Next Article in Journal
Once upon a Time, There Was a Piece of Wood: Present Knowledge and Future Perspectives in Fungal Deterioration of Wooden Cultural Heritage in Terrestrial Ecosystems and Diagnostic Tools
Next Article in Special Issue
Control of Fusarium Head Blight of Wheat with Bacillus velezensis E2 and Potential Mechanisms of Action
Previous Article in Journal
Development of a Combat-Relevant Murine Model of Wound Mucormycosis: A Platform for the Pre-Clinical Investigation of Novel Therapeutics for Wound-Invasive Fungal Diseases
Previous Article in Special Issue
The Transcription Factors AcuK and AcuM Influence Siderophore Biosynthesis of Aspergillus fumigatus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Discovery of Acremochlorins O-R from an Acremonium sp. through Integrated Genomic and Molecular Networking

1
College of Chemical Engineering, Nanjing Forestry University, Nanjing 210037, China
2
Key Laboratory of Marine Drugs, Chinese Ministry of Education, School of Medicine and Pharmacy, Ocean University of China, 5 Yushan Road, Qingdao 266003, China
3
Ningbo Customs District Technology Center, Ningbo 315100, China
4
Ningbo Institute of Marine Medicine, Peking University, Ningbo 315832, China
5
Department of Pharmacology and Toxicology, College of Pharmacy, University of Utah, Salt Lake City, UT 84112, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
J. Fungi 2024, 10(5), 365; https://doi.org/10.3390/jof10050365
Submission received: 1 April 2024 / Revised: 16 May 2024 / Accepted: 17 May 2024 / Published: 20 May 2024
(This article belongs to the Special Issue Fungal Metabolism in Filamentous Fungi: 2nd Edition)

Abstract

:
The fermentation of a soil-derived fungus Acremonium sp. led to the isolation of thirteen ascochlorin congeners through integrated genomic and Global Natural Product Social (GNPS) molecular networking. Among the isolated compounds, we identified two unusual bicyclic types, acremochlorins O (1) and P (2), as well as two linear types, acremochlorin Q (3) and R (4). Compounds 1 and 2 contain an unusual benzopyran moiety and are diastereoisomers of each other, the first reported for the ascochlorins. Additionally, we elucidated the structure of 5, a 4-chloro-5-methylbenzene-1,3-diol with a linear farnesyl side chain, and confirmed the presence of eight known ascochlorin analogs (613). The structures were determined by the detailed interpretation of 1D and 2D NMR spectroscopy, MS, and ECD calculations. Compounds 3 and 9 showed potent antibacterial activity against Staphylococcus aureus and Bacillus cereus, with MIC values ranging from 2 to 16 μg/mL.

Graphical Abstract

1. Introduction

Ascochlorins (ASCs) and their congeners are primarily produced by filamentous fungi and represent a unique class of polyketide–terpenoid hybrid natural products. They are generally characterized by the presence of an orsellinic acid unit combined with a sesquiterpene (C15) moiety [1]. They have attracted extensive attention from pharmacologists. Owing to their distinctive structural diversity, they have been reported to exhibit a range of promising biological activities including antitumor [2], anti-inflammatory [3], antimicrobial [4,5], and anti-trypanosome [6,7]. Recent studies indicated that they showed potent hDHODH inhibitory activity, which means they have been involved in the treatment of cancers.
A recent review summarized their structure isolation and identification, biosynthesis, and biological activity in 2023. To date, about 71 ASCs have been reported from filamentous fungi and are classified into three main chemotypes, the linear type, the monocyclic type, and the bicyclic type, which account for about 30%, 65%, and 0.6% of all the ASCs, respectively [8]. In addition, the gene cluster and biosynthesis of the representative products of ASCs, such as ascofuranone and ascochlorin, have been characterized, which are conducive to the discovery of more ASC congeners by genomic mining [1].
The filamentous fungus Acremonium sp. was isolated from soil collected on the University of Utah campus, UT, USA, and was shown to produce a rare class of 15-residue peptaibols [9]. In that study, the 38 Mbp genome of the strain was also reported to harbor 44 putative biosynthetic gene clusters, including 1 predicted for ASCs’ biosynthesis. As no ASCs had been previously reported from the target strain, we implemented genomic and GNPS molecular networking to determine if this orphan biosynthetic cluster was indeed responsible for the production of ASCs. Guided by UV absorption and MS data, we identified four undescribed ASCs (14), a newly natural product (5), and eight known ASC analogs (613). Herein, details of their isolation, structure elucidation, and antibiotic activities are described.

2. Material and Methods

2.1. General Experimental Procedures

Optical rotations were measured on an Anton Paar MCP 5500 polarimeter equipped with a sodium lamp (589 nm) and a 25 mm microcell. A Jasco J1500 spectrometer (Jasco Inc., Tokyo, Japan) was used to obtain the electric circular dichroism (ECD) spectra. The 1D and 2D NMR experiments (1H, 13C, NOESY, COSY, HSQC, and HMBC) were performed at 300 K in CDCl3 on a Bruker Avance Neo 600MHz spectrometer (Bruker BioSpin, GmBH) equipped with a Bruker 5 mm PI HR-BBO600s3 Probe. HR-ESIMS was utilized on an LTQ Orbitrap XL mass spectrometer or an Orbitrap Eclipse mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) equipped with Xcalibur 4.0 software. The 200–300 mesh silica gel (Shanghai, China), ODS (12 nm, YMC*GEL), and TOYOPEARL HW-40F were employed for column chromatography. HPLC separations were conducted on an Agilent 1260 series pumping system equipped with an Agilent DAD-G7115A refractive index detector on an X-bridge C18 column (4.6 × 250 mm, 5 µm, flow rate 1.0 mL/min). RP-HPLC separations were conducted on a Shimadzu LC-20AP series pumping system equipped with a Shimadzu SPD-20A refractive index detector and Shimadzu injector on an X-bridge C18 column (10 × 250 mm, 5 µm, flow rate 4.0 mL/min). TLC analyses were carried out using precoated HF254 (0.20 mm thickness) plates (Nuotai, Shanghai, China); compounds were detected by 10% sulfuric acid/ethanol (Sigma-Aldrich, St. Louis, MO, USA). All MS/MS data were converted to mzXML format files by ProteoWizard 3.0 software [10]. Molecular networking was performed using the GNPS data analysis workflow and the spectral clustering algorithm. The spectral networking was imported into Cytoscape (ver. 3.6.1) for visualization.

2.2. Fungal Material

The fungal strain Acremonium sp. (GenBank accession number MT053262) was originally isolated from soil collected on the University of Utah campus, UT, USA, and formally characterized as an Acremonium sp. in a previous publication from our group [9].

2.3. Incubation and Extraction

Acremonium sp. was cultured on the seed medium Potato Dextrose Agar plates (PDA medium: 20 g of potato extract powder, 20 g of glucose, 18 g of agar in 1 L of tap distilled H2O) at 28 °C for four to five days. Subsequently, the large-scale fermentation of Acremonium sp. was performed using modified rice solid medium (80 g of rice, 3.0 g/L of NaNO3, and 120 mL of H2O). Briefly, 120 mL of rice medium was added to a 48 × 1 L Erlenmeyer flask and inoculated using two 5 × 5 mm2 agar plugs from the PDA plates. The cultures were incubated at room temperature under static conditions, and after 30 days, they were harvested for chemical analysis. The fermented cultures were extracted three times with equal volumes of EtOAc, and the organic extracts were combined and concentrated under vacuo to provide a crude extract (87.1 g).

2.4. Isolation and Purification

Using normal silica gel chromatography, the crude extract was fractionated into nine fractions using different concentrations of petroleum ether, dichloromethane, and methanol. Fr.2 (19.1 g) eluted with 100% dichloromethane was further separated into 7 subfractions (Frs.2-1∼2-7) via ODS silica gel elution using a mixture of H2O/MeOH. Fr.2-6 eluted with MeOH/H2O (v/v, 80:20) was separated into twelve subfractions (Frs.2-6-1∼2-6-12) via preparative HPLC (85:15 MeCN–H2O with 0.1% formic acid, 4 mL/min, 205 nm and 254 nm) using an ODS column. Fr.2-6-6 was purified by preparative HPLC (60:40 MeCN–H2O with 0.1% formic acid, 4 mL/min, 205 nm and 254 nm) to afford 10 (7 mg, tR 27 min), 12 (69 mg, tR 32 min), 13 (2.2 mg, tR 43 min). Fr.2-6-7 was purified by preparative HPLC (65:35 MeCN–H2O with 0.1% formic acid, 4 mL/min, 205 nm and 254 nm) to yield 3 (48.9 mg, tR 43 min). Fr.2-6-8 was purified by preparative HPLC (70:30 MeCN–H2O with 0.1% formic acid, 4 mL/min, 205 nm and 254 nm) to afford 11 (27.3 mg, tR 27 min) and 4 (1.1 mg, tR 38 min). Fr.2-7 was further separated into seven subfractions (Frs.2-7-1∼2-7-7) via HW-40F silica gel elution with CH2Cl2/MeOH (v/v, 1:1) according to HPLC profiles. Among them, Fr.2-7-6 was purified by preparative HPLC (70–100% MeCN–H2O with 0.1% formic acid, 4 mL/min, 205 nm and 254 nm) to yield 9 (2.9 mg, tR 60 min), 1 (1.3 mg, tR 41 min), 2 (1.2 mg, tR 42 min), 5 (4.2 mg, tR 50 min), 6 (4.2 mg, tR 51.7 min), 7 (2.1 mg, tR 53.5 min), 8 (2.1 mg, tR 58.5 min).
Acremochlorin O (1): yellow oil; [α]25D + 8 (c 0.1, MeOH); UV (MeOH) λmax (log ε) 270 (4.06), 322 (1.09), 198 (2.33); ECD (0.15 mg/mL, MeOH) λmaxε) 207 (+14.40), 232 (−8.99), 269 (+0.88), 318 (−7.15); 1H and 13C NMR data, Table 1, Figures S6–S17; HR-ESIMS m/z 405.1840/407.1804 (3:1) ([M+H]+/[M+2+H]+, calcd for C23H30ClO4, 405.1827/407.1798).
Acremochlorin P (2): yellow oil; [α]25D—35.6 (c 0.18, MeOH); UV (MeOH) λmax (log ε) 270 (4.81), 322 (1.99), 200 (2.82); ECD (0.15 mg/mL, MeOH) λmaxε) 204 (−20.44), 232 (+3.41), 272 (−10.58), 322 (+6.25); 1H and 13C NMR data, Table 1, Figures S18–S29; HR-ESIMS m/z 405.1843/407.1807 (3:1) ([M+H]+/[M+2+H]+, calcd for C23H30ClO4, 405.1827/407.1798).
Acremochlorin Q (3): brown amorphous powder; [α]25D—42.8 (c 0.29, CH2CL2); UV (MeOH) λmax (log ε) 270 (4.06), 322 (1.09), 198 (2.33); ECD (0.15 mg/mL, MeOH) λmaxε) 205 (+9.01), 305 (−5.25); UV (MeOH) λmax 298 (3.52), 240 (3.18), 338 (1.43); 1H and 13C NMR data, Table 1, Figures S30–S37; HR-ESIMS m/z 421.1775/423.1741 (3:1) ([M-H2O+H]+/[M+2-H2O+H]+, calcd for C23H30ClO5, 421.1776/423.1747).
Acremochlorin R (4): yellow oil; [α]25D—3.6 (c 0.1, CH2CL2); UV (MeOH) λmax (log ε) 294 (4.06), 336 (3.87), 232 (4.16); 1H and 13C NMR data, Table 1; HR-ESIMS m/z 465.2039/467.2012 (3:1) ([M-H2O+H]+/[M+2-H2O+H]+, calcd for C25H34ClO6, 465.2038/467.2009).
4-chloro-5-methyl-2-((2E,6E)-3,7,11-trimethyldodeca-2,6,10-trien-1-yl)benzene-1,3-diol (5): brown oil; UV (MeOH) λmax (log ε) 198 (1.22), 282 (2.43), 342 (1.70); 1H and 13C NMR data, Table 1, Figures S46–S51; HR-ESIMS m/z 363.2085/365.2055 (3:1) ([M+H]+/[M+2+H]+, calcd for C22H32ClO2, 363.2085/365.2056).

2.5. Computation Section

Conformational searches were carried out using Spartan’14 (Wavefunction Inc., Irvine CA USA), based on the MMFF94. All conformers were optimized with DFT calculations at the B3LYP/6-31+G(d) level using the Gaussian 09 program [11,12]. For ECD calculations, TDDFT calculations were performed on the two lowest-energy conformations for 1 and 2 (>5% population) at the B3LYP/6-31+G(d) levels. In addition, the four lowest-energy conformations for 3 (>5% population) were calculated.

2.6. Antimicrobial Activities

The minimum inhibitory concentrations (MICs) were determined in 96-well plates using the microdilution method to screen compounds 19 and 11 for bioactivity. To prepare the inoculum for susceptibility testing, bacteria were streaked independently onto LB agar plates and incubated overnight at 37 °C. Individual colonies were then isolated and transferred to 50 mL of LB liquid medium and incubated at 37 °C for 4–6 h. The culture density was adjusted with LB liquid medium so that a concentration of 5 × 106 cfu/mL was achieved. Compounds 19 and 11 were tested for their individual activity against S. aureus, MRSA, MRCNS, B. Subtilis, and B. cereus using chloramphenicol as a positive control (64 μg/mL dissolved in DMSO). Briefly, 19 and 11 were dissolved in DMSO to generate 128 mg/mL stock solutions. The stock solutions were then serially diluted with LB liquid medium to afford working concentrations of 128 to 2 μg/mL. More specifically, to a 96-well microtiter plate, 2 μL stock solutions mixed with 98 μL of LB liquid medium was added to well A1. From this mixture, 50 μL was transferred to well A2 and mixed with 50 μL of fresh LB media. This process was repeated across the 96-well plate, and 50 μL of the appropriate bacterial cultures was then added to each well. The plates were incubated at 37 °C for 16–20 h. MIC values were determined by visual inspection and verified with an OD600 measurement using a BioTek Neo2 plate reader (Agilent, Winooski, VT, USA). The respective MIC values for 19 and 11 are reported in Table 2. All the pathogenic strains were clinical isolates and donated by the Marine Medicinal Biological Resources Center, Ocean University of China. Specific strain information can be found in Table S2.

2.7. Hydroxyl Radical Scavenging Activity

The Fenton reaction was used to produce hydroxyl radicals, which reacted with salicylic acid to form 2,3-dihydroxybenzoic acid with special absorption at 510 nm.
The test compound was prepared into a 200 mM solution with DMSO as the solvent. Then, 25 μL of the 200 mM sample solution, 25 μL of 9 mM FeSO4·7H2O, 25 μL of 9 mM salicylic acid, and 25 μL of 8.8 mM H2O2 were added into the 96-well plate successively and mixed well. After heating in a 37 °C water bath for 30 min, it was taken out and its absorbance at 510 nm was measured using a BioTek Neo2 plate reader.

3. Results and Discussion

The 38 Mbp genome of Acremonium sp. was previously sequenced and assembled, and the antiSMASH analysis revealed that a gene cluster, hereby named ascw, showed high similarity (87%) at the amino acid level to the characterized asc-1 gene cluster (Figure 1 and Figure S1). A more detailed bioinformatic analysis of the ascw gene cluster revealed that all eight of the genes encoding enzymatic machinery responsible for ASCs’ assembly in A. egyptiacum were present in ascw, suggesting the ability of Acremonium sp. to produce ASCs [1].
To evaluate whether the strain significantly produced ASCs or not, the fungal strain was cultured in rice media (80 g of rice, 3.0 g/L of NaNO3, and 120 mL of H2O) for 30 days. The EtOAc extract was evaluated by LC-MS/MS in the positive mode, and the data were processed through GNPS (http://gnps.ucsd.edu, accessed on 22 December 2023). The obtained molecular networking featured 13 clusters and 91 nodes, with GNPS analysis uncovering a cluster of 23 nodes matching ASC compounds, which displayed typical isotopic peaks for monochloride compounds in the grouped structure (Figure 2, Figure S2 and Figure S3). Known compounds 8 and 13 were directly identified by molecular networking. Based on the UV absorption of known compounds, further targeted isolation resulted in a total of 13 Ascochlorin (ASC) derivatives (Figure 2). Four compounds were new, including 1 and 2 (m/z: 405. 1840 and 405.1843), 3 (m/z: 421. 1775), and 4 (m/z: 465. 2039).
Compound 1 was obtained as a yellow oil. The HR-ESIMS exhibited a characteristic pseudomolecular ion at m/z 405.1840/407.1804 in a ratio of 3:1 ([M+H]+/[M+2+H]+, calcd for C23H30ClO4, 405.1827/407.1798), suggesting the presence of a chlorine atom in 1 and supporting the molecular formula of C23H31ClO5 containing nine degrees of unsaturation. Further analysis of the 1D NMR and HSQC data showed the presence of a hexasubstituted benzene moiety (δC 113.6, 158.9, 107.9, 156.3, 116.2, and 140.8), a non-conjugated ketone carbon (δC 213.8), one aldehyde carbon (δC/H 193.5/10.13), a double bond (δC/H 116.2/6.73 and 126.5/5.53), five methyls, four sp3 methylenes, two sp3 methines, and two quaternary carbons (one oxygenated carbon δC 82.0 and one sp3 δC 43.3), indicating the existence of two additional ring systems in the structure of 1 (Table 1).
The comparison of NMR data between 1 and co-isolated ilicicolin C (9) revealed that both compounds share identical monochlorinated benzaldehyde and cyclohexanone moieties (Figure 3). In 1, the double bond is located between C-9 and C-10, whereas it is in position between C-10 and C-11 in 9. This was confirmed by the 1H-1H COSY correlations of H-9/H-10. In addition, an oxygenated quaternary carbon was confirmed at C-11 (δC 82.0), supported by the HMBC correlations from Me-23 (δC/H 27.6/1.49) to C-10 (δC 126.5), C-11 and C-12 (δC 34.6), and H-9 (δC/H 116.2/6.73) to C-10, and C-11 (Figure 4). Subsequently, the HMBC correlations of H-9 with C-2 (δC 158.9), C-3 (δC 107.9), C-4 (δC 156.3), and Me-20 (δC/H 15.6/0.58) with C-13 (δC 30.8) allowed us to establish the connections of the monochlorinated benzaldehyde and cyclohexanone group by a single bond between C-3 and C-9. To satisfy the degree of unsaturation, the molecular formula and downfield chemical shift of C-11 (δC 82.0), a benzopyran moiety, was proposed, thus assigning the planar structure of 1.
Compound 2 was obtained as yellow oil and was determined by HR-ESIMS data to be at 405.1843/407.1807 (3:1) ([M+H]+/[M+2+H]+, calcd for C23H30ClO4, 405.1827/407.1798) and to have the same molecular formula of C23H29ClO4 as 1. When isolated using HPLC, compounds 1 and 2 eluted as adjacent peaks (Figure S4). The detailed inspection of 1D and 2D NMR data revealed that compounds 1 and 2 share identical planar structures. Slight differences in chemical shifts, primarily within the cyclohexanone moiety between 1 and 2, were observed, including Me-21 (δC/H 15.1/0.85 for 1 vs. 15.0/0.90 for 2), Me-22 (δC/H 7.5/0.93 for 1 vs. 7.6/0.88 for 2), H-19 (δC/H 50.6/2.44 for 1 vs. 50.6/2.40 for 2), H-15 (δC/H 36.3/1.94 for 1 vs. 36.4/1.97 for 2), and H-12 (δC/H 34.6/1.70 for 1 vs. 34.6/1.86 for 2) (Figure S5). Thus, 1 and 2 are isomers of each other.
The relative configurations of 1 and 2 were assigned by key NOESY correlations and coupling constants (Table 1 and Figure 5). The signal intensity of H-15 and H-13 (δC/H 30.8/1.61 for 1 and 30.6/1.57 for 2) increased after the irradiation of H-19, indicating a similar relative configuration of the cyclohexanone moiety in 1 and 2. The Z configuration of the Δ9(10) double bond of 1 and 2 was deduced through a strong NOESY correlation between H-9 and H-10 and relatively small coupling constants (JH-9/H-10 = 10.1 Hz). The relative configuration of C-11 was not deduced.
The ECD calculation and biosynthetic origin were involved in the assignment of the absolute configuration of 1 and 2. Surprisingly, the experimental ECD curves of 1 and 2 showed almost opposite cotton effects (Figure 6). We propose that the observed differences in the ECD cotton effects are primarily due to the benzopyran moiety, rather than the cyclohexanone group, as shown in a previous study, and the ECD method is suitable for the assignment of C-11 [13]. Thus, theoretical ECD calculations were performed using the time-dependent density functional theory (TD-DFT) approach. As shown in Figure 6, the experimental ECD spectrum of 1 displayed a good match with the calculated spectrum of 11S, while the curve of 2 was in accord with that of the 11R. The 14S, 15R, and 19R of 1 and 2 were deduced on the basis of the enzyme-mediated formation of the cyclohexanone group [1,2] and the same relative configuration of this moiety to that of co-isolated compound 9. In fact, the literature survey revealed that all the cyclohexanone groups in ASCs share ommon stereochemistry without exception [8]. Thus, the absolute configurations were finally determined to be 9Z, 11S, 14S, 15R, 19R-1 and 9Z, 11R, 14S, 15R, 19R-2, respectively, indicating that compounds 1 and 2 are diastereoisomers.
Compound 3 was obtained as a brown, amorphous powder. The molecular formula C23H31ClO6 was established by HR-ESIMS at m/z 421.1775/423.1741 ([M-H2O+H]+/[M+2-H2O+H]+, calcd for C23H30ClO5, 421.1776/423.1747). The NMR data of 3 were highly similar to those of the co-isolated chlorocylindrocarpol (6), suggesting that 3 contained an acyclic sesquiterpene moiety (Figure 3) [14]. The only difference between 3 and 6 is that the double bond group of C-16 and C-18 was substituted by a hydroxyl group and a non-conjugated ketone moiety in 3. The differences were supported by the COSY correlations of H-16/H-17 and HMBC correlations of Me-20 (δC/H 24.4/1.28) with C-18 (δC 218.1) and Me-22 (δC/H 11.3/1.63) with C-14 (δC/H 128.7/5.5), C-15 (δC 133.1), C-16 (δC/H 77.9/4.52) (Figure 4). The E configuration of both Δ10 and Δ14 double bonds was assigned by NOESY correlations between H2-9 and Me-23, H-10 and H2-12, H2-13 and Me-22, and H-14 and H-16, respectively (Figure 5). The ECD calculation was used to address the absolute configuration of C-16; the calculated ECD curve of 3 showed positive Cotton effects at around 200–250 nm and negative Cotton effects at around 280–320 nm, coinciding well with the experimental ECD spectrum and suggesting a 16R-configuration in 3 (Figure 7).
Compound 4 was obtained as a yellow oil. Analysis of the HR-ESIMS data showed a characteristic pseudomolecular ion indicative of a monochloroinated compound at m/z 465.2039/467.2012 in a ratio of 3:1 ([M-H2O+H]+/[M+2-H2O+H]+) and gave the molecular formula of C25H35ClO7. The high similarity of NMR spectroscopic data of 4 to compound 3 suggested that both structures were closely related. The difference between compound 3 and 4 was the O-acetylation of the ketone moiety at C-18 in 3, which was confirmed by COSY correlations between H-16, H-17 and H-18, and the key HMBC between Me-21 (δC/H 25.5/1.23) and H-18 (δC/H 79.6/4.99). Additionally, the acetyl group was determined to be attached to OH-18 through HMBC correlations between H-1′ (δC/H 21.2/2.07) and H-18 with C-2′ (δC 170.8). Both Δ10 and Δ14 double bonds were also assigned as E configuration by NOESY correlations (Figure 5). Attempts to obtain crystals for further analysis were unsuccessful.
Compound 5 was isolated as a newly natural product, which was initially reported as a chemically synthesized product, named 4-chloro-5-methyl-2-((2E,6E)-3,7,11-trimethyldodeca-2,6,10-trien-1-yl)benzene-1,3-diol (5). The structure of 5 was determined by comparing NMR data [15]. In addition to the five new structures, eight known ascochlorin derivatives, chlorocylindrocarpol (6) [14], grifolic acid (7) [16], ilicicolin A (8) [17], ilicicolin C (9) [18], LL-Z 1272e (10) [19], cylindrochlorin (11) [20], ilicicolin F (12) [19], and ascofuranone (13) [21], were identified by comparison with published NMR data.
Compounds 19 and 11 were assayed for their antimicrobial activities against Gram-positive pathogenic bacteria Staphylococcus aureus ATCC29213, methicillin-resistant Staphylococcus aureus (MRSA), methicillin-resistant coagulase-negative staphylococci (MRCNS), and Bacillus cereus, as well as the plant pathogenic fungi Botrytis cinerea, Fusarium graminearum, Colletotrichum, Fusarium oxysporum, and Exobasidium vexans. In summary, all compounds lacked inhibitory effects against the plant pathogenic fungi, whereas compound 3 exhibited potent inhibitory antibacterial effects with MIC values ranging from 2 to 8 μg/mL (Table 2), which exceeded the positive control, chloramphenicol [22,23]. The preliminary analysis of the structure–activity relationship revealed that the ketone moiety, rather than the acetoxyl group at C-18, is helpful to improve the antibacterial activity. In addition, compounds 15 also underwent testing for antioxidant activity. They displayed moderate antioxidant properties with hydroxyl radical clearance rates of 64.14%, 65.77%, 67.04%, 68.61%, and 69.32%, respectively, while the positive control, vitamin C, exhibited an 81.69% hydroxyl radical clearance rate at a concentration of 50 μM.

4. Conclusions

In summary, using integrated genomic and GNPS molecular networking, four undescribed ASC congeners, a newly natural product, and seven known ones were discovered from the soil-derived fungus Acremonium sp. Particularly, Acremochlorin O (1) and Acremochlorin P (2) possessed an unusual benzopyran moiety and were diastereoisomers of each other that had not been discovered in ASCs to date. Our finding indicated that ASCs have promising potential as lead compounds for developing new antibacterial agents.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/jof10050365/s1: The GNPS, NMR, HRESIMS, IR, and ECD data for 15 are included.

Author Contributions

G.C.: data curation, formal analysis, investigation, writing—original draft. L.Z.: methodology, software, data curation, investigation. X.Q.: data curation, formal analysis, investigation. P.Y.: methodology, validation. L.C.: formal analysis. P.W.: formal analysis. D.L.: resources, supervision. H.L.: methodology, supervision, writing—review and editing. J.M.W.: resources, supervision, writing—review and editing. G.W.: funding acquisition, methodology, project administration, resources, supervision, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Start-up Research Fund from the Nanjing Forestry University, China (163030196), the Student Practice Innovation and Training Program of Nanjing Forestry University (Grant Nos. 2021NFUSPITP0035, 2022NFUSPITP0061, and 2023NFUSPITP0063).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Araki, Y.; Awakawa, T.; Matsuzaki, M.; Cho, R.; Matsuda, Y.; Hoshino, S.; Shinohara, Y.; Yamamoto, M.; Kido, Y.; Inaoka, D.K.; et al. Complete biosynthetic pathways of ascofuranone and ascochlorin in Acremonium egyptiacum. Proc. Natl. Acad. Sci. USA 2019, 116, 8269–8274. [Google Scholar] [CrossRef]
  2. Luo, X.; Cai, G.; Guo, Y.; Gao, C.; Huang, W.; Zhang, Z.; Lu, H.; Liu, K.; Chen, J.; Xiong, X.; et al. Exploring Marine-Derived Ascochlorins as Novel Human Dihydroorotate Dehydrogenase Inhibitors for Treatment of Triple-Negative Breast Cancer. J. Med. Chem. 2021, 64, 13918–13932. [Google Scholar] [CrossRef]
  3. Lee, S.; Kwak, C.; Lee, S.; Ha, S.; Park, J.; Chung, T.; Ha, K.; Suh, S.; Chang, Y.; Chang, H.W.; et al. Anti-Inflammatory Effect of Ascochlorin in LPS-Stimulated RAW 264.7 Macrophage Cells Is Accompanied with the Down-Regulation of iNOS, COX-2 and Proinflammatory Cytokines Through NF-κB, ERK1/2, and p38 Signaling Pathway. J. Cell. Biochem. 2016, 117, 978–987. [Google Scholar] [CrossRef]
  4. Wu, B.; Oesker, V.; Wiese, J.; Malien, S.; Schmaljohann, R.; Imhoff, J.F. Spirocyclic drimanes from the marine fungus Stachybotrys sp. strain MF347. Mar. Drugs 2014, 12, 1924–1938. [Google Scholar] [CrossRef] [PubMed]
  5. Suzuki, T.; Yoshida, S.; Koseki, T.; Aboshi, T.; Murayama, T.; Supratman, U.; Shiono, Y. New Metabolites Produced by Cylindrocarpon sp. SY-39 from a Driftwood. Chem. Biodivers. 2018, 15, e1700493. [Google Scholar] [CrossRef]
  6. Fukai, Y.; Amino, H.; Hirawake, H.; Yabu, Y.; Ohta, N.; Minagawa, N.; Sakajo, S.; Yoshimoto, A.; Nagai, K.; Takamiya, S.; et al. Functional expression of the ascofuranone-sensitive Trypanosoma brucei brucei alternative oxidase in the cytoplasmic membrane of Escherichia coli. CBPC 1999, 124, 141–148. [Google Scholar] [CrossRef]
  7. Otoguro, K.; Ishiyama, A.; Namatame, M.; Nishihara, A.; Furusawa, T.; Masuma, R.; Shiomi, K.; Takahashi, Y.; Yamada, H.; Ōmura, S. Selective and Potent In Vitro Antitrypanosomal Activities of Ten Microbial Metabolites. J. Antibiot. 2008, 61, 372–378. [Google Scholar] [CrossRef] [PubMed]
  8. Gao, H.; Zhou, L.; Zhang, P.; Wang, Y.; Qian, X.; Liu, Y.; Wu, G. Filamentous Fungi-Derived Orsellinic Acid-Sesquiterpene Meroterpenoids: Fungal Sources, Chemical Structures, Bioactivities, and Biosynthesis. Planta Med. 2023, 89, 1110–1124. [Google Scholar] [CrossRef] [PubMed]
  9. Wu, G.; Dentinger, B.; Nielson, J.; Peterson, R.; Winter, J. Emerimicins V-X, 15-Residue Peptaibols Discovered from an Acremonium sp. through Integrated Genomic and Chemical Approaches. J. Nat. Prod. 2021, 84, 1113–1126. [Google Scholar] [CrossRef]
  10. Chambers, M.C.; Maclean, B.; Burke, R.; Amodei, D.; Ruderman, D.L.; Neumann, S.; Gatto, L.; Fischer, B.; Pratt, B.; Egertson, J.; et al. A cross-platform toolkit for mass spectrometry and proteomics. Nat. Biotechnol. 2012, 30, 918–920. [Google Scholar] [CrossRef]
  11. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Revision A.1; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  12. Bruhn, T.; Schaumlöffel, 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] [PubMed]
  13. Ma, R.; Liu, H.; Zhao, X.; Shan, P.; Sun, P.; Xue, J.; Wei, G.; Zhang, H. Meroterpenoids from Daphne genkwa shows promising in vitro antitumor activity via inhibiting PI3K/Akt/mTOR signaling pathway in A549 cells. Bioorg. Chem. 2023, 140, 106803. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, P.; Bao, B.; Dang, H.T.; Hong, J.; Lee, H.J.; Yoo, E.S.; Bae, K.S.; Jung, J. Anti-inflammatory sesquiterpenoids from a sponge-derived Fungus Acremonium sp. J. Nat. Prod. 2009, 72, 270–275. [Google Scholar] [CrossRef] [PubMed]
  15. Grabovyi, G.A.; Mohr, J.T. Total Synthesis of Grifolin, Grifolic Acid, LL-Z1272α, LL-Z1272β, and Ilicicolinic Acid A. Org. Lett. 2016, 18, 5010–5013. [Google Scholar] [CrossRef] [PubMed]
  16. Zechlin, L.; Wolf, M.; Steglich, W.; Anke, T. Antibiotika aus Basidiomyceten, XII. Cristatsäure, ein modifiziertes Farnesylphenol aus Fruchtkörpern von Albatrellus cristatus. Liebigs Ann. Chem. 1981, 1981, 2099–2105. [Google Scholar] [CrossRef]
  17. Hayakawa, S.; Minato, H.; Katagiri, K. The ilicicolins, antibiotics from Cylindrocladium ilicicola. J. Antibiot. 1971, 24, 653–654. [Google Scholar] [CrossRef] [PubMed]
  18. Gutiérrez, M.; Theoduloz, C.; Rodríguez, J.; Lolas, M.; Schmeda-Hirschmann, G. Bioactive Metabolites from the Fungus Nectria galligena, the Main Apple Canker Agent in Chile. J. Agri. Food Chem. 2005, 53, 7701–7708. [Google Scholar] [CrossRef]
  19. Ellestad, G.A.; Evans, R.H., Jr.; Kunstmann, M.P. Some new terpenoid metabolites from an unidentified fusarium species. Tetrahedron 1969, 25, 1323–1334. [Google Scholar] [CrossRef] [PubMed]
  20. Kato, A.; Ando, K.; Tamura, G.; Arima, K. Cylindrochlorin, a new antibiotic produced by Cylindrocladium. J. Antibiot. 1970, 23, 168–169. [Google Scholar] [CrossRef]
  21. Sasaki, H.; Hosokawa, T.; Sawada, M.; Ando, K. Isolation and structure of ascofuranone and ascofranol, antibiotics with hypolipidemic activity. J. Antibiot. 1973, 26, 676–680. [Google Scholar] [CrossRef]
  22. Liang, M.; Ge, X.; Xua, H.; Ma, K.; Zhang, W.; Zan, Y.; Efferth, T.; Xue, Z.; Hua, X. Phytochemicals with activity against methicillin-resistant Staphylococcus aureus. Phytomedicine 2022, 100, 154073. [Google Scholar] [CrossRef] [PubMed]
  23. Ríos, J.L.; Recio, M.C. Medicinal plants and antimicrobial activity. J. Ethnopharmacol. 2005, 100, 80–84. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Genomic analysis of Acremonium sp. Organization of the ASCs’ biosynthetic gene cluster identified in Acremonium sp. (ascw) (GenBank Accession number PP795974) in comparison to the ASC gene cluster from A. egyptiacum (asc-1) (GenBank Accession number LC406756).
Figure 1. Genomic analysis of Acremonium sp. Organization of the ASCs’ biosynthetic gene cluster identified in Acremonium sp. (ascw) (GenBank Accession number PP795974) in comparison to the ASC gene cluster from A. egyptiacum (asc-1) (GenBank Accession number LC406756).
Jof 10 00365 g001
Figure 2. Ascochlorin (ASC) derivatives observed. (A) GNPS molecular networking highlighting the cluster associated with ASCs. Acremochlorins O–R (14) are shown as red nodes, known analogs ilicicolin A (8) and ascofuranone (13) are shown as blue nodes, and unknown compounds are shown in gray. (B) UV profile of ASCs.
Figure 2. Ascochlorin (ASC) derivatives observed. (A) GNPS molecular networking highlighting the cluster associated with ASCs. Acremochlorins O–R (14) are shown as red nodes, known analogs ilicicolin A (8) and ascofuranone (13) are shown as blue nodes, and unknown compounds are shown in gray. (B) UV profile of ASCs.
Jof 10 00365 g002
Figure 3. Structures of compounds 113 isolated from Acremonium sp.
Figure 3. Structures of compounds 113 isolated from Acremonium sp.
Jof 10 00365 g003
Figure 4. Key HMBC and COSY correlations of compounds 14.
Figure 4. Key HMBC and COSY correlations of compounds 14.
Jof 10 00365 g004
Figure 5. Key NOESY correlations of compounds 14.
Figure 5. Key NOESY correlations of compounds 14.
Jof 10 00365 g005
Figure 6. Experimental and calculated ECD spectra for 1 and 2.
Figure 6. Experimental and calculated ECD spectra for 1 and 2.
Jof 10 00365 g006
Figure 7. Experimental and calculated ECD spectra of compound 3.
Figure 7. Experimental and calculated ECD spectra of compound 3.
Jof 10 00365 g007
Table 1. The 1H NMR and 13C NMR data of compounds 14 in CDCl3.
Table 1. The 1H NMR and 13C NMR data of compounds 14 in CDCl3.
Compound1234
PositionδCδH (J in Hz)δCδH (J in Hz)δCδH (J in Hz)δCδH (J in Hz)
1113.6 113.6 113.7 113.7
2158.9 158.9 162.3 162.3
3107.9 107.9 114.5 114.5
4156.3 156.3 156.5 156.6
5116.2 116.2 113.3 113.4
6140.8 140.8 137.8 137.8
714.72.60, s14.72.60, s14.62.61, s14.6
8193.510.13, s193.510.13, s193.510.14, s193.410.14, s
9116.26.73, d
(10.1)
116.26.73, d
(10.1)
22.13.39, d
(7.1)
22.13.39, d
(7.1)
10126.55.53, d
(10.1)
126.65.53, d
(10.1)
121.45.21, t
(7.3)
121.25.20, t
(7.2)
1182.0 82.0 136.1 136.5
1234.61.70, m
overlapping
34.61.86, m
overlapping
39.12.04, m39.22.01, m
1330.81.61, 1.44, m
overlapping
30.81.57 1.44, m
overlapping
26.12.16, m26.22.11, m
1443.3 43.3 128.75.50, t
(7.2)
127.55.41, t
(7.0)
1536.31.94, m36.41.97, m133.1 134.3
1631.01.84, 1.62, m
overlapping
31.01.84, 1.63, m
overlapping
77.94.52, m80.94.31, t
(7.7)
1741.72.33, m41.72.33, m40.12.38, m37.11.73, 2.44, m
18213.8 213.7 218.1 79.64.99, dd
(4.4, 7.0)
1950.62.44, q
(6.7)
50.62.40, q
(6.8)
80.9 81.9
2015.60.58, s15.60.58, s24.41.28, s22.81.22, s
2115.10.85, d
(6.8)
15.00.90, d
(7.0)
22.11.22, s25.51.23, s
227.50.93, d
(6.7)
7.60.88, d
(7.1)
11.31.63, s11.11.59, s
2327.61.49, s27.51.49, s16.31.79, s16.31.77, s
1′-OAc 21.22.07, s
2′ 170.8
2-OH 12.71, s 12.70, s 12.69, s 12.69, s
4-OH 6.43, s 6.48, s
Table 2. Antimicrobial activities of compounds 19 and 11 (MIC, µg/mL).
Table 2. Antimicrobial activities of compounds 19 and 11 (MIC, µg/mL).
StrainS. aureus
ATCC29213
S. aureus
MRSA
S. aureus
MRCNS
B. cereus
Compounds
1>128>128>128>128
2>128>128>128>128
34824
432643232
5>128>128>12816
632323216
764643216
8>128>128>128>128
9416416
11>128643232
DMSO>128>128>128>128
chloramphenicol8844
All assays were performed in triplicate.
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

Cui, G.; Zhou, L.; Liu, H.; Qian, X.; Yang, P.; Cui, L.; Wang, P.; Li, D.; Winter, J.M.; Wu, G. The Discovery of Acremochlorins O-R from an Acremonium sp. through Integrated Genomic and Molecular Networking. J. Fungi 2024, 10, 365. https://doi.org/10.3390/jof10050365

AMA Style

Cui G, Zhou L, Liu H, Qian X, Yang P, Cui L, Wang P, Li D, Winter JM, Wu G. The Discovery of Acremochlorins O-R from an Acremonium sp. through Integrated Genomic and Molecular Networking. Journal of Fungi. 2024; 10(5):365. https://doi.org/10.3390/jof10050365

Chicago/Turabian Style

Cui, Ge, Luning Zhou, Hanwei Liu, Xuan Qian, Pengfei Yang, Leisha Cui, Pianpian Wang, Dehai Li, Jaclyn M. Winter, and Guangwei Wu. 2024. "The Discovery of Acremochlorins O-R from an Acremonium sp. through Integrated Genomic and Molecular Networking" Journal of Fungi 10, no. 5: 365. https://doi.org/10.3390/jof10050365

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