key: cord-1036972-46de2xq2 authors: Omar, Abdelsattar M.; Mohamed, Gamal A.; Ibrahim, Sabrin R. M. title: Chaetomugilins and Chaetoviridins—Promising Natural Metabolites: Structures, Separation, Characterization, Biosynthesis, Bioactivities, Molecular Docking, and Molecular Dynamics date: 2022-01-27 journal: J Fungi (Basel) DOI: 10.3390/jof8020127 sha: 18f775e1ad0570584b78df0908446526c864a140 doc_id: 1036972 cord_uid: 46de2xq2 Fungi are recognized as luxuriant metabolic artists that generate propitious biometabolites. Historically, fungal metabolites have largely been investigated as leads for various therapeutic agents. Chaetomugilins and the closely related chaetoviridins are fungal metabolites, and each has an oxygenated bicyclic pyranoquinone core. They are mainly produced by various Chaetomaceae species. These metabolites display unique chemical features and diversified bioactivities. The current review gives an overview of research about fungal chaetomugilins and chaetoviridins regarding their structures, separation, characterization, biosynthesis, and bioactivities. Additionally, their antiviral potential towards the SARS-CoV-2 protease was evaluated using docking studies and molecular dynamics (MD) simulations. We report on the docking and predictive binding energy estimations using reported crystal structures of the main protease (PDB ID: 6M2N, 6W81, and 7K0f) at variable resolutions—i.e., 2.20, 1.55, and 1.65 Å, respectively. Chaetovirdin D (43) exhibited highly negative docking scores of −7.944, −8.141, and −6.615 kcal/mol, when complexed with 6M2N, 6W81, and 7K0f, respectively. The reference inhibitors exhibited the following scores: −5.377, −6.995, and −8.159 kcal/mol, when complexed with 6M2N, 6W81, and 7K0f, respectively. By using molecular dynamics simulations, chaetovirdin D’s stability in complexes with the viral protease was analyzed, and it was found to be stable over the course of 100 ns. Fungi are a wealthy and substantial pool of many secondary metabolites with many different structures and diversified bioactivities [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] . These metabolites attract much attention as lead metabolites for pharmaceutical agents, and for plant protection [1, [8] [9] [10] [11] [12] [13] [14] [15] [16] . Fungal polyketides (FPKs) represent one of the largest and most structurally diverse groups of fungal metabolites. They range from simple and aromatic to highly macrocyclic and complex [10, 13, 17] . Their backbone is biosynthesized by the condensation of acyl-CoA thioesters [18] . Their structural variations originate from differences in the starting and extending units, methylation pattern, chain length, degree of reduction, and modifications by tailoring enzymes [19] . Mycotoxins and pigments are among the FPKs that have had remarkable contributions in the field of drug discovery [20] . Azaphilones (azaphilonoids or isochromenes) are fungal pigments belonging to FPKs. Structurally, they have an isochromene skeleton that contains an oxygenated bicyclic pyrano-quinone core and a quaternary carbon center [21] . Biosynthetically, the O atom in the pyran chromophore could be exchanged by an N atom in the existence of primary amines, and accordingly, the pigment color will shift to red [22] . They are produced by various basidiomycetous and ascomycetous fungi, including Chaetomium, Penicillium, Aspergillus, Talaromyces, Phomopsis, Monascus, Emericella, Epicoccum, Hypoxylon, and Pestalotiopsis, where they are accountable for the green, red, or yellow color of mycelia and/or fruiting bodies [23] . They possess myriad bioactivities: antitumor, cytotoxic, antimicrobial, anti-inflammatory, antioxidant, enzyme inhibitory, antiviral, insecticidal, and antileishmanial [24] . Chaetomugilins and closely related chaetoviridins are azaphilones featuring a C-7-methyl group and C-5 chlorineexcept for chaetomugilins T (29) and U (30)-and a C-3-branched pentenyl chain ( Figure 1 ). However, chaetomugiline P (24) differs from the others in that it has no substituent at C-7 and a methyl group at C-5. 3-Methyl-4-hydroxy-1-pentyl chains at C-3 are found in some chaetomugilins and chaetoviridins. Sometimes, they bear a five-membered lactone and/or a fused tetrahydrofuran/δ-lactone [25] . The 7-OH group can have (S) or (R) configuration, though (7S) isomers are the most common among these metabolites. On the other hand, the 7-hydroxyl can be part of a furanone ring [26] . These fungal metabolites are produced by various Chaetomium species. Chaetoviridins were firstly reported by Takahashi et al. from Chaetomium globosum var. flavoviride [27] . Chaetomugilins are known as cytotoxic metabolites, whereas chaetoviridins have antifungal and antibiotic activities [28] . Recently, these metabolites have been recognized as a unique family of fungal metabolites in view of their interesting structural features and prominent bioactivities, which could provoke enormous attention from natural products chemists and pharmacologists. The current review focuses on chaetomugilins and chaetoviridins from fungal sources, including isolation, structural characterization, biosynthesis, and bioactivities (Tables 1 and 2 ). Some of the metabolites have been reported with the same names, despite having different structures and molecular formulae-e.g., chaetomugilin S, chaetoviridin B, and chaetoviridin G. Moreover, the structures of some compounds have been revised and renamed: in such cases, both structures have been drawn in the figures, highlighting the new names and corresponding references. Additionally, the emergence of the COVID-19 pandemic motivated us to investigate the potential of these metabolites as antiviral agents towards SARS-CoV-2 using docking studies and molecular dynamics (MD) simulations. Literature searching was carried out using diverse databases-Web of Science, PubMed (MedLine), GoogleScholar, Scopus, and SciFinder-and different publishers-Springer-Link (Cham, For isolation of these metabolites, the fungal materials were extracted with E MeOH. Their presence in the extract can be detected by TLC using the solvent s EtOAc:toluene:AcOH:formic acid 2.5:7.5:1:1, CHCl3:MeOH:AcOH 95:5:0.1, or pe spirit/EtOAc 40:60. Upon development, the spots on the SiO2 TLC plates can be vi as dark spots under ultraviolet light (365 nm and 254 nm), or by using ammonia (re or ceric ammonium molybdate stain (dark spots; H2O (90 mL), H2SO4 (10 mL), amm molybdate (2.5 g), cerium (IV) sulfate (1.0 g)) [27, [29] [30] [31] . Moreover, they gave Beilstein reactions [27] . The EtOAc or MeOH extract can be separated by column tography using SiO2 CC (PE:acetone 9:1-6.5:3.5 or 40:1-2:1; CH2Cl2:MeOH; cy ane:MeOH; EtOAc:CH2Cl2; cyclohexane:EtOAc; CHCl3:MeOH gradient), Sephade (MeOH; CH2Cl2:MeOH 6:4; CH2Cl2:MeOH 1:1; CHCl3:MeOH 1:1), or MPLC (C-1 (MeOH:H2O gradient) [25, [31] [32] [33] . Finally, the fractions or isolated compounds can ther purified by preparative HPLC (MeOH:H2O 90%, 85%, 75%, 65%, or 50 CH3CN/H2O (60:40, 65:35, or 75:25, v/v) [31] [32] [33] [34] [35] , or normal phase HPLC (n-hexan 7:3, 4:1, 1:4, 2:1, 1: 3, or 3:2) [36] . Their isolation can also be achieved by loading the onto TLC SiO2 plates in the form of uniform bands. For isolation of these metabolites, the fungal materials were extracted with EtOAc or MeOH. Their presence in the extract can be detected by TLC using the solvent systems: EtOAc:toluene:AcOH:formic acid 2.5:7.5:1:1, CHCl 3 :MeOH:AcOH 95:5:0.1, or petroleum spirit/EtOAc 40:60. Upon development, the spots on the SiO 2 TLC plates can be visualized as dark spots under ultraviolet light (365 nm and 254 nm), or by using ammonia (red color) or ceric ammonium molybdate stain (dark spots; H 2 O (90 mL), H 2 SO 4 (10 mL), ammonium molybdate (2.5 g), cerium (IV) sulfate (1.0 g)) [27, [29] [30] [31] . Moreover, they gave positive Beilstein reactions [27] . The EtOAc or MeOH extract can be separated by column chromatography using SiO 2 CC (PE:acetone 9:1-6.5:3.5 or 40:1-2:1; CH 2 Cl 2 :MeOH; cyclohexane:MeOH; EtOAc:CH 2 Cl 2 ; cyclohexane:EtOAc; CHCl 3 :MeOH gradient), Sephadex LH-20 (MeOH; CH 2 Cl 2 :MeOH 6:4; CH 2 Cl 2 :MeOH 1:1; CHCl 3 :MeOH 1:1), or MPLC (C-18 ODS) (MeOH:H 2 O gradient) [25, [31] [32] [33] . Finally, the fractions or isolated compounds can be further purified by preparative HPLC (MeOH:H 2 O 90%, 85%, 75%, 65%, or 50% v/v), CH 3 CN/H 2 O (60:40, 65:35, or 75:25, v/v) [31] [32] [33] [34] [35] , or normal phase HPLC (n-hexane:EtOAc 7:3, 4:1, 1:4, 2:1, 1: 3, or 3:2) [36] . Their isolation can also be achieved by loading the sample onto TLC SiO 2 plates in the form of uniform bands. Katsuura, Nachi, Wakayama, Japan [37] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [35] Chaetomugilin B (6) 464 C 24 H 29 ClO 7 C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [38] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [39] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [40] C. globosum Z1 Broussonetia papyrifera (Barks, Moraceae) Nanjing, Jiangsu, China [44] Chaetomugilin C (7) 432 C 23 H 25 ClO 6 C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [38] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [39] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [40] C. globosum HDN151398 Sediment sea South China Sea [22] Chaetomugilin D (8) C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [39] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [40] C. globosum Ginkgo bilob (Leaves, Ginkgoaceae) Linyi, Shandong province, China [42] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [41] C. globosum OUPS-T106B- 6 Mugil cephalus (Marine fish, Mugilidae) Katsuura Bay, Japan [43] C. globosum DAOM 240359 Indoor air samples or building materials Ontario, Alberta, Saskatchewan, Nova Scotia, Canada [30] C. globosum DAOM 240359 Indoor air samples or building materials Ontario, Alberta, Saskatchewan, Nova Scotia, Canada [50] The plates were developed by using toluene/EtOAc/formic acid (7:3:1), CH 2 Cl 2 / MeOH (20:1), benzene/ethyl acetate (8:2), EtOAc/CH 2 Cl 2 (5:95), n-hexane/ethyl acetate (4:1), or EtOAc/CH 2 Cl 2 (2:8) [29, 42, 60, 63] . The isolated metabolites can be purified by recrystallization from MeOH or CHCl 3 :MeOH until they show constant melting points. The structures of isolated metabolites were determined through extensive spectroscopic analyses, including UV, IR, MS, and 1D ( 1 H, 13 C NMR, and DEPT) and 2D NMR (COSY, NOESY, ROESY, HMQC, HSQC, or HMBC). The absolute configurations of these metabolites have been established with the aid of optical rotation sign, X-ray crystallography, CD (circular dichroism), the modified Mosher's method, and chemical transformation studies, including derivatization and degradation [22, 27, 35, 41, 48, 60, 75] . It has been reported that the absolute configuration at C(7) controls signs of the specific rotation [35] . Compounds with (S) C-11 and C-7 had negative optical rotation values; however, when C-7 was (R), the sign switched to positive with the same magnitude [50] . The absolute (S) configuration at C-7 was determined by the negative Cotton effect in the CD spectrum [58] . Mass spectra of these compounds displayed an isotopic peak [M+H] + /[M+H+2] + in a ratio 3:1, characterizing the presence of a single chlorine atom. Moreover, their IR spectrum exhibited characteristic bands for a hydroxyl group (3405-3450 cm −1 ), lactone (1718-1780 cm −1 ), and α,β-unsaturated ketone (1616-1684 cm −1 ). Characteristic UV bands of a highly extended conjugation system were observed at 283-429 nm. Table 2 ). It is noteworthy that compounds 7 and 13 exhibited remarkable cytotoxicity towards P388 and HL-60 cell lines (IC 50 3.6 and 3.3 µM and 2.7 and 1.3 µM, respectively), nearly equal to that of 5-fluorouracil (IC 50 1.7 and 2.7 µM, respectively). While other compounds had moderate to weak cytotoxicity (IC 50 ranging from 6.8 to 24.1 µM) [38, 39] . Further, 2, 7, and 13 displayed selective cytotoxicity towards a panel of 39 disease-related human cell lines, including breast, CNS, colon, lung, melanoma, ovary, kidney, stomach, and prostate cancer cells with range and delta values of 2 (1.24 and 1.13, respectively), 7 (1.19 and 0.71, respectively), and 13 (1.21 and 1.97, respectively) [38, 40] . It was suggested that the existence of C-12-hydroxyl and C-3-methoxyl groups had little effect on the activity [40] . Evaluation of the differential cytotoxicity patterns using COMPARE revealed that the modes of action for 2, 7, and 13 might be different from those of other anticancer drugs [40] . On the other hand, chaetomugilins A (2) biosynthesized by C. globosum Z1 isolated from Broussonetia papyrifera bark had no in vitro effectiveness towards SMMC-7721, MG-63 and A-549 cell lines (IC 50 > 50 µg/mL) in the MTT assay in comparison to doxorubicin [44] . C. globosum OUPS-T106B-6 isolated from M. cephalus yielded two new metabolites that demonstrated moderate cytotoxicity towards HL-60 and P388 cell lines (IC 50 ranged from 10.3 to 24.1 µM, respectively), compared to 5-FU (IC 50 2.7 and 1.7 µM) in the MTT assay [41] : chaetomugilins G (14) and H (15) . In another study by Yamada et al. on the same fungus, two new compounds named seco-chaetomugilins A (3) and D (8) were separated. Compound 8 exhibited weak activity (IC50 38.6-53.6 µM) towards HL-60, P388, KB, and L1210, compared with 5-FU (IC 50 1.7-6.0 µM); however, 8 was inactive. It was suggested that the C-12-hydroxyl group decreased the activity [48] . Chaetomugilins I (16), J (18), 11-epichaetomugilin A (4), 4 -epichaetomugilin A (5), and 106B-6 XXVIII (1) were separated from C. globosum 106B-6 and assessed for cytotoxic activity towards P388, HL-60, L1210, and KB cell lines. Interestingly, compounds 16 and 18 had significant cytotoxicity (IC 50 [58] . Furthermore, the new metabolites, 11-and 4 -epichaetomugilin A (4 and 5) purified from C. globosum isolated from M. cephalus, displayed moderate to weak cytotoxicity toward KB, P388, HL-60, and L1210 cell lines [35] . The new metabolites, chaetomugilins P-R (24-27) and 11-epi-chaetomugilin I (17), along with the formerly separated chaetomugilin I (16), were purified by the marine fish-associated C. globosum (Figures 4 and 5 ) [59] . Compounds 24, 17, and 16 possessed stronger cytotoxicity towards HL-60, P388, KB, and L1210 (IC 50 1.1, 1.1, 2.3, and 1.9 µM, respectively, for 16; 1.0, 0.7, 1.2, and 1.6 µM, respectively, for 17; and 1.2, 0.7, 1.8, and 1.5 µM, respectively, for 24) cell lines than 5-FU (IC 50 2.7, 1.7, 7.7, and 1.1 µM, respectively). The results indicated that the C-2 -C-4 enone moiety is essential for activity. On the other hand, compounds 25-27 were weakly to moderately active towards all tested cancer cell lines (IC 50 32.0-80.2 µM) [59] . The new metabolites, chaetomugilin S (28), T (29) , and U (30), separated from C. globosum derived from M. cephalus, revealed moderate to high growth inhibition towards HL-60, P388, KB, and L1210 cell lines (IC 50 ranging from 34.4 to 94.9 µM), relative to 5-FU (IC 50 1.9-20.6 µM) [43] . In 2019, Sun et al. evaluated the cytotoxicity of the new glutamine-containing derivatives, N-glutarylchaetoviridins A-C (35, 39, and 42), in addition to chaetomugilins A (2) and C (7) from the extract of deep-sea sediment-associated C. globosum HDN151398 toward BEL-7402, HeLa, HCT-116, L-02, HO8910, MGC-803, SH-SY5Y, U87, NCI-H1975, and MDA-MB-231 cancer cells using the SRB method and towards HL-60 and K562 using MTT method ( Figure 6 ). Chaetomugilins A (2) and C (7) and N-glutarylchaetoviridin C (42) exhibited powerful cytotoxicity towards all tested cell lines (IC 50 5.7-27.1 µM for 2, 6.6-26.6 µM for 7, and 6.6-26.5 µM for 42) compared to adriamycin (IC 50 0.1-0.6 µM) [22] . Among them, N-glutarylchaetoviridin C (42) had a remarkable cytotoxicity toward MGC-803 and HO-8910 (IC 50 6.6 and 9.7 µM, respectively) [22] . Wani et al. reported that seco-chaetomugilin D (9) isolated from C. cupreum had cytotoxicity towards MCF-7 (inhibition ranging from 25.25% to 75.25% after 24 h and from 41.5% to 99% after 38 h at 3.12-50 µg/mL). Further, it increased mitochondrial membrane depolarization (16.45% and 32.25% at 5 and 15 µg/mL, respectively) and induced ROS production (19.6% and 26.2% at 5 and 15 µg/mL, respectively) in comparison to untreated Chaetomugilin A (2), 11-epi-chaetomugilin A (4), and chaetomugilin D (8) displayed no noticeable cytotoxic activity (IC 50 > 40 µM) toward HepG-2, A549, and HeLa in the MTT assay compared to etoposide (IC 50 16.11, 16.46, and 15 .00 µM, respectively) [46] . Wani et al. reported that seco-chaetomugilin D (9) isolated from C. cupreum had cytotoxicity towards MCF-7 (inhibition ranging from 25.25% to 75.25% after 24 h and from 41.5% to 99% after 38 h at 3.12-50 µg/mL). Further, it increased mitochondrial membrane depolarization (16.45% and 32.25% at 5 and 15 µg/mL, respectively) and induced ROS production (19.6% and 26.2% at 5 and 15 µg/mL, respectively) in comparison to untreated cells. Therefore, it caused cell death via induction of mitochondrial ROS production and membrane depolarization [29] . C. globosum isolated from Ginkgo biloba leaves yielded chaetomugilin D (8) and chaetomugilin A (2) , which demonstrated significant toxicity toward brine shrimp (Artemia salina) larvae after 24 h (mortality rates 75.2 and 78.3%, respectively, at 10 µg/mL) [42] . Hu et al. proved that chaetomugilin J (18) combined with low-dose cisplatin decreased cell viability and boosted cisplatin-produced apoptosis in ovarian A2780 cells independently of the endoplasmic reticulum apoptotic pathway. It significantly induced mitochondrial dysfunction and apoptosis via increasing the intracellular and mitochondrial ROS levels and decreasing mitochondrial membrane potential. It also prohibited parkin/PINK1 induced mitophagy, resulting in weakening the mitophagy protective effect that led to apoptosis and increased sensitivity to cisplatin [34] . Chaetomugilin D (8), chaetoviridin A (31), and chaetoviridin E (44) purified from a deep sea-derived Chaetomium sp. NA-S01-R1 displayed moderate to weak cytotoxicity toward HeLa, A549, and HepG2, in comparison to doxorubicin (IC 50 0.1-1.1 µM) using the CCK-8 assay [25] (Figure 7) . Additionally, chaetoviridin A (31) and chaetoviridin E (44) were inactive towards AGS and MGC803 [71] . In addition, chaetoviridins E and F (44 and 47) had cytotoxicity toward NCI-H187, KB, and BC1 (IC 50 ranging from 3.5 to 13.4 µg/mL), in comparison to ellipticine (IC 50 ranging from 0.26 to 0.36 µg/mL) [60] . Chaetoviridin A (31) (2 µM, ID 50 0.6 µM) inhibited the inflammatory activity of TPA (12-O-tetradecanoylphorbol-13-acetate, 1 µg) in mice (ID 50 0.6 µM). Furthermore, it markedly suppressed the promoting effect of TPA on skin tumor formation in mice initiated with 7,12-dimethylbenz[a]an-thracene (50 µg). It was proposed that it inhibited TPAtumor promotion in a two-stage carcinogenesis model in mice due to its anti-inflammatory potential [76] . C. globosum TY1 associated with Ginkgo biloba barks yielded chaetoviridins B (38) and E (44) and 5 -epi-chaetoviridin A (33). Compounds 44 and 33 had moderate cytotoxicity towards HepG-2 (IC 50 40.6 and 35.3 µM, respectively) compared to camptothecin (IC 50 32.3 µM) in the SRB assay, and 38 was inactive [72] . C. globosum isolated from Artemisia desterorum roots yielded chaetoviridin E (44) and chaetoviridin A (31). They had no cytotoxic activity toward A549, HCT116, and HepG2 cancer cells [68] . The new metabolite, epi-chetomugilin D (10), along with chaetovirdins A (31), B (37) , and E (44) and chetomugilins D (8) and J (18) were purified from C. globosum associated with Adiantum capillus-veneris. Chaetovirdin E (5 µg/mL) exhibited cytotoxicity towards CaCO 2 and HepG2 cancers cells with 30% and 59% inhibition, respectively [49] . In addition, chaetoviridins E and F (44 and 47) had cytotoxicity toward NCI-H187, KB, and BC1 (IC50 ranging from 3.5 to 13.4 µg/mL), in comparison to ellipticine (IC50 ranging from 0.26 to 0.36 µg/mL) [60] . Chaetoviridin A (31) (2 µM, ID50 0.6 µM) inhibited the inflammatory activity of TPA (12-O-tetradecanoylphorbol-13-acetate, 1 µg) in mice (ID50 0.6 µM). Furthermore, it markedly suppressed the promoting effect of TPA on skin tumor formation in mice initiated with 7,12-dimethylbenz[a]an-thracene (50 µg). It was proposed that it inhibited TPA- The liquid culture of C. globosum DAOM-240359 isolated from an indoor air sample collected from Ottawa, Ontario, Canada produced new nitrogen-containing chaetoviridins, 4 -epi-N-2-hydroxyethyl-azachaetoviridin A (36) [61] . They also exhibited strong in vivo antifungal activity against M. grisea and Puccnicia recondite (wheat leaf rust) [36] . Chaetoviridin A (Conc. 62.5 µg/mL) inhibited rice blast development by >80%, but it had moderate control (50%) of tomato late blight at 125 µg/mL. Therefore, they can control wheat leaf rust rice blast and tomato late blight [36, 61] . Chaetoviridins B (38) and E (44) displayed antibacterial activity towards E. faecalis and S. aureus [74] . Chaetoviridins A (31) and B (34) also had antimicrobial activity against B. subtilis, Rhizoctonia solani, and E. coli (IZDs 15 and 14 mm, respectively, towards all strains) [63] . Yan et al. reported that chaetoviridin A (31) exhibited significant antifungal potential (EC 50 1.97 µg/mL) towards Sclerotinia sclerotiorum, which causes rape Sclerotinia rot (RSR). Further, 31 displayed in vivo protective efficacy (64.3%, dose 200 µg/mL) towards RSR, comparable to that of carbendazim (69.2%). Additionally, it had antifungal activity towards Botrytis cinerea, Phytophthora capsici, Fusarium graminearum, and F. moniliforme (inhibition rates 69.1, 60.7, 77.0, and 59.2%, respectively) [73] . C. globosum CEF-082, isolated from cotton plants, produced chaetoviridin A (31), which possesses significant antifungal activity towards Verticillium dahlia, which causes cotton Verticillium wilt (CVW). It induced mycelial deformation and cell necrosis, increased NO and ROS production, prohibited the germination of microsclerotia of V. dahliae, and boosted the cotton defensive response [69] . Chaetoviridin A (31), 5 -epichaetoviridin A (33) , and chaetoviridin E (44) were separated from the soil-associated C. globosum 22-10. They showed significant inhibitory effectiveness (inhibition 32.31%, 15.38%, and 13.85%, respectively) at 100 µg/mL towards Bipolaris sorokiniana, a soil-borne pathogen that commonly causes wheat root rot. It is noteworthy that chaetoviridin A had the same inhibitory efficiency as the carbendazim (3 mg/mL), suggesting its potential to be a biocontrol agent for B. sorokiniana [67] . Further, chaetoviridin A identified from an EtOAc extract of C. globosum F211_UMNG isolated from P. heptaphyllum was active towards F. oxysporum [66] . Chaetomugilin D (8) and chaetoviridin A (31) (200 µM) significantly reduced the growth of B. subtilis and Psuedomonas putida [55] . Chaetoviridins B (38) and E (44) displayed antibacterial activity towards E. faecalis and S. aureus [74] . Chaetoviridins A (31) The EtOAc extract of C. globosum associated with Amaranthus viridis yielded chaetomugilin D (8) and chaetomugilin J (18) , which exhibited phytotoxic potential against lettuce (Lactuca sativa) seed germination with IC 50 24.2 and 22.6 ppm, respectively, for root growth inhibition, and IC 50 27.8 and 21.9 ppm, respectively, for shoot growth inhibition. The results revealed the potential of these metabolites as herbicides or weedicides that can replace hazardous synthetic compounds [51] . Chaetomugilin A (2), D (8), S (28), I (16), J (18), Q (25) , and O (23) isolated from C. globosum TY1 exhibited allelopathic activity towards Brassica campestris, Cucumis sativus, Eruca sativa, Daucus carota, Lactuca sativa, Scrophularia ningpoensis, Brassica rapa, and Spinacia oleracea. Among them, 23 exhibited higher germination and root and shoot elongation inhibitory potential with lower IC 50 values and higher response indexes than glyphosate (positive control). Moreover, 2, 8, and 28 exhibited similar or better inhibitory effects than glyphosate. At the same time, 2, 8, and 23 were more powerful growth inhibitors than 16, 18, and 25, which could be attributed to the existence of a tetrahydrofuran moiety. On the other hand, 23 had a higher growth-suppression effect than those of 2, 8, and 28, suggesting that the lactone rings may reduce the inhibitory effects [45] . Phonkerd et al. isolated new derivatives, chaetoviridins E and F (44 and 47) and 5 -epi-chaetoviridin A (33), together with chaetoviridin A (31) from C cochliodes VTh01 and C. cochliodes CTh05. Compound 44 showed antimalarial activity against P. falciparum (IC 50 2.9 µg/mL)-and was compared to artemisinin (IC 50 0.001 µg/mL)-using the microculture radioisotope technique [60] . Additionally, 44 and 47 displayed weak antimycobacterial potential towards M. tuberculosis (MIC 50 and 100 µg/mL, respectively), in comparison to isoniazid and kanamycin sulfate in the microplate Alamar Blue assay (MABA) [60] . Chaetomugilin D (8) from C. globosum isolated from damp building materials notably increased TNF-α production in RAW 264.7 murine macrophages. Therefore, 8 contributes to the non-allergy-linked respiratory disorders such as non-allergic asthma and rhinitis for people working and living in damp buildings [52] . Two new derivatives, chaetoviridins J (53) and K (54), along with 11-epi-chaetomugilin I (17) and chaetomugilins E (11), F (13), I (16), J (18) , and N (22) biosynthesized by C. globosum were isolated from Wikstroemia uva-ursi leaves. Their structures were verified by NMR, Mosher's method, X-ray diffraction, and CD. Chaetoviridin K (54) was separated as a mixture of diastereoisomers that could not be purified by chiral columns. Their potential to inhibit TNF-α-induced NF-κB and NO (nitric oxide) production in the LPS-stimulated RAW 264.7 cells was assessed. Compounds 16 and 17 remarkably suppressed TNF-α-induced NF-κB activity (IC 50 0.9 µM, at 50 µM), in comparison to BAY-11 (IC 50 2.0 µM) and TPCK (IC 50 3.8 µM), whereas 11, 13, and 18 possessed moderate inhibitory activity (IC 50 ranging from 5.1 to 11.6 µM, conc. 50 µM). In addition, 11, 13, 16-18, and 53 strongly prohibited NO production (74.2-99.9%). It is noteworthy that 16 and 17 powerfully suppressed NO release (IC 50 0.3 and 0.8 µM, respectively) more than N-monomethyl-L-arginine (IC 50 25.1 µM). On the other hand, 11, 13, and 18 also considerably inhibited NO production (IC 50 values ranging from 1.9 to 5.8 µM); however, 54 exhibited a weak effect [56] . Chaetomugilins A (2), I (16), J (18), and Q (25) and chaetoviridins B (38) and J (53) purified from the EtOAc extract of C. globosum TY-1 isolated from Ginkgo biloba bark were tested for α-amylase and α-glucosidase inhibitory activity. Chaetoviridin B (38) had promising α-glucosidase inhibitory potential (IC 50 6.328 µM), compared to acarbose (IC 50 54.7 µM). The other compounds displayed no α-amylase or α-glucosidase inhibitory activity (IC 50 > 50 µM) compared to acarbose (IC 50 13 .7 and 54.7 µM, respectively) [47] . Chaetoviridin B (38) and chaetomugilin A (2) have the same skeleton except for the 2 having one more hydroxyl group in the side chain, revealing that the group is detrimental for α-glucosidase inhibition [47] . Chaetoviridins A (31) and B (37) purified from an EtOAc extract of C. globosum exhibited noticeable antioxidant potential on TLC using DPPH [63] . Additionally, chetomugilin D (8) and its analog epi-chetomugilin D (10) possessed antiviral activity towards HSV-2 (inhibition 33.3% and 40.7%, respectively, at 25µg/mL [49] . Caspase-3 (cysteine aspartyl-specific protease-3) is one of the executioners in caspaselinked apoptosis that is activated in nearly every apoptosis model [77] . It is a prominent therapeutic target for excessive apoptosis-associated disorders, such as ischemic damage and neurodegenerative disorders (e.g., Huntington's and Alzheimer's diseases) and autoimmune disorders [78] . New azaphilones, chaetomugilin S (27), 7,5 -bis-epi-chaetoviridin A (34), and 7-epi-chaetoviridin E (45), purified from the lichen-associated C. elatum 89-1-3-1, were isolated from Ramalina calicaris. Their absolute configurations were assigned by CD experiments and X-ray crystallography. They exhibited caspase-3 inhibitory potential (IC 50 20.6, 10.9, and 7.9 µM, respectively) in the cysteine aspartyl-specific protease-3 enzymatic assay compared with Ac-DEVD-CHO (IC 50 13.7 nM) [31] . On the other hand, 31 possessed weak MAO inhibitory potential (IC50 1.2 × 10 −2 g/mL) [27] . CETP allows the transfer and exchange of neutral lipids such as CE (cholesteryl ester) and TG (triacylglycerol) between plasma and lipoproteins. It is proven to play important role in atherosclerosis [79] . Tomoda et al. reported that chaetoviridin B (37) showed CETP (cholesteryl ester transfer protein) inhibitory activity with an IC 50 < 6.3 µM, whereas chaetoviridin A (31) had moderate inhibitory activity (IC 50 31.6 µM) [80] . It was indicated that the existence of an electrophilic enone(s) and/or ketone(s) at both C-8 and C-6 of isochromane core is substantial for eliciting activity [80] . The COVID-19 pandemic has affected global health since 2019. COVID-19 can lead to acute respiratory distress syndrome [81] . It is produced by a novel type of coronavirus (CoV) called SARS-CoV-2 that was first found in Wuhan City, China, and then spread worldwide [82, 83] . It is considered a highly pathogenic CoV in the human population. The SARS-CoV-2 genome encodes two polyproteins which are processed by a 3C-like protease (3CLpro) and a papain-like protease [84] . 3CLpro (3C-like protease) and PLpro (papain-like protease) are needed for processing the polyproteins into mature nonstructural proteins, such as helicase and RdRp (RNA-dependent RNA polymerase), which are substantial for viral replication and transcription [85] . 3CLpro has high substrate specificity and is also referred to as Mpro (main protease) [86] . 3CLpro's substrate specificity makes this enzyme an ideal target for developing broad-spectrum antiviral agents [87] . Its inhibitors are expected to have selective toxicity towards the virus [88] . Fungi are a treasure that can provide a remarkable pool of secondary metabolites with antiviral activities [89] . The characterization and discovery of antiviral fungal metabolites is an emerging and promising research field. Recently, many reports have been published on the structure-based virtual screening approach for the repurposing of natural metabolites, hoping to accelerate and assist in the discovery of agents for COVID-19 treatment [82] . We carried out a computational study on the reported fungal chaetomugilins and chaetoviridins was carried out to identify their 3CLpro inhibitory potential, using docking calculations and MD simulations (Tables 3-7) . Three crystal structures containing non-covalent inhibitors for the protease (PDB entry: 6W81, 6M2N, and 7K0F) were selected. All the listed metabolites were docked with extra precision for maximum accuracy. The docking method was validated by redocking the inhibitors that co-crystallized with 6W81, 6M2N, and 7K0F; and RMSD values were within an acceptable range and less than 1.50 Å. All the redocked inhibitors revealed the same binding interaction with the active site in the original pose. Further, in silico ADMET (drug absorption, distribution, metabolism, excretion, and toxicity) predictions of the properties of the investigated compounds were carried out. Finally, a molecular dynamics simulation was conducted to evaluate the nature of the ligand-target interaction under simulated physiological conditions for the most compatible drug-like molecule that could be used in pursuit of a truly adequate medication for COVID-19. By using LigPrep, the conversion of 2D structures to 3D, tautomerization, and ionization yielded 254 minimized 3D structures. The minimized 3D structures were used for docking with the crystal structure of the 3CL hydrolase (Mpro). Preparation of the viral protease (6M2N, 6W81, and 7K0f) by the Protein Preparation Wizard tool optimized the H-bonding network and minimized the geometry. Assurance of assigning the proper formal charges and force field treatments was achieved by adding missing hydrogens and correct ionization states (Figure 9 ). After defining the grid box in the prepared viral protease via the Receptor Grid Generation tool of Glide in Maestro, the prepared 3D molecular structures were docked into the co-crystallized inhibitor binding site of the viral protease. Table 3 shows the results of the top-score docked ligands chosen based on the most negative docking scores. These scores represent the best bound ligand conformations and relative binding affinities. Chaetovirdin D (43) exhibited the highly negative docking scores of −7.944, −8.141, and −6.615 kcal/mol, complexed with 6M2N, 6W81, and 7K0f, respectively. The reference inhibitor exhibited the following scores: −5.377, −6.995, and −8.159 kcal/mol, complexed with 6M2N, 6W81, and 7K0f, respectively. The docking analysis was updated by re-docking the selected chaetomugilins and chaetoviridins with the three crystal structures of the 3CL pro-hydrolase (PDB ID: 6M2N, 6W81, and 7K0f) at variable resolutions, 2.20, 1.55, and 1.65A, respectively. Analysis of the docking scores of these compounds with the inhibitor binding sites of 2019-nCoV main protease (PDB ID: 6M2N, 6W81, and 7K0f) revealed different docking scores, and hence binding affinities (Tables 4-6 ). These differences can be attributed to the differences in the After defining the grid box in the prepared viral protease via the Receptor Grid Generation tool of Glide in Maestro, the prepared 3D molecular structures were docked into the co-crystallized inhibitor binding site of the viral protease. Table 3 shows the results of the top-score docked ligands chosen based on the most negative docking scores. These scores represent the best bound ligand conformations and relative binding affinities. Chaetovirdin D (43) exhibited the highly negative docking scores of −7.944, −8.141, and −6.615 kcal/mol, complexed with 6M2N, 6W81, and 7K0f, respectively. The reference inhibitor exhibited the following scores: −5.377, −6.995, and −8.159 kcal/mol, complexed with 6M2N, 6W81, and 7K0f, respectively. The docking analysis was updated by re-docking the selected chaetomugilins and chaetoviridins with the three crystal structures of the 3CL pro-hydrolase (PDB ID: 6M2N, 6W81, and 7K0f) at variable resolutions, 2.20, 1.55, and 1.65A, respectively. Analysis of the docking scores of these compounds with the inhibitor binding sites of 2019-nCoV main protease (PDB ID: 6M2N, 6W81, and 7K0f) revealed different docking scores, and hence binding affinities (Tables 4-6 ). These differences can be attributed to the differences in the grid formation due to the presence of different inhibitors in the binding sites of the three crystal structures. Analysis of the docking of 43 revealed that it interacted through hydrogen bonds ( Figure 10 ) with the binding site residues of SARS-CoV-2 main protease (6W81). The binding site residues Asn141, Gly142, and Thr189 of the viral protease exhibited hydrogen bonding with the various hydroxyl groups of 43. Additionally, 43 interacted through hydrogen bonds ( Figure 11 ) with the binding site residues of SARS-CoV-2 main protease (6M2N). The binding site residues Asn142 and Thr190 of the viral protease showed hydrogen bonding with the various hydroxyl groups of 43. Its interactions through hydrogen bonds with the binding site residues of SARS-CoV-2 main protease (7K0F) are shown in Figure 12 Table 7 . ADMET analysis describes and determines the biological function, drug-likeness, physicochemical characters, and expected toxicity of compounds. This is meant to evaluate the usefulness of the molecules. The examined descriptors, such as drug-likeness, molecular weight, solvent accessible surface area, dipole moment, hydrogen bond acceptors, donor traits, aqueous solubility, octanol/water coefficient, binding to hu-man serum albumin, number of likely metabolic reactions, brain-blood partition coefficient, human oral absorption, IC 50 value for blockage of HERG K + channels, central nervous system activity, and number of reactive functional groups, were predicted for the reported metabolites. The values obtained for all the compounds are in the recommended ranges. The docking studies took a static view for the binding of each molecule in the active site of the protein. A molecular dynamics (MD) simulation computes the atoms movements over time. By using Desmond software, the stability and frequency of the 43 complex with the proteases-PDB ID 6M2N, 6W81, and 7K0f-were studied. Three MD simulations were run for complexes with 43 for 100 ns of simulated time; the complexes' structures were optimized at pH 7.0 ± 2.0. Complex stability was checked by analysis of the interaction map and the RMSD (root mean square deviation) plot of the ligand and protein. The RMSD plots in Figure 13a for the chaetovirdin D-SARS-CoV-2 main protease (PDB ID 6W81) complex, Figure 13b for the chaetovirdin D-SARS-CoV-2 main protease (PDB ID 7K0F) complex, and Figure 13c for the chaetovirdin D-SARS-CoV-2 main protease (PDB ID 6M2N) indicate that the complexes tended to stabilize during the simulations (100 ns) with respect to a reference frame at time 0 ns. There were slight fluctuations during the simulations, but within the permissible range of 1-3 Å; hence, they can be considered non-significant. Since the RMSD plots of 43 and protein backbone lie over each other, the formation of a stable complex can be inferred. The simulated time of 100 ns shows the formation of a stable complex without any significant conformational changes in protein structure. Figure 14a shows the residue interactions of chaetovirdin D with SARS-CoV-2 main protease (PDB ID 6W81) (the docked poses were retained during the simulation of 100 ns)i.e., molecular interactions with His41, Asp141, and Glu165 residues. Moreover, a hydrophobic interaction was also established between the Leu165 and aliphatic hydrocarbons of 43. In Figure 14b , the viral protease (PDB ID 7K0F)-chaetovirdin D contacts over the course of 100 ns are categorized into water bridges, hydrogen bonds, and hydrophobic interactions. The initial docked pose of 43 shows that the important hydrogen bonds (Thr24, His41, Asn142, Gly143, Ser144, Cys145) did not change during the MD simulation. Hydrogen bonding with residue Cys145 was retained for more than 70% of the simulation time. Figure 14c reveals that the viral protease (PDB ID 6M2N)-chaetovirdin D contacts over the course of 100 ns were categorized into water bridges, hydrogen bonds, and hydrophobic interactions. The bar chart shows that hydrogen bonds, hydrophobic contacts, and water bridges prevailed during the course of the simulation. The important hydrogen bonds observed in the initial docked pose of 43 (Thr26, His41, Glu166, Thr190, Gln192) did not change during the MD simulation. Hydrogen bonding with residue Thr26 was retained for more than 5% of the simulation time. The three PDB structures (PDB ID: 6M2N, 6W81, and 7K0f) were downloaded from the Protein Data Bank (Protein Data Bank; available online, prepared, and optimized by using "Protein preparation wizard" tool of Schrödinger suite (Schrödinger Release 2021-4: LigPrep, Schrödinger, LLC, New York, NY, USA, 2021) [90] . For this purpose, the bond orders for untemplated residues and known HET groups were assigned and hydrogens were added. Bonds to metals were broken, zero-order bonds between metals and nearby atoms were added, and formal charges to metals and neighboring atoms were corrected. Disulfide bonds were created. Water molecules beyond 5 Å from HET groups were deleted. For ligands, cofactors and metals het states were generated at pH 7.0 ± 2.0 using LigPrep (Schrödinger Release 2021-4: LigPrep, Schrödinger, LLC, New York, NY, USA, 2021). Finally, H-bonds were optimized by using PROPKA [91] at pH 7.0, water molecules beyond 3 Å from HET groups were removed, and restrained minimization was done using the OPLS4 force field. The ADME properties and drug-likeness of selected compounds were determined in terms of distribution, absorption, metabolism, excretion, etc., via the QikProp module of Maestro Schrodinger (Schrödinger Release 2021-4: QikProp, Schrödinger, LLC, New York, NY, USA, 2021). Glide (Schrödinger Release 2021-4: Glide, Schrödinger, LLC, New York, NY, USA, 2021) was utilized for both grid generation and ligands docking. For docking of fiftythree fungal chaetomugilins and chaetoviridins, three grids were generated using the PDB: 6M2N, 6W81, and 7K0f. For the first grid of PDB 6M2N, the binding region was defined by selecting 3WL. For the second grid PDB 6W81, the binding region was defined by selecting X77. For the third grid PDB 7K0f, the binding region was defined by selecting VR4. The non-polar atoms were set for the VdW radii scaling factor to 1.0, and the partial charge cut-off was 0.25. The ligands docking was performed by using the "ligand docking" tool of Schrödinger suite [92] . The selected protocol was standard precision (SP), the ligand sampling method was flexible, and all the other settings were maintained as default. 6M2N , 6W81, and 7K0f were retrieved from docking results and first tuned through the "System Builder" tool. The solvent model TIP3P and then orthorhombic shape box shape were selected. The side distances box was set to 10 Å, and the system was neutralized by adding Na + ions. The MD calculations were run for 100 ns per trajectory, with the number of atoms, pressure, and the temperature kept maintained constant (NPT ensemble). Pressure was set to 1.01325 bar and temperature 300.0 K, and the force field was set as OPLS4. The biosynthetic studies revealed that these metabolites are generated via a polyketide pathway [93] [94] [95] [96] . Their main polyketide chain was produced from malonate and acetate units in a conventional way [93, 95] . Briefly, a reduced triketide chain (I) was the starting unit to give the aromatic intermediate II, which was processed by the halogenase, followed by hydroxylation-stimulated cyclization catalyzed by monooxygenase to yield cazisochromene (III, chloropyranoquinone) [96] . Then, the addition of an oxidized triketide unit (IV) to III by acyltransferase formed the pyranoquinone V. The latter underwent a Knoevenagel reaction, resulting in chaetoviridin A (31) . Chaetoviridin A (31) was converted into other chaetomugilins and chaetoviridins through reductions, oxidations, rearrangements, or reactions with amines [93] (Scheme 1). converted into other chaetomugilins and chaetoviridins through reductions, oxidations, rearrangements, or reactions with amines [93] (Scheme 1). Scheme 1. Proposed biosynthesis of chaetomugilins and chaetoviridins [93] [94] [95] [96] . Fungi are important sources of natural polyketide pharmaceuticals with structural complexities that make them interesting and beneficial biometabolites. Chaetomugilins and chaetoviridins are azaphilone derivatives that are mainly sourced from various Chaetomaceae species. In this work, fifty-six metabolites were isolated from four species, C. globosum, C. cochliodes, C. siamense, C. elatum, and C. subafine, in addition to unidentified Chaetomium species. Most of them were from C. globosum, as shown in Figure 15 . The largest quantities of these metabolites were found in 2011 (18 compounds), 2009 (16 compounds), and 2017 (14 compounds) (Figure 16 ). Fungi are important sources of natural polyketide pharmaceuticals with structural complexities that make them interesting and beneficial biometabolites. Chaetomugilins and chaetoviridins are azaphilone derivatives that are mainly sourced from various Chaetomaceae species. In this work, fifty-six metabolites were isolated from four species, C. globosum, C. cochliodes, C. siamense, C. elatum, and C. subafine, in addition to unidentified Chaetomium species. Most of them were from C. globosum, as shown in Figure 15 . The largest quantities of these metabolites were found in 2011 (18 compounds These metabolites have been evaluated for diverse bioactivities, such as cytotoxic, antimicrobial, phytotoxic, antimalarial, anti-mycobacterial, anti-inflammatory, antidiabetic, antioxidant, and antiviral ones; and caspase-3, cholesteryl ester transfer protein, and monoamine oxidase (MAO) inhibitory activities ( Figure 17 ). It was found that some chaetomugilins possessed remarkable cytotoxicity towards certain cancer cell lines, equal to or stronger than the effects of control anticancer drugs, such as chaetomugilins C (7), F (13), I (16), and J (18). These metabolites have been evaluated for diverse bioactivities, such as cytotoxic, antimicrobial, phytotoxic, antimalarial, anti-mycobacterial, anti-inflammatory, antidiabetic, antioxidant, and antiviral ones; and caspase-3, cholesteryl ester transfer protein, and monoamine oxidase (MAO) inhibitory activities ( Figure 17 ). It was found that some chaetomugilins possessed remarkable cytotoxicity towards certain cancer cell lines, equal to or stronger than the effects of control anticancer drugs, such as chaetomugilins C (7), F (13), I (16), and J (18). These metabolites have been evaluated for diverse bioactivities, such as cytotoxic, antimicrobial, phytotoxic, antimalarial, anti-mycobacterial, anti-inflammatory, antidiabetic, antioxidant, and antiviral ones; and caspase-3, cholesteryl ester transfer protein, and monoamine oxidase (MAO) inhibitory activities ( Figure 17 ). It was found that some chaetomugilins possessed remarkable cytotoxicity towards certain cancer cell lines, equal to or stronger than the effects of control anticancer drugs, such as chaetomugilins C (7), F (13), I (16), and J (18) . Some studies revealed that the use of some additional compounds alongside anticancer drugs increased the sensitivity of the cancer cell lines towards these drugs. These metabolites could be further developed into anticancer agents; however, extensive in vivo studies and explorations of their mechanisms of action are needed. Chaetoviridins, particularly chaetoviridin A (31), have substantial activity towards different plant pathogens; thus, they might be used as biocontrol agents. Additionally, some reports revealed that chaetomugilins (e.g., A (2), D (8) , O (23), and S (28)) have serious phytotoxicity, more than the positive controls; therefore, they could be utilized for developing natural eco-friendly herbicides or weedicides that can replace hazardous synthetic compounds. However, extensive studies and field trials should be conducted. In molecular docking studies, chaetovirdin D exhibited the highly negative docking scores of −7.944, −8.141, and −6.615 kcal/mol, in complexes with 6M2N, 6W81, and 7K0f respectively. The reference inhibitor exhibited the following scores: −5.377, −6.995, and −8.159 kcal/mol, in complexes with 6M2N, 6W81, and 7K0f, respectively. By using molecular dynamics simulations, chaetovirdin D's stability in complexes with the viral protease was analyzed, and it was found to be stable over the course of 100 ns simulation time. Undoubtedly, chaetomugilins and chaetoviridins are fungi-derived metabolites that have multiple biological activities. They meet all the requirements for becoming drug leads in their respective therapeutic categories. They have varied chemical compositions that might provide the basis for the synthesis and design of novel and effective pharmaceutical agents. The different substituents at various positions in their skeleton play critical roles in the determination of some of their bioactivities. Finally, studies of the possible mechanisms, biosynthetic pathways, structure-activity relationships, and/or derivatization of these metabolites should be the focus of future research. Some studies revealed that the use of some additional compounds alongside anticancer drugs increased the sensitivity of the cancer cell lines towards these drugs. These metabolites could be further developed into anticancer agents; however, extensive in vivo studies and explorations of their mechanisms of action are needed. Chaetoviridins, particularly chaetoviridin A (31), have substantial activity towards different plant pathogens; thus, they might be used as biocontrol agents. Additionally, some reports revealed that chaetomugilins (e.g., A (2), D (8) , O (23), and S (28)) have serious phytotoxicity, more than the positive controls; therefore, they could be utilized for developing natural eco-friendly herbicides or weedicides that can replace hazardous synthetic compounds. However, extensive studies and field trials should be conducted. In molecular docking studies, chaetovirdin D exhibited the highly negative docking scores of −7.944, −8.141, and −6.615 kcal/mol, in complexes with 6M2N, 6W81, and 7K0f respectively. The reference inhibitor exhibited the following scores: −5.377, −6.995, and −8.159 kcal/mol, in complexes with 6M2N, 6W81, and 7K0f, respectively. By using molecular dynamics simulations, chaetovirdin D's stability in complexes with the viral protease was analyzed, and it was found to be stable over the course of 100 ns simulation time. Undoubtedly, chaetomugilins and chaetoviridins are fungi-derived metabolites that have multiple biological activities. They meet all the requirements for becoming drug leads in their respective therapeutic categories. They have varied chemical compositions that might provide the basis for the synthesis and design of novel and effective pharmaceutical agents. The different substituents at various positions in their skeleton play critical roles in the determination of some of their bioactivities. Finally, studies of the possible mechanisms, biosynthetic pathways, structure-activity relationships, and/or derivatization of these metabolites should be the focus of future research. Terretonin as a new protective agent against sepsis-induced acute lung injury: Impact on SIRT1/Nrf2/NF-κBp65/NLRP3 signaling Untapped potential of marine associated Cladosporium species: An overview on secondary metabolites, biotechnological relevance, and biological activities Genus Thielavia: Phytochemicals, industrial importance, and biological relevance Biologically active secondary metabolites and biotechnological applications of species of the family Chaetomiaceae (Sordariales): An updated review from Humicola genus: Chemical constituents, industrial importance, and biological activities Bright Side of Fusarium oxysporum: Secondary Metabolites Bioactivities and Industrial Relevance in Biotechnology and Nanotechnology Fusarithioamide A, a new antimicrobial and cytotoxic benzamide derivative from the endophytic fungus Fusarium chlamydosporium New tetracyclic triterpenoids from the endophytic fungus Fusarium sp Integracides F and G: New tetracyclic triterpenoids from the endophytic fungus Fusarium sp Biologically active fungal depsidones: Chemistry, biosynthesis, structural characterization, and bioactivities Fusarithioamide B, a new benzamide derivative from the endophytic fungus Fusarium chlamydosporium with potent cytotoxic and antimicrobial activities Terretonins from Aspergillus Genus: Structures, biosynthesis, bioactivities, and structural elucidation Fungal depsides naturally inspiring molecules: Biosynthesis, structural characterization, and biological activities I. γ-Butyrolactones from Aspergillus species: Structures, biosynthesis, and biological activities Biologically active secondary metabolites from the fungi Fungal metabolites with anticancer activity Fungal type I polyketide synthases Fungal polyketide biosynthesis-A personal perspective Natural products of filamentous fungi: Enzymes, genes, and their regulation Advances in cloning, functional analysis and heterologous expression of fungal polyketide synthase genes Fungal Pigments and Their Roles Associated with Human Health New glutamine-containing azaphilone alkaloids from deep-sea-derived fungus Chaetomium globosum HDN151398 Colorimetric characterization for comparative analysis of fungal pigments and natural food colorants Chaetolactam A, an azaphilone derivative from the endophytic fungus Chaetomium sp. g1 Chlorinated azaphilone pigments with antimicrobial and cytotoxic activities isolated from the deep sea derived fungus Chaetomium sp. NA-S01-R1 Total Synthesis and Structural Revision of Chaetoviridins, A Four New Azaphilones from Chaetomium globosum var. flavo-viridae Azaphilones: Chemistry and biology Evaluation of anticancer activity of Chaetomium cupreum extracts against human breast adenocarcinoma cell lines Chaetoglobosins and azaphilones produced by Canadian strains of Chaetomium globosum isolated from the indoor environment New azaphilones and chlorinated phenolic glycosides from Chaetomium elatum with caspase-3 inhibitory activity Identification of a unique azaphilone produced by Chaetomium globosum isolated from Polygonatum sibiricum Chaephilones A and B, two new azaphilone derivatives isolated from Chaetomium globosum Chaetomugilin J enhances apoptosis in human ovarian cancer A2780 cells induced by cisplatin through inhibiting pink1/parkin mediated mitophagy 11-and 4-epimers of chaetomugilin a, novel cytostatic metabolit marine fish-derived fungus Chaetomium globosum Azaphilones from the endophyte Chaetomium globosum Absolute stereostructures of novel azaphiloness produced by a marine fish-derived fungus Absolute stereostructures of cytotoxic metabolites, chaetomugilins A-C, produced by a Chaetomium species separated from a marine fish The structures of the cytotoxic metabolites produced by a marine fish-derived fungus Chaetomugilins, new selectively cytotoxic metabolites, produced by a marine fish derived Chaetomium species Absolute stereostructures of chaetomugilins G and H produced by a marine-fish-derived Chaetomium species Bioactive metabolites produced by Chaetomium globosum, an endophytic fungus isolated from Ginkgo biloba Three new azaphilones produced by a marine fish derived Chaetomium globosum Chemical constituents from an endophytic fungus Chaetomium globosum Z1 Potential allelopathic azaphilones produced by the endophytic Chaetomium globosum TY1 inhabited in Ginkgo biloba using the one strain-many compounds method Molecular epigenetic approach activates silent gene cluster producing dimeric bis-spiro-azaphilones in Chaetomium globosum CBS148.51 New metabolite with inhibitory activity against α-glucosidase and α-amylase from endophytic Chaetomium globosum New azaphilones, seco-chaetomugilins A and D, produced by a marine-fish-derived Chaetomium globosum Biological Evaluation and Chemical Identification of Secondary Metabolites of Endophytic Fungi from Egyptian Flora New azaphilones from Chaetomium globosum isolated from the built environment Two phytotoxic azaphilone derivatives from Chaetomium globosum, a fungal endophyte isolated from Amaranthus viridis leaves Inflammation-associated gene expression in RAW 264.7 macrophages induced by toxins from fungi common on damp building materials Detection of Chaetomium globosum, Ch. Cochliodes and Ch. Rectangulare during the diversity tracking of mycotoxin-producing Chaetomium-like isolates obtained in buildings in Finland Cytochalasan and Azaphilone Derivatives from a Marine-Derived Fungus Chaetomium globosum Structural Characterization of Secondary Metabolites Produced by Fungi Obtained from Damp Canadian Buildings Determination of the absolute configuration of chaetoviridins and other bioactive azaphilones from the endophytic fungus Chaetomium globosum Chemo-profiling of bioactive metabolites from Chaetomium globosum for biocontrol of Sclerotinia rot and plant growth promotion Chaetomugilins I-O, new potent cytotoxic metabolites from a marine-fish-derived Chaetomium species. Stereochemistry and biological activities New class azaphilone produced by a marine fish derived Chaetomium globosum. The stereochemistry and biological activities Bio-spiro-azaphilones and azaphilones from the fungi Chaetomium cochliodes VTh01 and C. cochliodes Cth05 Antifungal activity against plant pathogenic fungi of chaetoviridins isolated from Chaetomium globosum Chaetomium siamense sp. nov., a soil isolate from Thailand Bioassays guided isolation of compounds from Chaetomium globosum Isolation and identification of secondary metabolites from Chaetomium globosum CIB-160 and their immunological activity Production of mycotoxins by Chaetomium species Fungal endophytes isolated from Protium heptaphyllum and Trattinnickia rhoifolia as antagonists of Fusarium oxysporum The screening and identification of the biological control fungi Chaetomium spp. against wheat common root rot Bioactive metabolites from the desert plantassociated endophytic fungus Chaetomium globosum (Chaetomiaceae) Antifungal activity of chaetoviridin A from Chaetomium globosum CEF-082 metabolites against Verticillium dahliae in cotton Chaetoglobosins and azaphilones from Chaetomium globosum associated with Apostichopus japonicus Cytotoxic nitrogenated azaphilones from the deep-sea-derived fungus Chaetomium globosum MP4-S01-7 Cytotoxic azaphilone alkaloids from Chaetomium globosum TY1 New metabolites from endophytic fungus Chaetomium globosum CDW7. Molecules Identification of chaetoviridin E from a cultured microfungus, Chaetomium sp. and structural reassignment of chaetoviridins B and D The Structure of Dihydrodeoxy-8-epi-austdiol and the Absolute Configuration of the Azaphilones Azaphilones inhibit tumor promotion by 12-O-tetradecanoylphorbol-13-acetate in two-stage carcinogenesis in mice Caspases find a new place to hide Caspases as therapeutic targets Cholesteryl ester transfer protein inhibitors and cardiovascular outcomes: A systematic review and meta-analysis of randomized controlled trials Structure-specific inhibition of cholesteryl ester transfer protein by azaphilones Evolution of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) as coronavirus disease 2019 (COVID-19) pandemic: A global health emergency Repurposing of some natural product isolates as SARS-COV-2 main protease inhibitors via in vitro cell-free and cell-based antiviral assessments and molecular modeling approaches COVID-19: A promising cure for the global panic A review on coronavirus family persistency and considerations of novel type, COVID-19 features Coronavirus main proteinase (3CLpro) structure: Basis for design of anti-SARS drugs ALG-097111, a potent and selective SARS-CoV-2 3-chymotrypsin-like cysteine protease inhibitor exhibits in vivo efficacy in a Syrian Hamster model Design of wide-spectrum inhibitors targeting coronavirus main proteases Computational drug repurposing for the identification of SARS-CoV-2 main protease inhibitors Antiviral agents from fungi: Diversity, mechanisms and potential applications Protein and ligand preparation: Parameters, protocols, and influence on virtual screening enrichments PROPKA3: Consistent Treatment of Internal and Surface Residues in Empirical pK a Predictions Glide: A New Approach for Rapid, Accurate Docking and Scoring. 1. Method and Assessment of Docking Accuracy Biosynthesis of azaphilones: A review Identification and characterization of the chaetoviridin and chaetomugilin gene cluster in Chaetomium globosum reveals dual functions of an iterative highly reducing polyketide synthase Biochemical and structural basis for controlling chemical modularity in fungal polyketide biosynthesis Combinatorial generation of chemical diversity by redox enzymes in chaetoviridin biosynthesis Informed Consent Statement: Not applicable.Data Availability Statement: Not applicable. The authors declare no conflict of interest.