Next Article in Journal
Lactobacillus paracasei HY7015 and Lycopus lucidus Turcz. Extract Promotes Human Dermal Papilla Cell Cytoprotective Effect and Hair Regrowth Rate in C57BL/6 Mice
Previous Article in Journal
Supramolecular Diversity, Theoretical Investigation and Antibacterial Activity of Cu, Co and Cd Complexes Based on the Tridentate N,N,O-Schiff Base Ligand Formed In Situ
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Meroterpenoid and Isocoumarins from the Fungus Talaromyces amestolkiae MST1-15 Collected from Coal Area

1
State Key Laboratory of Functions and Applications of Medicinal Plants, School of Pharmacy, Guizhou Medical University, Guian New District, Guiyang 550025, China
2
Engineering Research Center for the Development and Application of Ethnic Medicine and TCM, Ministry of Education & Guizhou Provincial Key Laboratory of Pharmaceutics, Guiyang 550004, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2022, 27(23), 8223; https://doi.org/10.3390/molecules27238223
Submission received: 30 September 2022 / Revised: 11 November 2022 / Accepted: 22 November 2022 / Published: 25 November 2022
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
Three new compounds including a meroterpenoid (1) and two isocoumarins (8 and 9), together with thirteen known compounds (2–7, 10–16) were isolated from the metabolites of Talaromyces amestolkiae MST1-15. Their structures were identified by a combination of spectroscopic analysis. The absolute configuration of compound 1 was elucidated on the basis of experimental and electronic circular dichroism calculation, and compounds 8 and 9 were determined by Mo2(OAc)4-induced circular dichroism experiments. Compounds 7–16 showed weak antibacterial activities against Stenotrophomonas maltophilia with MIC values ranging from 128 to 512 μg/mL (MICs of ceftriaxone sodium and levofloxacin were 128 and 0.25 μg/mL, respectively).

1. Introduction

The genus Talaromyces was initially established by Benjamin and over 71 species have been reported [1,2]. Secondary metabolites from Talaromyces are rich in structural diversity including anthraquinones, alkaloids, terpenes, and azaphilones, and had a wide range of bioactivities, such as MAO-inhibitory, nematicidal, cytotoxic, antiviral, and anti-inflammatory activities [3,4,5]. Talaromyces amestolkiae belongs to the genus Talaromyces, species of which are found to widely inhabit plants, soil, sponges, and foods [6]. This fungus is increasingly attracting attention for its ability to produce high levels of xylanases, cellulases, active compounds, and natural colorants with potential applications in the fields of industry, medicine, and food [7,8]. Recently, a series of different structural types of compounds including meroterpenoids, isocoumarins, and benzofurans were isolated as metabolites of T. amestolkiae for activity screening [5,6,9,10].
Extremophiles have been proven to be a promising source of bioactive compounds with diverse structures since their first discovery in the mid-20th century. Most fungi that inhabit extreme environments are categorized as Ascomycota covering a range of genera, including Acremonium, Alternaria, Aspergillus, Cladosporium, and Penicillium, and have been demonstrated to produce numerous worthwhile compounds with antimicrobial, anticancer, and/or antidiabetic activities [11,12,13]. Coal areas are deemed as multi-extreme environments with high metal contents and/or extreme pH, in which fungi survive through producing extremolytes, extremozymes, as well as small useful molecules to cope with the extreme environments [14]. In our ongoing research on the discovery of bioactive compounds from fungi, a chemical study of the secondary metabolites of T. amestolkiae MST1-15 collected from the Xingren coal area was carried out. Sixteen compounds (116, Figure S1) including new meroterpenoid (1) and isocoumarins (8 and 9) were isolated. Compounds 716 showed weak antibacterial activities against Stenotrophomonas maltophilia with MIC values ranging from 128 to 512 μg/mL. Here, the isolation, structure elucidation, and antibacterial activities of these compounds are described.

2. Results and Discussion

Structures Elucidation

Compound 1 was obtained as a white amorphous powder. The molecular formula was determined as C26H32O8 according to the HR ESIMS data with a [M−H] ion peak at m/z 471.2025 (calcd for 471.2024), implying 11 degrees of unsaturation. The 1H NMR spectrum of 1 exhibited five singlet methyl signals (δH 1.77, 1.57, 1.48, 1.33, and 1.14), a methoxyl peak (δH 3.80, s), and one doublet methyl peak (δH 1.42 (3H, d, J = 6.8 Hz)). With the aid of HSQC, 13C NMR data (Table 1) analysis revealed the presence of seven methyls (δC 53.2, 27.9, 27.4, 27.3, 26.5, 18.3, and 17.7), four methylenes (δC 49.1, 34.0, 31.6, and 26.7), three methines (98.5, 72.3 and 46.0), four sp3 quaternary carbons (δC 85.6, 63.2, 44.1, and 41.0), three ester carbonyls (δC 179.8, 169.4, and 168.2), one keto carbonyl (δC 209.4), four olefinic carbons (δC 137.0, 136.3, 132.8, and 126.9). The aforementioned data suggested that compound 1 was a meroterpenoid [6,15], and these spectroscopic features were very similar to those of berkeleyacetal A [16]; the major difference being the replacement of the olefinic methine (CH-14) by an additional olefinic quaternary carbon in 1, suggesting 1 to be an olefinic positional isomer of berkeleyacetal A. Detailed 2D NMR (1H-1H COSY, HSQC, HMBC, and NOESY) data analysis (Figure 1 and Figure 2) further confirmed this deduction and figured out the position of the olefinic bonds. In particular, a Δ3,15 double bond was established by the HMBC correlations from H-2 (δH 3.32) to C-1 (δC 169.4), C-3 (δC 126.9), and C-15 (δC 136.3), while a Δ4,5 double bond was revealed by the HMBC correlations from H3-25 (δH 1.77) to C-3, C-4 (δC 132.8), and C-5 (δC 137.0) (Figure 1). The NOESY correlations of H-22 (δH 3.69) with H3-19 (δH 1.14) and H3-24 (δH 1.33) indicated that CH3-19, H-22, and CH3-24 are co-facially oriented and was assigned α orientation. In addition, a strong NOESY correlation observed for H-22 and H-23 (δH 6.09) suggested the cis-fusion of the furan and pyran rings. These observations showed that both berkeleyacetal A and 1 shared the same stereochemistry at almost all stereocenters. Moreover, the NOESY correlation of H3-26 (δH 3.80) with H3-21 (δH 1.42) indicated that both CH3-26 and CH3-21 were also on the same side of the hexa-ring. However, due to the absence of NOESY correlations between H3-21 and H-23, H-22, or H3-19, the configurations of C-9 and C-11 remained unclear.
To further establish the relative configuration of C-9 and C-11, the chemical shifts of four isomers (1a1d, Figure S35) were predicted at the B3LYP/6-311+G(d,p) with the PCM solvent model. The calculated chemical shifts of isomer 1a were assigned to be in agreement with the experimental values according to the DP4+ analyses (Table S1). The absolute configuration of compound 1 was finally determined by a comparison of its experimental ECD spectrum with that calculated for the proposed structure by quantum chemical TDDFT. The predicted ECD spectrum of 7R, 9R, 11S, 12S, 22R, and 23R was in good agreement with that of the experimental one that showed a negative Cotton effect around 243 nm (Figure 3). Therefore, the absolute configuration of 1 was established as (7R, 9R, 11S, 12S, 22R, and 23R) and was named amestolkolide E.
Compound 8 was obtained as a white amorphous powder with the molecular formula of C14H18O5 based on its HR-ESI MS peak at m/z 265.1081 [M − H]. Analysis of 1H and 13C NMR spectra suggested that compound 8 was an isocoumarin. The 1H NMR spectrum exhibited signals for three aromatic protons at δH 7.42 (1H, dd, J = 8.3, 7.3 Hz), 6.81 (1H, d, J = 8.3 Hz), and 6.76 (1H, d, J = 7.3 Hz), suggesting a 1,2,3-trisubstituted benzene motif in the structure, three oxygenated methines (δH 4.63, 3.55, and 3.35), one methyl (δH 1.16, d, J = 6.3 Hz), and three methylenes (δH 2.94 (1H, dd, J = 16.2, 11.2 Hz), 3.01 (1H, dd, J = 16.2, 3.0 Hz); 2.05 (1H, m), 1.77 (2H, m), 1.57 (1H, m)). With the aid of HSQC, the 13C NMR spectrum revealed 14 carbon signals for six aromatic carbons (δC 163.2, 141.6, 137.4, 116.7, 119.3, and 109.5), one ester carboxyl (δC 171.5), one methyl (δC 18.8), three methylenes (δC 33.6, 32.3, and 28.9), and three sp3 methines (δC 81.2, 76.2, and 71.8). All these spectroscopic data were very similar to those of aspergillumarin B [17] except for those for an extra oxymethine (δHC 3.35/76.2 or 3.55/71.8) in 8. The isochroman fragment was further confirmed by HMBC (Figure 1). The C1′-C5′ located at C-3 was demonstrated by HMBC correlations from H-1′ to C-3 and C-2′, from H-3′ to C-1′, C-2′, and C-4′, and from H-5′ to C-3′ and C-4′. Therefore, the planar structure of compound 8 was identified as 3,4-dihydroxypentyl-8-hydroxyisochroman-1-one. The 1H and 13C NMR data of compound 9 were nearly the same as compound 8 (Table 2) suggesting 9 to be an isomer of 8, which was further confirmed by the HMBC correlation analysis (Figure 1).
Since the relative stereochemistry of the vicinal hydroxyl groups in compounds 8 and 9 cannot be determined by their coupling constants, compounds 8 and 9 were reacted with 2,2-dimethoxypropane to yield their di-O-isopropylidene derivatives 8a and 9a (Scheme 1) [18]. ROESY experiment displayed correlations of H3-7′ with H-3′ and H-4′, suggesting CH3-7′, H-3′, and H-4′ to be cofacial in 8a. An erythro configuration was therefore deduced for the vicinal diol 8. Similarly, a threo configuration was established for 9 by the ROESY correlations of H3-7′/H-3′ and H3-8′/H-4′in 9a. Subsequently, the dimolybdenum tetraacetate [Mo2(OAc)4]-induced CD (ICD) spectra were further applied to determine the absolute configurations for 8 and 9. The ICD spectrum of 8 with [Mo2(OAc)4] in DMSO showed a positive Cotton effect at 317 nm (Figure 4), corresponding to a positive chirality for the O-C-C-O moiety according to Snatzke’s rule [19]; thus, an absolute configuration of 3S and 4R was figured out for conformation 8-A (Figure 5) and then for 8. Similarly, a positive Cotton effect at 310 nm in the ICD spectrum of 9-A revealed the absolute stereochemistry of 9 to be 3′S and 4′S. The remaining absolute configurations of C-3 for 8 and 9 were determined by ECD, where the negative Cotton effects at 258 and 257nm (Figure 6) were consistent with a 3R in 8 and 9, respectively [20].
Thirteen known compounds were isolated and identified as berkeleyacetals A (2) and B (3) [16], amestolkolide B (4) [6], chrysogenolide C (5) [21], purpurogenolides C (6) and E (7) [11], peniciisocoumarins F (10) and H (15) [22], 3R-((R)-4,5-dihydroxypentyl)-8-hydroxyisochroman-1-one (11) [5], aspergillumarins A (13) [17] and B(12) [23], 3R-8-methoxy-3-(4-oxopentyl) isochroman-1-one (14) [20], and penicimarin B (16) [23] by comparison of their optical values and spectroscopic data with those reported. The genera Talaromyces, a sexual state of the Penicillium species, can be found in a wide range of habitat conditions including extreme environments. At present, over fifty compounds of meroterpenoid, isocoumarin, azaphilone, and rare chemical skeletons (e.g., amestolkin) have been isolated from the metabolites of T. amestolkiae [5,6,9,10,24]. Meroterpenoids are the representative skeleton compounds isolated from the genus Penicillium and Talaromyces, biosynthesized from 3,5-dimethylorsellinic acid (DMOA) via a serial reaction such as epoxide-initiated cyclization, oxidative transformation, retro-claisen cleavage, and acetal lactonization [6,25,26,27]. Here, the biosynthetic pathway of compound 1 was also proposed (Figure 7).
Stenotrophomonas maltophilia is an aerobic Gram-negative opportunistic pathogen and one of the most common non-fermenting Gram-negative bacilli [28]. As an ongoing effort to explore new antibiotics against S. maltophilia, all isolated compounds (116) were tested for their antibacterial activities against S. maltophilia. Compounds 716 showed weak antibacterial activities against S. maltophilia with MICs of 256, 512, 512, 256, 512, 256, 128, 512, 256, and 512 μg/mL, respectively (MICs of ceftriaxone sodium and levofloxacin were 128 and 0.25 μg/mL, respectively).

3. Materials and Methods

3.1. General Experimental Procedures

High-resolution electrospray ionization mass spectra (HR ESIMS) were recorded on a Shimadzu LC MS-IT-TOF mass spectrometer equipped with an ESI interface. Optical rotation was determined on an Autopol VI automatic polarimeter. A Chirascan-plus CD spectrometer was employed for the UV and ECD spectra detection. IR spectrum was recorded by the FT-IR-650 spectrometer as KBr disk. Nuclear magnetic resonance (NMR) spectra were measured by a Bruker AVANCE NEO 600M NMR spectrometer (1H: 600 MHz; 13C: 150 MHz). Reversed-phase semi-preparative high-performance liquid chromatography (RP-HPLC) isolation was performed on an LC 3000 system equipped with a UV detector and a Kromasil C18 column (10 mm × 250 mm, 10 μm) with flow rate of 3 mL/min. Normal phase semi-preparative HPLC (NP-HPLC) isolation was used with an AS20005 system equipped with a UV detector and silica gel column (10 mm × 250 mm, 5 μm, Hanbon Sci & Tech (Suzhou, China), with flow rate of 3 mL/min. Sephadex LH-20 was purchased from GE healthcare company. Silica gel (200–300, 300–400 mesh) for column chromatography (CC) and silica gel GF254 (10–40 μm) for thin layer chromatography (TLC) and preparative TLC were obtained from Qingdao Haiyang Chemical Co., Ltd., Qingdao, China. All solvents were of analytical grade.

3.2. Fungal Material

Talaromyces amestolkiae MST1-15 was isolated from the soil collected in the Xingren coal area of Guizhou province in China (25°21′54″ N, 105°9′9″ E), in May 2020. The soil sample was suspended in sterile water and stirred for 10 min to obtain the suspension. A serial dilution method was subsequently adopted and the aliquot from each dilution was inoculated on potato dextrose agar (PDA) plates at 28 °C for 72 h. Isolates that appeared morphologically different were selected and purified, and maintained on a PDA slant stored at 4 °C. The strain was finally identified on the basis of the morphology and ITS analysis (see Supplementary Materials S1). The fungus was deposited at the Microbiological Collection Center of Guizhou Medical University (GMU-2020-MST 1-15).

3.3. Fermentation, Extraction, and Isolation

The fungus was firstly cultured in potato dextrose broth in aerobic closed conical flask shaking with 140 r/min at 28 °C for 4 days. A 10% volume (vs. rice weight) of the fungi suspension was subsequently inoculated into sterile rice medium with 0.3% peptone and cultured in a 28 °C incubator for 30 days to obtain the fermentation substance.
The fermentation substance (8 kg) of T. amestolkiae MST1-15 was extracted with EtOAc (30 L × 2, 2 days for each time) at room temperature. The extract was evaporated under reduced pressure to obtain an ethyl acetate extract (71.0 g). The extract was subjected over a silica gel column and eluted with petroleum ether (PE)/acetone (10/1, 5/1, 3/1, 1/1, 0/1, v/v, each 2000 mL) to give five fractions (Fr. A–E) based on TLC analysis. Fr. B was further purified by Sephadex LH-20 column eluted with CHCl3/MeOH (3/2, v/v) to yield Fr. B1-B3. Subfraction B2 was subsequently separated by NP-HPLC eluted with PE/EtOAc (65/45, v/v) to produce compounds 1 (13 mg, tR = 15 min), 2 (6 mg, tR = 38 min), and 6 (0.8 mg, tR = 45 min). B3 was further purified by NP-HPLC using gradient elution with PE/EtOAc (65/35→55/45, v/v) to give compounds 13 (49 mg, tR = 30 min) and 14 (7 mg, tR = 35 min). Fraction C was chromatographed over a silica gel column eluted with PE/EtOAc (10/1→1/1, v/v) to afford subfractions C1–C4. Compounds 3 (8 mg, tR = 53 min) and 4 (0.8 mg, tR = 41 min) were isolated with RP-HPLC eluting with MeOH/H2O (50/50, v/v) from C1. C2 was further isolated by RP-HPLC eluting with MeOH-H2O (45/55, v/v) to give compound 15 (13 mg, tR = 36 min). C3 was subjected to RP-HPLC eluted with MeOH/H2O (45/55, v/v) to give compound 5 (6 mg, tR = 25 min). Fraction D was purified over a silica gel column gradient elution with CH2Cl2/MeOH (40/1→5/1, v/v) to produce subfractions D1–D4. Compound 11 (3 mg) was obtained by preparative TLC developed with CH2Cl2/EtOAc (v/v = 1/1) from D1. D2 was purified by RP-HPLC eluted with MeOH/H2O (43/57, v/v) to give 16 (10 mg, tR = 43 min). Fraction D3 was subjected to RP-HPLC using MeOH/H2O (47/53, v/v) as mobile phase to afford compounds 7 (7 mg, 26 min) and 18 (6 mg, 32 min). Compounds 8 (6 mg, tR = 41 min), 9 (7 mg, tR =50 min), and 12 (3 mg, tR = 25 min) were isolated from D4 using RP-HPLC eluted with MeOH/H2O (40/60, v/v).

3.4. Spectral and Physical Data of Compounds 1, 8, and 9

Amestolkolide E (1): White amorphous powder; [α] D 21 −132.6 (c 0.11, CH3OH); UV (MeOH) λmax (log ε): 245 (3.67) nm; IR (KBr) υmax: 2962, 2923, 2805, 1671, 1619, 1459, 1382, 1234, 1112, and 790 cm−1; HR ESIMS (neg.) m/z 471.2025 [M − H] (calcd. 471.2024). CD (0.93 mM, MeOH) λmax nm (Δε) 244 (−9.98).
(R)-3-((3S,4R)-3,4-dihydroxypentyl)-8-hydroxyisochroman-1-one (8): White amorphous powder; [α] D 21 −12.6 (c 0.23, CH3OH); UV (MeOH) λmax (log ε): 210 (3.95), 247 (3.44), 315 (2.70) nm; IR (KBr) υmax: 3310, 2896, 2857, 1629, 1581, 1425, 1355, 1201, 1089, and 790 cm−1; HR ESIMS (neg.) m/z 265.1081 [M − H] (calcd. 265.1081); CD (1.11 mM, MeOH) λmax nm (Δε) 216 (−5.46), 239 (+3.19), 257 (−4.94).
(R)-3-((3S,4S)-3,4-dihydroxypentyl)-8-hydroxyisochroman-1-one (9): White amorphous powder; [α] D 21 −26.0 (c 0.37, CH3OH); UV (MeOH) λmax (log ε): 208 (3.98), 245 (3.40), 315 (3.22) nm; IR (KBr) υmax: 3232, 2911, 2865, 1633, 1585, 1428, 1303, 1201, 1085, 960, and 786 cm−1; HR ESIMS (neg.) m/z 265.1084 [M − H] (calcd. 265.1081); CD (0.98 mM, MeOH) λmax nm (Δε) 217 (−1.56), 239 (+0.88), 258 (−1.48).

3.5. Antibacterial Assay

The minimum inhibitory concentrations (MICs) were evaluated according to the reported procedure [29,30] with minor modifications. Briefly, Stenotrophomonas maltophilia ATCC 13637 was cultured on 0002 medium (1% peptone, 0.3% beef extract, and 0.5% NaCl in distilled water with pH 7.0) at 30 °C and prepared for its suspension with a final concentration of 5 × 105 CFU/mL. Samples (116) were severally added into 96-well plates employing serial two-fold dilution from concentrations of 512 to 2 μg/mL. Whereafter, 100 μL bacterial suspension was transferred to the wells to incubate for 16 h at 30 °C. The final concentration of 0.25% TTC (2,3,5-triphenyl tetrazolium chloride, Sigma) in wells as microorganism growth indicator was added and the microorganism was continuously incubated for 3 h, and the concentration in well that showed no red color indicating complete growth inhibition of bacteria was determined as MIC. Equivalent amounts of DMSO (2.5%) and medium were successively used as negative and blank controls. Ceftriaxone sodium and levofloxacin were used as positive controls. Each experiment was performed in triplicate.

3.6. Preparation of Acetonides of 8a and 9a

To a solution of compounds 8 and 9 (3.0 mg) in 200 μL DMF was added 2,2-dimethoxypropane (2 equiv) and p-TsOH (0.5 equiv), respectively [18], and the mixtures were stirred at room temperature for 6 h. The reaction was quenched by saturated aqueous NaHCO3 (200 μL) and extracted with EtOAc (500 μL × 3). The organic layer was dried with Na2SO4 and evaporated under reduced pressure. The residue was subjected to silica gel column chromatography eluted with petroleum ether/ethyl acetate (4:1) to afford 8a (2.6 mg) and 9a (2.9 mg).
Compound 8a: 1H NMR (600 MHz, CD3OD) δ 7.46 (1H, dd, J = 8.0, 7.4 Hz, H-6), 6.85 (1H, d, J = 8.0 Hz, H-7), 6.80 (1H, d, J = 7.4 Hz, H-5), 4.69 (1H, m, H-3), 4.30 (1H, m, H-4′), 4.12 (1H, m, H-3′), 3.05 (1H, dd, J = 16.4, 3.4 Hz, H-4), 2.98 (1H, dd, J = 16.4, 11.2 Hz, H-4), 2.05 (1H, m, H-1′), 1.81 (1H, m, H-1′), 1.74 (1H, m, H-2′), 1.62(1H, m, H-2′), 1.42 (3H, s, CH3-8′), 1.32 (3H, s, CH3-7′), 1.18 (3H, d, J = 6.4 Hz, H-5′). 13C NMR (150 MHz, CD3OD) δ 171.4 (C-1), 163.2 (C-8), 141.6 (C-10), 137.4 (C-6), 119.4 (C-5), 116.7 (C-7), 109.5 (C-9), 108.7 (C-6′), 80.9 (C-3), 78.8 (C-3′), 75.1 (C-4′), 33.7 (C-4), 32.6 (C-1′), 28.8 (CH3-8′), 26.3 (C-2′), 26.0 (CH3-7′), 15.7 (C-5′). ESI MS (neg.) m/z 305.1 [M − H].
Compound 9a: 1H NMR (600 MHz, CD3OD) δ 7.43 (1H, dd, J = 8.0, 7.4 Hz, H-6), 6.82 (1H, d, J = 8.0 Hz, H-7), 6.77 (1H, d, J = 7.4 Hz, H-5), 4.62 (1H, m, H-3), 3.73 (1H, m, H-4′), 3.55 (1H, m, H-3′), 3.02 (1H, dd, J = 16.4, 3.5 Hz), 2.95 (1H, dd, J = 16.4, 11.4 Hz, H-4), 1.95 (1H, m, H-1′), 1.87 (2H, m, H-1′, 2′), 1.60 (1H, m, H-2′), 1.33 (3H, s, CH3-7′), 1.32 (3H, s, CH3-8′), 1.23 (3H, d, J = 6.0 Hz, H-5′). 13C NMR (150 MHz, CD3OD) δ 171.4 (C-1), 163.2 (C-8), 141.6 (C-10), 137.4 (C-6), 119.4 (C-5), 116.8 (C-7), 109.5 (C-9), 109.1 (C-6′), 83.6 (C-3′), 81.4 (C-3), 78.2 (C-4′), 33.7 (C-4), 32.8 (C-1′), 28.9 (C-2′), 27.6 (CH3-7′), 27.5 (CH3-8′), 17.7 (C-5′). ESI MS (neg.) m/z 305.0 [M − H].

4. Conclusions

A chemical study of Talaromyces amestolkiae MST1-15 collected from coal areas led to the isolation and identification of seven meroterpenoids (17) and eleven isocoumarins (816), including three new compounds named amestolkolide E (1), (R)-3-((3S,4R)-3,4-dihydroxypentyl)-8-hydroxyisochroman-1-one (8), and (R)-3-((3S,4S)-3,4-dihydroxypentyl)-8-hydroxyisochroman-1-one (9). Compounds 716 showed weak antibacterial activities against S. maltophilia with MIC values ranging from 256 to 512 μg/mL. This study enriched the chemical structure and biological activity from the secondary metabolites of T. amestolkiae.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27238223/s1. References [31,32] are cited in the Supplementary Materials.

Author Contributions

K.-Y.L. and Q.-F.Z. performed the isolation and identification, and bioactivities experiments; X.-M.L. performed the MS and IR experiments; F.-R.W. and J.-L.A. performed the NMR experiment and analyzed the data; Q.-F.Z. identified the absolute configuration; S.-G.L. and G.-B.X. conceived and designed the research project, and wrote this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (No. 81903496), and Science and Technology Project of Guizhou Province ((2020)1Y394), and the Scientific Research Foundation for Innovative Talent of Guizhou Province ((2020)6011), and the Project of Science and Technology Department of Guizhou Province ((2020)5006).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Benjamin, C.R. Ascocarps of Aspergillus and Penicillium. Mycologia 1955, 47, 669–687. [Google Scholar] [CrossRef]
  2. Samson, R.A.; Yilmaz, N.; Houbraken, J.; Spierenburg, H.; Seifert, K.A.; Peterson, S.W.; Varga, J.; Frisvad, J.C. Phylogeny and nomenclature of the genus Talaromyces and taxa accommodated in Penicillium subgenus Biverticillium. Stud. Mycol. 2011, 70, 159–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zang, Y.; Genta-Jouve, G.; Escargueil, A.E.; Larsen, A.K.; Guedon, L.; Nay, B.; Prado, S. Antimicrobial oligophenalenone dimers from the soil fungus Talaromyces stipitatus. J. Nat. Prod. 2016, 79, 2991–2996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chen, C.; Sun, W.; Liu, X.; Wei, M.; Liang, Y.; Wang, J.; Zhu, H.; Zhang, Y. Anti-inflammatory spiroaxane and drimane sesquiterpenoids from Talaromyces minioluteus (Penicillium minioluteum). Bioorg. Chem. 2019, 91, 103166. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, S.; Liu, Y.; Liu, Z.; Cai, R.; Lu, Y.; Huang, X.; She, Z. Isocoumarins and benzofurans from the mangrove endophytic fungus Talaromyces amestolkiae possess α-glucosidase inhibitory and antibacterial activities. RSC Adv. 2016, 6, 26412. [Google Scholar] [CrossRef]
  6. Chen, S.; Ding, M.; Liu, W.; Huang, X.; Liu, Z.; Lu, Y.; Liu, H.; She, Z. Anti-inflammatory meroterpenoids from the mangrove endophytic fungus Talaromyces amestolkiae YX1. Phytochemistry 2018, 146, 8–15. [Google Scholar] [CrossRef]
  7. Oliveira, F.; Rocha, I.L.D.; Pinto, D.C.G.A.; Ventura, S.P.M.; Santos, A.G.; Crevelin, E.J.; Ebinuma, V.C.S. Identification of azaphilone derivatives of Monascus colorants from Talaromyces amestolkiae and their halochromic properties. Food Chem. 2022, 372, 131214. [Google Scholar] [CrossRef]
  8. Prieto, A.; de Eugenio, L.; Méndez-Líter, J.A.; Nieto-Domínguez, M.; Murgiondo, C.; Barriuso, J.; Bejarano-Muñoz, L.; Martínez, M.J. Fungal glycosyl hydrolases for sustainable plant biomass valorization: Talaromyces amestolkiae as a model fungus. Int. Microbiol. 2021, 24, 545–558. [Google Scholar] [CrossRef]
  9. Fu, Y.; Li, C.; Zhu, J.; Zhang, L.; Wang, Y.; Chen, Q.; Xu, L.; Zhang, S.; Fang, Y.; Liu, T. A new meroterpenoid from endophytic fungus Talaromyces amestolkiae CS-O-1. Biochem. Syst. Ecol. 2020, 100, 104377. [Google Scholar] [CrossRef]
  10. EI-Elimat, T.; Figueroa, M.; Raja, H.A.; Alnabulsi, S.; Oberlies, N.H. Coumarins, dihydroisocoumarins, a dibenzo-α-pyrone, a meroterpenoid, and a merodrimane from Talaromyces amestolkiae. Tetrahedron Lett. 2021, 72, 153067. [Google Scholar] [CrossRef]
  11. Rateb, M.E.; Ebel, R. Secondary metabolites of fungi from marine habitats. Nat. Prod. Rep. 2011, 28, 290–344. [Google Scholar] [CrossRef] [PubMed]
  12. Yogabaanu, U.; Faizal Weber, J.-F.; Peter Convey, P.; Rizman-Idid, M.; Alias, S.A. Antimicrobial properties and the influence of temperature on secondary metabolite production in cold environment soil fungi. Polar Sci. 2017, 14, 60–67. [Google Scholar] [CrossRef]
  13. Orwa, P.; Mugambi, G.; Wekesa, V.; Mwirichia, R. Isolation of haloalkaliphilic fungi from Lake Magadi in Kenya. Heliyon 2020, 6, e02823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kochhar, N.; Kavya, I.K.; Shrivastava, S.; Ghosh, A.; Rawat, V.S.; Sodhi, K.K.; Kumar, M. Perspectives on the microorganism of extreme environments and their applications. Curr. Res. Microb. Sci. 2022, 3, 100134. [Google Scholar] [CrossRef] [PubMed]
  15. Sun, J.; Zhu, Z.X.; Song, Y.L.; Dong, D.; Zheng, J.; Liu, T.; Zhao, Y.-F.; Ferreira, D.; Zjawiony, J.K.; Tu, P.F.; et al. Nitric oxide inhibitory meroterpenoids from the fungus Penicillium purpurogenum MHZ 111. J. Nat. Prod. 2016, 79, 1415–1422. [Google Scholar] [CrossRef]
  16. Stierle, D.B.; Stierle, A.A.; Patacini, B. The berkeleyacetals, three meroterpenes from a deep water acid mine waste Penicillium. J. Nat. Prod. 2007, 70, 1820–1823. [Google Scholar] [CrossRef] [Green Version]
  17. Li, S.D.; Wei, M.Y.; Chen, G.Y.; Lin, Y.C. Two new dihydroisocoumarins from the endophytic fungus Aspergillus sp. Collected from the south China sea. Chem. Nat. Compd. 2012, 48, 371–373. [Google Scholar] [CrossRef]
  18. Xu, L.; He, Z.; Xue, J.; Chen, X.; Wei, X. β-Resorcylic Acid Lactones from a Paecilomyces Fungus. J. Nat. Prod. 2010, 73, 885–889. [Google Scholar] [CrossRef]
  19. Di Bari, L.; Pescitelli, G.; Carmela, P.; Pini, D.; Salvadori, P. Determination of absolute configuration of acyclic 1,2-diols with Mo2(OAc)4. 1. Snatzke’s method revisited. J. Org. Chem. 2001, 66, 4819–4825. [Google Scholar] [CrossRef]
  20. Xu, G.-B.; Yang, F.-Y.; Wu, X.-Y.; Li, R.; Zhou, M.; Wang, B.; Yang, X.-S.; Zhang, T.-T.; Liao, S.-G. Two new dihydroisocoumarins with antimicrobial activities from the fungus Penicillium sp. XR046 collected from Xinren coal area. Nat. Prod. Res. 2021, 35, 1445–1451. [Google Scholar] [CrossRef]
  21. Qi, B.; Liu, X.; Mo, T.; Zhu, Z.; Li, J.; Wang, J.; Shi, X.; Zeng, K.; Wang, X.; Tu, P.; et al. 3,5-Dimethylorsellinic acid derived meroterpenoids from Penicillium chrysogenum MT-12, an endophytic fungus isolated from Huperzia serrata. J. Nat. Prod. 2017, 80, 2699–2707. [Google Scholar] [CrossRef] [PubMed]
  22. Bai, M.; Zheng, C.-J.; Huang, G.-L.; Mei, R.-Q.; Wang, B.; Luo, Y.-P.; Zheng, C.; Niu, Z.-G.; Chen, G.-Y. Bioactive meroterpenoids and isocoumarins from the mangrove-derived fungus Penicillium sp. TGM112. J. Nat. Prod. 2019, 82, 1155–1164. [Google Scholar] [CrossRef] [PubMed]
  23. Qi, J.; Shao, C.-L.; Li, Z.-Y.; Gan, L.-S.; Fu, X.-M.; Bian, W.-T.; Zhao, H.-Y.; Wang, C.-Y. Isocoumarin derivatives and benzofurans from a sponge-derived Penicillium sp. fungus. J. Nat. Prod. 2013, 76, 571–579. [Google Scholar] [CrossRef] [PubMed]
  24. Huang, L.-J.; Li, X.-A.; Jin, M.-Y.; Guo, W.-X.; Lei, L.-R.; Liu, R.; Zhang, M.-Z.; Guo, D.-L.; Wang, D.; Zhou, Y.; et al. Two previously undescribed phthalides from Talaromyces amestolkiae, a symbiotic fungus of Syngnathus acus. J. Asian Nat. Prod. Res. 2022. [Google Scholar] [CrossRef] [PubMed]
  25. Geris, R.; Simpson, T.J. Meroterpenoids produced by fungi. Nat. Prod. Rep. 2009, 26, 1063–1094. [Google Scholar] [CrossRef] [Green Version]
  26. Matsuda, Y.; Abe, I. Biosynthesis of fungal meroterpenoids. Nat. Prod. Rep. 2015, 33, 26–53. [Google Scholar] [CrossRef] [Green Version]
  27. Matsuda, Y.; Iwabuchi, T.; Fujimoto, T.; Awakawa, T.; Nakashima, Y.; Mori, T.; Zhang, H.; Hayashi, F.; Abe, I. Discovery of key dioxygenases that diverged the paraherquonin and acetoxydehydroaustin pathways in Penicillium brasilianum. J. Am. Chem. Soc. 2016, 138, 12671–12677. [Google Scholar] [CrossRef]
  28. Barsky, E.E.; Williams, K.A.; Priebe, G.P.; Sawicki, G.S. Incident Stenotrophomonas maltophilia infection and lung function decline in cystic fibrosis. Pediatr. Pulm. 2017, 52, 1276–1282. [Google Scholar] [CrossRef]
  29. Assaw, S.; Mohd Amir, M.I.H.; Khaw, T.T.; Bakar, K.; Mohd Radzi, S.A.; Mazlan, N.W. Antibacterial and antioxidant activity of naphthofuranquinones from the twigs of tropical mangrove Avicennia officinalis. Nat. Prod. Res. 2020, 34, 2403–2406. [Google Scholar] [CrossRef]
  30. Jabrane, A.; Jannet, H.B.; Mastouri, M.; Mighri, Z.; Casanova, J. Chemical composition and in vitro evaluation of antioxidant and antibacterial activities of the root oil of Ridolfia segetum (L.) Moris from Tunisia. Nat. Prod. Res. 2010, 24, 491–499. [Google Scholar] [CrossRef]
  31. Jiang, X.-Z.; Yu, Z.-D.; Ruan, Y.-M.; Wang, L. Three new species of Talaromyces sect. Talaromyces discovered from soil in China. Sci. Rep. 2018, 8, 4932. [Google Scholar] [CrossRef] [Green Version]
  32. Xu, G.-B.; He, G.; Bai, H.-H.; Yang, T.; Zhang, G.-L.; Wu, L.-W.; Li, G.-Y. Indole Alkaloids from Chaetomium globosum. J. Nat. Prod. 2015, 78, 1479–1485. [Google Scholar] [CrossRef]
Figure 1. The Key HMBC correlations of compounds 1, 8, and 9.
Figure 1. The Key HMBC correlations of compounds 1, 8, and 9.
Molecules 27 08223 g001
Figure 2. The Key NOESY (Molecules 27 08223 i001) and 1H-1H COSY (Molecules 27 08223 i002) correlations of 1.
Figure 2. The Key NOESY (Molecules 27 08223 i001) and 1H-1H COSY (Molecules 27 08223 i002) correlations of 1.
Molecules 27 08223 g002
Figure 3. Experimental and calculated ECD spectra of 1.
Figure 3. Experimental and calculated ECD spectra of 1.
Molecules 27 08223 g003
Scheme 1. Synthesis of di-O-isopropylidene derivatives 8a and 9a, and NOESY correlations (Molecules 27 08223 i003) of 8a and 9a.
Scheme 1. Synthesis of di-O-isopropylidene derivatives 8a and 9a, and NOESY correlations (Molecules 27 08223 i003) of 8a and 9a.
Molecules 27 08223 sch001
Figure 4. ICD spectra of compounds 8-A (for 8) and 9-A (for 9) and Mo2(OAc)4 in DMSO.
Figure 4. ICD spectra of compounds 8-A (for 8) and 9-A (for 9) and Mo2(OAc)4 in DMSO.
Molecules 27 08223 g004
Figure 5. Favored conformers 8-A and 9-A for compounds 8 and 9 reacted with Mo2(OAc)4.
Figure 5. Favored conformers 8-A and 9-A for compounds 8 and 9 reacted with Mo2(OAc)4.
Molecules 27 08223 g005
Figure 6. Experimental ECD spectra of 8 and 9.
Figure 6. Experimental ECD spectra of 8 and 9.
Molecules 27 08223 g006
Figure 7. Plausible biosynthetic pathway of compound 1.
Figure 7. Plausible biosynthetic pathway of compound 1.
Molecules 27 08223 g007
Table 1. 1H (600 MHz) and 13C NMR (150 MHz) data of 1.
Table 1. 1H (600 MHz) and 13C NMR (150 MHz) data of 1.
No.1 (in CDCl3) a
δH (J in Hz)δC
1 169.4
23.32 (1H, d, 21.1)
2.88 (1H, dd, 21.1, 1.3)
31.6
3 126.9
4 132.8
5 137.0
62.74 (1H, d, 16.2)
2.14 (1H, d, 16.2)
34.0
7 41.0
8 179.8
94.18 (1H, q, 6.8)72.3
10 209.4
11 63.2
12 44.1
133.22 (1H, dt, 13.7, 3.6)
1.54 (1H, dd, 13.7, 3.2)
49.1
142.70 (1H, t like)
1.95 (1H, dt, 13.7, 3.2)
26.7
15 136.3
16 85.6
171.57 (3H, s)27.4
181.48 (3H, s)27.9
191.14 (3H, s)27.3
20 168.2
211.42 (3H, d, 6.8)17.7
223.69 (1H, d, 8.2)46.0
236.09 (1H, d, 8.2)98.5
241.33 (3H, s)26.5
251.77 (3H, s)18.3
263.80 (3H, s)53.2
a All assignments are based on HSQC and HMBC experiments.
Table 2. 1H (600 MHz) and 13C NMR (150 MHz) data of compounds 8 and 9 (in methanol-d4) a.
Table 2. 1H (600 MHz) and 13C NMR (150 MHz) data of compounds 8 and 9 (in methanol-d4) a.
No.δH (J in Hz)δC
8989
1 171.5171.5
34.63 (1H, m)4.64 (1H, m)81.281.7
43.01 (1H, dd, 16.2, 3.0)
2.94 (1H, dd, 16.2, 11.2)
3.05 (1H, dd, 16.3, 3.2)
2.96 (1H, dd, 16.3, 11.4)
33.633.7
4a 141.6141.6
56.76 (1H, d, 7.3)6.80 (1H, d, 7.4)119.3119.3
67.42 (1H, dd, 8.3, 7.3)7.45 (1H, dd, 8.4, 7.4)137.4137.4
76.81 (1H, d, 8.3)6.84 (1H, d, 8.4)116.7116.7
8 163.2163.2
8a 109.5109.5
1′2.05 (1H, m)
1.77 (1H, m)
2.00 (1H, m)
1.89 (1H, m)
32.332.6
2′1.77 (1H, m)
1.57 (1H, m)
1.82 (1H, m)
1.53 (1H, m)
28.929.0
3′3.35 (1H, m)3.38 (1H, m)76.276.3
4′3.55 (1H, m)3.63 (1H, m)71.871.4
5′1.16 (3H, d, 6.3)1.17 (3H, d, 6.4)18.818.9
a All assignments are based on HSQC and HMBC experiments.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, K.-Y.; Zhu, Q.-F.; Ao, J.-L.; Wang, F.-R.; Long, X.-M.; Liao, S.-G.; Xu, G.-B. New Meroterpenoid and Isocoumarins from the Fungus Talaromyces amestolkiae MST1-15 Collected from Coal Area. Molecules 2022, 27, 8223. https://doi.org/10.3390/molecules27238223

AMA Style

Li K-Y, Zhu Q-F, Ao J-L, Wang F-R, Long X-M, Liao S-G, Xu G-B. New Meroterpenoid and Isocoumarins from the Fungus Talaromyces amestolkiae MST1-15 Collected from Coal Area. Molecules. 2022; 27(23):8223. https://doi.org/10.3390/molecules27238223

Chicago/Turabian Style

Li, Kai-Yu, Qin-Feng Zhu, Jun-Li Ao, Fu-Rui Wang, Xing-Mei Long, Shang-Gao Liao, and Guo-Bo Xu. 2022. "New Meroterpenoid and Isocoumarins from the Fungus Talaromyces amestolkiae MST1-15 Collected from Coal Area" Molecules 27, no. 23: 8223. https://doi.org/10.3390/molecules27238223

Article Metrics

Back to TopTop