Next Article in Journal
Exploring the Efficacy of Hydroxybenzoic Acid Derivatives in Mitigating Jellyfish Toxin-Induced Skin Damage: Insights into Protective and Reparative Mechanisms
Previous Article in Journal
Direct Degradation of Fresh and Dried Macroalgae by Agarivorans albus B2Z047
Previous Article in Special Issue
Marine-Fungi-Derived Gliotoxin Promotes Autophagy to Suppress Mycobacteria tuberculosis Infection in Macrophage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Antibacterial Polyketides from the Deep-Sea Cold-Seep-Derived Fungus Talaromyces sp. CS-258

1
CAS and Shandong Province Key Laboratory of Experimental Marine Biology, Institute of Oceanology, Chinese Academy of Sciences, Nanhai Road 7, Qingdao 266071, China
2
University of Chinese Academy of Sciences, Yuquan Road 19A, Beijing 100049, China
3
Laboratory for Marine Biology and Biotechnology, Qingdao Marine Science and Technology Center, Wenhai Road 1, Qingdao 266237, China
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2024, 22(5), 204; https://doi.org/10.3390/md22050204
Submission received: 29 March 2024 / Revised: 25 April 2024 / Accepted: 25 April 2024 / Published: 28 April 2024
(This article belongs to the Special Issue Bioactive Compounds from the Deep-Sea-Derived Microorganisms 2.0)

Abstract

:
Thirty-two fungal polyketide derivatives, including eleven new compounds, namely (3R,5′R)-5-hydroxytalaroflavone (1), talaroisochromenols A–C (3, 5, and 11), (8R,9R,10aR)-5-hydroxyaltenuene (13), (8R,9R,10aS)-5-hydroxyaltenuene (14), (8R,9S,10aR)-5-hydroxyaltenuene (15), nemanecins D and E (25 and 26), 2,5-dimethyl-8-iodochromone (27), and talarofurolactone A (29), together with one new naturally occurring but previously synthesized metabolite, 6-hydroxy-4-methoxycoumarin (28), were isolated and identified from the deep-sea cold-seep-derived fungus Talaromyces sp. CS-258. Among them, racemic ((±)-11) or epimeric (1315, 25, and 26) mixtures were successfully separated by chiral or gradient elution HPLC. Meanwhile, compound 27 represents a rarely reported naturally occurring iodinated compound. Their planar structures as well as absolute configurations were determined by extensive analysis via NMR, MS, single-crystal X-ray diffraction, Mosher’s method, and ECD or NMR calculation (with DP4+ probability analysis). Possible biosynthetic routes of some isolated compounds, which are related to chromone or isochromone biosynthetic pathways, were put forward. The biological analysis results revealed that compounds 7, 9, 10, 1822, 24, 30, and 31 showed broad-spectrum antibacterial activities against several human and aquatic pathogens with MIC ranges of 0.5–64 μg/mL.

1. Introduction

Polyketides are a big family of secondary metabolites generally produced from a common origin of acetyl-CoA polymerization by connecting acetic acid units via condensation reactions and intermediated by a non-reducing group of iterative polyketide synthases (PKSs) [1,2]. Fungal polyketides have attracted considerable attention due to their unique molecular scaffolds and diverse pharmacological activities [3,4,5]. Among them, a series of fungal polyketides, such as griseofulvin, brefeldin, and hypomycetin, have been selected as promising candidates for new antibiotic lead compounds and possess great commercial values [1,3,5].
Deep-sea cold-seep-derived fungi have formed a unique physiological process and metabolic mechanism during their adaptation to the typical chemosynthetic-driven ecosystems with hydrocarbon-rich fluid, strikingly emerging as a promising source for their prolific production of structurally diverse secondary metabolites with various biological properties in very recent years [6,7,8]. Increasing efforts have been made on deep-sea cold-seep fungi to seek and develop biologically active pharmaceutical molecules with antibiotic [7,9,10], antioxidant [11], phytotoxic [12], anti-inflammatory [13], cytotoxic [14], proangiogenic [15,16], and enzyme inhibitory [17] properties.
In the course of forwarding our program on discovering bioactive metabolites from deep-sea cold seep-derived fungi [6,15,16], a fungal strain Talaromyces sp. CS-258, which was isolated from a mussel sample collected in a cold seep area in the northeast of the South China Sea at a depth of −1200 m, displayed diverse classes of polyketide derivatives and exhibited antimicrobial activity in a preliminary screening. Intensive chemical investigations on the fermentation broth of the fungus resulted in the isolation and identification of 32 polyketides (Figure 1), including 11 new chromone- or isochromone-derived analogs (compounds 1, 3, 5, 11, 1315, 2527, and 29) and 1 new naturally occurring coumarin (28) that was previously obtained by chemical synthesis [18]. The isolated compounds were assayed for their antimicrobial activities against pathogenic microbes, which demonstrated that eleven polyketides, including 7, 9, 10, 1822, 24, 30, and 31, possessed broad-spectrum antibiotic activities against human and aquatic pathogenic bacteria with MIC values ranging from 0.5 to 64 μg/mL. Herein, the details of isolation, structural elucidation, plausible biosynthetic pathways, and bioactivity assays of these compounds are described. The structure–activity relationships (SARs) for these antibacterial polyketides have been briefly discussed in this paper.

2. Results and Discussion

2.1. Structure Elucidation

The molecular formula of compound 1 was given as C13H10O6 from HRESIMS data with nine degrees of unsaturation. A comprehensive analysis of its 1H, 13C, and DEPT NMR spectra (Table 1) revealed great similarity to talaroflavone (2), a previously described compound isolated from the sponge-derived fungus Alternaria sp. F49 [19]. However, the signals of a methoxy group at δC/H 55.8/3.75 at C-5 in the NMR spectra of 2 were absent in those of 1. In comparison to 2, obvious upfield shifts for C-5 in 1 were detected. These differences suggested that compound 1 was a 5-demethylation derivative of 2, which was further supported by the HMBC correlations of H-4/H-6 to C-5 (Figure 2). From a biosynthetic point of view, the stereoscopic configuration of 1 was assumed to be the same as that of 2. After slow evaporation of the MeOH solvent, single crystals of 2 were obtained. The relative configuration of 2 was thus assigned by the result of a single-crystal X-ray diffraction experiment using Cu Kα radiation as 3S*, 5′S* (Figure 3). Accordingly, the relative configuration of 1 was also established as 3S*, 5′S*, the same as 2. The absolute configuration of 1 was further studied by the time-dependent density functional (TDDFT) ECD calculation. The calculated ECD curve of (3R,5′R)-1 agreed well with the experimental ECD curve of 1 (Figure 4), which ascertained compound 1 as (3R,5′R)-5-hydroxytalaroflavone.
The protonated molecular ion peak at m/z 261.0398 [M − H] in HRESIMS revealed the molecular formula of C13H10O6 for 3 with nine degrees of unsaturation. As shown in Table 1, the 1H NMR spectrum exhibited signals for two singlet aromatic protons (δH 3.26 and 2.94), one doublet methylene (δH 3.12 and 2.88), and one singlet methyl (δH 1.70). The 13C and DEPT NMR spectra data of 3 indicated the presence of 13 carbon signals, which were sorted into one methyl, one methylene, two methines, and nine quaternary carbons. Detailed analysis of the NMR data revealed that the structure of 3 was similar to that of compound 4, an altenusin analog bearing a 6/6/5 tricyclic ring skeleton obtained from the mangrove endophytic fungus Alternaria sp. SK6YW3L [20]. However, resonances for a methoxy group of 4 were absent in the NMR spectra of 3, which suggested that 3 was a 5-demethylated derivative of 4. The above deduction was supported by HMBC correlation from H-4 to C-5 (Figure 2). The absolute configuration of 3 was established as 9aS based on the quantum chemical calculation of ECD (Figure 4) and nominated as talaroisochromenol A.
Compound 5 was afforded as a yellow oil, and the molecular formula of C13H12O6 for 5 was assigned by its positive HRESIMS. The NMR spectra of 5 were identical to those of 6, a known altenusin derivative with a 6/6/5 tricyclic ring [20], except for the obvious differences in the chemical shifts of CH-8, CH-9, and CH3-10. The above observation suggested that 5 was a new epimer of 6 (epimeric at the C-9 position), which was further proved by the NOESY correlations from H-7 to H-9 and from H3-10 to H-8 (Figure 5). Hence, the absolute configuration of 5 was assigned as 7R, 8S, 9R, which was further confirmed by comparison of its calculated ECD for 7R, 8S, 9R-5 with the measured ECD spectra (Figure 4). Thus, compound 5 was acknowledged as a new compound, namely talaroisochromenol B.
Compound 11 possessed a molecular formula of C12H10O6 as established from its HRESIMS spectrum, accounting for eight degrees of unsaturation. In the 1H NMR spectra, resonances for two phenolic hydroxyl groups (δH 11.08 and 10.99), two meta-substituted aromatic protons (δH 6.62, d, J = 2.2 Hz; δH 6.46, d, J = 2.2 Hz), one singlet oxymethine (δH 5.49), a chemically nonequivalent methylene (δH 3.26 and 2.94), and one singlet methyl (δH 1.51) were observed. The investigation of 13C NMR and HSQC spectral data displayed attribution signals of two lactone carbonyl groups, six aromatic carbons (including two methines and four quaternary carbons), one oxygenated quaternary carbon, one oxygenated methine, one methylene, and one methyl. These data demonstrated great similarity to those of 12, a known polyketide yielded from an endolichenic fungus Ulocladium sp. [21], except for the absence of a methoxy group at C-8 (Table 2). Compared to 12, the obvious upfield shift of C-8 in 11 was observed. Thus, 11 was determined as an 8-hydroxylated derivative of 12.
The relative configuration of 11 was determined by the NOESY correlation between H-9b and H3-10 as 3aS*, 9aS* (Figure 5). The planar structure and relative configuration of 11 were further verified by an X-ray diffraction experiment (Figure 3). However, compound 11 was acquired as a racemate with an optical rotation value near zero, and no obvious Cotton effect was observed from the ECD spectrum. The chiral HPLC separation of 11 on a Chiralcel IG column (hexane/isopropanol = 80:20, flow rate of 1 mL/min) succeeded in separating two isomers (Figure S2). Finally, as shown in Figure 4, the ECD computation allowed the assignment of absolute configurations for two isomers as (+)-11-(3aR, 9aR) and (−)-11-(3aS, 9aS), respectively. Thus, 11 was named talaroisochromenol C.
Compounds 1315 were originally isolated as yellow oils and owned the same molecular formula of C14H14O6 based on their positive HRESIMS data. The UV spectra of 1315 showed similar characteristic absorptions for an isocoumarin chromophore at λmax 242, 282, and 323 nm [22]. Subsequent interpretation of their NMR data (Table 3) revealed that compounds 1315 were characterized as three new stereoisomers of a known analog 5-hydroxyaltenuene, which were isolated from an endophytic fungus Penicillium sp.FJ-1 of Ceriops tagal [22].
The relative configurations of 1315 were assigned by NOESY, X-ray single-crystal diffraction experiment, or NMR calculations with DP4+ probability analysis. As shown in Figure 5, NOESY correlations from Hα-10 to H-8 and H3-11 implied the α-orientation of CH3-11 and β-orientation of OH-8 in 13. Meanwhile, NOESY correlations from H3-11 to H-9/Hβ-10 and from H-8 to Hα-10 suggested a syn orientation of CH3-11 and H-9 and an opposite orientation of H-9 and H-8 in 14. The relative configurations of 13 and 14 were further confirmed by single-crystal X-ray diffraction experimental data as (8R*,9R*, 10aR*)-13 and (8S*,9S*,10aR*)-14 (Figure 3). Moreover, NOESY correlations of 15 from H3-11 to H-9 and Hα-10 established its relative configuration of 9S*, 10aR*. The experimental 1H and 13C NMR data of 15 were compared with the calculated NMR data of 15a and 15b (two possible isomers of 15, Figure 6) and matched well with those calculated for the isomer 15a (8R*,9S*,10aR*) with a DP4+ probability of 100% (Table S16). The absolute configurations of 1315 were determined by comparing the experimental ECD data with the calculated ones (Figure 4), suggesting 1315 to be (8R,9R,10aR)-, (8R,9R, 10aS)- and (8R,9S,10aR)-5-hydroxyaltenuenes, respectively. Notably, the orientation of the methyl group at C-11 significantly determined the ECD Cotton effects observed for 5-hydroxyaltenuene. This was supported by the observation that compounds 13 and 15, with 10aR absolute configurations, displayed positive Cotton effects at 236 nm and negative ones at 282 nm, in contrast to 14 with a 10aS absolute configuration. This was the first isolation of three new epimers of 5-hydroxyaltenuene from the fungus Talaromyces sp.
Compounds 25 and 26 were both obtained as yellow oily substances with similar chromatographic properties in the semipreparative HPLC separation [tR (25) = 14 min, tR (26) = 12 min]. The HRESIMS analysis displayed pseudomolecular ion peaks at m/z 255.1228 [M + H]+ (calcd for 255.1227) of 25 and at m/z 255.1226 [M + H]+ (calcd for 255.1227) of 26, assigning their molecular formula both as C13H18O5 holding five degrees of unsaturation. Their UV spectra exhibited absorption maxima at 209 and 320 nm, revealing structural analogies between 25 and 26. A detailed inspection of 1H and 13C NMR spectra of 25 and 26 (Table 4) prominently indicated practically identical signals, suggesting that they shared the same planar structural scaffold. Both 1H NMR spectral data of 25 and 26 exhibited characteristic resonances ascribed to two olefinic protons (represented for two trisubstituted double bands), three sp3-hybridized methines (including two oxygenated), two sp3-hybridized methylenes (including one chemically nonequivalent O-substituted), two methyl groups (as a singlet and a doublet), and three exchangeable protons. Analyses of their 13C, DEPT, and HSQC spectrum confirmed the presence of a carbonyl group, two pairs of olefin carbons, two oxygenated methines, one sp3-hybridized methine, two methylene groups, two methyls, and one quaternary carbon.
The 1H-1H COSY and HMBC correlations of 25 and 26 shown in Figure 2 established the same planar structure, which was similar to nemanecins A–C, three azaphilone analogs isolated from the culture broth of the fungus Nemania sp. BCC 30850 [23], indicating that 25 and 26 are a pair of stereoisomers. However, the signal of a singlet methyl in the 1H and 13C NMR spectra of nemanecins A–C were absent in those of 25 and 26. Instead, signals for a 2-hydroxypropyl group were observed in the NMR spectra of 25 and 26. The above inference was further confirmed by the key 1H-1H COSY correlation from H-10 to H-9 and H-11 as well as the observed HMBC correlations from H-9 to C-3, C-4, and C-10 and from H-11 to C-9 and C-10 (Figure 2). Thus, the planar structures of 25 and 26 were assigned.
The relative configurations of compounds 25 and 26 of C-7, C-8, and C-8a were both assigned as 7S*,8R*, and 8aR* based on the NOESY analysis and scalar coupling constant data (Figure 5). Both of them displayed a NOESY correlation from H3-10 to H-8a, which implied that H3-10 and H-8a were co-facial and arbitrarily assigned in β-orientation. The large coupling constant between H-8a and H-8 (J = 10.1 Hz) confirmed the trans-relationship of these protons.
The modified Mosher’s method using an NMR tube [24,25] was used to determine the absolute configuration of C-10 in 25 and 26, leading to the assignments of S- for 25 and R-absolute configuration for 26 based on Δδ values (Figure 7). ECD calculations were performed to further solve the absolution configurations of C-7, C-8, and C-8a in 25 and 26. The experimental ECD spectra of 25 and 26 both exhibited a strong positive Cotton effect at 214 nm and a negative one at 345 nm, which was in good agreement with the calculated ECD curve of (7S,8R,8aR,10S)-25 and (7S,8R,8aR,10R)-26 (Figure 5). The results of ECD calculations indicated that the configuration of C-10 made no contribution to the ECD Cotton effects. Thus, the structures of compounds 25 and 26 were finally identified and named nemanecins D and E, respectively.
An analysis of the HRESIMS spectrum for m/z 316.9668 [M + H]+ (calcd for 316.9669) determined the molecular formula of compound 27 (yellow oil) as C11H10O3I requiring seven degrees of unsaturation. Its 1H and 13CNMR data (Table 5) were in good accord with the NMR spectral information of the known chromone derivative 2,5-dimethychromone [26], except that the signal of the methine of C-8 resonating at δC 100.4 in the NMR spectra of 2,5-dimethychromone was replaced by a deprotonated carbon resonating at δC 75.3 in those of 27. The chemical shift of C-8 at δC 75.3 suggested the substitution by an iodine atom according to the previously reported compounds 2-iodo-5-methoxyphenol (δC 74.4) [27] and (aS)-6-iodofonsecinone A (δC 79.3) [28]. Thus, the structural assignment of 27 was assigned as 2,5-dimethy-8-iodochromone as shown in Figure 1, which was further confirmed by the HMBC correlations from H3-9 to C-2/C-3, H-3 to C-2/C-4a, and H-10 to C-4a/C-5/C-6 (Figure 2).
Compound 28 was obtained as a yellow oil, and its 1H and 13C NMR spectra data (Table 5) exactly matched those of the known compound 6-hydroxy-4-methoxycoumarin synthesized in the previous literature [18]. It is now isolated for the first time from a natural source. We supplemented its relevant NMR, MS, and UV data here.
The HRESIMS data established the molecular formula of compound 29 as C11H16O4 with four degrees of unsaturation. As shown in Table 5, the 1H and 13C NMR spectrum displayed signals corresponding to two quaternary carbons (one lactone and one olefinic), five methines (three olefinic and one oxygenated), three methylenes (one oxygenated), and one methyl. The 1H-1H COSY correlations from H-4 to H-5 and HMBC correlations from H2-4 to C-2 and C-3 as well as H-5 to C-4 established a dihydrofuran-2(3H)-one moiety. The assignment for a spin coupling system of [=CHCH2CH3] was identified by 1H-1H COSY correlations from H-2” to H-1” and H-3”. Furthermore, a 3,4-dihydroxybut-1-en-1-yl group was assigned by 1H-1H COSY correlations from H-2′ to H-1 and H-3′, from H-3′ to H-4′ and 3′-OH, and from H-4′ to 4′-OH. The moiety of [=CHCH2CH3] spin system was attached to C-3 outside the lactone ring with regard to the HMBC correlations from H-1″ to C-2 and C-4 and from H-2″ to C-3. In addition, the 3,4-dihydroxybut-1-en-1-yl group was bonded to C-5, as proven by the 1H-1H COSY correlation from H-1′ to H-5 and HMBC correlations from H-1′ to C-4 and C-5 and from H-2′ to C-5. The E-configuration of double bonds Δ1′, 2′ was assigned based on the large coupling constants of J = 15.5 Hz, while the cis relation of Δ3, 1″ was ascertained by NOESY correlation between H2-4 and H-1″ (Figure 8A).
The relative configuration of 29 was determined by NMR calculation with DP4+ analysis. As a result, the experimental NMR data of 29 corresponded to the computed NMR data for (5R*,3′S*)-isomer 29b (99.12% probability, Table S19 in the supplementary material). The absolute configuration of 29 was assigned by ECD calculation. The calculated ECD curves of (5R,3′S)-29 coincided with the measured curve, which demonstrated the absolute configuration of 29 as 5R, 3′S (Figure 8B). Thus, the structure of 29 was established, and it was named talarofuranone A trivially.
In addition, based on the comparison of their NMR and optical rotation data with those reported in the literature, the structures of the other sixteen polyketides were identified as phialophoriol (7) [29], (1S,3S)-2,3-dihydro-3,6,8-trihydroxy-1-methylcyclopenta[c][2]-benzopyran-5(1H)-one (8) [20], (S)-1,2-dihydro-6,8-dihydroxy-1-methylcyclo-penta[c][2]-benzopyran-3,5-dione (9) [20], (S)-1-deoxyrubralactone (10) [30], (8R,9R,10aR)-altenuene (16) [31], (8S,9S,10aR)-altenuene (17) [31], alternariol 9-O-methyl ether (18) [32,33], alternariol (19) [33,34], 3′-hydroxyalternariol-5-O-methyl ether (20) [34], 3-O-demethylaltenuisol (21) [35], 4-hydroxyalternariol (22) [36,37], alterlactone (23) [34], alteryulactone (24) [38], purpactin A (30) [39], penicillide (31) [39,40], and dehydroisopenicillide (32) [41]. Among them, compounds 124 are altenusin or alternariol derivatives with a similar polyketide origin. Based on the previously reported references, the possible biosynthetic routes of these polyketides are proposed here (Figure 9) [20,21,29,42,43,44,45]. All of them are derived from the heptaketide intermediate through the iterative condensation of acetyl-CoA and malonyl-CoA via PKSs, followed by a multi-step reaction of aldol-type cyclization, lactonization, methylation, hydroxylation, reduction, transmethylation, etc.

2.2. Biological Activity

Partially isolated compounds were evaluated for their in vitro antimicrobial activities against some pathogenic bacteria and fungi. Eleven bacterial strains including four human pathogens (methicillin-resistant Staphylococcus aureus, Pseudomonas aeruginosa, Escherichia coli, and Klebsiella pneumonia) and seven aquatic bacteria (Vibrio alginolyticus, Aeromonas hydrophilia, Micrococcus luteus, Vibrio anguillarum, Vibrio parahaemolyticus, Vibrio vulnificus, and Vibrio harveyi) together with six plant-pathogenic fungi (Ceratobasidium cornigerum, Penicillium digitatum, Physalospora piricola, Valsa mali, Colletotrichum gloeosporioides, and Fusarium oxysporum) were used as the tested strains. As shown in Table 6, Tables S1 and S2, eleven compounds, 7, 910, 1822, 24, 30, and 31, exhibited comparable broad-spectrum antibiotic effects with MIC values ranging from 0.5 to 64 μg/mL.
Notably, compound 10 displayed significant inhibitory activities against four bacteria, namely E. coli, A. hydrophilia, V. parahaemolyticus, and V. harveyi, with MIC values in the range of 0.5–1 μg/mL, better than or equivalent to the positive control chloramphenicol (MIC = 0.25–2 μg/mL). Compound 18 showed considerable antibacterial activities against human pathogen E. coli and aquatic pathogenic bacteria V. parahaemolyticus that were comparable to those of the positive control, chloramphenicol. Moreover, compounds 20, 22, and 24 displayed inhibitory effects on the growth of P. aeruginosa with V. vulnificus, A. hydrophilia with V. parahaemolyticus, and A. hydrophilia, respectively, with MIC values like those of chloramphenicol from 0.5 to 4 μg/mL. Unfortunately, all measured compounds did not show any significant antifungal activity at a concentration of 64 µg/mL.
The structure–activity relationship (SAR) for the above polyketides is discussed here. The addition of a ketonic carbonyl or methoxy group in the structures of methylcyclopenta[c][2]-benzopyran could improve their antibacterial effects (10 vs. 79). Then, a comparison of the inhibitory activities of compounds 1824 indicated that their activities were influenced by the number and position of methyl, hydroxyl, and methoxy groups, as well as the expansion of the lactone ring. Furthermore, a correlation was observed between the acetylation of the side chain and antibacterial efficacy for penicillide-type compounds (30 vs. 31).

3. Materials and Methods

3.1. General Experimental Procedures

The general experimental procedures in this study were similar to those previously reported [6,14,15,46].

3.2. Fungal Material

The fungus strain CS-258 was isolated from the marine mussel sample collected from a cold seep in the northeast of the South China Sea in May 2020, and its strain identification was conducted by a BLAST search in GenBank. This strain sequence data had been uploaded to GenBank to obtain an accession number of No. PP065775. The strain was deposited at the Key Laboratory of Experimental Marine Biology, Institute of Oceanology, Chinese Academy of Sciences (IOCAS).

3.3. Fermentation, Extraction, and Isolation

The fresh mycelium of Talaromyces sp. CS-258 was cultured on potato dextrose broth medium for 3 days and then incubated in autoclaved 1 L Erlenmeyer flasks with rice culture medium (70 g rice, 0.2 g corn steep liquor, 0.5 g yeast extract, 0.3 g peptone, 0.6 g gourmet powder, and 100 mL seawater collected from the Huiquan Gulf of the Yellow Sea near the campus of IOCAS) for 30 days at room temperature. After incubation, the EtOAc crude extract was obtained by exhaustive concentration with MeOH and successive extraction with EtOAc/H2O four times.
The crude extract was fractionated by silica gel column vacuum liquid chromatography (VLC) elution with increasing polarity gradient of petroleum ether (PE)/EtOAc and CH2Cl2/MeOH to yield nine fractions (Fr.A–Fr.I). Fr.E was further split by column chromatography (CC) using Sephadex LH-20 to produce three subfractions from E-1 to E-3. Fr.E-1 was rechromatographed over silica gel elution with a slow CH2Cl2/MeOH gradient from 200:1 to 20:1 to yield three mixtures of E.1.1–E.1.3. Then, mixture E.1.1 was followed by semipreparative HPLC separation (Elite ODS-BP column, 5 μm, 10 × 250 mm, MeOH-H2O, 75:25, 3 mL/min) to afford compounds 12 (6.0 mg, tR 8 min) and 10 (10.2 mg, tR 10 min). Mixtures of E.1.2 and E.1.3 were also subjected to semipreparative HPLC purification to yield 7 (3.3 mg, tR 20 min, 70% MeOH-H2O), 30 (2.0 mg, tR 18 min, 75% MeOH-H2O), and 31 (5.1 mg, tR 28 min, 65% MeOH-H2O). Compound 4 (21.1 mg) was obtained via recrystallization from Fr.E-2. Another semipreparative HPLC collection with 75% MeOH-H2O was applied for Fr.E-3 to give compounds 18 (4.2 mg, tR 22 min), 19 (3.2 mg, tR 16 min), and 20 (4.0 mg, tR 18 min).
Fr.F was subjected to reversed-phase CC using Lobar LiChroprep RP-18 from 10% to 90% MeOH-H2O to produce three subfractions, Fr.F-1 to F-3. Fr.F-1 was sequentially recrystallized to obtain compound 2 (100.0 mg). Fr.F-2 was subjected to Sephadex LH-20 CC to afford compound 24 (65.9 mg) and two mixed components. Then, the former component was purified by HPLC separation with MeOH-H2O (52:48) and identified as 8 (13.7 mg, tR 25 min). The latter was subjected to preparative TLC (developing solvents: CH2Cl2−MeOH, 20:1) to obtain 27 (3.1 mg). Treated with successive CC on Sephadex LH-20 and HPLC with 58% MeOH-H2O, Fr.F-3 was found to produce compounds 9 (16.5 mg, recrystallization), 21 (15.3 mg, tR 16 min), 22 (9.5 mg, tR 18 min), and 23 (4.9 mg, recrystallization) totally.
Fr.G was subjected to further CC over Lobar LiChroprep RP-18 in the MeOH−H2O solvent system and thus gave 6 subfractions, Fr.G-1 to Fr.G-6. Subsequently, compounds 25 (4.0 mg, tR 14 min) and 26 (2.5 mg, tR 12 min) were provided from Fr.G-1 by Sephadex LH-20 (MeOH) and semipreparative HPLC elution with 42% MeOH−H2O; meanwhile, compound 1 (11.0 mg, tR 16 min) was obtained from another mixture of Fr.G-1 via HPLC in 30% MeOH−H2O. Fr.G-2 was subjected to a series of CC on Sephadex LH-20 and the HPLC separation system of 42% MeOH/H2O and 45% MeOH/H2O to afford 11 (6.7 mg, tR 18 min) and 5 (24.6 mg, tR 15 min). Fr.G-3 was further purified by Sephadex LH-20 and then by preparative TLC (developing solvents: CH2Cl2−MeOH, 15:1) and by HPLC with 80% MeOH/H2O to obtain compounds 13 (90.0 mg, tR 27 min), 14 (14.6 mg, tR 35 min), and 15 (4.8 mg, tR 25 min). The remaining subfractions, Fr.G-4 to Fr.G-6, were partitioned using nearly the same method through CC over Sephadex LH-20, silica gel, and semipreparative HPLC to produce compounds 16 (5.5 mg, from Fr.G-4, tR 10 min, 52% MeOH/H2O), 17 (5.0 mg, from Fr.G-6, tR 28 min, 51% MeOH/H2O), and 6 (6.0 mg, from Fr.G-5).
After CC on reversed-phase RP-18 and Sephadex LH-20, Fr.I was fractionated and purified to give compound 29 (3.2 mg, by silica gel column and following Sephadex LH-20), 28 (2.0 mg, by preparative TLC), 3 (13.0 mg, by semipreparative HPLC collection of 46% MeOH-H2O at tR 35 min), and 32 (2.4 mg, by preparative TLC and Sephadex LH-20).

3.4. Spectroscopic Data

(3R,5R)-5-hydroxytalaroflavone (1): brown oil; [ α ] D 25   +42.9 (c 0.44, MeOH); UV (MeOH, 0.20 mg/mL) λmax (log ε) 216 (3.39), 259 (2.94), 297 (2.58) nm; ECD (0.20 mg/mL, MeOH) λmaxε) 208 (−11.78), 222 (+12.43), 240 (+7.56) nm; 1H and 13C NMR data, see Table 1; (+)-HRESIMS m/z 263.055012 [M + H]+ (calcd for C13H11O6, 263.055014).
Talaroisochromenol A (3): brown oil; [ α ] D 25   +105.9 (c 0.06, MeOH); UV (MeOH, 0.20 mg/mL) λmax (log ε) 209 (3.13), 261 (2.92), 310 (2.64) nm; ECD (0.20 mg/mL, MeOH) λmax (Δε) 213 (−3.47), 261 (−1.62) nm; 1H and 13C NMR data, see Table 1; (+)-HRESIMS m/z 261.0398 [M + H]+ (calcd for C13H9O6, 261.0405).
Talaroisochromenol B (5): yellow oil; [ α ] D 25   +33.3 (c 0.24, MeOH); UV (MeOH, 0.20 mg/mL) 248 (3.08), 330 (2.18) nm; ECD (0.20 mg/mL, MeOH) λmaxε) 210 (+1.54), 238 (+1.24), 291 (−0.38) nm; 1H and 13C NMR data, see Table 1; (+)-HRESIMS m/z 265.0714 [M + H]+ (calcd for C13H13O6, 265.0707).
Talaroisochromenol C (11): white solid; [ α ] D 25   +4.3 (c 0.24, MeOH); UV (MeOH, 0.25 mg/mL) λmax (log ε) 213 (3.20), 229 (2.96), 265 (2.68), 305 (2.56) nm; 1H and 13C NMR data, see Table 2; (+)-HRESIMS m/z 251.0556 [M + H]+ (calcd for C12H11O6, 251.0550).
(+)-(11): [ α ] D 25   +60.0 (c 0.20, MeOH); ECD (0.30 mg/mL, MeOH) λmaxε) 206 (+6.83), 226 (–1.93), 268 (−3.68) nm.
(−)-(11): [ α ] D 25   −45.0 (c 0.20, MeOH); ECD (0.25 mg/mL, MeOH) λmaxε) 208 (−4.41), 226 (+1.70), 268 (+2.69) nm.
(8R,9R,10aR)-5-hydroxyaltenuene (13): brown oil; [ α ] D 25   −23.7 (c 0.30, MeOH); UV (MeOH, 0.25 mg/mL) λmax (log ε) 242 (3.24), 282 (2.73), 323 (2.47) nm; ECD (0.25 mg/mL, MeOH) λmaxε) 236 (+10.43), 282 (−5.59) nm; 1H and 13C NMR data, see Table 3; (+)-HRESIMS m/z 279.0870 [M + H]+ (calcd for C14H15O6, 279.0863).
(8R,9R,10aS)-5-hydroxyaltenuene (14): brown oil; [ α ] D 25   +43.6 (c 0.18, MeOH); UV (MeOH, 0.20 mg/mL) λmax (log ε) 244 (3.34), 282 (2.81), 325 (2.49) nm; ECD (0.20 mg/mL, MeOH) λmaxε) 237 (−2.85), 284 (+1.31) nm; 1H and 13C NMR data, see Table 3; (+)-HRESIMS m/z 279.0869 [M + H]+ (calcd for C14H15O6, 279.0863).
(8R,9S,10aR)-5-hydroxyaltenuene (15): brown oil; [ α ] D 25   −85.7 (c 0.21, MeOH); UV (MeOH, 0.30 mg/mL) λmax (log ε) 244 (3.17), 282 (2.64), 325 (2.32) nm; ECD (0.30 mg/mL, MeOH) λmaxε) 236 (+12.30), 282 (−7.29) nm; 1H and 13C NMR data, see Table 3; (+)-HRESIMS m/z 279.0861 [M + H]+ (calcd for C14H15O6, 279.0863).
Nemanecin D (25): yellow oil; [ α ] D 25   −11.5 (c 0.26, MeOH); UV (MeOH, 0.40 mg/mL) λmax (log ε) 209 (2.98), 320 (3.10) nm; ECD (0.40 mg/mL, MeOH) λmaxε) 214 (+1.73), 345 (−1.67) nm; 1H and 13C NMR data, see Table 4; (+)-HRESIMS m/z 255.1228 [M + H]+ (calcd for C13H19O5, 255.1227).
(S)-MTPA Ester (25a): 1H NMR (Figure S65, selected signals, pyridine-d5, 500 MHz) δH: 5.60 (1H, m, H-10), 2.63 (1H, dd, H-9α), 2.60 (1H, dd, H-9β), 1.31 (3H, d, H-11); 1H-1H COSY and HRMS spectra, see Figures S66 and S67.
(R)-MTPA Ester (25b): 1H NMR (Figure S68, selected signals, pyridine-d5, 500 MHz) δH: 5.51 (1H, m, H-10), 2.54 (2H, m, H-9), 1.36 (3H, d, H-11); 1H-1H COSY and HRMS spectra, see Figures S69 and S70.
Nemanecin E (26): yellow oil; [ α ] D 25   −48.0 (c 0.40, MeOH); UV (MeOH, 0.25 mg/mL) λmax (log ε) 209 (2.89), 319 (2.76) nm; ECD (0.40 mg/mL, MeOH) λmaxε) 214 (+0.70), 345 (−1.18) nm; 1H and 13C NMR data, see Table 4; (+)-HRESIMS m/z 255.1226 [M + H]+ (calcd for C13H19O5, 255.1227).
(S)-MTPA Ester (26a): 1H NMR (Figure S79, selected signals, pyridine-d5, 500 MHz) δH: 5.58 (1H, m, H-10), 2.62 (1H, dd, H-9α), 2.50 (1H, dd, H-9β), 1.38 (3H, d, H-11); 1H-1H COSY and HRMS spectra, see Figures S80 and S81.
(R)-MTPA Ester (26b): 1H NMR (Figure S82, selected signals, pyridine-d5, 500 MHz) δH: 5.53 (1H, m, H-10), 2.68 (1H, dd, H-13α), 2.54 (1H, dd, H-9β), 1.28 (3H, d, H-15); 1H-1H COSY and HRMS spectra, see Figures S83 and S84.
2,5-dimethy-8-iodochromone (27): yellow oil; UV (MeOH, 0.05 mg/mL) λmax (log ε) 210 (3.56), 247 (3.52), 296 (3.19) nm; 1H and 13C NMR data, see Table 5; (+)-HRESIMS m/z 316.9668 [M + H]+ (calcd for C11H10O3I, 316.9669).
6-hydroxy-4-methoxycoumarin (28): yellow oil; UV (MeOH, 0.05 mg/mL) λmax (log ε) 222 (3.33), 270 (2.96), 331 (2.67) nm; 1H and 13C NMR data, see Table 5; (+)-HRESIMS m/z 193.0493 [M + H]+ (calcd for C10H9O4, 193.0495).
Talarofurolactone A (29): yellow oil; [ α ] D 25 −110.3 (c 0.15, MeOH); UV (MeOH, 0.25 mg/mL) λmax (log ε) 220 (3.01) nm; ECD (0.25 mg/mL, MeOH) λmaxε) 211 (−4.27) nm; 1H and 13C NMR data, see Table 5; (+)-HRESIMS m/z 213.1122 [M + H]+ (calcd for C11H17O4, 213.1121), m/z 230.1389 [M + NH4]+ (calcd for C11H20O4N, 230.1387), m/z 235.0942 [M + Na]+ (calcd for C11H16O4Na, 235.0941).

3.5. X-ray Crystallographic Analysis of Compounds 2, 11, 13, and 14

Compounds 2, 11, 13, and 14 were dissolved in MeOH and crystallized after slow evaporation in a refrigerator. All X-ray single-crystal diffraction data were recorded at 293(2) K on an Agilent Xcalibur Eos Gemini CCD plate diffractometer equipped with graphite monochromatic CuKα radiation. The data were revised for absorption with the program SADABS [47]. Direct methods of the SHELXTL software and subsequent refinement via full-matrix least-squares difference Fourier techniques were applied to solve the structures accurately [48,49]. Additional refinements for non-hydrogen, H, and O atoms were carried out as previously reported [9,14]. All crystallographic data have been deposited in the Cambridge Crystallographic Data Centre (CCDC) and can be obtained on application to the director free of charge.
Crystal data for 2: m.p. 224.0–248.0 °C, C14H12O6, fw = 276.24, monoclinic, space group C2/c, unit cell dimensions a = 23.8978(11) Å, b = 6.9420(3) Å, c = 14.9300(7) Å, V = 2467.83(19) Å3, α = 90, β = 94.895(2), γ = 90, Z = 8, dcalcd = 1.487 mg/m3, crystal size 0.16 × 0.15 × 0.14 mm3, μ = (CuKα) 1.002 mm−1, F(000) = 1152.0. Independent reflections: 2260 [Rint = 0.0427, Rsigma= 0.0248]. The final indices gave R1 = 0.0405 and wR2 = 0.1032 [I > 2σ(I)]. CCDC number was deposited as 2327879.
Crystal data for 11: m.p. 181.3–184.0 °C, C12H10O6, fw = 250.20, triclinic, space group P-1, unit cell dimensions a = 7.300(2) Å, b = 7.4441(2) Å, c = 9.9268(2) Å, V = 516.92(2) Å3, α = 70.6080(10), β = 86.7490(10), γ = 87.6150(10), Z = 2, dcalcd = 1.607 mg/m3, crystal size 0.16 × 0.15 × 0.14 mm3, μ = (CuKα) 1.126 mm−1, F(000) = 260.0. Independent reflections: 1869 [Rint = 0.0291, Rsigma= 0.0280]. The final indices gave R1 = 0.0347 and wR2 = 0.0996 [I > 2σ(I)]. CCDC number was deposited as 2327875.
Crystal data for 13: m.p. 244.1–247.2 °C, C14H14O6, fw = 278.25, monoclinic, space group P21/n, unit cell dimensions a = 9.0780(2) Å, b = 6.90080(10) Å, c = 19.2001(4) Å, V = 1202.71(4) Å3, α = 90, β = 90.6920(10), γ = 90, Z = 4, dcalcd = 1.537 mg/m3, crystal size 0.18 × 0.16 × 0.15 mm3, μ = (CuKα) 1.028 mm−1, F(000) = 584.0. Independent reflections: 2185 [Rint = 0.0319, Rsigma= 0.0248]. The final indices gave R1 = 0.0328 and wR2 = 0.0873 [I > 2σ(I)]. CCDC number was deposited as 2327876.
Crystal data for 14: m.p. 273.0–278.0 °C, C14H14O6 +(CH3)2SO, fw = 356.38, monoclinic, space group P21/c, unit cell dimensions a = 8.8828(3) Å, b = 20.8250(7) Å, c = 8.9113(3) Å, V = 1595.86(9) Å3, α = 90, β = 104.511(2), γ = 90, Z = 2, dcalcd = 1.483 mg/m3, crystal size 0.16 × 0.15 × 0.12 mm3, μ = (CuKα) 2.143 mm−1, F(000) = 752.0. Independent reflections: 2917 [Rint = 0.0489, Rsigma = 0.0318]. The final indices gave R1 = 0.0315 and wR2 = 0.0877 [I > 2σ(I)]. CCDC number was deposited as 2327877.

3.6. Antimicrobial Activity Assay

The antimicrobial evaluations against four human pathogens (methicillin-resistant S. aureus, P. aeruginosa, E. coli, and K. pneumonia) and seven aquatic bacteria (V. alginolyticus, A. hydrophilia, M. luteus, V. anguillarum, V. parahaemolyticus, V. vulnificus, and V. harveyi) as well as six plant-pathogenic fungi (C. cornigerum, P. digitatum, P. piricola, V. mali, C. gloeosporioides and F. oxysporum) were carried out in a 96-well microplate as described in our previous reports [6,46,50,51]. These microbial strains were provided by the Institute of Oceanology, Chinese Academy of Sciences, by purchase or isolation. Chloramphenicol and amphotericin B served as positive controls against bacteria and fungi, respectively, while DMSO was treated as the negative control. All measurements at various concentrations were performed in triplicate.

3.7. Details of Computational Methods

The computational methods for ECD calculation and DP4+ probability analysis were similar to those in the previous papers from our group [6,10,51], with some modifications detailed below. Molecular mechanics using the MM+ method was carried out for conformational searches in HyperChem software (Version 8.0, Hypercube, Inc., Gainesville, FL, USA). The energy-minimized conformers were generated and further optimized using DFT calculations at the B3LYP/6-31G(d) level in Gaussian 09 software (Version D.01; Gaussian, Inc.: Wallingford, CT, USA). Frequency calculations were carried out at the same level of theory to confirm the absence of imaginary frequencies and to obtain thermal corrections to the Gibbs free energies. These obtained conformers were subjected to ECD calculations using the TDDFT method at the CAM-B3LYP/TZVP, BH&HLYP/TZVP, or PBE0/TZVP level. The solvent effects of MeOH were evaluated at the same DFT level using the self-consistent reaction field (SCRF) method with the polarizable continuum model (PCM). The ECD spectrum was generated by the SpecDis program and finally drawn using Origin Pro 8.5 software [52,53,54,55,56,57]. The NMR shielding tensors were calculated by the DFT method at the mPW1PW91\6-31+G(d) PCM level in DMSO and then weighted according to Boltzmann’s distribution. GIAO (gauge-independent atomic orbital) NMR chemical calculations were performed using an equation described previously. The theoretical shielding tensors (1H and 13C) and experimental chemical shifts were finally analyzed and compared using DP4+ probability [58,59,60].

3.8. The Modified Mosher’s Method

The modified Mosher’s method was conducted in NMR tubes following the details described in the previous papers [25,61].

4. Conclusions

In this study, the chemical investigation of cold-seep-derived fungus Talaromyces sp. CS-258 has resulted in the isolation and identification of 32 fungal polyketides including 11 new compounds and 1 new naturally isolated metabolite. Notably, a pair of enantiomers ((±)-11) and two groups of diastereoisomers (1315 and 25/26) were obtained, the isolation of which presented challenges due to their difficulty in separation from each other. Furthermore, compound 27 was identified as a rare occurring natural iodo-chromone analog. The plausible biogenetic pathways of compounds 124, which belong to altenusin or alternariol families with polyketide origin bearing variable tricyclic ring skeletons, were proposed. The results of antimicrobial activities displayed that 11 polyketides exhibited broad-spectrum antibiotic effects against human and aquatic pathogenic bacteria. Among them, compounds 10 and 18 exhibited potent inhibition against E. coli, A. hydrophilia, and V. parahaemolyticus, while 22 and 24 effectively suppressed the growth of A. hydrophilia. These findings highlight the cold-seep-derived fungus Talaromyces sp. CS-258 as a promising source for bioactive metabolites, especially for fungal polyketides, with potential applications as antibiotic agents in medicinal development and agriculture.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/md22050204/s1. Figures S1–Figures S120: Chemical structures, HRESIMS, UV, 1D and 2D NMR spectra of new compounds; 1D/2D NMR and HRMS data for the modified Mosher’s product of compounds 25 and 26, the chiral HPLC separation of compound 11; antibacterial and antifungal activities of isolated compounds and optimized low-energy conformers for ECD calculations. Tables S1–S23: Energy analysis for optimized conformers of new compounds, the DP4+ probability analysis and calculated shielding tensors of each conformer for the candidate isomers of 15 and 29, and the experimental as well as calculated chemical shifts of 15 and 29.

Author Contributions

Z.W. performed chemical and bioactive experiments, analyzed the data, and wrote the manuscript. X.-M.L. measured and recorded the NMR spectra. S.-Q.Y. and X.L. isolated and identified the fungus Talaromyces sp. CS-258. X.L. and B.-G.W. designed the experiments and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Key R&D Program of China (2022YFC2804700), by the National Natural Science Foundation of China (42076090 and U2006203), by the Shandong Provincial Natural Science Foundation (ZR2021ZD28), and by the Senior User Project of the Research Vessel KEXUE of the National Major Science and Technology Infrastructure from the Chinese Academy of Sciences (KEXUE2020GZ02). X.L. acknowledges the Youth Innovation Promotion Association CAS (2023216).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data are included in the article and the Supplementary Material.

Acknowledgments

The authors acknowledge the support of the RV KEXUE of the National Major Science and Technology Infrastructure from the Chinese Academy of Sciences (for sampling) and the Oceanographic Data Center at IOCAS (for CPU time).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chooi, Y.H.; Tang, Y. Navigating the fungal polyketide chemical space: From genes to molecules. J. Org. Chem. 2012, 77, 9933–9953. [Google Scholar] [CrossRef] [PubMed]
  2. Crawford, J.M.; Korman, T.P.; Labonte, J.W.; Vagstad, A.L.; Hill, E.A.; Kamari-Bidkorpeh, O.; Tsai, S.C.; Townsend, C.A. Structural basis for biosynthetic programming of fungal aromatic polyketide cyclization. Nature 2009, 461, 1139–1143. [Google Scholar] [CrossRef] [PubMed]
  3. Crawford, J.M.; Townsend, C.A. New insights into the formation of fungal aromatic polyketides. Nat. Rev. Microbiol. 2010, 8, 879–889. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, Y.-R.; Dong, Y.-L.; Li, X.-M.; Shi, X.-S.; Li, H.-L.; Meng, L.-H.; Xu, R.; Wang, B.-G. Curvularin derivatives from the marine mangrove derived fungus Penicillium sumatrense MA-325. Phytochemistry 2024, 220, 114000. [Google Scholar] [CrossRef] [PubMed]
  5. Weissman, K.J.; Leadlay, P.F. Combinatorial biosynthesis of reduced polyketides. Nat. Rev. Microbiol. 2005, 3, 925–936. [Google Scholar] [CrossRef] [PubMed]
  6. Yan, L.H.; Du, F.Y.; Li, X.M.; Yang, S.Q.; Wang, B.G.; Li, X. Antibacterial indole diketopiperazine alkaloids from the deep-sea cold seep-derived fungus Aspergillus chevalieri. Mar. Drugs 2023, 21, 195. [Google Scholar] [CrossRef] [PubMed]
  7. Jin, E.; Li, H.; Liu, Z.; Xiao, F.; Li, W. Antibiotic dixiamycins from a cold-seep-derived Streptomyces olivaceus. J. Nat. Prod. 2021, 84, 2606–2611. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, Y.; Su, X.; Chen, F.; Wang, Y.; Jiao, L.; Dong, H.; Huang, Y.; Jiang, H. Microbial diversity in cold seep sediments from the northern South China Sea. Geosci. Front. 2012, 3, 301–316. [Google Scholar] [CrossRef]
  9. Chi, L.P.; Li, X.M.; Wan, Y.P.; Li, X.; Wang, B.G. Ophiobolin sesterterpenoids and farnesylated phthalide derivatives from the deep sea cold-seep-derived fungus Aspergillus insuetus sd-512. J. Nat. Prod. 2020, 83, 3652–3660. [Google Scholar] [CrossRef]
  10. Li, Y.; Li, X.; Li, X.; Yang, S.; Wang, B.; Li, H. Verrucosidin derivatives from the deep sea cold-seep-derived fungus Penicillium polonicum CS-252. Int. J. Mol. Sci. 2022, 23, 5567. [Google Scholar] [CrossRef]
  11. Song, Q.; Yang, S.Q.; Li, X.M.; Hu, X.Y.; Li, X.; Wang, B.G. Aromatic polyketides from the deep-sea cold-seep mussel associated endozoic fungus Talaromyces minioluteus CS-138. Mar. Drugs 2022, 20, 529. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, Y.P.; Fang, S.T.; Shi, Z.Z.; Wang, B.G.; Li, X.N.; Ji, N.Y. Phenylhydrazone and quinazoline derivatives from the cold-seep-derived fungus Penicillium oxalicum. Mar. Drugs 2020, 19, 9. [Google Scholar] [CrossRef] [PubMed]
  13. Cong, M.; Zhang, Y.; Feng, X.; Pang, X.; Liu, Y.; Zhang, X.; Yang, Z.; Wang, J. Anti-inflammatory alkaloids from the cold-seep-derived fungus Talaromyces helicus SCSIO41311. 3 Biotech 2022, 12, 161. [Google Scholar] [CrossRef] [PubMed]
  14. Hu, X.Y.; Wang, C.Y.; Li, X.M.; Yang, S.Q.; Li, X.; Wang, B.G.; Si, S.Y.; Meng, L.H. Cytochalasin derivatives from the endozoic Curvularia verruculosa CS-129, a fungus isolated from the deep-sea squat lobster Shinkaia crosnieri living in the cold seep environment. J. Nat. Prod. 2021, 84, 3122–3130. [Google Scholar] [CrossRef] [PubMed]
  15. Yan, L.H.; Li, P.H.; Li, X.M.; Yang, S.Q.; Liu, K.C.; Wang, B.G.; Li, X. Chevalinulins A and B, proangiogenic alkaloids with a spiro[bicyclo[2.2.2]octane-diketopiperazine] skeleton from deep-sea cold-seep-derived fungus Aspergillus chevalieri CS-122. Org. Lett. 2022, 24, 2684–2688. [Google Scholar] [CrossRef] [PubMed]
  16. Hu, X.; Li, X.; Yang, S.; Li, X.; Wang, B.; Meng, L. Vercytochalasins A and B: Two unprecedented biosynthetically related cytochalasins from the deep-sea-sourced endozoic fungus Curvularia verruculosa. Chin. Chem. Lett. 2023, 34, 107516. [Google Scholar] [CrossRef]
  17. Che, Y.-H.; Wang, J.-F.; Shi, X.-F.; Ding, W.-P.; Xiao, Z.-H.; Wu, J.-M.; Wang, F.-Z.; Zhang, S. The 8R-methoxy-9R-hydroxyl-fumitremorgin C, a new diketopiperazine alkaloid from Haima cold seep-derived fungus Aspergillus fumigatus CYH-5. Nat. Prod. Res. 2023. [Google Scholar] [CrossRef] [PubMed]
  18. Fall, Y.; Santana, L.; Uriarte, E. Synthesis and characterization of some coumarins with two hydroxy or methoxy substituents. J. Heterocycl. Chem. 2001, 38, 1231. [Google Scholar] [CrossRef]
  19. Tian, Y.; Liu, S.; Liu, Y. A new polyketide from a sponge-derived fungus. Acta Pharm. Sinica 2018, 53, 1134–1138. [Google Scholar]
  20. Liu, Y.; Wu, Y.; Zhai, R.; Liu, Z.; Huang, X.; She, Z. Altenusin derivatives from mangrove endophytic fungus Alternaria sp. SK6YW3L. RSC Adv. 2016, 6, 72127–72132. [Google Scholar] [CrossRef]
  21. Wang, Q.X.; Bao, L.; Yang, X.L.; Guo, H.; Yang, R.N.; Ren, B.; Zhang, L.X.; Dai, H.Q.; Guo, L.D.; Liu, H.W. Polyketides with antimicrobial activity from the solid culture of an endolichenic fungus Ulocladium sp. Fitoterapia 2012, 83, 209–214. [Google Scholar] [CrossRef]
  22. Jin, P.F.; Dai, H.F.; Zuo, W.J.; Zeng, Y.B.; Guo, Z.K.; Mei, W.L. Metabolites from the endophytic fungus of Ceriops tagal. Chin. J. Med. Chem. 2013, 23, 309. [Google Scholar]
  23. Kornsakulkarn, J.; Saepua, S.; Suvannakad, R.; Supothina, S.; Boonyuen, N.; Isaka, M.; Prabpai, S.; Kongsaeree, P.; Thongpanchang, C. Cytotoxic tropolones from the fungus Nemania sp. BCC 30850. Tetrahedron 2017, 73, 3505–3512. [Google Scholar] [CrossRef]
  24. Dale, J.A.; Mosher, H.S. Nuclear magnetic resonance enantiomer regents. Configurational correlations via nuclear magnetic resonance chemical shifts of diastereomeric mandelate, O-methylmandelate, and α-methoxy-α-trifluoromethylpheny-lacetate (MTPA) esters. J. Am. Chem. Soc. 1973, 95, 512–519. [Google Scholar] [CrossRef]
  25. Su, B.N.; Park, E.J.; Mbwambo, Z.H.; Santarsiero, B.D.; Mesecar, A.D.; Fong, H.H.S.; Pezzuto, J.M.; Kinghorn, A.D. New chemical constituents of Euphorbia quinquecostata and absolute configuration assignment by a convenient mosher ester procedure carried out in NMR tubes. J. Nat. Prod. 2002, 65, 1278–1282. [Google Scholar] [CrossRef] [PubMed]
  26. Kashiwada, Y.; Nonaka, G.; Nishioka, I. Chromone glucosides from Rhubarb. Phytochemistry 1990, 29, 1007–1009. [Google Scholar] [CrossRef]
  27. Zhang, H.; Ma, C.; Zheng, Z.; Sun, R.; Yu, X.; Zhao, J. Synthesis of 2-arylbenzofuran-3-carbaldehydes via an organocatalytic [3+2] annulation/oxidative aromatization reaction. Chem. Commun. 2018, 54, 4935–4938. [Google Scholar] [CrossRef]
  28. Li, C.-P.; Song, Y.-P.; Wang, B.-G.; Ji, N.-Y. Sulfurated and iodinated metabolites from the cold-seep fungus Cladosporium cladosporioides 8-1. Tetrahedron Lett. 2022, 93, 153689. [Google Scholar] [CrossRef]
  29. Ye, F.; Chen, G.-D.; He, J.-W.; Li, X.-X.; Sun, X.; Guo, L.-D.; Li, Y.; Gao, H. Xinshengin, the first altenusin with tetracyclic skeleton core from Phialophora spp. Tetrahedron Lett. 2013, 54, 4551–4554. [Google Scholar] [CrossRef]
  30. Naganuma, M.; Nishida, M.; Kuramochi, K.; Sugawara, F.; Yoshida, H.; Mizushina, Y. 1-deoxyrubralactone, a novel specific inhibitor of families X and Y of eukaryotic DNA polymerases from a fungal strain derived from sea algae. Bioorg. Med. Chem. 2008, 16, 2939–2944. [Google Scholar] [CrossRef]
  31. He, J.W.; Chen, G.D.; Gao, H.; Yang, F.; Li, X.X.; Peng, T.; Guo, L.D.; Yao, X.S. Heptaketides with antiviral activity from three endolichenic fungal strains Nigrospora sp., Alternaria sp. and Phialophora sp. Fitoterapia 2012, 83, 1087–1091. [Google Scholar] [CrossRef]
  32. Hua, Y.; Pan, R.; Bai, X.; Wei, B.; Chen, J.; Wang, H.; Zhang, H. Aromatic polyketides from a symbiotic strain Aspergillus fumigatus D and characterization of their biosynthetic gene D8.T287. Mar. Drugs 2020, 18, 324. [Google Scholar] [CrossRef] [PubMed]
  33. Hildebrand, A.A.; Kohn, B.N.; Pfeiffer, E.; Wefers, D.; Metzler, M.; Bunzel, M. Conjugation of the mycotoxins alternariol and alternariol monomethyl ether in tobacco suspension cells. J. Agric. Food Chem. 2015, 63, 4728–4736. [Google Scholar] [CrossRef] [PubMed]
  34. Aly, A.H.; Edrada-Ebel, R.; Indriani, I.D.; Wray, V.; Muller, W.E.G.; Totzke, F.; Zirrgiebel, U.; Schachtele, C.; Kubbutat, M.H.G.; Lin, W.H.; et al. Cytotoxic metabolites from the fungal endophyte Alternaria sp. and their subsequent detection in its host plant Polygonum senegalense. J. Nat. Prod. 2008, 71, 972–980. [Google Scholar] [CrossRef] [PubMed]
  35. Kong, F.D.; Yi, T.F.; Ma, Q.Y.; Xie, Q.Y.; Zhou, L.M.; Chen, J.P.; Dai, H.F.; Wu, Y.G.; Zhao, Y.X. Biphenyl metabolites from the patchouli endophytic fungus Alternaria sp. PfuH1. Fitoterapia 2020, 146, 104708. [Google Scholar] [CrossRef] [PubMed]
  36. Xu, L.; Ma, H.B.; Wu, T.; Liu, L.F.; Xie, M.M.; Hu, M.Y.; Gai, Y.B.; Zhong, T.H.; Yang, X.W. Hyposterolactone A, a 3α-hydroxy steroidal lactone from the deep-sea-derived fungus Hypocrea sp. ZEN14. Chem. Biodiversity 2023, 20, e202300753. [Google Scholar] [CrossRef]
  37. Pfeiffer, E.; Schebb, N.H.; Podlech, J.; Metzler, M. Novel oxidative in vitro metabolites of the mycotoxins alternariol and alternariol methyl ether. Mol. Nutr. Food Res. 2007, 51, 307–316. [Google Scholar] [CrossRef]
  38. Wu, Y.Y.; Zhang, T.Y.; Zhang, M.Y.; Cheng, J.; Zhang, Y.X. An endophytic fungi of Ginkgo biloba L. produces antimicrobial metabolites as potential inhibitors of FtsZ of Staphylococcus aureus. Fitoterapia 2018, 128, 265–271. [Google Scholar] [CrossRef] [PubMed]
  39. Komai, S.-i.; Hosoe, T.; Itabashi, T.; Nozawa, K.; Yaguchi, T.; Fukushima, K.; Kawai, K.-i. New penicillide derivatives isolated from Penicillium simplicissimum. J. Nat. Med. 2006, 60, 185–190. [Google Scholar] [CrossRef]
  40. Jeon, H.; Shim, S.H. Chemical constituents of the endophyte Penicillium sp. isolated from Artemisia princeps. Chem. Nat. Compd. 2020, 56, 122–124. [Google Scholar] [CrossRef]
  41. Kazumi, S.; Koohei, N.; Shun-ichi, U.; Shoichi, N.; Ken-ichi, K. Penicillide and dehydroisopenicillide from Talaromyces derxii. Phytochemistry 1991, 30, 2096–2098. [Google Scholar]
  42. Ayer, W.A.; Racok, J.S. The metabolites of Talaromyces flavus: Part 2. biological activity and biosynthetic studies. Can. J. Chem. 1990, 68, 2095–2101. [Google Scholar] [CrossRef]
  43. Yang, H.; Qi, B.; Ding, N.; Jiang, F.; Jia, F.; Luo, Y.; Xu, X.; Wang, L.; Zhu, Z.; Liu, X.; et al. Polyketides from Alternaria alternata MT-47, an endophytic fungus isolated from Huperzia serrata. Fitoterapia 2019, 137, 104282. [Google Scholar] [CrossRef] [PubMed]
  44. Thomas, M.H.; James, V.H. Biogenetically modeled syntheses of heptaacetate metabolites. Alternariol and lichexanthone. J. Am. Chem. Soc. 1977, 99, 1631–1637. [Google Scholar]
  45. Sun, J.; Awakawa, T.; Noguchi, H.; Abe, I. Induced production of mycotoxins in an endophytic fungus from the medicinal plant Datura stramonium L. Bioorg. Med. Chem. Lett. 2012, 22, 6397–6400. [Google Scholar] [CrossRef]
  46. Chi, L.P.; Liu, D.; Li, X.M.; Wan, Y.; Wang, B.G.; Li, X. Aspertides A-E: Antimicrobial pentadepsipeptides with a unique p-methoxycinnamoyl amide group from the marine isolates Aspergillus tamarii MA-21 and Aspergillus insuetus SD-512. J. Agric. Food Chem. 2023, 71, 13316–13324. [Google Scholar] [CrossRef] [PubMed]
  47. Blessing, R.H. An empirical correction for absorption anisotropy. Acta Crystallogr. A 1995, 51, 33–38. [Google Scholar] [CrossRef]
  48. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  49. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  50. Shi, X.-S.; Wang, D.-J.; Li, X.-M.; Li, H.-L.; Meng, L.-H.; Li, X.; Pi, Y.; Zhou, X.-W.; Wang, B.-G. Antimicrobial polyketides from Trichoderma koningiopsis QA-3, an endophytic fungus obtained from the medicinal plant Artemisia argyi. RSC Adv. 2017, 7, 51335–51342. [Google Scholar] [CrossRef]
  51. Li, H.L.; Li, X.M.; Mandi, A.; Antus, S.; Li, X.; Zhang, P.; Liu, Y.; Kurtan, T.; Wang, B.G. Characterization of cladosporols from the marine algal-derived endophytic fungus Cladosporium cladosporioides EN-399 and configurational revision of the previously reported cladosporol derivatives. J. Org. Chem. 2017, 82, 9946–9954. [Google Scholar] [CrossRef]
  52. Bruhn, T.; Schaumloeffel, A.; Hemberger, Y.; Bringmann, G. Specdis: Quantifying the comparison of calculated and experimental electronic circular dichroism spectra. Chirality 2013, 25, 243–249. [Google Scholar] [CrossRef]
  53. Dai, P.; Jiang, N.; Tan, R.-X. Assignment of absolute stereostructures through quantum mechanics electronic and vibrational circular dichroism calculations. J. Asian Nat. Prod. Res. 2016, 18, 72–91. [Google Scholar] [CrossRef]
  54. Pecul, M.; Ruud, K.; Helgaker, T. Density functional theory calculation of electronic circular dichroism using London orbitals. Chem. Phys. Lett. 2004, 388, 110–119. [Google Scholar] [CrossRef]
  55. Li, X.-C.; Ferreira, D.; Ding, Y. Determination of absolute configuration of natural products: Theoretical calculation of electronic circular dichroism as a tool. Curr. Org. Chem. 2010, 14, 1678–1697. [Google Scholar] [CrossRef] [PubMed]
  56. Pescitelli, G.; Bruhn, T. Good computational practice in the assignment of absolute configurations by TDDFT calculations of ECD spectra. Chirality 2016, 28, 749. [Google Scholar] [CrossRef] [PubMed]
  57. Ding, Y.; Li, X.-C.; Ferreira, D. Theoretical calculation of electronic circular dichroism of the rotationally restricted 3,8″-biflavonoid morelloflavone. J. Org. Chem. 2007, 72, 9010–9017. [Google Scholar] [CrossRef] [PubMed]
  58. Grimblat, N.; Zanardi, M.M.; Sarotti, A.M. Beyond DP4: An improved probability for the stereochemical assignment of isomeric compounds using quantum chemical calculations of NMR shifts. J. Org. Chem. 2015, 80, 12526–12534. [Google Scholar] [CrossRef] [PubMed]
  59. Waters, A.L.; Oh, J.; Place, A.R.; Hamann, M.T. Stereochemical studies of the karlotoxin class using nmr spectroscopy and DP4 chemical-shift analysis: Insights into their mechanism of action. Angew. Chem. Int. Ed. Engl. 2015, 54, 15705–15710. [Google Scholar] [CrossRef]
  60. Smith, S.G.; Goodman, J.M. Assigning stereochemistry to single diastereoisomers by Giao NMR calculation: The DP4 probability. J. Am. Chem. Soc. 2010, 132, 12946–12959. [Google Scholar] [CrossRef]
  61. Jongrungruangchok, S.; Kittakoop, P.; Yongsmith, B.; Bavovada, R.; Tanasupawat, S.; Lartpornmatulee, N.; Thebtaranonth, Y. Azaphilone pigments from a yellow mutant of the fungus Monascus kaoliang. Phytochemistry 2004, 65, 2569–2575. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of compounds 132.
Figure 1. Chemical structures of compounds 132.
Marinedrugs 22 00204 g001
Figure 2. The 1H–1H COSY and key HMBC correlations of compounds 1, 3, 5, 11, 1315, and 2529.
Figure 2. The 1H–1H COSY and key HMBC correlations of compounds 1, 3, 5, 11, 1315, and 2529.
Marinedrugs 22 00204 g002
Figure 3. X-ray crystallographic structures of compounds 2, 11, 13, and 14.
Figure 3. X-ray crystallographic structures of compounds 2, 11, 13, and 14.
Marinedrugs 22 00204 g003
Figure 4. The experimental and calculated ECD spectra of 1, 3, 5, 11, 1315, 25, and 26.
Figure 4. The experimental and calculated ECD spectra of 1, 3, 5, 11, 1315, 25, and 26.
Marinedrugs 22 00204 g004
Figure 5. The NOESY correlations of compounds 5, 11, 1315, 25, and 26.
Figure 5. The NOESY correlations of compounds 5, 11, 1315, 25, and 26.
Marinedrugs 22 00204 g005
Figure 6. Two possible isomers of compound 15 for DP4+ probability analysis.
Figure 6. Two possible isomers of compound 15 for DP4+ probability analysis.
Marinedrugs 22 00204 g006
Figure 7. The Δδ values [Δδ = δSδR] for the MTPA esters 25 and 26.
Figure 7. The Δδ values [Δδ = δSδR] for the MTPA esters 25 and 26.
Marinedrugs 22 00204 g007
Figure 8. The NOESY correlations (A), experimental and calculated ECD spectra (B), and two possible isomers for DP4+ probability analysis (C) of 29.
Figure 8. The NOESY correlations (A), experimental and calculated ECD spectra (B), and two possible isomers for DP4+ probability analysis (C) of 29.
Marinedrugs 22 00204 g008
Figure 9. The possible biosynthetic routes of compounds 124.
Figure 9. The possible biosynthetic routes of compounds 124.
Marinedrugs 22 00204 g009
Table 1. 1H and 13C NMR spectroscopic data of 1, 3, and 5 in DMSO-d6 (δ in ppm, J in Hz).
Table 1. 1H and 13C NMR spectroscopic data of 1, 3, and 5 in DMSO-d6 (δ in ppm, J in Hz).
No.135
1 166.9 C
2 168.2 C 166.9 C
2a 99.1 C 1 97.8 C
3 91.4 C 165.0 C 166.3 C
3a 149.9 C
45.76, d (1.6)100.8 CH6.44, s103.2 CH6.33, s101.4 CH
5 164.6 C 163.9 C 163.8 C
66.34, s103.3 CH6.98, s106.9 CH6.37, s102.1 CH
6a 133.1 C 137.2 C
7a 103.5 C 130.1 C 118.1 C
7 158.7 C 150.8 C 16.89, dd (6.9, 2.6)42.0 CH
81.76, s12.9 CH3 197.0 C3.67, m83.3 CH
9 3.12, d (18.0)
2.88, d (18.0)
46.8 CH24.47, d (2.4)78.1 CH
9a 81.4 C 153.1 C
10 1.70, s27.3 CH31.29, d (6.9)18.5 CH3
2′ 169.2 C
3′6.36, s129.8 CH
4′ 200.1 C
5′4.61, s78.2 CH
1 Observed in HMBC spectrum.
Table 2. 1H and 13C NMR spectroscopic data of 11 in DMSO-d6 (δ in ppm, J in Hz).
Table 2. 1H and 13C NMR spectroscopic data of 11 in DMSO-d6 (δ in ppm, J in Hz).
No.11
2 173.2 C
33.26, d (17.6)
2.94, d (17.6)
42.2 CH2
3a 85.9 C
5 166.0 C
5a 97.8 C
6 163.5 C
76.46, d (2.2)104.3 CH
8 165.2 C
96.62, d (2.2)110.9 CH
9a 133.3 C
9b5.49, s76.9 CH
101.51, s19.9 CH3
6-OH10.99, s
8-OH11.08, s
Table 3. 1H and 13C NMR spectroscopic data of 1315 in DMSO-d6 (δ in ppm, J in Hz).
Table 3. 1H and 13C NMR spectroscopic data of 1315 in DMSO-d6 (δ in ppm, J in Hz).
No.131415
2 168.3 C 167.8 C 167.7 C
2a 98.8 C 98.3 C 98.1 C
3 163.1 C 163.3 C 163.2 C
46.29, s102.3 CH6.26, d (2.1)102.6 CH6.28, overlap102.7 CH
5 165.1 C 165.5 C 165.8 C
66.53, s103.4 CH6.47, d (2.1)103.3 CH6.53, d (2.2)103.8 CH
6a 139.4 C 137.7 C 138.2 C
7a 132.2 C 131.4 C 133.5 C
76.14, d (3.1)130.3 CH6.05, d (2.5)130.3 CH6.28, overlap127.8 CH
83.95, t (4.0)69.7 CH4.07, dd (7.9, 2.5)72.5 CH4.08, t (4.6)64.9 CH
93.68, dt (7.9, 3.9) 68.8 CH3.59, ddd (12.0, 7.9, 3.7) 70.2 CH3.66, dt (12.3, 3.7) 65.4 CH
10β 2.24, dd (14.0, 3.5)
α 1.93, dd (14.5, 8.1)
38.7 CH2β 2.14, dd (12.0, 3.7)
α 2.01, t (12.3)
43.5 CH2β 2.21, t (12.0)
α 1.90, dd (11.6, 3.3)
38.4 CH2
10a 80.9 C 82.0 C 81.6 C
111.45, s27.4 CH31.48, s26.2 CH31.43, s25.8 CH3
3-OH11.22, s 11.25, s 11.28, s
Table 4. 1H and 13C NMR spectroscopic data of 2526 in DMSO-d6 (δ in ppm, J in Hz).
Table 4. 1H and 13C NMR spectroscopic data of 2526 in DMSO-d6 (δ in ppm, J in Hz).
No.2526
14.56, dd (10.7, 5.4)
3.69, dd (13.4, 10.7)
68.6 CH24.56, dd (10.7, 5.4)
3.67, m
68.6 CH2
3 165.0 C 164.9 C
45.58, s101.8 CH5.59, s101.7 CH
4a 151.0 C 150.9 C
55.54, d (2.2)114.4 CH5.55, d (2.1)114.3 CH
6 200.6 C 200.5 C
7 77.1 C 77.0 C
83.45, dd (10.1, 4.5)73.7 CH3.45, dd (10.1, 4.6)73.6 CH
8a2.58, m37.7 CH2.59, m37.7 CH
92.25, m44.2 CH22.32, dd (14.0, 6.8)
2.17, dd (14.0, 6.2)
44.1 CH2
103.86, dq (12.1, 6.2)64.1 CH3.88, m64.0 CH
111.07, d (6.2)18.6 CH31.07, d (6.2)18.6 CH3
121.10, s23.4 CH31.10, s23.3 CH3
7-OH4.93, s 4.92, s
8-OH5.33, d (4.5) 5.31, d (4.6)
10-OH4.69, d (5.0) 4.69, d (5.2)
Table 5. 1H and 13C NMR spectroscopic data of 2729 in DMSO-d6 (δ in ppm, J in Hz).
Table 5. 1H and 13C NMR spectroscopic data of 2729 in DMSO-d6 (δ in ppm, J in Hz).
No.272829
2 160.9 C 161.9 C 170.5 C
35.75, s109.8 CH5.85, s90.3 CH 126.5 C
4 176.9 C 165.7 C3.09, dd (17.5, 8.6)
2.55, dd (17.0, 3.0)
31.5 CH2
4a 108.7 C 1 115.6 C
5 138.4 C7.10, d (2.8)106.8 CH5.03, q (7.1)77.4 CH
66.19, s120.0 CH 153.7 C
7 159.1 C7.06, dd (8.8, 2.9)120.7 CH
7a
8 75.3 C7.25, d (8.8)117.5 CH
8a - 2 146.0 C
92.22, s19.1 CH3
102.48, s20.4 CH3
1′ 5.73, m128.7 CH
2′ 5.83, m135.4 CH
3′ 3.98, brs71.6 CH
4′ 3.30, m66.2 CH2
5′
1″ 6.54, tt (7.6, 3.0)141.4 CH
2″ 2.16, m23.3 CH2
3″ 1.02, t (7.5)12.9 CH3
4-OMe 3.99, s57.0 CH3
6-OH 9.90, brs
3′-OH 4.92, brs
4′-OH 4.64, brs
1 Observed in HMBC spectrum, 2 missed signal.
Table 6. Antibacterial activity of isolated compounds (MIC, μg/mL).
Table 6. Antibacterial activity of isolated compounds (MIC, μg/mL).
BacteriaCompounds
Chloramphenicol79101819202122243031
MRSA864---6464---64-
PA416-32-164---3232
EC0.2532320.50.58432328464
KP8646464-6464-64646464
VAl1-64--32646432--64
AH0.532320.5484-0.50.58-
ML13232--84---8-
VAn164---864---16-
VP132321284640.5643232
VV4-3216-84-----
VH232160.532168-32648-
* MRSA: methicillin-resistant S. aureus, PA: P. aeruginosa, EC: E. coli, KP: K. pneumonia, VAl: V. alginolyticus, AH: A. hydrophilia, ML: M. luteus, VAn: V. anguillarum, VP: V. parahaemolyticus, VV: V. vulnificus, VH: V. harveyi. -: no activity.
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

Wu, Z.; Li, X.-M.; Yang, S.-Q.; Wang, B.-G.; Li, X. Antibacterial Polyketides from the Deep-Sea Cold-Seep-Derived Fungus Talaromyces sp. CS-258. Mar. Drugs 2024, 22, 204. https://doi.org/10.3390/md22050204

AMA Style

Wu Z, Li X-M, Yang S-Q, Wang B-G, Li X. Antibacterial Polyketides from the Deep-Sea Cold-Seep-Derived Fungus Talaromyces sp. CS-258. Marine Drugs. 2024; 22(5):204. https://doi.org/10.3390/md22050204

Chicago/Turabian Style

Wu, Zhenger, Xiao-Ming Li, Sui-Qun Yang, Bin-Gui Wang, and Xin Li. 2024. "Antibacterial Polyketides from the Deep-Sea Cold-Seep-Derived Fungus Talaromyces sp. CS-258" Marine Drugs 22, no. 5: 204. https://doi.org/10.3390/md22050204

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