and a moderate cytotoxicity against monocytes (THP-1) and peripheral blood cells (PBMC) (IC50 values of 33.65–40.01 µM).

Graphical Abstract

1. Introduction

Marine fungi, which usually produce metabolites different from their terrestrial counterparts, represent important sources of bioactive metabolites for drug discovery [1]. Indeed, reports of new natural products from marine-derived fungi have significantly increased over the last few decades [2]. Chaetomium and Chaetomium-like species (Chaetomiaceae) have been proven to be rich sources of specialized metabolites [3,4,5]. The taxonomy of this genus has recently been revised based on phylogenetic analyses and morpho-physiological characteristics [6]. Currently, the family Chaetomiaceae comprises 10 genera which include five newly established genera, including the genus Amesia [6].
Amesia nigricolor (L.M. Ames) X. Wei Wang and Samson (synonym Chaetomium nigricolor) has latterly attracted the attention of several researchers. Several metabolites recently isolated from A. nigricolor have demonstrated anti-inflammatory or cholesterol-lowering activities [4,7,8,9]. Kim et al. identified twelve metabolites from A. nigricolor isolated from soil [7], finding a bis-naphtho-γ-pyrone derivative ((aS)-asperpyrone A) as an important anti-inflammatory component. In another study, Dhayanithy et al. reported that the extract of the endophytic fungus A. nigricolor, which was isolated from Catharanthus roseus, exhibited potent cytotoxicity and antioxidant properties [10]. Chamiside A, a cytochalasin with a tricyclic core skeleton, was identified by Wang et al. and found to display a moderate antibacterial activity against Staphylococcus aureus [8]. Recently, Gu et al. continued the chemical investigation of the same strain of A. nigricolor F5 of the latter study, discovering five novel chytochalasans, Chamisides B-F [10]. Interestingly, Chamiside F showed a promising bioactivity, with inhibitory activity against the cholesterol transporter Niemann-Pick C1-like 1 (NPC1L1) protein, reducing the cholesterol absorption.
In this context, we investigated the metabolites produced by the marine fungus A. nigricolor MUT6601. Herein, we report the isolation, structural elucidation, and antimicrobial activities of a new bicyclic polyketide, amesilide (1), along with the previously reported metabolites, chamisides A (2), B (3), and E (4), chaetoconvosins B (5) and C (6), and chaetochromins A (7) and B (8) from the marine fungus A. nigricolor MUT6601. All of the isolated compounds were evaluated for their antimicrobial properties and cytotoxicity.

2. Results and Discussion

The EtOAc extract of A. nigricolor MUT6601 was fractionated and purified by a combination of C-18 solid-phase extraction, silica gel column chromatography, and semi-preparative HPLC to yield compounds 18 (Figure 1).
Amesilide (1) was isolated as a yellow powder. Its molecular formula, C25H32O4, was deduced from the HRESI(+)MS analysis which showed a pseudo-molecular ion peak at m/z 397.2369 [M+H]+ (calcd for C25H33O4+ 397.2373, Δ = 1.01 ppm, 10 degrees of unsaturation). The 1H, 13C, and 1H-13C HSQC NMR spectra indicated the presence of six methyl groups (δc 32.7, 20.1, 16.8, 14.6, 13.9, and 11.5), one sp3 methylene (δc 41.3), five sp3 methines (δc 50.9, 44.5, 41.2, 37.5, and 33.5), six sp2 methines (δc 152.9, 144.7, 139.3, 134.0, 122.9, and 115.6), and seven quaternary sp2 carbons (δc 213.8, 198.8, 172.0, 151.1, 135.4, 133.4, and 131.9), including one ketone (δc 213.8), one α,β-unsaturated ketone (δc 198.8), and one α,β-unsaturated carboxylic acid (δc 172.0) (Table 1). The assembly of the 1H-13C HMBC correlations and 1H-1H COSY spin systems revealed that 1 has a bicyclic decalin system. The structure of compound 1 was established based on key 1H-13C HMBC correlations, including H-23 to C-9, C-10, and C-11; H-24 to C-11, C-12, and C-13; H-25 to C-15, C-16, and C-17; H-7 to C-8, C-9, and C-17; and H-18 to C-8, C-16, and C-17. The COSY spectrum showed continuous spin systems (H-7 to H-23 and H-24 to H-15), further supporting this framework. The locations of a ketone group at C-11, three methyl groups attached to C-10, C-12, and C-16, and two substituted chains attached to C-8 and C-17 were also determined through the above correlations. One substituent was determined as 4,6-dimethylhepta-2,4,6-trienoic acid by the COSY correlation between H-2 and H-3, and the HMBC correlations from H3-21 to C-3, C-4, and C-5 and from H3-22 to C-5, C-6, and C-7. The other substituent was determined by the HMBC correlations from the methyl group H-20 to C-18 and C-19 and from H-18 to C-19 and C-20 (Table 1). The planar structure of 1 was then established as shown in Figure 2.
The relative configuration of 1 was determined by the analyses of vicinal coupling constants and the NOESY correlations. The geometry of the double bond at C-2/C-3 was determined as E, by the large vicinal coupling constant between H-2 and H-3 (J2,3 = 15.5 Hz), and the 1H-1H NOESY correlation of H-2/H-21. Then, the NOESY correlations of H-21/H-22, H-3/H-5, and H-5/H-7 indicated both double bonds C-4/C-5 and C-6/C-7 were E. The geometry of the olefin C-17/C18 was established as E by the key NOESY correlation of H-18/H-25. The NOESY correlations of H-8/H-23 and H-7/H-14 indicated a cis-fused ring junction. A small coupling constant of less than 1 Hz between H-9 and H-14 indicated that the angle between them was near 90°, providing stronger confirmation for the cis-fused decalin system. The NOESY correlation of H-10/H-12 suggested they were both axial and co-facial.
The most unusually distinctive feature of the NMR spectral dataset of 1 is the downfield shift of H-8 at δH 5.15, which is directly connected to a carbon at δC 37.5. However, this phenomenon has also been reported in polyketides with structures closely related to 1, isolated from a yew-associated Penicillium species by Stierle et al. [11]. This may be due to the anisotropy effect generated by the C-19 ketone group, which deshields H-8. The authors attempted to reduce the C-19 carbonyl group using NaBH4, yielding two products in which the chemical shifts of H-8 decreased to approximately δH 3.82–3.86, confirming the effect of the carbonyl group. However, preussilide C, another polyketide from the endophytic fungus Preussia similis, has an almost identical structure to 1, including a carbonyl group at C-19 but differing in the arrangement of its double bonds [12]. Despite this similarity, it exhibits an H-8 shift of only δH 3.26. This suggests that the C-15 to C-19 double bond system also plays a role in causing H-8 to shift toward the downfield region. Hence, the distinctive downfield signal of H-8 is a very coincidental effect of many factors, in which H-8 is both the allylic proton of the two conjugated double bond systems from C-1 to C-7 and C-15 to C-19, and is also influenced by the anisotropic effect from the carbonyl group at C-19. To provide further solid evidence for this phenomenon, the 1H and 13C chemical shifts of 1 were also calculated for comparison.
The experimental CD spectrum of 1 exhibited a negative Cotton effect (CE) at λmax = 278 nm and one positive CE at λmax = 310 nm (Figure 3). The Boltzmann-averaged TD-DFT calculated ECD spectrum for the most stable conformers of the enantiomer 1a (8R, 9R, 10R, 12S, and 14S), performed at the CAM-B3LYP/6–31G(d,p) level of theory (IEFPCM, acetonitrile), also showed one negative CE at λmax = 275 nm and one positive CE at λmax = 305 nm, which reproduced the signs and differences in amplitude of the experimental CEs. Thus, the absolute configuration of 1 was determined as 8R, 9R, 10R, 12S, and 14S (Figure 3, Tables S1–S25).
Seven already reported compounds were identified as chamisides A (2), B (3), and E (4) [8,9], chaetoconvosins B (5) [13] and C (6) [9], and chaetochromins A (7) and B (8) [14] by comparing their HRESIMS, NMR data, and specific rotations with those reported in the literature. The relative stereochemistry of chaetochromin B (8) remained unclarified.
All isolated compounds were evaluated for their antimicrobial activities against a panel of microorganisms, including Staphylococcus aureus S25, Escherichia coli UTI89, Candida krusei CK01, Candida albicans CA01, and Leishmania infantum (LI01). Only chaetochromins A (7) and B (8) exhibited strong activity against S. aureus S25, with MIC values of 6.25 µM (inhibiting bacterial growth) and MBC values of 12.50 µM (killing 99.9% of bacteria). The strong antibacterial activity of chaetochromin A against S. aureus is also consistent with previous reports [15]. More interestingly, despite showing strong antibacterial activity, these compounds showed no cytotoxicity against human erythrocytes (RBC), and moderate activities against human monocytes (THP-1) and peripheral blood mononuclear cells (PBMC) (IC50 values of 33.65–40.01 µM, Table S1). These results encourage further investigation into the potential clinical applications of compounds 7 and 8.

3. Materials and Methods

3.1. General Experimental Procedure

Optical rotations were determined using an Anton Paar MCP 150 polarimeter (Anton Paar, Graz, Austria). Fourier transform infrared (FT-IR) spectra were recorded on a Nicolet iS50 FT-IR spectrometer (Thermo Scientific, Waltham, MA, USA). Semi-preparative chromatography and high-performance liquid chromatography (HPLC) analyses were performed using a Waters Alliance e2695 HPLC system (Waters Corporation, Milford, MA, USA), equipped with a bifunctional Macherey-Nagel NUCLEODUR Sphynx RP column (250 × 4.6 mm or 250 × 10.0 mm, 5 µm) connected to both a Waters 2424 evaporative light scattering (ELS) detector and a Waters 2998 photodiode array (PDA) detector. HPLC-grade solvents were purchased from Sigma-Aldrich (Merck KGaA, Saint-Louis, MO, USA). NMR spectra were recorded on a 400 MHz Bruker Avance NMR spectrometer (Bruker Corporation, Billerica, MA, USA). Circular dichroism (CD) spectra were measured using a JASCO J-810 spectropolarimeter (JASCO International Co., Ltd., Tokyo, Japan). High-resolution mass spectra (HRMS) were acquired using a Thermo Q-Exactive Focus Orbitrap UPLC-HRMS system (Thermo Fisher Scientific, Waltham, MA, USA), and the data were processed with Thermo Xcalibur software version 2.2.44.

3.2. Microorganism Isolation and Identification

Amesia nigricolor MUT6601 was isolated from the sediments of the Harbor of Livorno, in the Tyrrhenian Sea in Tuscany (IT; 43°33′43.92″ N 10°17′42″ E). The harbor of Livorno is an area affected by several petrochemical and industrial activities, primarily due to contamination from metals and hydrocarbons [16].
As previously described [16], the dilution plate method was used to isolate fungi: the sediment suspensions were plated on Corn Meal Agar (CMA: 2 gL−1 corn meal infusion, 15 g L−1 agar) with antibiotics (Gentamicin 80 mgL−1 and Tazobactam 100 mgL−1) and 2% w/v of sea salts. All of the chemicals were purchased from Sigma-Aldrich. The identification at species level was carried out through morphological and molecular analyses. The internal transcribed spacer region (ITS) gene sequence obtained for A. nigricolor MUT6601 is available at GenBank NCBI under the accession number OP161804.
The strain is deposited at Mycotheca Universitatis Taurinensis (MUT, www.mut.unito.it) of the Department of Life Sciences and Systems Biology, University of Torino, Torino (Italy).

3.3. Fungal Solid State Fermentation

The fermentation was carried out following an adaptation of the method previously described by Yue et al. [17]. Briefly, A. nigricolor MUT6601 was fermented in solid rice media (RM), in 100 mL bottles containing 20 g of rice and 20 mL of mineral medium (MM: 10 mL/L mineral solution—MS and 1 mL/L trace metal solution—TMS. MS (100 mL): KCl 5 g, MgSO4•7H2O 5 g, and FeSO4•7H2O 0.1 g. TMS (100 mL): ZnSO4•7H2O 1 g and CuSO4 0.5 g.
Prior to the fermentation A. nigricolor, MUT6601 was cultured on CMA plates for 7–14 days at 24 °C. A mycelium homogenate was prepared as inoculum: the fungal biomass was collected from CMA plates with a sterilized scalpel and blended using distilled sterilized water (1 cm2/1 mL) using an Ultra-turrax® homogenizer (IKA) (IKA, Staufen, Germany) [18].
The solid-state fermentation was carried out in 80 bottles of RM in which 3 mL of mycelium homogenate was inoculated. The fermentation lasted 2 weeks.

3.4. Extraction and Purification

The colonized rice was extracted with ethyl acetate, and the resulting solution was concentrated in vacuo to yield 8.96 g of crude organic extract. The total extract was then fractionated by C-18 solid-phase extraction, eluting with water/methanol (100:0, 80:20, 60:40, 40:60, 20:80, 10:90, and 0:100, v/v), and then methanol/dichloromethane (50:50, v/v), to give 8 fractions (AN1 → AN8). Fraction AN4 (593.4 mg) was fractionated by silica gel column chromatography, eluting with cyclohexane/ethyl acetate (3:1 → 0:1, v/v) to yield seven sub-fractions (AN4.1 → AN4.7). Sub-fraction AN4.2 (181.2 mg) was purified by analytical HPLC using acetonitrile/water (55:45, 1 mL/min) containing 0.1% formic acid, to yield 3 (4.0 mg, tR = 8.9 min), 1 (6.4 mg, tR = 13.1 min), 5 (39.4 mg, tR = 14.7 min), and 2 (25.5 mg, tR = 16.2 min). Sub-fraction AN4.3 (43.9 mg) was purified by semi-preparative HPLC using the same mobile phase with a flow rate of 3 mL/min, to yield 6 (2.6 mg, tR = 12.5 min). Fraction AN5 (707.9 mg) was separated by silica gel column chromatography and eluting with cyclohexane/ethyl acetate (9:1 → 0:1, v/v) to yield 4 (38.2 mg) and eleven other sub-fractions (AN5.1, AN5.3 → AN5.12). Sub-fraction AN5.9 (27.5 mg) was purified by semi-preparative HPLC using acetonitrile/water (65:35, 3 mL/min) containing 0.1% formic acid, to yield 7 (2.0 mg, tR = 14.5 min) and 8 (3.6 mg, tR = 12.0 min).
Amesilide (1): yellow powder. [α]D20 -12.0 (c 0.6, CHCl3). Molecular formula: C25H32O4. HRESIMS m/z 397.2369 [M+H]+ (calcd for C25H33O4+ 397.2373). IR υmax: 2928.3, 1706.5, 1618.2, 1582.3, 1452.9, 1376.7, 1183.4, and 1024.1. CD (acetonitrile) λ (Δε): 214 (+6.11), 277 (−18.76), and 310 (+12.21). 1H and 13C NMR data see Table 1.

3.5. Computational Details

The conformational searches of each possible isomer were performed by employing MAESTRO software (Schrodinger LLC, New York, NY, USA), using an energy window of 5 kcal/mol, applying 10,000 steps of the Monte Carlo multiple minimum method with PRCG energy minimization using the Merck Molecular Force Field (MMFF) in gas phase, yielding 42 conformers of the isomer 1a (8S, 9S, 10S, 12R, and 14R) and 43 conformers of the enantiomer 1b (8R, 9R,10R, 12S, and 14S). Those occurring conformers were then subjected to geometrical optimization and vibrational frequencies calculation using the DFT/B3LYP/6-31G(d,p) (IEFPCM, acetonitrile) level with the Gaussian 16 A.03 package (Gaussian Inc., Wallingford, CT, USA) [19]. The low-energy conformers over 0.5% Boltzmann population were chosen for ECD calculation at TD-DFT/CAM-B3LYP/6-31G(d,p) (IEFPCM, acetonitrile) level. ECD curves were Boltzmann averaged and extracted by SpecDis v.1.7 software with a half-band of 0.3 eV.
For NMR calculation, all 43 conformers of 1b (8R, 9R,10R, 12S, and 14S) were further geometrically optimized using the DFT/B3LYP/6-31G(d,p) (IEFPCM, chloroform) level. They were then calculated for NMR shielding tensors at the IEFPCM/mpw1pw91/6-311G(d,p) level with chloroform as a solvent. The calculated NMR shielding tensors were averaged based on the Boltzmann populations, and the chemical shift values were calculated by the equation below (Equation (1)) where δx is the calculated NMR shift for nucleus x, and σ0 is the shielding tensor for the proton or carbon nuclei in tetramethylsilane calculated at the same condition.
δx = (σ0σx)/(1 − σ0/106)

3.6. Antimicrobial Assay

Stock solutions of all compounds were prepared in DMSO. Two bacterial (E. coli UTI89 and S. aureus S25) [20,21] and two yeast strains (C. krusei CK1 and C. albicans CA01) have been used for this study. For antibacterial assays, the two strains E. coli UTI89 and S. aureus S25 were inoculated in LB liquid medium and incubated at 37 °C with agitation until the culture had an optical density (OD600nm) of 1.2. Each culture was diluted 12 times to have a final OD value of 0.1. Then, 100 µL from the diluted (OD of 0.1) bacterial culture and 0.5 μL of each stock solution was added to the corresponding wells (in triplicate) in a 96-well plate to a final concentration of 50 μM. For the antifungal assay, the two strains C. krusei CK1 and C. albicans CA01 were spread on Petri plates with YPD (yeast-extract peptone dextrose) agar. Later, an isolated colony was collected in a tube containing 5mL of RGM (RPMI 1640® (21875, Invitrogen GIBCO®) + 34.56g/L MOPS) to have an OD of 0.05. The fungal culture and stock solutions of compounds were diluted and added to the wells as indicated for the antibacterial activity assay. To determine the potential inhibitory effect of extracts on fungal growth, absorbance at t = 0 h, t = 20 h (for antibacterial assay), and t = 48 h (for antifungal assay) were measured with the Biotek Synergy 2 Plate Reader at a wavelength of 600 nm and analyzed with the software Gen5 v3.08. Then, the differential OD was calculated (ODT48/20-ODT0) and compared to solvent control. A two-fold dilution concentration series of each compound that showed no growth (at 50 μM) was prepared to repeat the experiment to determine MIC/MBC. Ampicillin was used as positive control (MIC of 0.5 µg/mL for all bacteria and 1.0 µg/mL for all fungi).

3.7. Hemolytic and Cytotoxicity Assay

Each compound which showed antimicrobial activity at a concentration of 50 μM was selected for the cytotoxicity assay. Human erythrocytes (RBC), human monocytes (THP-1), and peripheral blood mononuclear cells (PBMC) were used for this study. THP-1 cells (TIB-202) were obtained from ATCC (https://www.atcc.org (accessed on 26 January 2025)) and activated with PMA (phorbol myristate acetate) (100 ng/mL) for 48 h. Cells were grown in RPMI 1640 medium supplemented with fetal bovine serum (FBS) 10% (10270-098, GIBCO®) and penicillin/streptomycin (1%) (15140-122 GIBCO®) at 37 °C under 5% CO2 atmosphere.
For the hemolytic assay, freshly obtained human erythrocytes (RBCs) from patients were washed twice with PBS (pH 7.4). This assay was performed at 10% hematocrit with two-fold serial dilutions from 200 μM of test compounds in a final volume of 100 μL, followed by incubation at 37 °C under 5% CO2 atmosphere for 30 min. Later, the plates were centrifugated at 3000g for 5 min and supernatant was transferred onto a new 96-well plate. Finally, RBC lysis was determined by the absorbance of the supernatant at 540 nm. Results were calculated with the following formula (Equation (2)):
%   hemolysis   =   100   ×   ODt     OD 100 OD 0     OD 100  
where ODt = the absorbance value of test compounds; OD100 = the absorbance value of the medium; and OD0 = the absorbance value of the wells treated with 1% Triton-X.
For the cytotoxicity assay, THP-1 and PBMC medium cultures were added into 96-well plates with 5 × 103 cells per well. A two-fold dilution concentration series of each compound was added to the plates to final concentrations of 200 − 6.25 μM. AlamarBlue (DAL1025, ThermoFisher, Waltham, MA, USA), a ready-to-use resazurin-based solution, was used to measure cell viability after treatment with the extracts. A total of 10 µL was added to each well and the plate was incubated for 6–8 h at 37 °C. Fluorescence was then measured with the Biotek Synergy 2 Plate Reader using excitation at 530 nm and emission at 590 nm. The percentage of cell death was then calculated using the following formula (Equation (3)):
%   Cellular   Death   =   100   ×   ODt     OD 100 OD 0     OD 100
where ODt = the fluorescence value of the test compound, OD100 = the fluorescence value for the culture medium, and OD0 = the fluorescence value for Triton 0.12%.

4. Conclusions

The marine fungus Amesia nigricolor MUT6601 has proven to be a promising source of bioactive compounds. In our study, a new bicyclic polyketide, amesilide (1), along with the previously reported metabolites, chamisides A (2), B (3), and E (4), chaetoconvosins B (5) and C (6), and chaetochromins A (7) and B (8), were isolated from the marine fungus Amesia nigricolor MUT6601. Notably, chaetochromins A (7) and B (8) exhibited significant antibacterial activity against Staphylococcus aureus, with minimal cytotoxicity, suggesting their potential for further development as therapeutic agents. These results highlight the potential of marine fungi as prolific sources of novel bioactive compounds, and lay the groundwork for future pharmacological development and optimization.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30153169/s1, Figure S1. HRESIMS of 1; Figure S2. IR spectrum of 1; Figure S3. 1H NMR (400 MHz) spectrum of 1 in CDCl3; Figure S4. 13C NMR (100 MHz) spectrum of 1 in CDCl3; Figure S5. 1H-1H COSY spectrum of 1 in CDCl3; Figure S6. 1H-13C HSQC spectrum of 1 in CDCl3; Figure S7. 1H-13C HMBC spectrum of 1 in CDCl3; Figure S8. 1H-1H NOESY spectrum of 1 in CDCl3; Figure S9. HRESIMS of 2; Figure S10. 1H NMR (400 MHz) spectrum of 2 in CDCl3; Figure S11. 13C NMR (100 MHz) spectrum of 2 in CDCl3; Figure S12. 1H-1H COSY spectrum of 2 in CDCl3; Figure S13. 1H-13C HSQC spectrum of 2 in CDCl3; Figure S14. 1H-13C HMBC spectrum of 2 in CDCl3; Figure S15. 1H-1H NOESY spectrum of 2 in CDCl3; Figure S16. HRESIMS of 3; Figure S17. 1H NMR (400 MHz) spectrum of 3 in CD3OD; Figure S18. 13C NMR (100 MHz) spectrum of 3 in CD3OD; Figure S19. 1H-1H COSY spectrum of 3 in CD3OD; Figure S20. 1H-13C HSQC spectrum of 3 in CD3OD; Figure S21. 1H-13C HMBC spectrum of 3 in CD3OD; Figure S22. 1H-1H NOESY spectrum of 3 in CD3OD; Figure S23. HRESIMS of 4; Figure S24. 1H NMR (400 MHz) spectrum of 4 in CDCl3; Figure S25. 13C NMR (100 MHz) spectrum of 4 in CDCl3; Figure S26. 1H-1H COSY spectrum of 4 in CDCl3; Figure S27. 1H-13C HSQC spectrum of 4 in CDCl3; Figure S28. 1H-13C HMBC spectrum of 4 in CDCl3; Figure S29. 1H-1H NOESY spectrum of 4 in CDCl3; Figure S30. HRESIMS of 5; Figure S31. 1H NMR (400 MHz) spectrum of 5 in CDCl3; Figure S32. 13C NMR (100 MHz) spectrum of 5 in CDCl3; Figure S33. 1H-1H COSY spectrum of 5 in CDCl3; Figure S34. 1H-13C HSQC spectrum of 5 in CDCl3; Figure S35. 1H-13C HMBC spectrum of 5 in CDCl3; Figure S36. 1H-1H NOESY spectrum of 5 in CDCl3; Figure S37. HRESIMS of 6; Figure S38. 1H NMR (400 MHz) spectrum of 6 in CDCl3; Figure S39. 13C NMR (100 MHz) spectrum of 6 in CDCl3; Figure S40. 1H-1H COSY spectrum of 6 in CDCl3; Figure S41. 1H-13C HSQC spectrum of 6 in CDCl3; Figure S42. 1H-13C HMBC spectrum of 6 in CDCl3; Figure S43. 1H-1H NOESY spectrum of 6 in CDCl3; Figure S44. HRESIMS of 7; Figure S45. 1H NMR (400 MHz) spectrum of 7 in CDCl3; Figure S46. 13C NMR (100 MHz) spectrum of 7 in CDCl3; Figure S47. 1H-1H COSY spectrum of 7 in CDCl3; Figure S48. 1H-13C HSQC spectrum of 7 in CDCl3; Figure S49. 1H-13C HMBC spectrum of 7 in CDCl3; Figure S50. 1H-1H NOESY spectrum of 6 in CDCl3; Figure S51. HRESIMS of 8; Figure S52. 1H NMR (400 MHz) spectrum of 8 in CDCl3; Figure S53. 13C NMR (100 MHz) spectrum of 8 in CDCl3; Figure S54. 1H-13C HSQC spectrum of 8 in CDCl3; Figure S55. 1H-13C HMBC spectrum of 8 in CDCl3; Figure S56. 1H-1H COSY spectrum of 8 in CDCl3; Figure S57. 1H-1H NOESY spectrum of 8 in CDCl3; Table S1: Cytotoxic activities of compounds 78 against RBC, THP-1, and PBMC cells. Table S2: Gibbs free energy and Boltzmann population of conformers of isomer 1a—8R, 9R, 10R, 12S, and 14S; Table S3. Gibbs free energy and Boltzmann population of conformers of isomer 1b—8S, 9S, 10S, 12R, and 14R; Table S4. Coordinates (Ångstroms) for conformer 1a-1; Table S5. Coordinates (Ångstroms) for conformer 1a-2; Table S6. Coordinates (Ångstroms) for conformer 1a-3; Table S7. Coordinates (Ångstroms) for conformer 1a-4; Table S8. Coordinates (Ångstroms) for conformer 1a-5; Table S9. Coordinates (Ångstroms) for conformer 1a-6; Table S10. Coordinates (Ångstroms) for conformer 1a-7; Table S11. Coordinates (Ångstroms) for conformer 1a-8; Table S12. Coordinates (Ångstroms) for conformer 1a-17; Table S13. Coordinates (Ångstroms) for conformer 1a-37; Table S14. Coordinates (Ångstroms) for conformer 1a-39; Table S15. Coordinates (Ångstroms) for conformer 1b-1; Table S16. Coordinates (Ångstroms) for conformer 1b-2; Table S17. Coordinates (Ångstroms) for conformer 1b-3; Table S18. Coordinates (Ångstroms) for conformer 1b-4; Table S19. Coordinates (Ångstroms) for conformer 1b-5; Table S20. Coordinates (Ångstroms) for conformer 1b-6; Table S21. Coordinates (Ångstroms) for conformer 1b-7; Table S22. Coordinates (Ångstroms) for conformer 1b-8; Table S23. Coordinates (Ångstroms) for conformer 1b-17; Table S24. Coordinates (Ångstroms) for conformer 1b-25; Table S25. Coordinates (Ångstroms) for conformer 1b-39.

Author Contributions

Conceptualization, L.B., G.C.V. and M.M.; Methodology, G.N.P., M.F.F., J.A.G.-S., P.M., F.M.A.-L. and M.M.; Validation, L.B., G.C.V. and M.M.; Investigations, G.N.P., M.F.F., J.A.G.-S., P.M. and M.M.; Data analysis, G.N.P., J.A.G.-S., P.M. and M.M.; Original draft preparation, G.N.P., L.B., M.F.F., J.A.G.-S., G.C.V. and M.M.; Manuscript final revision, M.M. All authors have read and agreed to the published version of the manuscript.

Funding

G.N. Pham has received funding from the European Union’s Horizon 2020 research and innovation program under the Marie Curie grant agreement n°847581 (COFUND BoostUrCAreer program), from the Région SUD Provence-Alpes-Côte d’Azur and IDEX UCAjedi. Part of the work was also funded by the European Commission Next Generation EU, Project SUS-MIRRI.IT, “Strengthening the MIRRI Italian Research Infrastructure for Sustainable Bioscience and Bioeconomy”, code n. IR0000005. We thank the Canceropôle Provence-Alpes-Côte d’Azur and the Provence-Alpes-Côte d’Azur Region for the financial support provided to the MetaboCell and MetaboPure projects.

Institutional Review Board Statement

Ethical review and approval were not required for this study, as no experiments were conducted directly on human subjects. Human blood from healthy donors was obtained from the Etablissement Français du Sang (EFS, agreement N13002), in compliance with French regulations and institutional guidelines.

Informed Consent Statement

Not applicable. The study did not involve human participants from whom informed consent would be required. Human blood samples were obtained from the Etablissement Français du Sang (EFS, agreement N13002) in accordance with ethical and regulatory requirements.

Data Availability Statement

The original data presented in the study are included in the article/Supplementary Materials; further inquiries can be directed to the corresponding author.

Acknowledgments

We thank Marc Gaysinski and Lionel Massi for recording the NMR and HRMS spectra, respectively. The authors are grateful for the support of the Microbial Resource Research Infrastructure—Italian Joint Research Unit (MIRRI-IT) for the technical and scientific support. Part of this work was funded by the European Commission—NextGenerationEU, Project SUS-MIRRI.IT, “Strengthening the MIRRI Italian Research Infrastructure for Sustainable Bioscience and Bioeconomy”, code n. IR0000005.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shin, H.J. Natural Products from Marine Fungi. Mar. Drugs 2020, 18, 230. [Google Scholar] [CrossRef] [PubMed]
  2. Gonçalves, M.F.; Esteves, A.C.; Alves, A. Marine fungi: Opportunities and challenges. Encyclopedia 2022, 2, 559–577. [Google Scholar] [CrossRef]
  3. Tian, Y.; Li, Y. A Review on Bioactive Compounds from Marine-Derived Chaetomium Species. J. Microbiol. Biotechnol. 2022, 32, 541–550. [Google Scholar] [CrossRef] [PubMed]
  4. Dwibedi, V.; Rath, S.K.; Jain, S.; Martinez–Argueta, N.; Prakash, R.; Saxena, S.; Rios-Solis, L. Key insights into secondary metabolites from various Chaetomium species. Appl. Microbiol. Biotechnol. 2023, 107, 1077–1093. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, Q.; Li, H.Q.; Zong, S.C.; Gao, J.M.; Zhang, A.L. Chemical and bioactive diversities of the genus Chaetomium secondary metabolites. Mini-Rev. Med. Chem. 2012, 12, 127–148. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, X.W.; Houbraken, J.; Groenewald, J.Z.; Meijer, M.; Andersen, B.; Nielsen, K.F.; Crous, P.W.; Samson, R.A. Diversity and taxonomy of Chaetomium and Chaetomium-like fungi from indoor environments. Stud. Mycol. 2016, 84, 145–224. [Google Scholar] [CrossRef] [PubMed]
  7. Kim, M.J.; Kim, D.C.; Kwon, J.; Ryu, S.M.; Kwon, H.; Guo, Y.; Hong, S.B.; Kim, Y.C.; Oh, H.; Lee, D. Anti-inflammatory Metabolites from Chaetomium nigricolor. J. Nat. Prod. 2020, 83, 881–887. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, H.H.; Li, G.; Qiao, Y.N.; Sun, Y.; Peng, X.P.; Lou, H.X. Chamiside A, a Cytochalasan with a Tricyclic Core Skeleton from the Endophytic Fungus Chaetomium nigricolor F5. Org. Lett. 2019, 21, 3319–3322. [Google Scholar] [CrossRef] [PubMed]
  9. Gu, C.C.; Zeng, J.; Peng, X.P.; Sun, Y.J.; Yuan, S.Z.; Wang, X.N.; Zhang, R.S.; Lou, H.X.; Li, G. Cytochalasans with Inhibitory Activity against NPC1L1 from the Endophytic Fungus Chaetomium nigricolor F5. J. Org. Chem. 2023, 88, 3185–3192. [Google Scholar] [CrossRef] [PubMed]
  10. Dhayanithy, G.; Subban, K.; Chelliah, J. Diversity and biological activities of endophytic fungi associated with Catharanthus roseus. BMC Microbiol. 2019, 19, 22. [Google Scholar] [CrossRef] [PubMed]
  11. Stierle, D.B.; Andrea, A.S.; Barbie, K.G. Isolation of two highly methylated polyketide derivatives from a yew-associated Penicillium species. J. Nat. Prod. 1999, 62, 1147–1150. [Google Scholar] [CrossRef] [PubMed]
  12. Noumeur, S.R.; Helaly, S.E.; Jansen, R.; Gereke, M.; Stradal, T.E.; Harzallah, D.; Stadler, M. Preussilides A–F, bicyclic polyketides from the endophytic fungus Preussia similis with antiproliferative activity. J. Nat. Prod. 2020, 80, 1531–1540. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, G.B.; Li, L.M.; Yang, T.; Zhang, G.L.; Li, G.Y. Chaetoconvosins A and B, alkaloids with new skeleton from fungus Chaetomium convolutum. Org. Lett. 2012, 14, 6052–6055. [Google Scholar] [CrossRef] [PubMed]
  14. Koyama, K.; Natori, S. Chaetochromins B, C and D, bis(naphtho–gamma–pyrone) derivatives from Chaetomium gracile. Chem. Pharm. Bull. 1987, 35, 578–584. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, G.-B.; Yang, T.; Bao, J.-K.; Fang, D.-M.; Li, G.-Y. Isochaetochromin A2, a new bis(naphthodihydropyran-4-one) with antimicrobial and immunological activities from fungus Chaetomium microcephalum. Arch. Pharm. Res. 2014, 37, 575–579. [Google Scholar] [CrossRef] [PubMed]
  16. Florio Furno, M.; Poli, A.; Ferrero, D.; Tardelli, F.; Manzini, C.; Oliva, M.; Pretti, C.; Campani, T.; Casini, S.; Fossi, M.C.; et al. The Culturable Mycobiota of Sediments and Associated Microplastics: From a Harbor to a Marine Protected Area, a Comparative Study. J. Fungi 2022, 8, 927. [Google Scholar] [CrossRef] [PubMed]
  17. Yue, Y.; Yu, H.; Li, R.; Hu, L.; Liu, S.; Xing, R.; Li, P. Isolation and identification of antimicrobial metabolites from sea anemone-derived fungus Emericella sp. SMA01. J. Oceanol. Limnol. 2021, 39, 1010–1019. [Google Scholar] [CrossRef]
  18. Spina, F.; Spini, G.; Poli, A.; Romagnolo, A.; Zanellati, A.; Bentivegna, N.; El-Azhari, N.; Regnier, T.; Blieux, A.; Echairi, A.; et al. Screening of Anionic Biosurfactants Production among Fungi and Bacteria. Chem. Eng. Trans. 2018, 64, 493–498. [Google Scholar]
  19. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16; Revision A.03; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  20. Mulvey, M.A.; Schilling, J.D.; Hultgren, S.J. Establishment of a persistent Escherichia coli reservoir during the acute phase of a bladder infection. Infect. Immun. 2001, 69, 4572–4579. [Google Scholar] [CrossRef] [PubMed]
  21. Boyer, L.; Doye, A.; Rolando, M.; Flatau, G.; Munro, P.; Gounon, P.; Clément, R.; Pulcini, C.; Popoff, M.R.; Mettouchi, A. Induction of transient macroapertures in endothelial cells through RhoA inhibition by Staphylococcus aureus factors. J. Cell Biol. 2006, 173, 809–817. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of compounds 18 isolated from the fungus A. nigricolor MUT6601.
Figure 1. Structures of compounds 18 isolated from the fungus A. nigricolor MUT6601.
Molecules 30 03169 g001
Figure 2. Key COSY, HMBC, and NOESY correlations of amesilide (1).
Figure 2. Key COSY, HMBC, and NOESY correlations of amesilide (1).
Molecules 30 03169 g002
Figure 3. Experimental CD spectrum for amesilide (1) in CH3CN and calculated ECD spectra for 1a (8R, 9R, 10R, 12S, and 14S) and 1b (8S, 9S, 10S, 12R, and 14R).
Figure 3. Experimental CD spectrum for amesilide (1) in CH3CN and calculated ECD spectra for 1a (8R, 9R, 10R, 12S, and 14S) and 1b (8S, 9S, 10S, 12R, and 14R).
Molecules 30 03169 g003
Table 1. 1H (400 MHz) and 13C NMR (100 MHz) data (in CDCl3) of amesilide (1) and its calculated chemical shifts.
Table 1. 1H (400 MHz) and 13C NMR (100 MHz) data (in CDCl3) of amesilide (1) and its calculated chemical shifts.
Position1Calculated Shifts
1H [δ, mult. (J in Hz)]13C (δ)1H (δ)13C (δ)
1-172.0-172.6
25.81 d (15.5)115.65.82118.5
37.37 d (15.5)152.97.68161.2
4-131.9-140.4
56.27 s144.76.68154.8
6-135.4-144.9
75.40 d (9.0)134.05.73144.0
85.15 dd (9.0, 2.6)37.56.0541.9
91.81 br d (11.8)50.91.8356.8
102.28 m44.52.2749.0
11-213.8-223.4
122.42 dq (12.1, 6.1)41.22.4046.0
132.10 ddd (13.5, 5.1, 2.2);
1.60 td (13.5, 5.1)
41.32.09
1.59
44.2
142.83 br s33.52.9138.9
156.04 s139.36.64153.1
16-133.4-142.5
17-151.1-163.5
186.27 s122.96.47125.3
19-198.8-206.2
202.25 s32.72.2736.1
211.90 br s13.91.8315.5
222.00 br s16.82.0719.7
231.07 d (6.4)11.51.0013.5
240.99 d (6.4)14.60.8916.5
251.96 br s20.12.0122.4
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

Pham, G.N.; Florio Furno, M.; Garcia-Sanchez, J.A.; Munro, P.; Abdoul-Latif, F.M.; Boyer, L.; Varese, G.C.; Mehiri, M. Amesilide, a New Bicyclic Polyketide from the Marine Fungus Amesia nigricolor MUT6601. Molecules 2025, 30, 3169. https://doi.org/10.3390/molecules30153169

AMA Style

Pham GN, Florio Furno M, Garcia-Sanchez JA, Munro P, Abdoul-Latif FM, Boyer L, Varese GC, Mehiri M. Amesilide, a New Bicyclic Polyketide from the Marine Fungus Amesia nigricolor MUT6601. Molecules. 2025; 30(15):3169. https://doi.org/10.3390/molecules30153169

Chicago/Turabian Style

Pham, Giang Nam, Matteo Florio Furno, Juan A. Garcia-Sanchez, Patrick Munro, Fatouma Mohamed Abdoul-Latif, Laurent Boyer, Giovanna Cristina Varese, and Mohamed Mehiri. 2025. "Amesilide, a New Bicyclic Polyketide from the Marine Fungus Amesia nigricolor MUT6601" Molecules 30, no. 15: 3169. https://doi.org/10.3390/molecules30153169

APA Style

Pham, G. N., Florio Furno, M., Garcia-Sanchez, J. A., Munro, P., Abdoul-Latif, F. M., Boyer, L., Varese, G. C., & Mehiri, M. (2025). Amesilide, a New Bicyclic Polyketide from the Marine Fungus Amesia nigricolor MUT6601. Molecules, 30(15), 3169. https://doi.org/10.3390/molecules30153169

Article Metrics

Back to TopTop