Next Article in Journal
Light Spectra, a Promising Tool to Modulate Ulva lacinulata Productivity and Composition
Previous Article in Journal
An Overview of Secondary Metabolites from Soft Corals of the Genus Capnella over the Five Decades: Chemical Structures, Pharmacological Activities, NMR Data, and Chemical Synthesis
Previous Article in Special Issue
Novel Metabolites from the Marine-Derived Fungus Peniophora sp. SCSIO41203 Show Promising In Vitro Antitumor Activity as Methuosis Inducers in PC-3 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Sesquiterpenoids from the Mangrove-Derived Fungus Talaromyces sp. as Modulators of Nuclear Receptors

1
NMPA Key Laboratory for Research and Evaluation of Drug Metabolism, Guangdong Provincial Key Laboratory of New Drug Screening, School of Pharmaceutical Sciences, Southern Medical University, Guangzhou 510515, China
2
CAS Key Laboratory of Tropical Marine Bio-Resources and Ecology/Guangdong Key Laboratory of Marine Materia Medica, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou 510301, China
3
University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Mar. Drugs 2024, 22(9), 403; https://doi.org/10.3390/md22090403
Submission received: 29 July 2024 / Revised: 16 August 2024 / Accepted: 27 August 2024 / Published: 3 September 2024

Abstract

:
Four new sesquiterpenoids, talaroterpenes A–D (14), were isolated from the mangrove-derived fungus Talaromyces sp. SCSIO 41412. The structures of compounds 14 were elucidated through comprehensive NMR and MS spectroscopic analyses. The absolute configurations of 14 were assigned based on single-crystal X-ray diffraction and calculated electronic circular dichroism analysis. Talaroterpenes A–D (14) were evaluated with their regulatory activities on nuclear receptors in HepG2 cells. Under the concentrations of 200 μM, 1, 3 and 4 exhibited varying degrees of activation on ABCA1 and PPARα, while 4 showed the strongest activities. Furthermore, 4 induced significant alterations in the expression of downstream target genes CLOCK and BMAL1 of RORα, and the in silico molecular docking analysis supported the direct binding interactions of 4 with RORα protein. This study revealed that talaroterpene D (4) was a new potential non-toxic modulator of nuclear receptors.

1. Introduction

Nuclear receptors (NRs) are an important class of drug targets because they are key regulators of many cardiovascular, metabolic and inflammatory diseases [1]. PPARs (peroxisome proliferator activated receptors), FXRs (farnesoid X receptors, FXRs), LXRs (liver X receptors) and RORs (retinal acid receptor-related orphan receptors) are important members of the nuclear receptor family, playing critical roles in regulating fat synthesis, cholesterol metabolism, fatty acid oxidation, inflammatory response and so on. In recent years, considerable research efforts have been devoted to the development of agonists or inhibitors targeting those receptors. Our previous review summarized 261 natural LXR/FXR modulators, originating from terrestrial plants and microorganisms, marine organisms and marine-derived microorganisms, in the recent two decades (2000–2020) [2]. Obeticolic acid is an FXR agonist that inhibits downstream CYP7A1 expression [3], suppresses bile acid synthesis, promotes bile secretion and so reduces intracellular bile acid concentration. It has been approved by the FDA for the treatment of primary biliary cirrhosis. Saroglitazar, a new PPAR agonist [4], which shows an evident effect on fat metabolism and insulin resistance, has been approved in India for two indications, including diabetes dyslipidemia and type 2 diabetes that cannot be controlled by statins.
Marine microorganisms produce metabolites with unique chemical structures that could target specific cellular physiological and biochemical processes, thus exhibiting significant biological activity and medicinal value [5,6,7]. In our previous studies, we discovered several LXRα agonists from marine microorganism-derived natural products [8,9,10,11], and some of them could promote lipid efflux by upregulating ABCA1 expression and reducing lipid synthesis, which could be considered as lead compounds for hypercholesterolemia [12].
Mangrove sediment-derived microbes, as a rich reservoir of natural product diversity, could be utilized in the research and development of new drugs [13], and many lead compounds have been obtained through our extensive research [14,15]. In our ongoing pursuit of pharmaceutically active molecules, especially natural modulators of nuclear receptors, regulatory activities on nuclear receptors were tested after the natural products were obtained from the mangrove sediment-derived microbes.
In this study, four new sesquiterpenoids, talaroterpenes A–D (14), were isolated from a mangrove sediment-derived fungus Talaromyces sp. SCSIO 41412 (Figure 1). Herein, the specifics of the isolation, structural elucidation and bioactive assessments of isolated compounds are reported.

2. Results and Discussion

2.1. Structural Determination

Compound 1 was obtained as colorless needles. Its molecular formula was determined to be C15H24O4 based on the high-resolution electrospray ionization mass spectroscopy (HRESIMS) data, corresponding to 4 degrees of unsaturation. Analysis of the 1H NMR data (Table 1) revealed the presence of one olefinic proton (δH 5.31 (1H, d, J = 5.1 Hz, H-9)), five sp3 methines with two of them being attached to oxygen atoms (δH 3.03 (1H, overlapped, H-2), 3.03 (1H, overlapped, H-3), 1.00 (1H, m, H-4), 1.74 (1H, dd, J = 7.0, 2.5 Hz, H-7), 2.15 (1H, overlapped, H-11)), three methylenes (δH 2.15 (1H, overlapped, H-1a), 2.06 (1H, overlapped, H-1b), 1.61 (1H, overlapped, H-6a), 0.89 (1H, m, H-6b), 1.61 (1H, overlapped, H-8a), 1.92 (1H, m, H-8b)) and three methyls (δH 1.01 (3H, d, J = 7.0 Hz, H3-13), 0.88 (3H, s, H3-14), 0.86 (3H, d, J = 6.7 Hz, H3-15)). In addition to the aforementioned 12 proton linked carbon signals, the 13C NMR and HSQC data also showed three additional carbon signals, including one carboxylic carbonyl (δC 177.0 (C-12)), one olefinic carbon (δC 141.0 (C-10)) and one quaternary carbon (δC 38.0 (C-5)). The NMR data were similar to those of the sesquiterpenoid, artefreynic acid A [16]. Combined with the two spin systems H2-1/H-2/H-3/H-4/H3-15 and H2-6/H-7/H2-8/H-9 observed in the 1H-1H COSY spectrum, as well as the HMBC correlations of H2-1/C-5, C-9, C-10 and H3-14/C-4, C-5, C-6 (Figure 2A), 1 was shown to be a sesquiterpenoid derivative with a 6/6 bicyclic skeleton, with the two methyl groups attached to C-4 and C-5, respectively. The HMBC correlation of H-7, H3-13/C-12 and 1H-1H COSY correlation of H-7/H-11/H3-13 confirmed the presence of a branched carboxylic acid moiety attached at C-7. Considering the chemical shifts of CH-2 (δH/C 3.03/74.7) and CH-3 (δH/C 3.03/75.5) as well as the molecular formula, it was deduced that the two hydroxyl groups were attached to C-2 and C-3, respectively. Based on these data, the planar structure of 1 was established as shown in Figure 1. The relative configuration of the rings in 1 was presumed to be rel-(2S, 3S, 4R, 5R, 7S) based on the NOESY correlations of H-2/H-4, H3-15/H-3, H3-15/H3-14 and H3-14/H-7 (Figure 2B). However, the peaks of H-2 and H-3 overlapped heavily and the configuration of the CH-11 side chain could not be unambiguously determined. Finally, the absolute configuration of 1 was established as 2S, 3S, 4R, 5R, 7S, 13R by X-ray diffraction analysis (Figure 3), and the compound was named talaroterpene A (1).
Compound 2 was obtained as colorless needles and its molecular formula was determined as C15H24O4 by HRESIMS data, indicating the same degree of unsaturation as 1. Analysis of the 1D NMR data (Table 1) showed that 2 was similar to 1. Careful analysis of the 2D NMR data of 2 and 1 (Figure 2) revealed that they were diastereomers with the same planar structure. Based on the NOESY correlations of H-2/H-4, H3-15/H-3, H3-15/H3-14 and H3-14/H-7, the relative configuration of 2 was presumed to be rel-(2S, 3S, 4R, 5R, 7S), which was the same as that of 1, indicating that the difference might lie in the configuration of the CH-11 side chain. The absolute configuration of 2 was finally established as 2S, 3S, 4R, 5R, 7S, 13S by X-ray diffraction analysis (Figure 3), and 2 was named talaroterpene B.
Compound 3 was obtained as colorless crystals. Its molecular formula was determined as C15H22O4 by HRESIMS data, indicating 5 degrees of unsaturation, which was one higher than that of 1/2. Analysis of the 1D NMR data of 3 (Table 2) revealed similarities with 1/2, except for the presence of two additional olefinic carbon–hydrogen signals (δH/C 6.04, 5.51/103.8; δC 146.2) and the absence of one methyl and one methine signal, suggesting that the -CH-CH3 moiety of the side chain had been oxidized to a double bond. The HMBC correlations of H2-13/C-7, 11, 12 (Figure 2A) of 3 confirmed this hypothesis. Based on the NOESY correlations of H-2/H-4, H3-15/H-3, H3-15/H3-14 and H3-14/H-7, the relative configuration of 3 was presumed to be rel-(2S, 3S, 4R, 5R, 7S) which was the same as that of 1/2, suggesting that they shared the same relative configuration. Finally, the absolute configuration of 3 was confirmed as 2S, 3S, 4R, 5R, 7S by X-ray diffraction analysis, and 3 was named talaroterpene C.
Compound 4 was obtained as a colorless oil, and its molecular formula was found to be two oxygen atoms less than that of 3. Analysis of the NMR data (Table 2) showed that 4 was similar to 3, except for the appreciable differences in the presence of CH2-2, CH-8, CH2-12 and a tetrahydrofuran ring in 4 versus the presence of CH-2, CH2-8 and CO-12 in 3. The 1H-1H COSY correlations of H2-1/H2-2/H-3/H-4/H3-15 indicated that 4 was one hydroxy group less at C-2 compared to 3. The HMBC correlations (Figure 2A) of H2-12/C-7, C-8, C-11 and H2-13/C-7, C-11, C-12 of 4 confirmed the formation of a tetrahydrofuran ring. The relative configuration of the rings in 4 was determined to be rel-(3R, 4R, 5R, 7S, 8S) based on the NOESY correlations (Figure 2B) of H3-15/H-3, H3-15/H3-14, H3-14/H-7, H3-14/H-8 and H-7/H-8. Based on the calculated electronic circular dichroism (ECD) for the 3R, 4R, 5R, 7S, 8S and 3S, 4S, 5S, 7R, 8R configuration (Figure 4), the experimental ECD spectrum of 4 showed a good match with the calculated one with the 3R, 4R, 5R, 7S, 8S configuration. Thus, 4 was named talaroterpene D.

2.2. Bioactivity Assay

For this work, compounds 14 were evaluated with their regulatory activities on nuclear receptors in HepG2 cells [10]. Firstly, CCK8 assay was taken to evaluate the toxicity against HepG2 cells (Figure 5). Compounds 2 and 4 are non-toxic to cells at concentrations of 200 μM and below. Compounds 1 and 3 showed weak cytotoxic effects on cells at concentration of 200 μM, with IC50 > 500 μM.
It was observed that talaroterpene D (4), with a non-toxic concentration of 200 μM, notably stimulated LXRα and its downstream target gene ABCA1, and significantly suppressed the expression of FXR and downstream CYP7A1. Compounds 1 and 3 (200 μM) could also stimulate expression of ABCA1, but there were no obvious activities with respect to LXRα, FXR and CYP7A1. Additionally, compounds 1, 3 and 4 (200 μM) exhibited varying degrees of activation on PPARα. Talaroterpene D (4) also showed activation on downstream target genes CPT1α and ACOX1. Furthermore, talaroterpene D (4) showed minimal impact on RORα itself; however, it induced significant alterations in the expression of downstream target genes CLOCK and BMAL1 of RORα, with a concentration of 200 μM, possibly due to direct binding interactions with RORα (Figure 6).
To further elucidate the binding mode of compounds 14 with RORα, a homology model of RORα (PDB code: 1N83) was selected for in silico molecular docking analysis [17]. The docking results demonstrated that compound 4 fits comfortably within the binding pocket of RORα (Figure 7A), yielding a binding score of −8.613. The 2D binding models of 4 with RORα (Figure 7B) revealed that the hydroxy group forms hydrogen bonds with the active-site residue TYR380. Similarly, the hydroxyl groups of compounds 13 form hydrogen bond interactions with TYR380 in the RORα (Figures S41–S43). The above results indicate that the presence of the hydroxyl group is beneficial for enhancing the binding interactions between the compounds and RORα. Compounds 13 all contain a carboxyl group, which under normal physiological pH conditions, typically exists in an ionized form, leading to a decrease in lipophilicity, thus hindering drug absorption and ultimately resulting in reduced biological activity, suggesting that the presence of the carboxyl group is not favorable for enhancing the RORα activity. Additionally, 3 experiences spatial conflict between the alkene group on the side chain and the RORα active site, leading to a reduction in its interactions with RORα. In contrast, compound 4, which also contains an alkene group, has a larger gap between the cyclic alkene and the RORα active site due to the cyclization, thereby eliminating the spatial conflict and enhancing its interactions with RORα. For future structural optimization and modification, it is promising to explore the possibility of cyclizing the carboxyl-containing side chains, which may further improve the biological activity of talaroterpenes.

3. Materials and Methods

3.1. General Experimental Procedures

Optical rotations were determined using an Anton Paar MPC 500 polarimeter (Anton, Graz, Austria). UV and IR spectra were recorded using a Shimadzu UV-2600 PC spectrometer (Shimadzu, Beijing, China) and an IR Affinity-1 spectrometer (Shimadzu), respectively. A Quantum-I Plus 500 Hz NMR spectrometer (Q-one Instrument Co., Ltd., Wuhan, China) operating at 500 MHz for 1H and 125 MHz for 13C was used. HRESIMS were acquired on a Bruker maXis Q-TOF mass spectrometer (Bruker BioSpin International AG, Fällanden, Switzerland). HPLC was performed on the Hitachi Primaide (Hitachi, Tokyo, Japan) with a DAD detector, using an ODS column (YMC-pack ODS-A, 10 × 250 mm, 5 μm). X-ray diffraction was performed on an XtalLAB PRO diffractometer (Rigaku, Akishima-shi, Japan) with Cu Kα radiation.

3.2. Fungal Material

A fungal strain identified as Talaromyces sp. SCSIO 41412 was isolated from a sediment sample obtained from the Gaoqiao Mangrove in Zhanjiang City, Guangdong Province, China, in August 2021. The taxonomic assignment of this fungus was based on analysis of the internally transcribed spacer (ITS) region of the ribosomal DNA (rDNA), and the ITS sequence has been deposited in GenBank under accession number PP001498. The fungal strain was stored on malt extract agar (15 g malt extract, 18 g agar, 10 g sea salt and 1 L water) at 4 °C, and deposited in the CAS Key Laboratory of Tropical Marine Bioresources and Ecology, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou, China.

3.3. Fermentation, Extraction and Isolation

The fungal strain Talaromyces sp. SCSIO 41412 was cultured in 200 mL of seed medium (15 g malt extract, 10 g sea salt and 1 L water) on a rotary shaker (180 rpm) at 28 °C for 3 days. This seed culture was then used to inoculate a large-scale fermentation, which was incubated statically at 26 °C for 28 days using a rice medium (200 g rice, 2% sea salt and 230 mL water) in 47 Erlenmeyer flasks. The entire fermented culture was extracted with ethyl acetate (EtOAc) three times, yielding a total extract of 326.1 g. The EtOAc extract was chromatographed over a silica gel column eluted with CH2Cl2/petroleum ether (0:1, 1:1, 1:0) and CH3OH/CH2Cl2 (1:99, 2:98, 3:97, 5:95, 10:90, 20:80, 50:50) to yield ten fractions (Frs.1–10). Fr.4 was subjected to ODS silica gel chromatography and eluted with CH3OH/H2O (5–100%) to yield 14 subfractions. Fr.4-3 was further separated by semi-preparative HPLC (50% CH3CN/H2O, 3 mL/min, YMC-pack ODS-A, 10 × 250 mm, 5 μm) to afford Fr.4-3-3, which was then purified by HPLC (68% CH3OH/H2O, 2.5 mL/min) to yield 4 (6.8 mg, tR = 15.5 min). Fr.6 was subjected to ODS silica gel chromatography and eluted with CH3OH/H2O (5–100%) to yield 10 subfractions. Fr.6-4 was separated by semi-preparative HPLC (33% CH3CN/H2O, 3.0 mL/min, YMC-pack ODS-A, 10 × 250 mm, 5 μm) to afford Fr.6-4-1, which was further separated by semi-preparative HPLC (50% CH3CN/H2O, 3.0 mL/min, YMC-pack ODS-A, 10 × 250 mm, 5 μm) to yield 2 (9.9 mg, tR = 16.0 min), 1 (119.8 mg, tR = 16.2 min) and 3 (49.1 mg, tR = 16.9 min).

3.4. Spectroscopic Data of Compounds

Talaroterpene A (1): colorless needles; [ α ]   D 25 +14.5 (c 0.1, CH3OH); ECD (0.3 mg/mL, CH3OH) λmaxε) 224 (−2.36); UV (CH3OH) λmax (log ε) 200 (3.88), 320 (2.73) nm; IR νmax 3398, 1697, 1558, 1541, 1456, 1047 cm−1; 1H and 13C NMR, Table 1; HRESIMS m/z 286.2014 [M + NH4]+ (calcd for C15H28NO4+, 286.2013).
Talaroterpene B (2): colorless needles; [ α ]   D 25 +18.8 (c 0.1, CH3OH); ECD (0.3 mg/mL, CH3OH) λmaxε) 202 (−2.35); UV (CH3OH) λmax (log ε) 200 (3.81) nm; IR νmax 3375, 2968, 2914, 1703, 1550, 1448, 1373, 1251, 1199, 1041, 1024 cm−1; 1H and 13C NMR, Table 1; HRESIMS m/z 286.2018 [M + NH4]+ (calcd for C15H28NO4+, 286.2013).
Talar terpene C (3): colorless block crystal; [ α ]   D 25 +23.5 (c 0.1, CH3OH); ECD (0.3 mg/mL, CH3OH) λmaxε) 224 (−2.94); UV (CH3OH) λmax (log ε) 200 (3.89) nm; IR νmax 3338, 1681, 1541, 1373, 1253, 1041 cm−1; 1H and 13C NMR, Table 2; HRESIMS m/z 265.1451 [M-H] (calcd for C15H21O4, 265.1445).
Talaroterpene D (4): colorless oil; [ α ]   D 25 +3.0 (c 0.1, CH3OH); ECD (0.3 mg/mL, CH3OH) λmaxε) 203 (−27.70); UV (CH3OH) λmax (log ε) 200 (3.90) nm; IR νmax 3392, 1653, 1373, 1026 cm−1; 1H and 13C NMR, Table 2; HRESIMS m/z 235.1694 [M + H]+ (calcd for C15H23O2+, 235.1693).

3.5. X-ray Crystallographic Analysis

The X-ray diffraction data for 13 were collected using an XtaLAB PRO diffractometer with Cu Kα radiation, with the crystals grown from methanol by slow evaporation. The crystal structures were solved using SHELXS97, expanded through difference Fourier techniques, and then refined by full-matrix least-squares methods. All non-hydrogen atoms were refined anisotropically, and the hydrogen atoms were fixed at calculated positions. The crystallographic data for 13 have been deposited in the Cambridge Crystallographic Data Centre.
Crystal Data for Talaroterpene A (1): C15H24O4, M = 268.34, monoclinic, space group P21, a = 10.23266 (8) Å, b = 6.72249 (4) Å, c = 11.56363(9) Å, β = 115.5824(10)°, V = 717.467 (11) Å3, Z = 2, T = 100.00 (10) K, μ(Cu Kα) = 0.719 mm−1, Dcalc = 1.242 g/cm3, 13,788 reflections measured (8.478° ≤ 2Θ ≤ 148.13°), 2843 unique (Rint = 0.0420, Rsigma = 0.0236) which were used in all calculations. The final R1 was 0.0280 (I > 2σ(I)) and wR2 was 0.0724. The Flack parameter was 0.06 (7) (CCDC 2372513).
Crystal Data for Talaroterpene B (2): C15H24O4M = 268.34, monoclinic, space group P21, a = 9.09200 (10) Å, b = 7.93080 (10) Å, c = 9.95020 (10) Å, β = 96.4630 (10)°, V = 712.918 (14) Å3, Z = 2, T = 100.00 (10) K, μ (Cu Kα) = 0.724 mm−1, Dcalc = 1.250 g/cm3, 10,853 reflections measured (8.944° ≤ 2Θ ≤ 148.192°), 2834 unique (Rint = 0.0223, Rsigma = 0.0158) which were used in all calculations. The final R1 was 0.0289 (I > 2σ(I)) and wR2 was 0.0787. The Flack parameter was 0.09 (5) (CDCC 2372514).
Crystal Data for Talaroterpene C (3): C15H22O4, M = 266.32, orthorhombic, space group P212121, a = 7.24180 (10) Å, b = 11.42500 (10) Å, c = 16.6083 (2) Å, V = 1374.13 (3) Å3, Z = 4, T = 99.99 (10) K, μ(Cu Kα) = 0.751 mm−1, Dcalc = 1.287 g/cm3, 12,963 reflections measured (9.396° ≤ 2Θ ≤ 148.8°), 2750 unique (Rint = 0.0569, Rsigma = 0.0327) which were used in all calculations. The final R1 was 0.0433 (I > 2σ(I)) and wR2 was 0.1198. The Flack parameter was 0.09 (13) (CDCC 2372529).

3.6. ECD Computation Section

Compound 4 was subjected to conformational searching in Spartan’14 (v1.1.4, Wavefunction, Irvine, CA, USA) using the Molecular Merck force field. The stable conformers of 99% (the relative energy within 2 kcal/mol) were then optimized in methanol solvent at the B3LYP/6-31G (d) level of theory using Gaussian 09 (D.01, Pittsburgh, PA, USA). The optimized low-energy conformations were further analyzed using TDDFT with a polarizable continuum model at the B3LYP/6-311G (d, p) level [18]. The calculated ECD spectra were generated from GaussView 6.0 and Origin 2021, with a half-bandwidth of 0.3 eV, wavelength-corrected by the calculated UV curve, and Boltzmann-weighted to obtain the final calculated ECD spectra.

3.7. Cell Culture

HepG2 cells were purchased from the Shanghai Cell Bank, Chinese Academy of Sciences, and cultured in Dulbecco’s modified Eagle’s medium (DMEM, Gibco, New York, NY, USA) supplemented with 10% fetal bovine serum (FBS, ExCell Bio, Suzhou, China) at 37 °C under 5% CO2. Mother solutions of 20 mM 14 were dissolved in dimethyl sulfoxide (DMSO, Tianjin Fuyu Chemical and Industry Factory, Tianjin, China) and then diluted to 200 μM with the cell-culture medium. The control group was treated with a medium containing 1‰ DMSO.

3.8. Cytotoxic Bioassay

Cell viability was assessed using the CCK-8 assay (Dojindo, Kumamoto, Japan). Initially, cells were plated at a density of 3000 cells/well in 96-well plates and exposed to varying concentrations of compounds or a solvent control. Following a 24 h incubation period, CCK-8 reagent was introduced, and the absorbance of triplicate samples was measured at 450 nm using an Envision 2104 multilabel reader (PerkinElmer, Waltham, MA, USA). Dose-response curves were generated to determine the IC50 values with Prism 5.0 software (GraphPad, San Diego, CA, USA).

3.9. RT-qPCR

Total RNA was extracted using the RNAprep Kit (RE-03014, FOREGENE, Chengdu, China) and reverse transcribed to cDNA using RT Master Mix (RR037A, Takara, Shiga, Japan). RT-qPCR was performed using SYBR Green Master Mix (A6002, Promega, Madison, WI, USA). The data were normalized to the housekeeping gene GAPDH. Primers used are listed in Table S1.

3.10. Statistical Analysis

Prism 9 software (GraphPad, San Diego, CA, USA) was used to perform statistical tests. Groups (n = 6) were compared by one-way analysis of variance (ANOVA). Nonparametric data were log-transformed for statistical analysis, or when this failed to normalize the data, Mann–Whitney tests were used. p < 0.05 was significant (* p < 0.05, ** p < 0.01, *** p < 0.001, and **** p < 0.0001 compared to the control group).

4. Conclusions

In summary, this study identified four new sesquiterpenoid compounds, talaroterpenes A–D (14), isolated from the mangrove-derived fungus Talaromyces sp. SCSIO 41412. Through detailed structural characterization, the absolute configurations of these compounds were elucidated. Those new sesquiterpenoids were evaluated with their regulatory activities on nuclear receptors in HepG2 cells. Under the non-toxic concentration of 200 μM, 1, 3 and 4 exhibited varying degrees of activation on ABCA1 (downstream target of LXRα) and PPARα, while 4 showed the strongest activities. Furthermore, 4 induced significant alterations in the expression of downstream target genes CLOCK and BMAL1 of RORα, and the in silico molecular docking analysis supported the direct binding interactions of 4 with RORα protein. This study revealed that the sesquiterpenoid talaroterpene D (4), as a new potential non-toxic modulator of nuclear receptors, holds promise as a lead compound for the development of candidate drugs for metabolic or cardiovascular diseases.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/md22090403/s1, the NMR, HRESIMS, UV, IR, and ECD spectra of 14; Primers of RT-Qpcr; The docking results of 13 with RORα.

Author Contributions

Funding acquisition, L.T.; Investigation, T.G., J.C. and J.S.; Project administration, X.Z.; Resources, Y.L.; Supervision, X.Z. and L.T.; Writing—original draft, T.G., D.X. and J.C.; Writing—review & editing, L.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Special Project for Marine Economic Development of the Department of Natural Resources of Guangdong Province (GDNRC [2024]25).

Data Availability Statement

Data are contained within the article.

Acknowledgments

We are grateful to Z. Xiao, X. Zheng, A. Sun, Y. Zhang and X. Ma in the analytical facility at SCSIO for recording spectroscopic data.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yang, C.; Li, Q.; Li, Y. Targeting Nuclear Receptors with Marine Natural Products. Mar. Drugs 2014, 12, 601–635. [Google Scholar] [CrossRef] [PubMed]
  2. She, J.; Gu, T.; Pang, X.; Liu, Y.; Tang, L.; Zhou, X. Natural Products Targeting Liver X Receptors or Farnesoid X Receptor. Front. Pharmacol. 2022, 12, 772435. [Google Scholar] [CrossRef] [PubMed]
  3. Sun, L.; Cai, J.; Gonzalez, F.J. The Role of Farnesoid X Receptor in Metabolic Diseases, and Gastrointestinal and Liver cancer. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 335–347. [Google Scholar] [CrossRef] [PubMed]
  4. Siddiqui, M.S.; Parmar, D.; Sheikh, F.; Sarin, S.K.; Cisneros, L.; Gawrieh, S.; Momin, T.; Duseja, A.; Sanyal, A.J. Saroglitazar, a Dual PPAR α/γ Agonist, Improves Atherogenic Dyslipidemia in Patients with Non-cirrhotic Nonalcoholic Fatty Liver Disease: A Pooled Analysis. Clin. Gastroenterol. Hepatol. 2023, 21, 2597–2605. [Google Scholar] [CrossRef] [PubMed]
  5. Liang, J.; She, J.; Fu, J.; Wang, J.; Ye, Y.; Yang, B.; Liu, Y.; Zhou, X.; Tao, H. Advances in Natural Products from the Marine-Sponge-Associated Microorganisms with Antimicrobial Activity in the Last Decade. Mar. Drugs 2023, 21, 236. [Google Scholar] [CrossRef] [PubMed]
  6. Zhou, X. Mangrove Soil-derived Streptomyces: An Important Resource of Pharmaceutical Active Natural Products. J. Holist. Integr. Pharm. 2022, 3, 300–314. [Google Scholar] [CrossRef]
  7. Luo, Z.; Yin, F.; Wang, X.; Kong, L. Progress in Approved Drugs from Natural Product Resources. Chin. J. Nat. Med. 2024, 22, 195–211. [Google Scholar] [CrossRef] [PubMed]
  8. Liang, Z.; Chen, Y.; Gu, T.; She, J.; Dai, F.; Jiang, H.; Zhan, Z.; Li, K.; Liu, Y.; Zhou, X.; et al. LXR-Mediated Regulation of Marine-Derived Piericidins Aggravates High-Cholesterol Diet-Induced Cholesterol Metabolism Disorder in Mice. J. Med. Chem. 2021, 64, 9943–9959. [Google Scholar] [CrossRef] [PubMed]
  9. She, J.; Zhou, X. New Insights into the Antitumor Potential of Natural Piericidins. J. Holist. Integr. Pharm. 2021, 2, 153–162. [Google Scholar]
  10. Liang, Z.; Gu, T.; Wang, J.; She, J.; Ye, Y.; Cao, W.; Luo, X.; Xiao, J.; Liu, Y.; Tang, L.; et al. Chromene and Chromone Derivatives as Liver X Receptors Modulators from a Marine-Derived Pestalotiopsis neglecta Fungus. Bioorg. Chem. 2021, 112, 104927. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, J.; Liang, Z.; Li, K.; Yang, B.; Liu, Y.; Fang, W.; Tang, L.; Zhou, X. Ene-yne Hydroquinones from a Marine-derived Strain of the Fungus Pestalotiopsis neglecta with Effects on Liver X Receptor Alpha. J. Nat. Prod. 2020, 83, 1258–1264. [Google Scholar] [CrossRef] [PubMed]
  12. Fang, Y.; She, J.; Zhang, X.; Gu, T.; Xie, D.; Luo, X.; Yi, X.; Gao, C.; Liu, Y.; Zhang, C.; et al. Discovery of Anti-Hypercholesterolemia Agents Targeting LXRα from Marine Microorganism-Derived Natural Products. J. Nat. Prod. 2024, 87, 322–331. [Google Scholar] [CrossRef] [PubMed]
  13. Li, K.; Chen, S.; Pang, X.; Cai, J.; Zhang, X.; Liu, Y.; Zhu, Y.; Zhou, X. Natural Products from Mangrove Sediments-Derived Microbes: Structural Diversity, Bioactivities, Biosynthesis, and Total Synthesis. Eur. J. Med. Chem. 2022, 230, 114117. [Google Scholar] [CrossRef] [PubMed]
  14. Cai, J.; Gao, L.; Wang, Y.; Zheng, Y.; Lin, X.; Zhou, P.; Chen, C.; Liu, K.; Tang, L.; Liu, Y.; et al. Discovery of a Novel Anti-Osteoporotic Agent from Marine Fungus-Derived Structurally Diverse Sirenins. Eur. J. Med. Chem. 2023, 265, 116068. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, C.; Xiao, L.; Luo, X.; Cai, J.; Huang, L.; Tao, H.; Zhou, X.; Tan, Y.; Liu, Y. Identifying Marine-Derived Tanzawaic Acid Derivatives as Novel Inhibitors against Osteoclastogenesis and Osteoporosis via Downregulation of NF-kappaB and NFATc1 Activation. J. Med. Chem. 2024, 67, 2602–2618. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, C.; Wen, R.; Ma, X.L.; Zeng, K.W.; Xue, Y.; Zhang, P.M.; Zhao, M.B.; Jiang, Y.; Liu, G.Q.; Tu, P.F. Nitric Oxide Inhibitory Sesquiterpenoids and Its Dimers from Artemisia freyniana. J. Nat. Prod. 2018, 81, 866–878. [Google Scholar] [CrossRef] [PubMed]
  17. Slominski, A.T.; Kim, T.K.; Takeda, Y.; Janjetovic, Z.; Brozyna, A.A.; Skobowiat, C.; Wang, J.; Postlethwaite, A.; Li, W.; Tuckey, R.C.; et al. RORα and ROR γ are Expressed in Human Skin and Serve as Receptors for Endogenously Produced Noncalcemic 20-Hydroxy- and 20,23-Dihydroxyvitamin D. FASEB J. 2014, 28, 2775–2789. [Google Scholar] [CrossRef] [PubMed]
  18. Luo, X.; Lin, X.; Tao, H.; Wang, J.; Li, J.; Yang, B.; Zhou, X.; Liu, Y. Isochromophilones A-F, Cytotoxic Chloroazaphilones from the Marine Mangrove Endophytic Fungus Diaporthe sp. SCSIO 41011. J. Nat. Prod. 2018, 81, 934–941. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of compounds 14.
Figure 1. Structures of compounds 14.
Marinedrugs 22 00403 g001
Figure 2. Key 1H-1H COSY, HMBC (A), and NOESY (B) correlations of 14.
Figure 2. Key 1H-1H COSY, HMBC (A), and NOESY (B) correlations of 14.
Marinedrugs 22 00403 g002
Figure 3. Perspective ORTEP drawing of X-ray structures of 13.
Figure 3. Perspective ORTEP drawing of X-ray structures of 13.
Marinedrugs 22 00403 g003
Figure 4. Experimental and calculated ECD spectra of 4.
Figure 4. Experimental and calculated ECD spectra of 4.
Marinedrugs 22 00403 g004
Figure 5. The cytotoxic effects of 14 against HepG2 cells. ** p < 0.01; *** p < 0.001; **** p < 0.0001.
Figure 5. The cytotoxic effects of 14 against HepG2 cells. ** p < 0.01; *** p < 0.001; **** p < 0.0001.
Marinedrugs 22 00403 g005
Figure 6. The regulatory activities of 14 (200 μM) on nuclear receptors in HepG2 cells. The results were mean ± SEM (n = 3, * p < 0.05, ** p < 0.01, *** p < 0.001 and **** p < 0.0001 compared to the control group).
Figure 6. The regulatory activities of 14 (200 μM) on nuclear receptors in HepG2 cells. The results were mean ± SEM (n = 3, * p < 0.05, ** p < 0.01, *** p < 0.001 and **** p < 0.0001 compared to the control group).
Marinedrugs 22 00403 g006
Figure 7. The docking results of 4 with RORα. (A) Three-dimensional and (B) two-dimensional binding mode of 4 with RORα (PDB code: 1N83) predicted by in silico molecular docking.
Figure 7. The docking results of 4 with RORα. (A) Three-dimensional and (B) two-dimensional binding mode of 4 with RORα (PDB code: 1N83) predicted by in silico molecular docking.
Marinedrugs 22 00403 g007
Table 1. 1H (500 MHz) and 13C (125 MHz) NMR data of 1 and 2 in DMSO-d6.
Table 1. 1H (500 MHz) and 13C (125 MHz) NMR data of 1 and 2 in DMSO-d6.
Pos.1 2
δC TypeδH (J in Hz)δC TypeδH (J in Hz)
139.3, CH22.15, overlapped
2.06, m
39.3, CH22.16, overlapped
2.06, m
274.7, CH3.03, overlapped74.7, CH3.02, overlapped
375.5, CH3.03, overlapped75.5, CH3.02, overlapped
447.3, CH1.00, m47.4, CH0.98, m
538.0, C 37.9, C
642.3, CH21.61, overlapped
0.89, m
40.5, CH21.67, m
0.83, m
732.4, CH1.74, m32.7, CH1.72, overlapped
828.5, CH21.92, m
1.61, overlapped
29.7, CH21.89, m
1.72, overlapped
9120.1, CH5.31, d (5.1)120.2, CH5.30, d (5.0)
10141.0, C 140.9, C
1144.4, CH2.15, overlapped43.8, CH2.16, overlapped
12177.0, C 176.9, C
1313.9, CH31.01, d (7.0)13.9, CH31.03, d (6.9)
1419.1, CH30.88, s19.1, CH30.89, overlapped
1510.8, CH30.86, d (6.7)10.8, CH30.89, overlapped
Table 2. 1H (500 MHz) and 13C (125 MHz) NMR data of 3 and 4 in DMSO-d6.
Table 2. 1H (500 MHz) and 13C (125 MHz) NMR data of 3 and 4 in DMSO-d6.
Pos.34
δC TypeδH (J in Hz)δC TypeδH (J in Hz)
139.3, CH22.19, dd (13.3, 4.8)
2.11, d (10.3)
30.4, CH22.22, m
2.10, dt (14.0, 4.0)
274.8, CH3.04, overlapped36.0, CH21.88, m
1.13, m
375.5, CH3.04, overlapped69.6, CH3.29, m
447.3, CH1.05, overlapped49.7, CH0.96, m
538.5, C 38.3, C
643.4, CH21.70, d (12.5)
1.05, overlapped
40.1, CH21.65, dd (12.9, 5.0)
1.06, m
731.2, CH2.06, m38.2, CH2.59, m
831.1, CH22.64, m
1.81, m
74.6, CH3.97, m
9120.3, CH5.36, d (5.4)117.3, CH5.45, s
10140.9, C 149.1, C
11146.2, C 153.7, C
12168.4, C 69.6, CH24.35, d (13.4)
4.10, d (13.4)
13121.9, CH26.04, s
5.51, s
103.8, CH25.00, d (1.25)
4.90, s
1419.1, CH30.94, s17.7, CH30.90, s
1510.8, CH30.89, d (6.7)11.2, CH30.93, d (4.7)
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

Gu, T.; Cai, J.; Xie, D.; She, J.; Liu, Y.; Zhou, X.; Tang, L. New Sesquiterpenoids from the Mangrove-Derived Fungus Talaromyces sp. as Modulators of Nuclear Receptors. Mar. Drugs 2024, 22, 403. https://doi.org/10.3390/md22090403

AMA Style

Gu T, Cai J, Xie D, She J, Liu Y, Zhou X, Tang L. New Sesquiterpenoids from the Mangrove-Derived Fungus Talaromyces sp. as Modulators of Nuclear Receptors. Marine Drugs. 2024; 22(9):403. https://doi.org/10.3390/md22090403

Chicago/Turabian Style

Gu, Tanwei, Jian Cai, Danni Xie, Jianglian She, Yonghong Liu, Xuefeng Zhou, and Lan Tang. 2024. "New Sesquiterpenoids from the Mangrove-Derived Fungus Talaromyces sp. as Modulators of Nuclear Receptors" Marine Drugs 22, no. 9: 403. https://doi.org/10.3390/md22090403

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