Next Article in Journal
Functional Genomics of a Collection of Gammaproteobacteria Isolated from Antarctica
Next Article in Special Issue
New Cyclic Pentapeptides from the Mangrove-Derived Aspergillus fumigatus GXIMD 03099
Previous Article in Journal
Genomics- and Transcriptomics-Guided Discovery of Clavatols from Arctic Fungi Penicillium sp. MYA5
Previous Article in Special Issue
New Phenol Derivatives from the Haima Cold Seep-Derived Fungus Aspergillus subversicolor CYH-17
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Secondary Metabolites from Marine-Derived Fungus Talaromyces minnesotensis BTBU20220184

1
Key Laboratory of Marine Mineral Resources and Polar Geology, Ministry of Education, School of Ocean Sciences, China University of Geosciences, Beijing 100083, China
2
Key Laboratory of Geriatric Nutrition and Health, Ministry of Education of China, School of Light Industry Science and Engineering, Beijing Technology and Business University, Beijing 100048, China
3
Key Laboratory of Tropical Marine Ecosystem and Bioresource, Guangxi Key Laboratory of Beibu Gulf Marine Resources, Environment and Sustainable Development, Fourth Institute of Oceanography, Ministry of Natural Resources, Beihai 536000, China
4
State Key Laboratory of Mycology, Institute of Microbiology, Chinese Academy of Sciences, Beijing 100101, China
5
CAS Key Laboratory of Experimental Marine Biology, Center for Ocean Mega-Science, Institute of Oceanology, Chinese Academy of Sciences, Qingdao 266071, China
6
Laboratory for Marine Biology and Biotechnology, Qingdao National Laboratory for Marine Science and Technology, Qingdao 266237, China
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2024, 22(6), 237; https://doi.org/10.3390/md22060237
Submission received: 30 April 2024 / Revised: 16 May 2024 / Accepted: 22 May 2024 / Published: 23 May 2024
(This article belongs to the Special Issue Bioactive Secondary Metabolites of Marine Fungi 2.0)

Abstract

:
Six new compounds, talamitones A and B (1 and 2), demethyltalamitone B (3), talamiisocoumaringlycosides A and B (4 and 5), and talaminaphtholglycoside (6), together with six known compounds (712), were isolated from the marine-derived fungus Talaromyces minnesotensis BTBU20220184. The new structures were characterized by using HRESIMS and NMR. This is the first report of isocoumaringlycoside derivatives from a fungus of the Talaromyces genus. Compounds 5, 6, and 9 showed synergistic antibacterial activity against Staphylococcus aureus.

1. Introduction

The marine ecosystem is a rich pool for filamentous fungi. With the quick development in sampling using marine and molecular taxonomy technology, the fungal diversity of the marine environment has been greatly improved [1]. Over 1800 species of marine-derived fungi, belonging to 769 genera, have been studied, such as Aspergillus, Penicillium, Trichoderma, Eurotium, Talaromyces, and so on [2,3]. The section of Talaromyces was first introduced by Stolk and Samson in 1972 [4], and over 200 species have been reported in this genus [5,6]. Species of Talaromyces are commonly distributed in various environments, such as soil [6,7], marine [8], food products [9], leaf litter [10], and atmospheric environments [11]. A chemical investigation of Talaromyces species revealed the potential for new bioactive chemical entries [12]. A panel of new structural classes have been reported from this fungus, including ergosteroids [13,14], meroterpenoids [15], and pyridone derivatives [16,17]. Nicoletti reviewed the chemical diversity of Talaromyces species from marine environments. Over 500 natural products have been identified from fungi of the Talaromyces genus, and 45% of the compounds were new structures. These compounds displayed a wide spectrum of biological activities and chemical diversity, attracting researchers to screening new pharmaceutical entries [18].
In our ongoing chemical investigation of marine-derived microorganisms, new antibacterial compounds have been identified from marine-derived Talaromyces strains [19,20]. Further research on a marine-derived fungus from Qinzhou Bay, Guangxi Province, China, resulted in the identification of twelve compounds, including six new compounds, talamitones A and B (1 and 2), demethyltalamitone B (3), talamiisocoumaringlycosides A and B (4 and 5), and talaminaphtholglycoside (6), along with six reported compounds, mucorisocoumarin B (7), diaportinol (8), peniisocoumarin E (9), dichlorodiaportin (10), orsellinic acid (11), and 4-(hydroxymethyl)-5-hydroxy-2H-pyran-2-one (12) (Figure 1), from T. minnesotensis BTBT20220184. Compounds 5, 6, and 9 showed synergistic antibacterial activity against Staphylococcus aureus. This paper focuses on the isolation, detailed structure elucidation, and antimicrobial activity evaluation of these compounds.

2. Results

2.1. Structure Determination

Compound 1 was isolated as a light-yellow oil. The molecular formula of 1 was determined to be C20H16O8 based on the HRESIMS spectrum (m/z [M + H]+ 385.0923, calcd for C20H17O8, 385.0918), indicating thirteen degrees of unsaturation (Figure S1). The 1H NMR data of 1 (Table 1, Figure S2) showed five aromatic protons at δH 7.50 (d, J = 1.5 Hz, H-5), 7.51 (d, J = 1.5 Hz, H-7), 6.39 (d, J = 8.0 Hz, H-3′, 5′), and 7.04 (t, J = 8.0 Hz, H-4′); two methylenes at δH 2.44 (m, H2-8) and 2.46 (m, H2-9); and one methoxy at δH 3.60 (s, 10-OCH3). The 13C and HSQC spectra of 1 (Figures S3 and S4) displayed twenty carbon signals (Table 1), including one carbonyl at δC 198.5 (C-1); two carboxyl carbons at δC (175.3, C-10) and (168.9, C-11); eight sp2 quaternary carbons at δC 136.7 (C-2), 154.0 (C-3), 134.8 (C-3a), 152.6 (C-4), 134.4 (C-6), 133.1 (C-7a), 110.7 (C-1′), and 156.7 (C-2′, 6′); five sp2 methine carbons at δC 126.8 (C-5), 115.4 (C-7), 107.6 (C-3′, 5′), and 130.9 (C-4′); two methylene carbons at δC 20.7 (C-8) and 32.9 (C-9); and one methoxy carbon at δC (52.1, 10-OCH3). The 1H-1H COSY correlations (Figure 2 and Figure S5) revealed the fragments of C-8/C-9 and C-3′/C-4′/C-5′. In the HMBC spectrum (Figure S6), the correlations from H-5 to C-3, C-3a, C-4, C-6, and C-7; from H-7 to C-1, C-3a, and C-5; and from H-5 and H-7 to C-11, along with the chemical shift of C-4 (δC 152.6), suggested that the hydroxyl and carboxyl substituted the inden-1-one moiety (Figure 2). The HMBC correlations from H-4′ to C-2′ and C-6′ and from H-3′ and H-5′ to C-1′ indicated the 1,2,3-trisubstituted benzene fragment. The HMBC correlations from H2-8, H2-9, and H3-OMe to C-10 confirmed the presence of a C-8/C-9/C-10/OMe moiety. The HMBC correlations from H2-8 to C-1, C-2, and C-3; from H2-9 to C-2; and from H-3′ and H-5′ to C-1′ and C-3 revealed the connection from C-2 to C-8 and from C-3 to C-1′, respectively. Therefore, the structure of 1 was assigned as shown in Figure 1 and named talamitone A.
Compound 2 was isolated as a light-yellow powder. The molecular formula of 2 was determined to be C20H18O9 based on the HRESIMS spectrum (m/z [M + H]+ 403.1011, calcd for C20H19O9, 403.1024), indicating twelve degrees of unsaturation (Figure S7). The analysis of the 1H, 13C, HSQC, and 1H-1H COSY spectra (Table 1, Figures S8–S11) revealed the presence of two substituted benzene rings (δC 136.8, C-3a; δC 155.2, C-4; δC 121.7/δH 7.68, C-5/H-5; δC 132.8, C-6; δC 122.0/δH 8.07, C-7/H-7; δC 137.3, C-7a; δC 112.2, C-1′; δC 163.3, C-2′,6′; δC 108.1/δH 6.28, C-3′,5′/H-3′,5′; δC 137.4/δH 7.21, C-4′/H-4′); three sp3 methylenes (δC 38.8/δH 3.06, C-2/H-2; δC 20.4/1.90, C-8/H-8; δC 33.7/δH 2.35, C-9/H-9); one methoxy (δC 52.1/δH 3.64, 10-OCH3); two carbonyls at δC 201.0 (C-1) and 202.8 (C-3); as well as two carboxyls at δC 175.5 (C-10) and δC 168.6 (C-11). All of the above NMR data indicated that the structure of 2 is similar to that of 1. A detailed comparison of the NMR data of compounds 1 and 2 showed that the double bond C-2/C-3 was modified. The double bonds of C-2/C-3 in 1 were replaced by one sp3 methylene (δH 3.06, 2H, t, J = 7.0 Hz; δC 38.8, C-2) and one carbonyl (δC 202.8, C-3). In the HMBC spectrum (Figure 2 and Figure S12), the long-range correlation from H-5 to C-3 confirmed the carbonyl (C-3)’s bearing on C-3a. The HMBC correlations from H-3′ and H-5′ to C-3 and from H2-2 to C-1 and C-7a indicated that the α,β-unsaturated moiety of C-1/C-2/C-3 in 1 changed to the moiety of C-1 (carbonyl)/C-2 (methylene) and the other carbonyl of C-3 in 2. Therefore, the structure of 2 was assigned as shown in Figure 1 and named talamitone B.
Compound 3 was isolated as a light-yellow powder. The molecular formula of 3 was determined to be C19H16O9 based on the HRESIMS spectrum (m/z [M + H]+ 389.0871, calcd for C19H17O9, 389.0867), indicating twelve degrees of unsaturation (Figure S13). Compound 3 showed similar NMR data to those of 2. By comparing the molecular formula and NMR data (Table 1, Figures S14 and S15), 3 was deduced to be the product of the demethylation of 2. The structure of 3 was confirmed by detailed analysis of the HSQC, 1H-1H COSY, and HMBC spectra (Figures S16–S18). The long-range correlation from H-5 to C-3 confirmed the carbonyl (C-3)’s bearing on C-3a. The presence of HMBC correlations from H2-8 and H2-9 to C-10 indicated the carboxyl of C-10. Therefore, the structure of 3 was assigned as shown in Figure 1 and named demethyltalamitone B.
Compound 4 was isolated as a light-yellow powder. The molecular formula of 4 was determined to be C19H24O11 based on the HRESIMS spectrum (m/z [M + Na]+ 451.1204, calcd for C19H24O11Na, 451.1211), indicating eight degrees of unsaturation (Figure S19). The 1H and 13C NMR data of 4 (Table 2, Figures S20 and S21) showed signals for one isocoumarin substructure at δC 167.9 (C-1, carboxyl), 157.9 (C-3, oxygenated), δC 101.2 (C-4)/δH 6.94 (s, H-4), 133.5 (C-4a), 132.9 (C-5), 161.1 (C-6, oxygenated), δC 99.7 (C-7)/δH 6.63 (s, H-7), 161.8 (C-8, oxygenated), and 99.3 (C-8a); one hydroxypropyl at δC 30.7 (C-9)/δH 2.63 (m, H2-9), δC 30.8 (C-10)/δH 1.90 (tt, J = 7.0 Hz, H2-10), and δC 61.7, C-11/δH 3.63 (m, H2-11); one glucose at δC 105.9 (C-1′)/δH 4.70 (d, J = 7.5 Hz, H-1′), δC 75.6 (C-2′)/ δH 3.50 (m, H-2′), δC 77.8 (C-3′)/δH 3.42 (m, H-3′), δC 71.2 (C-4′)/δH 3.42 (m, H-4′), δC 78.2, C-5′/δH 3.17 (m, H-5′), δC 62.5, C-6′)/ δH 3.76 (dd, J = 11.5, 2.5 Hz, H-6′a), and 3.67 1H (dd, J = 11.5, 4.5 Hz, H-6′b); and one methoxy at δC 56.8/δH 3.94 (s, 6-OCH3). The protonated carbons and correlations were confirmed by the HSQC and 1H-1H COSY spectra (Figures S22 and S23). In the HMBC spectrum (Figure 2 and Figure S24), the correlations from H3-6-OMe to C-6 revealed the methoxy’s bearing on C-6. The HMBC correlations from H-4 to C-3 and C-9 and from H2-9 and H2-10 to C-3 indicated the hydroxypropyl at C-3. The connection between C-5 and C-1′ through an oxygen atom was confirmed by the HMBC correlations from H-1′ to C-5. Therefore, the planar structure of 4 was assigned and named talamiisocoumaringlycoside A.
Compound 5 was isolated as a light-yellow powder. The molecular formula of 5 was determined to be C19H23ClO11 based on the HRESIMS spectrum (m/z [M + Na]+ 485.0830, calcd for C19H23ClO11Na, 485.0821), indicating eight degrees of unsaturation (Figure S25). The 1H and 13C NMR data of 5 (Table 2, Figures S26 and S27) showed similar signals to those of 4, except for the methylene of CH2-10, which was replaced by one sp3 methine. Through analysis of the molecular formula, chemical shift of 13C NMR, and 2D NMR data (Figures S28–S30), the chloro-substituted methine at δC 60.4 (C-4)/δH 4.32 (m, H-4) was deduced. By comparison of the optical rotations of 5 ( [ α ] D 25 + 26.0) and peniisocoumarin E ( [ α ] D 25 + 69.2), the configuration of C-10 was assigned to be the same as the known compound peniisocoumarin E (9) [21]. Therefore, the planar structure of 5 was assigned as shown in Figure 1 and named talamiisocoumaringlycoside B.
Compound 6 was isolated as a light-yellow powder. The molecular formula of 6 was determined to be C19H22O9 based on the HRESIMS spectrum (m/z [M + H]+ 395.1338, calcd for C19H23O9, 395.1337), indicating nine degrees of unsaturation (Figure S31). The 1H and 13C NMR data of 5 (Table 2, Figures S32 and S33) showed signals for one glucose moiety as in 4 or 5 at δC 107.1 (C-1′)/δH 4.78 (d, J = 7.5 Hz, H-1′), 75.8 (C-2′)/δH 3.69 (m, H-2′), 78.2 (C-3’)/δH 3.47 (m, H-13′), 71.4 (C-4′)/δH 3.47 (m, H-4’), 77.9 (C-5′)/δH 3.13 (m, H-5’), δC (62.6, C-6’)/δH 3.76 (dd, J = 2.5, 11.5 Hz, H-6’a), and 3.67 (m, H-6’b); three aromatic methines at δC 110.0 (C-1)/δH 6.89 (s, H-1), 131.3 (C-4)/δH 9.17 (s, H-4), and 104.4 (C-8)/δH 6.76 (s, H-8); two methyls at δC 27.2 (C-10)/δH 2.76 (s, H3-10) and 10.9 (C-11)/δH 2.34 (s, H3-11); one carbonyl at δC 207.0 (C-9); and seven sp2 quaternary carbons at δC 158.4 (C-2, oxygenated), 119.5 (C-3), 118.6 (C-4a), 154.0 (C-5, oxygenated), 120.7 (C-6), 160.5 (C-7, oxygenated), and 140.4 (C-8a). The protonated carbons were confirmed by the HSQC and 1H-1H COSY spectra (Figures S34 and S35). The substituted naphthalene substructure was confirmed by the HMBC correlations (Figure 2 and Figure S36) from H-1 to C-3, C-4a, and C-8; from H-4 to C-2, C-5, C-8a, and C-9; from H-8 to C-1, C-4a, and C-6; from H3-10 to C-3 and C-9; as well as from H3-11 to C-5, C-6, and C-7. The connection between C-5 and C-1′ through an oxygen atom was characterized by the HMBC correlation from H-1′ to C-5. Therefore, the planar structure of 6 was assigned and named talaminaphtholglycoside.
With almost identical NMR data, the sugar moieties of 46 were deduced to be the same configurations. The anomeric protons H-1′ of 46 showed large coupling constants (J = 7.5 Hz), indicating the presence of a β-glycosidic bond [22]. In the ROESY spectrum (Figure 2 and Figure S37), H-1′ showed crossing peaks with H-3′ and H-5′, which revealed the axial–axial relationship. Through a detailed comparison of the 13C NMR data for the sugar moiety of 1-hydroxy-3-methoxy-8-methyl-2-O-β-D-glucopyranosylnaphthaline [23] with those of 46, the sugar moieties were determined as β-D glucopyranose.
The known compounds were isolated from T. minnesotensis BTBT20220184 and determined as mucorisocoumarin B (7) [24], diaportinol (8) [25], peniisocoumarin E (9) [21], dichlorodiaportin (10) [25], orsellinic acid (11) [26], and 4-(hydroxymethyl)-5-hydroxy-2H-pyran-2-one (12) [27] by comparing their spectroscopic data with those reported in the literature.

2.2. Biological Activity

The biological evaluations were conducted to assess the antibacterial activities against Candida albicans ATCC 10231, Staphylococcus aureus ATCC 25923, and Escherichia coli ATCC 25923. None of the compounds showed markable antimicrobial activities, except for 8 inhabiting the growth of S. aureus at 200 μg/mL. When combing them with 0.125 μg/mL of methicillin, 5, 6, and 9 showed synergistic antibacterial activity against S. aureus at concentrations of 25, 50, and 12.5 μg/mL, respectively.

3. Materials and Methods

3.1. General Experimental Procedures

Optical rotations [ α ] D 25 were recorded on an Anton Paar MCP 200 Modular Circular Polarimeter (Austria) in a 100 × 2 mm cell. NMR spectra were measured on a Bruker Avance 500 spectrometer with residual solvent peaks as references (CD3OD: δH 3.31, δC 49.0). High-resolution HRESIMS measurements were obtained on a Xevo G2 XS QToF system (Manchester, UK) in positive-ion mode. HPLC was performed using an Agilent 1200 Series separation module equipped with an Agilent 1200 Series diode array and Agilent 1260 Series fraction collector and an Agilent ZORBAX RX-C8 column (250 × 9.4 mm, 5 µm).

3.2. Microbial Material, Fermentation, Extraction, and Purification

The strain T. minnesotensis BTBU20220184 was isolated from a sediment sample collected from Qinzhou Bay, Guangxi Province, China. It was cultured on a malt extract agar plate at 26 °C for 4 days. The genomic DNA of T. minnesotensis BTBU20220184 were extracted using a Fungi Genomic DNA Extraction Kit (Solarbio Life Sciences, Beijing, China). The β-tubulin gene sequence region was amplified using the conventional primer pair of Bt2a (5′-GGTAACCAAATCGGTGCTGCTTTC-3′) and Bt2b (5′-ACCCTCAGTGTAGTGACCCTTGGC-3′). PCR products were sent to Beijing Qingke Biotechnology Co., Ltd. (Beijing, China) for DNA sequencing. BTBU20220184 was identified as T. minnesotensis by comparing the β-tubulin gene sequence with the GenBank database using the BLAST program. A neighbor-joining (NJ) tree (Figure S38) was constructed using a software package, Mega version 5 [28]. The fungus was assigned the accession number BTBU20220184 in the culture collection at Beijing Technology and Business University, Beijing.
The strain T. minnesotensis BTBU20220184 was inoculated on a potato dextrose agar plate and cultured for 5 days. Then, a slit of agar with fungus was cut from the plate and inoculated into 250 mL conical flasks as seeds and cultured at 28 °C (180 rpm) for 3 days. Then, 5 mL of seed culture was added in to twenty 1 L conical flasks, each containing a solid medium consisting of rice (160 g) and artificial seawater (3.5%; 140 mL), and the flasks were incubated under static conditions at 26 °C for 29 days. The cultures were extracted three times with a mixture of EtOAc:MeOH (80:20), and the combined extracts were evaporated in vacuo to yield a residue. The residue was suspended in 800 mL distilled water and partitioned with isometric EtOAc. Then, the EtOAc layer was dried in vacuo to give a dark residue (57.30 g). The EtOAc fraction was fractionated by a vacuum liquid silica gel chromatography (90 × 400 mm column, Silica gel 60 H for thin-layer chromatography) using a stepwise gradient of 50–100% hexane/CH2Cl2 and then 0–50% MeOH/CH2Cl2 to afford 17 fractions (A–Q). Fraction I was subjected to a Sephadex LH-20 column using an isocratic elution of CH2Cl2:MeOH (2:1) to yield twelve subfractions (I1–I12), and subfraction I8 was further fractionated by HPLC (Agilent ZORBAX RX-C8, 9.4 × 250 mm, 5 μm column, 3.0 mL/min, elution with 30% to 45% acetonitrile/H2O) to yield 1 (2.5 mg) and 2 (4.2 mg). Subfraction I12 was further fractionated by HPLC (Agilent ZORBAX RX-C8, 9.4 × 250 mm, 5 μm column, 3.0 mL/min, elution with 25% to 45% acetonitrile/H2O) to yield 3 (1.7 mg) and 11 (2.3 mg). Fraction J was fractionated on a Sephadex LH-20 column using an isocratic elution of CH2Cl2:MeOH (2:1) to yield fifteen subfractions (J1–J15). Subfraction J6 was further fractionated by HPLC (Agilent ZORBAX RX-C8, 9.4 × 250 mm, 5 μm column, 3.0 mL/min, elution with 10% to 85% acetonitrile/H2O) to yield 8 (7.0 mg), 7 (11.0 mg), 10 (10.2 mg), and 9 (6.8 mg). Subfraction J11 was further fractionated by HPLC (Agilent ZORBAX RX-C8, 9.4 × 250 mm, 5 μm column, 3.0 mL/min, elution with 15% to 62% acetonitrile/H2O) to yield 12 (7.0 mg). Fraction K was fractionated on a Sephadex LH-20 column using an isocratic elution of CH2Cl2:MeOH (2:1) to give nine subfractions (K1–K9). Subfraction K4 was further fractionated by HPLC (Agilent ZORBAX RX-C8, 9.4 × 250 mm, 5 μm column, 3.0 mL/min, elution with 19% to 81% acetonitrile/H2O) to yield 4 (2.1 mg) and 5 (2.6 mg). Subfraction K9 was further fractionated by HPLC (Agilent ZORBAX RX-C8, 9.4 × 250 mm, 5 μm column, 3.0 mL/min, elution with 15% to 36% acetonitrile/H2O) to yield 6 (2.5 mg).
Talamitone A (1): Light-yellow oil; [ α ] D 25 + 63.5 (c 0.00625, MeOH); 1H and 13C NMR data, Table 1; HRESIMS m/z 385.0923 [M + H]+ (calcd for C20H17O8, 385.0918).
Talamitone B (2): Light-yellow powder; [ α ] D 25 + 16.0 (c 0.025, MeOH); 1H and 13C NMR data, Table 1; HRESIMS m/z 403.1011 [M + H]+ (calcd for C20H19O9, 403.1024).
Demethyltalamitone B (3): Light-yellow powder; [ α ] D 25 + 12.0 (c 0.025, MeOH); 1H and 13C NMR data, Table 1; HRESIMS m/z 389.0871 [M + H]+ (calcd for C19H17O9, 389.0867).
Talamiisocoumaringlycoside A (4): Light yellow powder; [ α ] D 25 + 8.0 (c 0.1, MeOH); 1H and 13C NMR data, Table 2; HRESIMS m/z 451.1204 [M + Na]+ (calcd for C19H24O11Na, 451.1211).
Talamiisocoumaringlycoside B (5): Light yellow powder; [ α ] D 25 + 26.0 (c 0.1, MeOH); 1H and 13C NMR data, Table 2; HRESIMS m/z 485.0830 [M + Na]+ (calcd for C19H23ClO11Na, 485.0821).
Talaminaphtholglycoside (6): Light yellow powder; [ α ] D 25 + 12.0 (c 0.025, MeOH); 1H and 13C NMR data, Table 2; HRESIMS m/z 395.1338 [M + H]+ (calcd for C19H23O9, 395.1337).
Mucorisocoumarin B (7): 1H NMR (500 MHz, DMSO-d6) δH: 11.02 (1H, s), 6.60 (1H, d, J = 7.0 Hz), 6.53 (1H, s), 6.51 (1H, d, J = 7.0 Hz), 3.96 (1H, m), 3.85 (3H, s), 3.80 (1H, m), 2.64 (1H, dd, J = 14.5, 4.0 Hz), 2.46 (1H, dd, J = 14.5, 8.5 Hz), 1.58 (1H, m), 1.43 (1H, m), 1.07 (3H, d, J = 6.0 Hz); 13C NMR (125 MHz, DMSO-d6) δC: 165.7 (C-1), 155.7 (C-3), 105.7 (C-4), 139.7 (C-4a), 101.1 (C-5), 166.5 (C-6), 100.3 (C-7), 162.5 (C-8), 99.4 (C-8a), 41.3 (C-9), 66.2 (C-10), 46.0 (C-11), 64.1 (C-12), 23.7 (C-13), 55.9 (6-OCH3); ESIMS m/z 293.1 [M − H].
Diaportinol (8): 1H NMR (500 MHz, DMSO-d6) δH: 10.97 (1H, s), 6.59 (1H, d, J = 2.5 Hz), 6.54 (1H, s), 6.51 (1H, d, J = 2.5 Hz), 4.89 (1H, d, J = 5.0 Hz), 4.73 (1H, t, J = 5.5 Hz), 3.85 (3H, s), 3.81 (1H, m), 3.40 (1H, m), 3.30 (1H, m), 2.74 (1H, dd, J = 14.5, 3.5 Hz), 2.39 (1H, dd, J = 14.5, 9.0 Hz); 13C NMR (125 MHz, DMSO-d6) δC: 165.6 (C-1), 155.9 (C-3), 105.5 (C-4), 139.7 (C-4a), 101.1 (C-5), 166.5 (C-6), 100.3 (C-7), 162.5 (C-8), 99.4 (C-8a), 37.8 (C-9), 69.0 (C-10), 65.4 (C-11), 55.9 (6-OCH3); ESIMS m/z 267.1 [M + H]+.
Peniisocoumarin E (9): 1H NMR (500 MHz, DMSO-d6) δH: 10.93 (1H, s), 6.65 (1H, s), 6.63 (1H, d, J = 2.5 Hz), 6.55 (1H, d, J = 2.5 Hz), 4.30 (1H, m), 3.85 (3H, s), 3.66 (2H, m), 3.17 (1H, dd, J = 15.0, 10.0 Hz), 2.78 (1H, dd, J = 15.0, 4.0 Hz); 13C NMR (125 MHz, DMSO-d6) δC: 165.1 (C-1), 153.7 (C-3), 106.2 (C-4), 139.2 (C-4a), 101.5 (C-5), 166.6 (C-6), 100.8 (C-7), 162.6 (C-8), 99.4 (C-8a), 37.8 (C-9), 60.2 (C-10), 65.0 (C-11), 56.0 (6-OCH3); ESIMS m/z 285.1 [M + H]+.
Dichlorodiaportin (10): 1H NMR (500 MHz, DMSO-d6) δH: 10.97 (1H, s), 6.62 (1H, s), 6.61 (1H, d, J = 2.5 Hz), 6.54 (1H, d, J = 2.5 Hz), 6.31 (1H, d, J = 3.0 Hz), 6.13 (1H, d, J = 6.0 Hz), 4.18 (1H, m), 3.85 (3H, s), 2.88 (1H, dd, J = 14.5, 3.0 Hz), 2.64 (1H, dd, J = 14.5, 9.5 Hz); 13C NMR (125 MHz, DMSO-d6) δC: 165.3 (C-1), 153.7 (C-3), 106.3 (C-4), 139.4 (C-4a), 101.4 (C-5), 166.5 (C-6), 100.5 (C-7), 162.5 (C-8), 99.5 (C-8a), 36.4 (C-9), 72.2 (C-10), 77.0 (C-11), 56.0 (6-OCH3); ESIMS m/z 319.1 [M + H]+.
Orsellinic acid (11): 1H NMR (500 MHz, CD3OD) δH: 6.19 (1H, d, J = 2.5 Hz), 6.13 (1H, d, J = 2.5 Hz), 2.49 (3H, s); 13C NMR (125 MHz, CD3OD) δC: 105.1 (C-1), 163.6 (C-2), 101.7 (C-3), 166.9 (C-4), 112.2 (C-5), 145.2 (C-6), 175.2 (C-7), 24.3 (8-CH3); ESIMS m/z 169.1 [M + H]+.
4-(Hydroxymethyl)-5-hydroxy-2H-pyran-2-one (12): 1H NMR (500 MHz, DMSO-d6) δH: 8.03 (1H, s), 6.33 (1H, s), 4.28 (2H, s); 13C NMR (125 MHz, DMSO-d6) δC: 174.0 (C-2), 109.8 (C-3), 145.7 (C-4), 168.1 (C-5), 139.3 (C-6), 59.5 (C-7).; ESIMS m/z 143.0 [M + H]+.

3.3. Biological Activity

Compounds 112 were evaluated against S. aureus, E. coli, and C. albicans for their antimicrobial activities in 96-well plates according to the Antimicrobial Susceptibility Testing Standards outlined by the Clinical and Laboratory Standards Institute document M07-A7 (CLSI) [29] and our previous report [30,31]. The MIC was defined as the minimum concentration of the compound that prevented visible growth of the microbes. The synergistic antimicrobial activities were determined by combing one-fourth of the minimum inhibitory concentration of the positive control (MIC of methicillin was 0.5 μg/mL) for S. aureus with two-fold diluted tested compounds. The synergistic antimicrobial activities were calculated based on a previous report [32]. Briefly, final concentrations were preprepared from 1.5625 to 100 μg/mL of isolated compounds. Methicillin was added by a column, while the tested compounds were diluted 2-fold by row in a 96-well microtiter plate. The fractional inhibitory concentration index (FICI) was calculated as the sum of the MIC of each drug when used in combination, divided by the MIC of the drug used alone. Synergistic antibacterial activity was defined by FICI ≤ 0.5.

4. Conclusions

In summary, 12 compounds, including 6 new compounds, talamitones A and B (1 and 2), demethyltalamitone B (3), talamiisocoumaringlycosides s A and B (4 and 5), and talaminaphtholglycoside (6), together with 6 previous reported compounds, mucorisocoumarin B (7), diaportinol (8), peniisocoumarin E (9), dichlorodiaportin (10), orsellinic acid (11), and 4-(hydroxymethyl)-5-hydroxy-2H-pyran-2-one (12), were isolated from the marine sediment-derived fungus T. minnesotensis BTBU20220184. The structures of the new compounds were characterized by HREIMS and NMR data analysis. All of the compounds were screened for activity against S. aureus, E. coli, and C. albicans. Compound 8 exhibited weak antibacterial activity against S. aureus at a concentration of 200 μg/mL. Compounds 5, 6, and 9 showed synergistic antibacterial activity against S. aureus at concentrations of 25, 50, and 12.5 μg/mL with 0.125 μg/mL of methicillin.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/md22060237/s1, Figures S1–S6: HRESIMS, 1H, 13C, HSQC, 1H–1H COSY, and HMBC spectra for compound 1; Figures S7–S12: HRESIMS, 1H, 13C, HSQC, 1H–1H COSY, and HMBC spectra for compound 2; Figures S13–S18: HRESIMS, 1H, 13C, HSQC, 1H–1H COSY, and HMBC spectra for compound 3; Figures S19–S24: HRESIMS, 1H, 13C, HSQC, 1H–1H COSY, and HMBC spectra for compound 4; Figures S25–S30: HRESIMS, 1H, 13C, HSQC, 1H–1H COSY, and HMBC spectra for compound 5; Figures S31–S37: HRESIMS, 1H, 13C, HSQC, 1H–1H COSY, HMBC, and ROESY spectra for compound 6; Figure S38: Phylogenetic tree of BTBU20220184..

Author Contributions

Conceptualization, F.S. and W.W.; methodology, X.X. and N.Y.; validation, W.W. and J.W.; formal analysis, W.W. and F.S.; investigation, W.W.; resources, L.W., R.J. and J.W.; data curation, F.S.; writing—original draft preparation, W.W. and X.X.; writing—review and editing, X.X., F.S., N.Y. and R.J.; visualization, X.X.; supervision, X.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the grants of the Guangxi Key Laboratory of Beibu Gulf Marine Resources, Environment and Sustainable Development (MRESD-2023-C02), the Qingdao Natural Science Foundation, China (23-2-1-174-zyyd-jch), and the National Natural Science Foundation of China (81973204).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Manohar, C.S.; Raghukumar, C. Fungal diversity from various marine habitats deduced through culture-independent studies. FEMS Microbiol. Lett. 2013, 341, 69–78. [Google Scholar] [CrossRef]
  2. Behera, A.D.; Das, S. Ecological insights and potential application of marine filamentous fungi in environmental restoration. Rev. Environ. Sci. Bio. 2023, 22, 281–318. [Google Scholar] [CrossRef]
  3. Jones, E.B.G.; Suetrong, S.; Sakayaroj, J.; Bahkali, A.H.; Abdel-Wahab, M.A.; Boekhout, T.; Pang, K.L. Classification of marine Ascomycota, Basidiomycota, Blastocladiomycota and Chytridiomycota. Fungal Divers. 2015, 73, 1–72. [Google Scholar] [CrossRef]
  4. Stolk, A.C.; Samson, R.A. The genus Talaromyces–studies on Talaromyces and related genera II. Stud. Mycol. 1972, 2, 1–65. [Google Scholar]
  5. Wang, X.-C.; Zhuang, W.-Y. New species of Talaromyces (Trichocomaceae, Eurotiales) from Southwestern China. J. Fungi 2022, 8, 647. [Google Scholar] [CrossRef]
  6. Zang, W.; Li, M.; Sun, J.Q.; Gao, C.H.; Wang, L. Two new species of Talaromyces Sect. Trachyspermi Discovered in China. Mycopathologia 2023, 188, 793–804. [Google Scholar] [CrossRef]
  7. 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]
  8. Han, P.J.; Sun, J.Q.; Wang, L. Two new sexual Talaromyces species discovered in etuary soil in China. J. Fungi 2021, 8, 36. [Google Scholar] [CrossRef]
  9. Wei, S.; Xu, X.; Wang, L. Four new species of Talaromyces section Talaromyces discovered in China. Mycologia 2021, 113, 492–508. [Google Scholar] [CrossRef]
  10. Yilmaz, N.; López-Quintero, C.A.; Vasco-Palacios, A.M.; Frisvad, J.C.; Theelen, B.; Boekhout, T.; Samson, R.A.; Houbraken, J. Four novel Talaromyces species isolated from leaf litter from Colombian Amazon rain forests. Mycol. Prog. 2016, 15, 1041–1056. [Google Scholar] [CrossRef]
  11. Luo, Y.; Lu, X.H.; Bi, W.; Liu, F.; Gao, W.W. Talaromyces rubrifaciens, a new species discovered from heating, ventilation and air conditioning systems in China. Mycologia 2016, 108, 773–779, Erratum in Mycologia 2019, 111, 1072. https://doi.org/10.1080/00275514.2019.1678303. [Google Scholar] [CrossRef]
  12. Lan, D.; Wu, B. Chemistry and bioactivities of secondary metabolites from the genus Talaromyces. Chem. Biodivers. 2020, 17, e2000229. [Google Scholar] [CrossRef] [PubMed]
  13. Zheng, M.J.; Xiao, Y.; Li, Q.; Lai, Y.X.; Dai, B.B.; Zhang, M.; Kang, X.; Tong, Q.Y.; Wang, J.P.; Chen, C.M.; et al. Cytotoxic ergosteroids from a strain of the fungus Talaromyces adpressus. J. Nat. Prod. 2023, 86, 2081–2090. [Google Scholar] [CrossRef]
  14. Li, Q.; Zhang, M.; Zhang, X.T.; Li, L.Q.; Zheng, M.J.; Kang, J.B.; Liu, F.; Zhou, Q.; Li, X.N.; Sun, W.G.; et al. Talaroclauxins A and B: Duclauxin-ergosterol and duclauxin-polyketide hybrid metabolites with complicated skeletons from Talaromyces stipitatus. Chin. Chem. Lett. 2024, 35, 108193. [Google Scholar] [CrossRef]
  15. Cai, J.; Zhou, X.M.; Wang, B.; Zhang, X.L.; Luo, M.Y.; Huang, L.T.; Wang, R.X.; Chen, Y.H.; Li, X.Y.; Luo, Y.P.; et al. Bioactive polyketides and meroterpenoids from the mangrove-derived fungus Talaromyces flavus TGGP35. Front. Microbiol. 2024, 15, 1342843. [Google Scholar] [CrossRef]
  16. Yang, W.; Zhang, B.; Tan, Q.; Chen, Y.; Chen, T.; Zou, G.; Sun, B.; Wang, B.; Yuan, J.; She, Z. 4-Hydroxy-2-pyridone derivatives with antitumor activity produced by mangrove endophytic fungus Talaromyces sp. CY-3. Eur. J. Med. Chem. 2024, 269, 116314. [Google Scholar] [CrossRef] [PubMed]
  17. Zheng, M.J.; Zhou, C.X.; Liao, H.; Li, Q.; Bao, A.L.; Chen, C.M.; He, F.; Wu, P.; Sun, W.G.; Zhu, H.C.; et al. Enantiomeric α-pyrone derivatives with immunosuppressive activity from Talaromyces adpressus. Phytochemistry 2024, 218, 113931. [Google Scholar] [CrossRef]
  18. Nicoletti, R.; Bellavita, R.; Falanga, A. The outstanding chemodiversity of marine-derived Talaromyces. Biomolecules 2023, 13, 1021. [Google Scholar] [CrossRef] [PubMed]
  19. Song, F.; Dong, Y.; Wei, S.; Zhang, X.; Zhang, K.; Xu, X. New antibacterial secondary metabolites from a marine-derived Talaromyces sp. Strain BTBU20213036. Antibiotics 2022, 11, 222. [Google Scholar] [CrossRef]
  20. Zhang, K.; Zhang, X.; Lin, R.; Yang, H.; Song, F.; Xu, X.; Wang, L. New secondary metabolites from the marine-derived fungus Talaromyces mangshanicus BTBU20211089. Mar. Drugs 2022, 20, 79. [Google Scholar] [CrossRef]
  21. Cai, R.; Wu, Y.; Chen, S.; Cui, H.; Liu, Z.; Li, C.; She, Z. Peniisocoumarins A–J: Isocoumarins from Penicillium commune QQF-3, an endophytic fungus of the mangrove plant Kandelia candel. J. Nat. Prod. 2018, 81, 1376–1383. [Google Scholar] [CrossRef] [PubMed]
  22. Pettit, G.R.; Bond, T.J.; Herald, D.L.; Penny, M.; Doubek, D.L.; Williams, M.D.; Pettit, R.K.; Hooper, J.N.A. Isolation and structure of spongilipid from the Republic of Singapore marine porifera Spongiacf. hispida. Can. J. Chem. 1997, 75, 920–925. [Google Scholar] [CrossRef]
  23. Feng, C.-C.; Chen, G.-D.; Li, X.-X.; Tang, J.-S.; Liu, X.-Z.; Li, Y.; Yao, X.-S.; Gao, H. Three new glucosides from a cold-adapted fungal strain Mucor sp. J. Asian Nat. Prod. Res. 2013, 15, 921–927. [Google Scholar] [CrossRef] [PubMed]
  24. Mallampudi, N.A.; Choudhury, U.M.; Mohapatra, D.K. Total synthesis of (−)-citreoisocoumarin, (−)-citreoisocoumarinol, (−)-12-epi-citreoisocoumarinol, and (−)-mucorisocoumarins A and B using a gold(I)-catalyzed cyclization strategy. J. Org. Chem. 2020, 85, 4122–4129. [Google Scholar] [CrossRef]
  25. Larsen, T.O.; Breinholt, J. Dichlorodiaportin, diaportinol, and diaportinic acid:  three novel isocoumarins from Penicillium nalgiovense. J. Nat. Prod. 1999, 62, 1182–1184. [Google Scholar] [CrossRef] [PubMed]
  26. Lopes, T.I.B.; Coelho, R.G.; Yoshida, N.C.; Honda, N.K. Radical-scavenging activity of Orsellinates. Chem. Pharm. Bull. 2008, 56, 1551–1554. [Google Scholar] [CrossRef] [PubMed]
  27. Lin, A.Q.; Lu, X.M.; Fang, Y.C.; Zhu, T.J.; Gu, Q.Q.; Zhu, W.M. Two new 5-hydroxy-2-pyrone derivatives isolated from a marine-derived fungus Aspergillus flavus. J. Antibiot. 2008, 61, 245–249. [Google Scholar] [CrossRef] [PubMed]
  28. Tamura, K.; Peterson, D.; Peterson, N.; Stecher, G.; Nei, M.; Kumar, S. MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol. Biol. Evol. 2011, 28, 2731–2739. [Google Scholar] [CrossRef]
  29. Clinical and Laboratory Standards Institute. Methods for Dilution Antimicrobial Susceptibility Tests for Bacteria That Grow Aerobically, 7th ed.; approved standard; Clinical and Laboratory Standards Institute: Wayne, PA, USA, 2008. [Google Scholar]
  30. Han, J.; Yang, N.; Wei, S.; Jia, J.; Lin, R.; Li, J.; Bi, H.; Song, F.; Xu, X. Dimeric hexylitaconic acids from the marine-derived fungus Aspergillus welwitschiae CUGBMF180262. Nat. Prod. Res. 2022, 36, 578–585. [Google Scholar] [CrossRef]
  31. Wang, Q.; Song, F.; Xiao, X.; Huang, P.; Li, L.; Monte, A.; Abdel-Mageed, W.M.; Wang, J.; Guo, H.; He, W.; et al. Abyssomicins from the South China Sea deep-sea sediment Verrucosispora sp.: Natural thioether Michael addition adducts as antitubercular prodrugs. Angew. Chem. Int. Ed. Engl. 2013, 52, 1231–1234. [Google Scholar] [CrossRef]
  32. Han, J.J.; Wang, H.Y.; Zhang, R.; Dai, H.Q.; Chen, B.S.; Wang, T.; Sun, J.Z.; Wang, W.Z.; Song, F.H.; Li, E.R.; et al. Cyclic tetrapeptides with synergistic antifungal activity from the fungus Aspergillus westerdijkiae using LC-MS/MS-based molecular networking. Antibiotics 2022, 11, 166. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of 112.
Figure 1. Chemical structures of 112.
Marinedrugs 22 00237 g001
Figure 2. Key COSY (bold lines) and HMBC (arrows) correlations in 16.
Figure 2. Key COSY (bold lines) and HMBC (arrows) correlations in 16.
Marinedrugs 22 00237 g002
Table 1. 1H (500 MHz) and 13C NMR (125 MHz) data of 13 (CD3OD).
Table 1. 1H (500 MHz) and 13C NMR (125 MHz) data of 13 (CD3OD).
Pos123
δC, TypeδH, (J in Hz)δC, TypeδH, (J in Hz)δC, TypeδH, (J in Hz)
1198.5, C 201.0, C 201.2, C
2136.7, C 38.8, CH23.06, t (7.0)38.8, CH23.06, t (7.0)
3154.0, C 202.8, C 202.9, C
3a134.8, C 136.8, C 136.5, C
4152.6, C 155.2, C 155.2, C
5126.8, CH7.50, d (1.5)121.7, C7.68, d (1.0)121.8, CH7.68, d (1.5)
6134.4, C 132.8, C 133.7, C
7115.4, CH7.51, d (1.5)122.0, C8.07, d (1.0)122.0, CH8.08, d (1.5)
7a133.1, C 137.3, C 137.3, C
820.7, CH22.44, m20.4, CH21.90, quint (7.0)20.5, CH21.90, quint (7.0)
932.9, CH22.46, m33.7, CH22.35, t (7.0)33.9, CH22.35, t (7.0)
10175.3, C 175.5, C 177.2, C
11168.9, C 168.6, C 169.3, C
1′110.7, C 112.2, C 112.2, C
2′156.7, C 163.3, C 163.3, C
3′107.6, CH6.39, d (8.0)108.1, CH6.28, d (8.5)108.1, CH6.28, d (8.0)
4′130.9, CH7.04, t (8.0)137.4, CH7.21, t (8.5)137.4, CH7.21, t (8.0)
5′107.6, CH6.39, d (8.0)108.1, CH6.28, d (8.5)108.1, CH6.28, d (8.0)
6′156.7, C 163.3, C 163.3, C
10-OCH352.1, CH33.60, s52.1, CH33.64, s
Table 2. 1H (500 MHz) and 13C NMR (125 MHz) data of 46 (CD3OD).
Table 2. 1H (500 MHz) and 13C NMR (125 MHz) data of 46 (CD3OD).
Pos456
δC, TypeδH, (J in Hz)δC, TypeδH, (J in Hz)δC, TypeδH, (J in Hz)
1167.9, C 167.5, C 110.0, CH6.89, s
2- - 158.4, C
3157.9, C 154.0, C 119.5, C
4101.2, CH6.94, s103.6, CH7.04, s131.3, CH9.17, s
4a133.5, C 133.0, C 118.6, C
5132.9, C 133.2, C 154.0, C
6161.1, C 161.1, C 120.7, C
799.7, CH6.63, s100.1, CH6.66, s160.5, C
8161.8, C 161.9, C 104.4, CH6.76, s
8a99.3, C 99.3, C 140.4, C
9a30.7, CH22.63, m39.4, CH23.15, dd (15.0, 4.0)207.0, C
9b 2.88, dd (15.0, 9.5)
1030.8, CH21.90, tt (7.0)60.4, CH4.32, m27.2, CH32.76, s
1161.7, CH23.63, m66.8, CH23.78, m10.9, CH32.34, s
1′105.9, CH4.70, d (7.5)106.0, CH4.70, d (7.5)107.1, CH4.78, d (7.5)
2′75.6, CH3.50, m75.6, CH3.51, dd (9.0, 7.5)75.8, CH3.69, m
3′77.8, CH3.42, m77.8, CH3.44, m78.2, CH3.47, m
4′71.2, CH3.42, m71.2, CH3.44, m71.4, CH3.47, m
5′78.2, CH3.17, m78.1, CH3.17, m77.9, CH3.13, m
6′a62.5, CH23.76, dd (11.5, 2.5)62.4, CH23.76, dd (11.5, 2.5)62.6, CH23.76, dd (11.5, 2.5)
6′b 3.66, dd (11.5, 5.0) 3.67, dd (11.5, 4.5) 3.67, m
6-OCH356.8, CH33.94, s56.9, CH33.94, s
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

Wang, W.; Wang, J.; Song, F.; Jia, R.; Wang, L.; Xu, X.; Yang, N. New Secondary Metabolites from Marine-Derived Fungus Talaromyces minnesotensis BTBU20220184. Mar. Drugs 2024, 22, 237. https://doi.org/10.3390/md22060237

AMA Style

Wang W, Wang J, Song F, Jia R, Wang L, Xu X, Yang N. New Secondary Metabolites from Marine-Derived Fungus Talaromyces minnesotensis BTBU20220184. Marine Drugs. 2024; 22(6):237. https://doi.org/10.3390/md22060237

Chicago/Turabian Style

Wang, Weiliang, Jingjing Wang, Fuhang Song, Renming Jia, Long Wang, Xiuli Xu, and Na Yang. 2024. "New Secondary Metabolites from Marine-Derived Fungus Talaromyces minnesotensis BTBU20220184" Marine Drugs 22, no. 6: 237. https://doi.org/10.3390/md22060237

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