Next Article in Journal
Characterization Variation of the Differential Coloring Substances in Rapeseed Petals with Different Colors Using UPLC-HESI-MS/MS
Next Article in Special Issue
Exploration of Binding Affinities of a 3β,6β-Diacetoxy-5α-cholestan-5-ol with Human Serum Albumin: Insights from Synthesis, Characterization, Crystal Structure, Antioxidant and Molecular Docking
Previous Article in Journal
Mass Spectra Fitting as Diagnostic Tool for Magnetron Plasmas Generated in Ar and Ar/H2 Gases with Tungsten Targets
Previous Article in Special Issue
Biological Activity and Structural Diversity of Steroids Containing Aromatic Rings, Phosphate Groups, or Halogen Atoms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fascinating Furanosteroids and Their Pharmacological Profile

by
Valery M. Dembitsky
Centre for Applied Research, Innovation and Entrepreneurship, Lethbridge College, 3000 College Drive South, Lethbridge, AB T1K 1L6, Canada
Current address: Bio-Geo-Chem Laboratories, 23615 El Toro Rd X, P.O. Box 049, Lake Forest, CA 92630, USA.
Molecules 2023, 28(15), 5669; https://doi.org/10.3390/molecules28155669
Submission received: 27 June 2023 / Revised: 21 July 2023 / Accepted: 24 July 2023 / Published: 26 July 2023
(This article belongs to the Special Issue Steroid Compounds with Potential Biological Activity)

Abstract

:
This review article delves into the realm of furanosteroids and related isoprenoid lipids derived from diverse terrestrial and marine sources, exploring their wide array of biological activities and potential pharmacological applications. Fungi, fungal endophytes, plants, and various marine organisms, including sponges, corals, molluscs, and other invertebrates, have proven to be abundant reservoirs of these compounds. The biological activities exhibited by furanosteroids and related lipids encompass anticancer, cytotoxic effects against various cancer cell lines, antiviral, and antifungal effects. Notably, the discovery of exceptional compounds such as nakiterpiosin, malabaricol, dysideasterols, and cortistatins has revealed their potent anti-tuberculosis, antibacterial, and anti-hepatitis C attributes. These compounds also exhibit activity in inhibiting protein kinase C, phospholipase A2, and eliciting cytotoxicity against cancer cells. This comprehensive study emphasizes the significance of furanosteroids and related lipids as valuable natural products with promising therapeutic potential. The remarkable biodiversity found in both terrestrial and marine ecosystems offers an extensive resource for unearthing novel biologically active compounds, paving the way for future drug development and advancements in biomedical research. This review presents a compilation of data obtained from various studies conducted by different authors who employed the PASS software 9.1 to evaluate the biological activity of natural furanosteroids and compounds closely related to them. The utilization of the PASS software in this context offers valuable advantages, such as screening large chemical libraries, identifying compounds for subsequent experimental investigations, and gaining insights into potential biological activities based on their structural features. Nevertheless, it is crucial to emphasize that experimental validation remains indispensable for confirming the predicted activities.

Graphical Abstract

1. Introduction

Natural and synthesized organic compounds that contain a furan ring(s) exhibit diverse cardiovascular activity [1,2,3]. These compounds find extensive use as antibacterial, antiviral, anti-inflammatory, antifungal, anticancer, antihyperglycemic, analgesic, and anticonvulsant medications [1,2,3,4,5,6]. According to ChemNetBase, the number of natural compounds containing a furan ring exceeds 12,000 per molecule. Additionally, compounds with a 2,5-dihydrofuran ring surpass 12,200, while those with a 2,3-dihydrofuran ring exceed 1850. Moreover, compounds with a tetrahydrofuran ring account for over 47,570 compounds. Furthermore, steroids and related isoprenoid lipids that possess the additional ring(s) can be referred to as furanosteroids and their analogues and derivatives, comprising a collection of over 1000 compounds [7].
Furanosteroids represent a class of pentacyclic isoprenoid lipids that are synthesized by fungi and various other organisms. These compounds possess additional fused furan, dihydrofuran, or tetrahydrofuran ring(s) attached to their pentacyclic backbone [8,9,10].
Initially, furanosteroids were identified in mushrooms nearly 80 years ago, characterized by the presence of an extra furan ring connecting positions 4 and 6 of the steroid skeleton. However, the term “furanosteroids” has since been expanded to include any steroids and related compounds that contain the furan ring, 2,3-dihydrofuran ring, 2,5-dihydrofuran ring, and/or tetrahydrofuran ring (s). It is believed that the strained furan cycle contributes to the diverse biological activities exhibited by these compounds. Several metabolites belonging to this class of natural products have garnered significant attention in pharmacological research due to their potent anti-inflammatory and antibiotic properties, potential anti-proliferative activity, and ability to inhibit inositide-3-kinase [8,9,11,12,13,14].
This review is dedicated to furanosteroids and related isoprenoid lipids, their occurrence in fungi, plants, and marine organisms, and the exploration of their biological activity.

2. Furanosteroids Produced by Fungi and Fungal Endophytes

The furanosteroid scaffold (1) represents a highly oxygenated framework comprising a [5,6,6,6]-tetracycle (A-B-C-D rings), along with the presence of a furan ring, as depicted in Figure 1. Furanosteroids, exemplified by compounds 29, 11, 12, 14, and 15, with their biological activities summarized in Table 1, encompass both the furan ring and steroids, viridin (2, 3D model is shown in Figure 2), 35, and 15.
These steroids can be classified as aromatic steroids due to the presence of the C aromatic ring. Among them, the steroid known as secovironolide (10) contains an additional 2,3-dihydrofuran ring, while tricholumin A (16) is an example that includes a 2,5-dihydrofuran ring. Furthermore, the most commonly encountered steroids, 13 and 1725, feature a tetrahydrofuran ring, as illustrated in Figure 1.
The data presented in Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8, Table 9 and Table 10 are taken from published data and obtained using the German computer software PASS (http://www.akosgmbh.de/mobile/pass.htm accessed on 1 April 2023). This program is in the public domain, and annually it is used by more than 26,000 scientists from around the world. The site of this program provides complete information on its use as well as the interpretation of the data obtained.
Viridin (2, activity is shown in Figure 3), a furano-steroidal antibiotic, represents the first identified member of the furanosteroid family. It was initially isolated in 1945 from a pigment-forming strain of the common soil fungus Trichoderma viride (as shown in Figure 4) [15,16]. Viridin, along with its derivatives 3, 4, 5, 6, and 11, have demonstrated remarkable antifungal and antibacterial activities. Moreover, these compounds are known as potent inhibitors of the lipid kinase PI-3K [16]. Phosphatidylinositol 3-kinases (PI3Ks) are lipid kinases that play a central role in cell cycle regulation, apoptosis, DNA repair, aging, angiogenesis, cellular metabolism, and motility [16,17,18]. These enzymes catalyze the synthesis of specific members of the lipid family collectively known as “phosphoinositides”. These PI3K products can in turn modulate the activation of many downstream proteins, ultimately regulating several cellular processes. Mammalian cells possess eight PI3Ks, which are grouped into three classes depending on their structure and substrate specificity [16,17,18].
More recently, viridin (2) and the phytotoxin viridiol (3) were discovered in liquid cultures produced by the fungus Gliocladium virens [19]. Additionally, demethoxyviridin (4) and demethoxyviridiol (5) were isolated for the first time from an unidentified fungus [20]. The ash dieback-causing fungus Hymenoscyphus pseudoalbidus yielded furanosteroids (17 and 18) along with known compounds like viridiol (6) and demethoxyviridiol (5) [14,20,21]. Epoxyvirone (11) was detected in Talaromyces species (as depicted in Figure 5) and produced by the marine sponge-associated fungus Talaromyces stipitatus KUFA 0207 [22,23]. Ding and co-workers [24] reported the isolation of wortmannolone (7), wortmannin (8), 11-desacetoxywortmannin-17β-ol (9), secovironolide (10, represented by the 3D graph in Figure 6), epoxyvirone (11), wortmannin C (13), and 3-dihydrovirone (12) from the culture broth of the endophytic fungus Talaromyces wortmannii LGT-4, which was derived from the Chinese medicinal plant Tripterygium wilfordii.
Furthermore, a furanosteroid (10), along with viridiol (3), and 8, 11, and 13 were isolated from the soil fungus Trichoderma virens. Notably, 9-epi-viridiol (23) and viridiol (2) exhibited cytotoxicity against HeLa and KB cells, with IC50 values of 19 and 50 μg/mL, respectively [25].
Wortmannin (8) was initially isolated in 1957 by Brian and his colleagues from the broth of Penicillium wortmanni and subsequently from other fungi such as P. funiculosum, Talaromyces wortmannii, Fusarium oxysporum, P. radicum, and Talaromyces sp. [17,26,27]. Wortmannin and its derivatives, including compounds 7, 8, 9, 13, and 23, exhibit antiproliferative properties and phosphatidylinositol 3-kinase activity [28,29,30,31]. Two furanosteroids, wortmannolone (7) and wortmannolol (14), were isolated from the fungal endophyte Talaromyces sp., which was obtained from Tripterygium wilfordii [31].
Figure 3. Percentage distribution of dominant and related biological activities using the widely recognized furanosteroid, viridin (2), which possesses unique pharmacological properties. The activities are indicated by the following numbers: 1. Antifungal (14.4%); 2. Anti-inflammatory (14.4%); 3. Antiprotozoal (Plasmodium) (12.6%); 4. Antibacterial (12.4%); 5. Antineoplastic (14.4%); 6. Apoptosis agonist (13.1%); 7. Prostate disorders treatment (9.5%); and 8. Proliferative diseases treatment (9.2%). The antifungal activity of the fungus Trichoderma viride has been observed to exhibit an inhibitory effect against various pathogens, such as Fusarium solani, Rhizoctonia solani, and Sclerotium rolfsii. Moreover, the extract obtained from raw mycelium has demonstrated notable antimicrobial activity, specifically displaying an antibacterial effect against Bacillus subtilis, Escherichia coli, and Pseudomonas fluorescens. Maximum antifungal effectiveness has also been recorded against Candida albicans, Rhizoctonia solani, Pythium ultimum, Fusarium solani, and F. oxysporium. The viridin metabolite present in the alcoholic mycelia extract of Trichoderma viride exerts antimicrobial, antifungal, and anticancer effects [14,32,33,34].
Figure 3. Percentage distribution of dominant and related biological activities using the widely recognized furanosteroid, viridin (2), which possesses unique pharmacological properties. The activities are indicated by the following numbers: 1. Antifungal (14.4%); 2. Anti-inflammatory (14.4%); 3. Antiprotozoal (Plasmodium) (12.6%); 4. Antibacterial (12.4%); 5. Antineoplastic (14.4%); 6. Apoptosis agonist (13.1%); 7. Prostate disorders treatment (9.5%); and 8. Proliferative diseases treatment (9.2%). The antifungal activity of the fungus Trichoderma viride has been observed to exhibit an inhibitory effect against various pathogens, such as Fusarium solani, Rhizoctonia solani, and Sclerotium rolfsii. Moreover, the extract obtained from raw mycelium has demonstrated notable antimicrobial activity, specifically displaying an antibacterial effect against Bacillus subtilis, Escherichia coli, and Pseudomonas fluorescens. Maximum antifungal effectiveness has also been recorded against Candida albicans, Rhizoctonia solani, Pythium ultimum, Fusarium solani, and F. oxysporium. The viridin metabolite present in the alcoholic mycelia extract of Trichoderma viride exerts antimicrobial, antifungal, and anticancer effects [14,32,33,34].
Molecules 28 05669 g003
Another set of furanosteroids, namely 9-epi-viridiol (15) and viridiol (3), were isolated from Trichoderma virens [25]. Tricholumin A (16) possesses a unique carbon skeleton and is a highly transformed ergosterol derivative. It was isolated from the alga-endophytic fungus Trichoderma asperellum. Tricholumin A has demonstrated inhibitory effects against several pathogenic microbes, including V. harveyi, V. splendidus, and Pseudoalteromonas citrea. Additionally, it has displayed antifungal activity against Glomerella cingulata and has shown inhibition towards various marine phytoplankton species such as Chattonella marina, Heterosigma akashiwo, Karlodinium veneficum, and Prorocentrum donghaiense [35].
An unusual steroid compound, 1,11-epoxy-3-hydroxy-methylenecycloartan-28-oic acid (17), was isolated from an unidentified fungus [36]. Similar steroid structures have been identified in species belonging to the Agaricaceae family. Notably, a series of rare 1,11-epoxy lanostane-type triterpenoids called lepiotaprocerins A–F (1823) were isolated from the fruiting bodies of the edible mushroom Macrolepiota procera collected in Poland. These compounds exhibited significant inhibitory effects on nitric oxide production, outperforming the positive control, L-NG-monomethyl arginine [37].
Figure 4. (a) Trichoderma viride is a fungus or mold that reproduces asexually through spores. The mycelium of this fungus can produce various enzymes, such as cellulases and chitinases. These enzymes are involved in the degradation of cellulose and chitin, respectively. Trichoderma viride commonly grows on wood, which predominantly consists of cellulose, as well as on fungi whose cell walls are primarily composed of chitin. It exhibits a parasitic nature towards the mycelium and fruiting bodies of other fungi, including cultivated ones. The fungus is known to cause “Green mold fungus disease” [38,39]. (b) Hymenoscyphus pseudoalbidus, also known as the Ash Wax Moth, is a saprotrophic fungus. It can be found growing on decaying trunks and branches of ash trees. The fruiting bodies of this fungus are hardly distinguishable from a macroscopic or microscopic perspective. It is commonly observed on the same substrate [40,41].
Figure 4. (a) Trichoderma viride is a fungus or mold that reproduces asexually through spores. The mycelium of this fungus can produce various enzymes, such as cellulases and chitinases. These enzymes are involved in the degradation of cellulose and chitin, respectively. Trichoderma viride commonly grows on wood, which predominantly consists of cellulose, as well as on fungi whose cell walls are primarily composed of chitin. It exhibits a parasitic nature towards the mycelium and fruiting bodies of other fungi, including cultivated ones. The fungus is known to cause “Green mold fungus disease” [38,39]. (b) Hymenoscyphus pseudoalbidus, also known as the Ash Wax Moth, is a saprotrophic fungus. It can be found growing on decaying trunks and branches of ash trees. The fruiting bodies of this fungus are hardly distinguishable from a macroscopic or microscopic perspective. It is commonly observed on the same substrate [40,41].
Molecules 28 05669 g004
Figure 5. (a) Talaromyces marneffei: Talaromyces is a genus of fungi belonging to the family Trichocomaceae. It was first described in 1955 by the American mycologist Chester Ray Benjamin, who specialized in the taxonomy of fungal molds in the Eurotiales and Mucorales orders. Talaromyces marneffei, previously known as Penicillium marneffei, was identified in 1956. This fungus is endemic to Southeast Asia and is a significant cause of opportunistic infections in individuals with immunodeficiency associated with HIV/AIDS [42,43]. (b) Penicillium funiculosum: P. funiculosum is a plant pathogenic fungus that infects pineapple fruits. The disease it causes is referred to as fruit core rot, which is characterized by the browning or blackening and rotting of the fruit’s core [44,45]. (c) Fusarium oxysporum: F. oxysporum is an ascomycete fungus belonging to the family Nectriaceae. F. oxysporum strains are commonly found in soil and can exist as saprophytes, deriving nutrients from decaying organic matter. They are capable of degrading lignin and complex carbohydrates present in soil residues [46,47].
Figure 5. (a) Talaromyces marneffei: Talaromyces is a genus of fungi belonging to the family Trichocomaceae. It was first described in 1955 by the American mycologist Chester Ray Benjamin, who specialized in the taxonomy of fungal molds in the Eurotiales and Mucorales orders. Talaromyces marneffei, previously known as Penicillium marneffei, was identified in 1956. This fungus is endemic to Southeast Asia and is a significant cause of opportunistic infections in individuals with immunodeficiency associated with HIV/AIDS [42,43]. (b) Penicillium funiculosum: P. funiculosum is a plant pathogenic fungus that infects pineapple fruits. The disease it causes is referred to as fruit core rot, which is characterized by the browning or blackening and rotting of the fruit’s core [44,45]. (c) Fusarium oxysporum: F. oxysporum is an ascomycete fungus belonging to the family Nectriaceae. F. oxysporum strains are commonly found in soil and can exist as saprophytes, deriving nutrients from decaying organic matter. They are capable of degrading lignin and complex carbohydrates present in soil residues [46,47].
Molecules 28 05669 g005
Endophytic fungi originating from Astragalus species were found to catalyze the biotransformation of cycloastragenol, leading to the synthesis of an uncommon meroterpenoid compound 24 [48]. Additionally, through the utilization of an Astragalus endophyte called Alternaria eureka 1E1BL1, the biotransformation of cyclocephagenol, a novel cycloartane-type sapogenin with a tetrahydropyran unit, resulted in the formation of rare metabolites (2527). The structures of these metabolites are depicted in Figure 7 [49].
Figure 6. Three-dimensional graph illustrates the predicted and calculated activity as inositol-3-kinase inhibitors for steroids 7, 8, and 10 produced by the endophytic fungus Talaromyces wortmannii. The graph demonstrates a high confidence level of over 94% in the activity predictions. According to the data from Table 1, the dominant activity of steroids 7, 8, and 10 in the graph is shown in red. The steroid wortmanolone (7) is an antineoplastic and inositol-3-kinase inhibitor. The steroid wortmannin A (8) is an inositol-4-kinase inhibitor and inositol-3-kinase inhibitor of activity, and steroid secovironolide (10) is an inositol-4-kinase inhibitor and inositol-3-kinase inhibitor of activity. The activity text is in red for the steroid wortmanolone; in black is the activity of the steroid wortmannin A; and in blue is the activity of the steroid secovironolide. Red color—strong activity, blue color—no activity.
Figure 6. Three-dimensional graph illustrates the predicted and calculated activity as inositol-3-kinase inhibitors for steroids 7, 8, and 10 produced by the endophytic fungus Talaromyces wortmannii. The graph demonstrates a high confidence level of over 94% in the activity predictions. According to the data from Table 1, the dominant activity of steroids 7, 8, and 10 in the graph is shown in red. The steroid wortmanolone (7) is an antineoplastic and inositol-3-kinase inhibitor. The steroid wortmannin A (8) is an inositol-4-kinase inhibitor and inositol-3-kinase inhibitor of activity, and steroid secovironolide (10) is an inositol-4-kinase inhibitor and inositol-3-kinase inhibitor of activity. The activity text is in red for the steroid wortmanolone; in black is the activity of the steroid wortmannin A; and in blue is the activity of the steroid secovironolide. Red color—strong activity, blue color—no activity.
Molecules 28 05669 g006
Various Ganoderma species, used in traditional Asian medicine to prevent and treat diseases [50], have yielded intriguing seco-steroids (2841), as depicted in Table 2. Extracts from the wood-decay fungus Ganoderma applanatum revealed the discovery of ganoapplanic acid A (28), ganodapplanoic acids A (29), and B (30). These rearranged lanotane-type triterpenoids possess a 6/6/5/6-fused tetracyclic structure, featuring an uncommon C-13/C-15 oxygen bridge moiety [51,52]. Furthermore, G. cochlear yielded cochlates A and B (31 and 32), isomeric compounds with a 3,4-seco-9,10-seco-9,19-cyclo skeleton [53], while fornicatin A (33), a 3,4-seco-trinortriterpenoid, was first isolated from the fruiting bodies of G. fornicatum [54].
An intriguingly rearranged hexanorlanostane, known as cochlate C (34), was obtained from the fruiting bodies of Ganoderma cochlear [55]. Additionally, Ganoderma tropicum and G. boninense, two medicinal mushrooms, yielded a steroid (35) and a series of secolanostanes (3641, 3D graph, see Figure 8) with anti-plasmodial activity [56,57,58]. Furthermore, the fungal strain Emericella sp. TJ29, an endophyte derived from the root of Hypericum perforatum, produced several extraordinary meroterpenoids named emeridones A–F (4247) [59].
Trichocitrin (48), a diterpene, was extracted from the alga-endophytic fungus Trichoderma citrinoviride. Its structure is depicted in Figure 9, and the biological activity is outlined in Table 3. This compound represents the first furan-bearing fusicoccane diterpene derived from Trichoderma sp. and has shown inhibitory effects against E. coli [60]. The fungus Aspergillus ustus, isolated from the Mediterranean sponge Suberites domuncula, produced the sesterterpenoids ophiobolin-types 49 and 50 [61].
Two phytotoxic sesterterpenoids, namely ophiobolin A lactone (51) and B (52), are synthesized by Pseudomonas aeruginosa, while compound 52 is also produced by the pathogenic fungi Drechslera maydis and D. sorghicola [62,63,64]. Additionally, the two mangrove fungi Aspergillus terreus H010 and Lophiostoma bipolare BCC25910 yielded compounds 5355; structures are shown in Figure 9, and a 3D graph is shown in Figure 10 [65,66].
Microbial transformation by the bacterial strain Bacillus sp. IMM-006 yielded two harziene-type diterpenoids, furanharzianones A (56) and B (57), featuring an unusual 4/7/5/6/5 ring system [67,68].
Dictyophora rubrovolvata, a saprophytic mushroom extensively cultivated in China, particularly in Guizhou Province, serves as a valuable source of two diterpenoids: 7,16,17-trihydroxy-19,6-kauranolide (58) and 2,7,11,14-tetrahydroxy-16-kauren-19,6-olide (59) [69]. Additionally, an Acrostalagmus fungus yielded LL-Z 1271a (60), a C16 terpenoid known for its antifungal properties [70,71]. Aspergilone A (61), an isopimarane compound, was isolated from a fungal species called Epicoccum [72], whereas xylarenolide (62), a pimarane derivative, has been identified in endophytic Tubercularia species [73] and certain Xylaria species [74,75].
From the solid culture of Aspergillus flocculosus 16D-1, two distinct steroids were isolated. The first one, asperflotone (63), is an unusual 8(14 → 15)-abeo-ergostane-type steroid. It possesses a unique ergosteroid structure characterized by a rearranged bicyclo [4.2.1]non-2-ene ring system, potentially resulting from α-ketol rearrangement during biosynthesis. Remarkably, both asperflotone and the second steroid, asperfloroid (64), exhibited inhibitory activity against IL-6 production in induced THP-1 cells [76]. In addition, the sponge-derived fungus Aspergillus flocculosus 16D-1 produced two other steroids, aspersecosteroids A (65) and B (66), which are 11(9→10)-abeo-5,10-secosteroids. Notably, both compounds demonstrated a potent inhibitory effect on the production of TNF-α and IL-6 [77].
Tricholumin A (67) was obtained from the alga-endophytic fungus Trichoderma asperellum. It retains cycle A, the final structural element of the original ergosterol, after undergoing deep oxidative transformations in the rest of the molecule, including the side chain fragment [35].
Xylarglycoside B (68), an antibacterial steroid, was isolated from the fungus Xylaria sp. KYJ-15, which was derived from the leaves of Illigera celebica. It displayed antibacterial activity against Staphylococcus aureus [78]. The endophytic fungus Emericella variecolor led to the isolation of emericellic acid (69), a meroterpenoid compound [79]. From the fruiting bodies of the mushroom Stropharia rugosoannulata, an unusual sterol with an unprecedented ether ring (70) was isolated. This mushroom is known as saketsubatake in Japanese and wine-cap stropharia in English [80]. Ergopyrone (71), an extraordinary styrylpyrone-fused ergosterol derivative, was isolated and structurally characterized from the mushroom Gymnopilus orientispectabilis. This steroid features a hexacyclic 6/5/6/6/6/5 skeleton that is formed via [3 + 2] cycloaddition between ergosterol and styrylpyrone precursors [81].
Physalis angulata yielded several C28 steroids of the withasteroid family, including physanolide A (72) with an unprecedented skeleton containing a seven-membered ring and various physalins (7380), as shown in Figure 11. These compounds displayed biological activity, as outlined in Table 4. Physalins B, D, and F exhibited potent cytotoxicity against multiple tumor cell lines, including KB, A431, HCT-8, PC-3, and ZR751, with EC50 values below 0.4 μM [82,83]. Antheridiol (81), the fungal sex hormone, was isolated from Achlya bisexualis, a water mold [84]. Physangulide B (82), a steroid, was identified in the calyxes of Physalis angulata, featuring an additional tetrahydrofuran ring in its structure [85].
Bioactive compounds 8399, known as withaphysalins, were isolated from Physalis minima. These compounds typically possess a hemiacetal or lactone linkage between C-18 and C-20, and they have demonstrated cytotoxic, antiproliferative, and anti-inflammatory activities [86,87].
Agaricus blazei (a picture of this fungus is shown in Figure 12), also known as ‘Cogumelo do Sol’ in Brazil or ‘Himematsutake’ in Japan, is a widely cultivated mushroom with medicinal uses.
It has been traditionally employed to treat various common ailments such as atherosclerosis, hepatitis, hyperlipidemia, diabetes, dermatitis, and cancer. The mushroom contains bioactive, highly oxygenated des-A-ergostane derivatives, including agariblazeispirols A (90) and B (91), as well as blazeispirols B (92), C (93), E (94), and F (95), which were isolated from cultured mycelia of Agaricus blazei [88,89]. Agariblazeispirols (9799, 3D graph, see Figure 13) exhibited a moderate circumvention of drug resistance in mouse leukemia P388/VCR cells [89]. Additionally, blazeispirol A (96), featuring an unprecedented skeleton, has been isolated from the cultured mycelia of the same fungus [90].
A triterpenoid named irpexolidal (97) with an unprecedented carbon skeleton, along with its biogenetic-related compound irpexolide A (98) were isolated from the fruiting bodies of the medicinal fungus Irpex lacteus [91]. Study of the extract of the fruiting bodies of the mushroom Leucopaxillus gentianeus, allowed the isolation of minor cucurbitane triterpene, leucopaxillone B (99, the 3D graph is shown in Figure 13). The antiproliferative activity of the isolated triterpene was determined against the NCI-H460 human tumour cell line [92]. Simplifusidic acids B (100) and D (101), fusidane-type nortriterpenoids, were isolated from the marine-derived fungus Simplicillium sp. SCSIO 41513 [93].
Figure 12. (a) Xylaria sp.: This ascomycete fungus typically thrives on dead wood and is known for producing steroids (68). The genus Xylaria comprises fungal endophytes associated with both vascular and non-vascular plants, and it serves as a valuable source of bioactive secondary metabolites. These metabolites include sesquiterpenoids, esters, alcohols, terpenoids, cytochalasins, mellein, alkaloids, polyketides, and aromatic compounds. Some of these compounds have demonstrated potential activity as herbicides, fungicides, insecticides, antibacterials, antimalarials, antifungals, or α-glucosidase inhibitors [94]. (b) Agaricus blazei: This medicinal fungus contains secosteroids (9095). With significant commercial value, Agaricus blazei offers a wide range of health benefits. The mushrooms are rich in biologically active substances such as polysaccharides, lipids, sterols, proteins, vitamins B, C, and D, as well as phenolic compounds. Polysaccharides from A. blazei have been shown to possess immunoregulatory, anti-inflammatory, hepatoprotective, and antitumor properties. Extracts from this fungus have been used to treat diabetes and bacterial infections, exhibiting anticarcinogenic and antimutagenic effects [95,96]. (c) Irpex lacteus: This medicinal fungus serves as a source of triterpenoids (97 and 98). Irpex lacteus, a white rot fungus, is widely employed in bioremediation and food biotechnology due to its exceptional lignin-degrading capabilities. The fungus produces various extracellular enzymes, including lignin peroxidase, laccase, glucose oxidase, proteases, and α-galactosidase, involved in oxidative, hydrolytic, and lignocellulose degradation processes. Irpex lacteus readily oxidizes steroids, triterpenoids, alkanes, and cyclic ketones [97,98].
Figure 12. (a) Xylaria sp.: This ascomycete fungus typically thrives on dead wood and is known for producing steroids (68). The genus Xylaria comprises fungal endophytes associated with both vascular and non-vascular plants, and it serves as a valuable source of bioactive secondary metabolites. These metabolites include sesquiterpenoids, esters, alcohols, terpenoids, cytochalasins, mellein, alkaloids, polyketides, and aromatic compounds. Some of these compounds have demonstrated potential activity as herbicides, fungicides, insecticides, antibacterials, antimalarials, antifungals, or α-glucosidase inhibitors [94]. (b) Agaricus blazei: This medicinal fungus contains secosteroids (9095). With significant commercial value, Agaricus blazei offers a wide range of health benefits. The mushrooms are rich in biologically active substances such as polysaccharides, lipids, sterols, proteins, vitamins B, C, and D, as well as phenolic compounds. Polysaccharides from A. blazei have been shown to possess immunoregulatory, anti-inflammatory, hepatoprotective, and antitumor properties. Extracts from this fungus have been used to treat diabetes and bacterial infections, exhibiting anticarcinogenic and antimutagenic effects [95,96]. (c) Irpex lacteus: This medicinal fungus serves as a source of triterpenoids (97 and 98). Irpex lacteus, a white rot fungus, is widely employed in bioremediation and food biotechnology due to its exceptional lignin-degrading capabilities. The fungus produces various extracellular enzymes, including lignin peroxidase, laccase, glucose oxidase, proteases, and α-galactosidase, involved in oxidative, hydrolytic, and lignocellulose degradation processes. Irpex lacteus readily oxidizes steroids, triterpenoids, alkanes, and cyclic ketones [97,98].
Molecules 28 05669 g012
Figure 13. Three-dimensional graph illustrating the predicted and calculated antineoplastic activity of steroids (74, 75, and 99) with a confidence level exceeding 91%. These steroids are derived from meroterpenoids produced by the fungus Physalis angulata, while the steroid is sourced from the mushroom Leucopaxillus gentianeus.
Figure 13. Three-dimensional graph illustrating the predicted and calculated antineoplastic activity of steroids (74, 75, and 99) with a confidence level exceeding 91%. These steroids are derived from meroterpenoids produced by the fungus Physalis angulata, while the steroid is sourced from the mushroom Leucopaxillus gentianeus.
Molecules 28 05669 g013

3. Furanosteroids Derived from Plant Species

Furanosteroids derived from plant species have been extensively studied. Around 50 years ago, Indian scientists isolated a hexacyclic tetranortriterpenoid called vilasinin (102) from the green leaves of the neem tree (Azadirachta indica). Since then, numerous related metabolites have been discovered in various plant species [99]. 1,3-Diacetylvilasinin (103) has been reported from Melia volkensii [100], Chisocheton paniculatus [101], Azadirachta indica [102], as well as two African Turraea species, Turraea holstii and T. parvifolia [103]. Another compound, 1,3-diacetyl-12α-hydroxy-7-tigloylvilasinin (104), has been found in Azadirachta indica [104] and Malleastrum antsingyense (depicted in Figure 14) [105].
The leaves of Trichilia gilgiana, extracted with CH2Cl2-MeOH, yielded vilasinin-type limonoids known as rubescin H (105), gilgianin A (106), gilgianin B (107), TS3 (108), and trichirubine A (109). Furanosteroids 105, 106, and 107 demonstrated potent anti-plasmodial activity alongside significant cytotoxicity. Compounds 108 and 109 exhibited the highest anti-plasmodial activity, with IC50 values of 1.1 and 1.3 μM, respectively [106]. The structures of compounds 105, 106, and 107 are shown in Figure 15, and their biological activity is detailed in Table 5. Furthermore, a 3D graph illustrating the activity of compounds 108 and 109 is displayed in Figure 16.
Figure 14. (a) Malleastrum: Malleastrum is a genus comprising over 20 species within the Meliaceae family. Native to Madagascar, the Comoros, and Aldabra, plants in the Meliaceae family are known to contain limonoids, terpenoids, alkaloids, flavonoids, and phenolic compounds as their primary chemical constituents. Many species within this family exhibit cytotoxic, antimicrobial, or antimalarial activity [107]. (b) Azadirachta indica (Neem): Azadirachta indica, commonly known as Neem, belongs to the Meliaceae family. The leaves of Neem are widely used in Chinese, Ayurvedic, and Unani medicines, particularly in the Indian subcontinent. Neem leaves have demonstrated antibacterial, anthelmintic, antiviral, and anticancer properties, and most notably, they act as an immunomodulatory agent [108]. (c) Annona squamosa (Sugar Apple): Annona squamosa, also known as sugar apple, is a member of the Annonaceae family. It has been traditionally used in Indian, Thai, and American medicine. The leaves of sugar apple are commonly used as a decoction to treat dysentery and urinary tract infections [109].
Figure 14. (a) Malleastrum: Malleastrum is a genus comprising over 20 species within the Meliaceae family. Native to Madagascar, the Comoros, and Aldabra, plants in the Meliaceae family are known to contain limonoids, terpenoids, alkaloids, flavonoids, and phenolic compounds as their primary chemical constituents. Many species within this family exhibit cytotoxic, antimicrobial, or antimalarial activity [107]. (b) Azadirachta indica (Neem): Azadirachta indica, commonly known as Neem, belongs to the Meliaceae family. The leaves of Neem are widely used in Chinese, Ayurvedic, and Unani medicines, particularly in the Indian subcontinent. Neem leaves have demonstrated antibacterial, anthelmintic, antiviral, and anticancer properties, and most notably, they act as an immunomodulatory agent [108]. (c) Annona squamosa (Sugar Apple): Annona squamosa, also known as sugar apple, is a member of the Annonaceae family. It has been traditionally used in Indian, Thai, and American medicine. The leaves of sugar apple are commonly used as a decoction to treat dysentery and urinary tract infections [109].
Molecules 28 05669 g014
Azadirachta indica, belonging to the family Meliaceae, is commonly known as neem or Indian lilac. It is utilized for its antimalarial, anti-inflammatory, antipyretic, antitumor, and anthelmintic properties. The leaves of Azadirachta indica contain tetranortriterpenoids such as 110 and 111 [110].
Sutherlandia frutescens, a plant from the Fabaceae family native to South Africa, is commonly referred to as Cancer bush. It is renowned for its multifunctional medicinal uses, and infusions and decoctions of the plant are widely employed in South Africa for treating cancer, inflammation, viral infections, and gastrointestinal diseases. An unusual cycloartane glycoside called sutherlandioside A (112) has been isolated from the water-methanol fraction of the plant [111].
Figure 15. Steroids and meroterpenoids derived from plant species.
Figure 15. Steroids and meroterpenoids derived from plant species.
Molecules 28 05669 g015
Chisocheton paniculatus twigs yielded two tetranortriterpenoids named chisonimbolinins A (113) and F (114) [112]. Furthermore, a triterpenoid 115 with insecticidal activity was isolated from a methanol extract of fresh leaves of Azadirachta indica [113].
Walsura cochinchinensis bark extract yielded two limonoids, walsucochinones B (116) and C (117). The ethyl acetate extract and walsucochinone C (117) displayed cytotoxic activity against MCF-7 human breast cancer cells [114]. Melia toosendan (Meliaceae) root bark provided spirosendan (118), a skeletal limonoid with a spiro-structure [115]. Additionally, an aqueous methanolic extract of Cedrela odorata leaves yielded four tetranortriterpenoids: cedrodorin (119), 6-acetoxycedrodorin (120), 6-deoxy-9R-hydroxycedrodorin (121), and 9R-hydroxycedrodorin (122) [116].
Dysoxylum gaudichaudianum, commonly known as ivory mahogany, yielded four tetranortriterpenoids named dysoxylins A–D (123126), which exhibited potent antiviral activity against the respiratory syncytial virus [117]. Xylocarpus rumphii heartwood provided a triterpenoid derivative identified as xylorumphiins E (127) [118].
Stem bark extracts of Khaya anthotheca contained three limonoids: anthothecanolide (128), 3-O-acetylanthothecanolide (129), and 2,3-di-O-acetyl-anthothecanolide (130) [119]. The structure of compound 130 is depicted in Figure 17, and its biological activities are outlined in Table 6.
A tetranortriterpenoid called kokosanolide D (131) has been isolated from the methanol extract of fruit peels of Lansium domesticum, found in West Java, Indonesia [120]. Furthermore, an ethanol extract of a plant belonging to the Malleastrum genus yielded three steroids known as malleastrones A–C (132, 133, and 134), respectively. Compounds 133 and 134 exhibited antiproliferative activity against a range of cancer cell lines, with IC50 values ranging from 0.19 to 0.63 µM [121,122].
Figure 16. Three-dimensional graph illustrates the predicted and calculated antiprotozoal (Plasmodium) activity of the vilasinin-type limonoids (106, 107, 108, and 109) with a confidence level exceeding 95%. These limonoids are derived from the tree Trichilia gilgiana, which is predominantly found in southern Nigeria and eastern Congo. In the Congo region, the extract of Trichilia gilgiana bark is utilized for its analgesic and stimulant properties. It has been traditionally used in traditional medicine to treat abdominal pain, chest pain, fever, and as a tonic. The juice of the young leaves is applied to circumcised wounds, while crushed leaves are added to drinking water for the treatment of respiratory diseases. Additionally, the juice of the leaves has demonstrated potent activity against the malarial Plasmodium [123,124,125].
Figure 16. Three-dimensional graph illustrates the predicted and calculated antiprotozoal (Plasmodium) activity of the vilasinin-type limonoids (106, 107, 108, and 109) with a confidence level exceeding 95%. These limonoids are derived from the tree Trichilia gilgiana, which is predominantly found in southern Nigeria and eastern Congo. In the Congo region, the extract of Trichilia gilgiana bark is utilized for its analgesic and stimulant properties. It has been traditionally used in traditional medicine to treat abdominal pain, chest pain, fever, and as a tonic. The juice of the young leaves is applied to circumcised wounds, while crushed leaves are added to drinking water for the treatment of respiratory diseases. Additionally, the juice of the leaves has demonstrated potent activity against the malarial Plasmodium [123,124,125].
Molecules 28 05669 g016
Figure 17. Bioactive Steroids and triterpenoids derived from plant species.
Figure 17. Bioactive Steroids and triterpenoids derived from plant species.
Molecules 28 05669 g017
Limbocinin (135) and compound 136 were isolated from ethanolic extracts of neem seeds and the leaves of Azadirachta indica and Annona squamosa [126,127]. Both compounds exhibited antifungal activity [128].
Furanosteroids (137 and 138) have been isolated from plants belonging to the Meliaceae and Simaroubaceae families [129,130,131,132]. Furthermore, a rearranged lanostane (139) has been isolated from Abies nephrolepis, commonly known as Khingan fir [133].
Two limonoids with an aromatic ring D, walsucochinoids A (140) and B (141) were detected in the air-dried plant Walsura cochinchinensis (Figure 18 depicts a picture of this plant).
From the leaves of Caloncoba glauca, a triterpenoid called caloncobalactone C (142) was isolated. It exhibited inhibitory activity against human and mouse 11β-hydroxysteroid dehydrogenase types 1 and 2 [140,141]. The 3D graph illustrating its structure is presented in Figure 19.
Within the EtOAc and n-BuOH extracts of the vines and leaves of Momordica aurantia, two steroids, octanorcucurbitacin (143) and kuguacin I (144), were discovered. These compounds demonstrated anti-HIV-1 activities in vitro [142]. Triterpenes gypensapogenin H (145) and I (146) were isolated from the hydrolyzate of the total saponin extract obtained from the dioecious herbaceous climber Gynostemma pentaphyllum, which is widely distributed in South and East Asia [143].
Two cucurbitane-type triterpenoids, namely (23E)-5β,19-epoxycucurbita-6,23,25-triene-3β-ol (147) and (19R,23E)-5β,19-epoxy-19-ethoxycucurbita-6,23-diene-3β,25-diol (148), have been isolated from the fruit of Momordica charantia. These compounds exhibited weak cytotoxic activity against cancer cell lines including MCF-7, HepG2, Du145, Colon205, and HL-60 [144].
The roots of Cimicifuga heracleifolia, a plant included in the Chinese Pharmacopoeia and used in traditional medicine in China for centuries, contain two unusual ring A cracking 9,19-cycloartane triterpenes (149) and (150) [145]. The dried fruit of Vitex negundo yielded the 3,10-epoxide (151) with antitumor activity [146].
From the seeds of Hovenia trichocarpa, two saponins named hoduloside XI and hoduloside XII were isolated. Both compounds shared the genin 20,26-epoxy-pseudojujubogenin (152) and displayed inhibitive activities against human cancer cell lines HL-60 and K562 [147]. In addition to these steroids, a furanosteroid named ceramicine J (153) was discovered in the hexane layer of Chisocheton ceramicus bark extract. This compound exhibited dose-dependent, moderate cytotoxicity against the HL-60 cell line [148]. Furthermore, a secosteroid called 6-O-deacetylseverinolide (154) was identified in the stem barks of Atalantia buxifolia extract [36].
The steroid inertogenin (155) is a compound containing a rare 7,15-tetrahydrofuran group. It is found in the leaves of Strophanthus amboensis, an erect deciduous shrub harvested for medical use in Southwest Africa [149]. The structure of inertogenin is depicted in Figure 20, and its biological activity is shown in Table 7. The leaves of Toona ciliata var. yunnanensis yielded several seco steroids: tooonayunnanins F (156), G (157), J (158), and K (159) [150].
Fritillaria pallidiflora bulbs are a source of unique jervinine-type alkaloids [151]. Peimissine (160), cycloparnine (161), and cycloposine (162) were isolated from the bulbs of F. pallidiflora, as well as the isosteroidal alkaloid yibeinone A (163) [152]. Puqienine F (164), a veratramine alkaloid with a 12,16-epoxy ring, was isolated from the bulbs of Fritillaria puqiensis [153]. A picture of Fritillaria pallidiflora is shown in Figure 21, and the 3D graph representing the activity of puqienine F (164) can be found in Figure 22.
From Buxus hyrcana, collected in Iran, two steroidal alkaloids were isolated: (+)-O6-buxafurandiene (165) and (+)-7-deoxy-O6-buxafurandiene (166). These compounds belong to the rare class of Buxus alkaloids with a tetrahydrofuran ring incorporated into their structures. Furthermore, they exhibited acetylcholinesterase enzyme inhibitory activity [154].
Steroidal alkaloids, solasodine (167), and tomatidine (168) were isolated from the aerial parts of Solanum leucocarpum, a plant belonging to the Solanaceae family. The collection was made at the regional natural park Ucumarí in Colombia. Both alkaloids have demonstrated biological activity. Solasodine exhibits DNA-damaging activity, while tomatidine displays activity through DNA topoisomerase II inhibition [155].
Two unique abeo-steroids, spirochensilides A (169) and B (170), were isolated from Abies chensiensis. These compounds represent the first example of triterpenoids with a distinctive 8,10-cyclo-9,10-seco and methyl-rearranged carbon skeleton [156]. Abies faxoniana, an endemic plant found in several provinces in China, serves as the source of lanostane and cycloartane derivatives A1/A2, 171174, which feature epimeric spiro-side chains [157].
Bungsteroid A (175), possessing an unreported carbon skeleton, was isolated from the pericarps of Zanthoxylum bungeanum. It represents a C34 steroid analogue with a unique 6/6/6/6/5-fused pentacyclic skeleton. This compound exhibited antiproliferative effects against HepG2, MCF-7, and HeLa cell lines, with IC50 values of 56.3, 64.2, and 74.2 μM, respectively [158].
The roots and stems of Cyathula officinalis, also known as Cyathula Root or Radix Cyathula, yielded two cyasterone stereoisomers: 28-epi-cyasterone (176) and 25-epi-28-epi-cyasterone (177) [159]. The biologically active plant steroid, 28-epi-cyasterone (176), has also been found in Eriophyton wallchii [160] and fronds of the fern Microsorum scolopendria [161].
Figure 20. Steroids, di-, and triterpenoids derived from plant species.
Figure 20. Steroids, di-, and triterpenoids derived from plant species.
Molecules 28 05669 g020
Figure 21. (a) The plant Fritillaria pallidiflora is a source of unusual steroids (160164). F. pallidiflora, also known as Siberian hazel grouse, is a species that was initially misnamed as it does not grow wild in Siberia. It was discovered in 1857 in the western regions of the Himalayas and Asia Minor. The plant is known for its medicinal uses, and an infusion of dried chopped onion from this plant is commonly used orally to treat cough, bronchitis, pneumonia, febrile illnesses, and abscesses [162,163]. Shaanxi fir (b) is a tree that grows in Gansu, Hubei, and Sichuan. Abies chensiensis, a species of Shaanxi fir, produces unusual triterpenoids (169 and 170). The essential oil of Shaanxi fir, as well as Siberian fir (Abies sibirica), has a pleasant, fresh pine aroma and contains bornyl acetate, which contributes to its soothing, balancing, and anti-inflammatory effects. The oil is used to relieve anger, promote contentment, alleviate intolerance in toxic relationships, promote self-connection, and foster a fearless attitude [164,165]. The pericarps of Zanthoxylum bungeanum (c) produce an unusual steroid (175). Extracts from Z. bungeanum fruit are widely used in the cosmetics industry to produce creams. Various parts of this plant, including the fruit, stems, leaves, and bark, have been utilized in local medical systems to treat fever, stomach pain, toothache, and inflammation [166,167].
Figure 21. (a) The plant Fritillaria pallidiflora is a source of unusual steroids (160164). F. pallidiflora, also known as Siberian hazel grouse, is a species that was initially misnamed as it does not grow wild in Siberia. It was discovered in 1857 in the western regions of the Himalayas and Asia Minor. The plant is known for its medicinal uses, and an infusion of dried chopped onion from this plant is commonly used orally to treat cough, bronchitis, pneumonia, febrile illnesses, and abscesses [162,163]. Shaanxi fir (b) is a tree that grows in Gansu, Hubei, and Sichuan. Abies chensiensis, a species of Shaanxi fir, produces unusual triterpenoids (169 and 170). The essential oil of Shaanxi fir, as well as Siberian fir (Abies sibirica), has a pleasant, fresh pine aroma and contains bornyl acetate, which contributes to its soothing, balancing, and anti-inflammatory effects. The oil is used to relieve anger, promote contentment, alleviate intolerance in toxic relationships, promote self-connection, and foster a fearless attitude [164,165]. The pericarps of Zanthoxylum bungeanum (c) produce an unusual steroid (175). Extracts from Z. bungeanum fruit are widely used in the cosmetics industry to produce creams. Various parts of this plant, including the fruit, stems, leaves, and bark, have been utilized in local medical systems to treat fever, stomach pain, toothache, and inflammation [166,167].
Molecules 28 05669 g021
Triterpenoids with medicinal properties have been identified from various plant sources. Abinukitrine A (178), a triterpenoid, was isolated from Abies nukiangensis extracts and exhibited a potent anti-hepatitis C virus (HCV) effect [62]. Malabaricol (179), another triterpenoid, was first reported by Indian chemists from the National Chemical Laboratory in 1967, isolated from the tropical tree Ailanthus malabarica [168]. Subsequently, malabaricol (179) and 180 were also found in the heartwood of Ailanthus excelsa [169]. Malabaricol and its derivatives [170] possess antibacterial effects, supporting their traditional use in folk medicine [171]. Ailanthus triphysa is a source of ailanthusins F (181) and G (182) [172], while two other derivatives, 183 and 184, were discovered in the leaves of Caloncoba echinata [173].
Furthermore, triterpene glycosides called taimordisins A (185) and B (186) were isolated from the fresh fruits of Taiwanese Momordica charantia [174]. Although these compounds demonstrate favorable anti-inflammatory activity, they do not exhibit any anti-cancer properties and have been determined to be safe for use.
Figure 22. Three-dimensional graph illustrating the predicted and calculated anti-tuberculosis activity of steroids (160, 161, 162, 163, and 164) with a confidence level exceeding 90%. These steroids include the unusual steroidal jervinine-type alkaloids (160163) derived from the bulbs of Fritillaria pallidiflora, as well as alkaloid (164) obtained from the bulbs of Fritillaria puqiensis. Notably, these compounds possess the rare property of exhibiting anti-tuberculosis activity. The genus Fritillaria has been widely used worldwide for medicinal and culinary purposes. For over 2000 years, decoctions made from the bulbs of various Fritillaria species have been utilized in traditional Chinese medicine to address diverse ailments such as asthma, pharyngitis, bronchitis, coughs, goiter, and hemoptysis. Additionally, it has been employed as an expectorant and antitussive agent [175,176].
Figure 22. Three-dimensional graph illustrating the predicted and calculated anti-tuberculosis activity of steroids (160, 161, 162, 163, and 164) with a confidence level exceeding 90%. These steroids include the unusual steroidal jervinine-type alkaloids (160163) derived from the bulbs of Fritillaria pallidiflora, as well as alkaloid (164) obtained from the bulbs of Fritillaria puqiensis. Notably, these compounds possess the rare property of exhibiting anti-tuberculosis activity. The genus Fritillaria has been widely used worldwide for medicinal and culinary purposes. For over 2000 years, decoctions made from the bulbs of various Fritillaria species have been utilized in traditional Chinese medicine to address diverse ailments such as asthma, pharyngitis, bronchitis, coughs, goiter, and hemoptysis. Additionally, it has been employed as an expectorant and antitussive agent [175,176].
Molecules 28 05669 g022

4. Furanosteroids Derived from Marine Sources

Furanosteroids, a fascinating group of natural compounds, are derived from marine sources and are produced by various marine invertebrates, including cyanobacteria, fungal endophytes associated with algae, sponges, soft corals, or molluscs [2,3,4,5,6,7,8].
Merosterol A (187), a cyanobacterial cytotoxin found in Scytonema sp. PCC 10023, has been exhibited against HeLa cells with IC50 values of 1.8 µM [177]. In an active organic extract obtained from an Okinawan marine sponge of the genus Dysidea (depicted in Figure 23), two polyoxygenated steroids, dysideasterols F (188) and G (189), were identified. These compounds demonstrated a similar cytotoxic effect, with IC50 values of 0.15 and 0.3 µM, respectively, against human epidermoid carcinoma A431 cells [178]. Additionally, a toxic polyoxygenated steroid named cholest-6-en-11β,19-epoxy-3β,5α,8α,9α-tetrol (190) has been isolated from the sponge Dysidea tupha [179]. Furthermore, a sponge species, Strongylophora sp., yielded furano-pregnanes, namely 3,4-dihydroxypregna-5,17-diene-10,2-carbolactone (191, whose structure is shown in Figure 24) and 3,4-dihydroxypregna-5,15-dien-20-one-10,2-carbolactone (192) [180].
Nakiterpiosin (193) and nakiterpiosinone (194, whose structures are shown in Figure 24), which are halogenated and rearranged norsteroids, were isolated from the Okinawan marine sponge Terpios hoshinota. These compounds have demonstrated cytotoxicity against murine P388 leukemia cells [181].
Unusual steroids known as erylosides T (195) and U (196), derived from the sea sponge Erylus goffrilleri, contain novel genins [182]. Lanostanes (197 and 198), whose structures are depicted in Figure 24 and whose biological activity is presented in Table 8, were obtained from the marine sponge Penares sp., found in Vietnamese waters [183]. These discoveries highlight the diverse range of bioactive compounds originating from marine sources.
A spiroketal steroid, 199, was obtained from a collection of Gorgonella umbraculum found in the Indian Ocean off the Tuticorin coast [184]. From a Japanese octocoral species, Dendronephthya sp., two secosteroids named isogosterones A (200) and B (201) were isolated. These compounds share common structures characterized as 12α-acetoxy-13,17-seco-cholesta-1,4-dien-3-ones with hemiacetal functionality. Isogosterones A and B exhibited inhibition of larval settlement in the barnacle Balanus amphitrite, with an EC50 value of 0.2 μM [185].
Furthermore, two steroids were isolated from the marine sponge Isis hippuris, also known as sea bamboo, collected from the Andaman Islands, India. These compounds are 3,11-diacetylhippurin-1 (202) and 22-epi-hippuri-stanol (203) [186]. Additionally, the marine sponge Lendenfeldia frondosa, collected from the Solomon Islands, yielded epihomoscalaralactone IIA (204) [187], while another marine sponge, Phyllospongia dendyi, from the Indian Ocean contained homoscalaralactone I1 B (205) [188]. Bioactive scalaranes (206208) have been identified in the bio-toxic extracts of the marine sponge Hyrtios erecta [189]. Phyllofolactone A (209), isolated from the marine sponge Phyllospongia (syn. Carteriospongia) foliascens found in the South China Sea, has demonstrated cytotoxicity against P-388 cells [190].
The marine sponge Hippospongia sp. collected from Taitung, Taiwan, serves as the source of a cytotoxic metabolite called hippospongide A (210) [191], as depicted in the 3D graph shown in Figure 25. Furthermore, the same compound, named salmahyrtisol A (211), has been isolated from the marine sponge Hyrtios erecta and exhibits significant cytotoxicity against murine leukemia (P-388), A-549, and HT-29 human cancer cells [192]. These findings underscore the potential of marine sponges as a valuable source of bioactive compounds with cytotoxic properties.
Steroidal anticancer alkaloids known as cortistatins A (212), B (213), C (214), and D (215) have been isolated from the marine sponge Corticium simplex. These alkaloids possess a unique 9 (10–19)-abeo-androstane and isoquinoline skeleton and have demonstrated the ability to inhibit the proliferation of human umbilical vein endothelial cells with high selectivity [193].
Metabolites containing the core structure of viridin have been extracted from marine invertebrates. Over 40 years ago, a research group led by Paul Scheuer at the University of Hawaii at Mānoa isolated demethoxyviridin (4) and its furano-quinone analogue called halenaquinone (216) from the marine sponge Xestospongia exigua [194]. Raspacionin A (217), a triterpenoid, was obtained from the red sponge Raspaciona aculeata found in the Mediterranean Sea. This compound exhibited cytotoxicity against the MCF-7 tumor cell line, with an IC50 value of 4 μM [195,196,197]. Highly polyoxygenated steroids (218221) containing three or four additional tetrahydrofuran fragments were discovered in the bamboo coral Isis hippuris, collected from the Southeast coast of Taiwan. These steroids displayed cytotoxic activity against Hep G2, Hep 3B, A549, MCF-7, and MDAMB-231 cells [198]. Additionally, two steroids (222 and 223), whose structures are depicted in Figure 26 and whose biological activity is presented in Table 9, were isolated from the soft coral Sarcophyton crassocaula found in the Indian Ocean [199]. Furthermore, the cytotoxic compound sinubrasone B (224) was identified in the reef soft coral Sinularia brassica [200].
Crellastatin A (225), a unique nonsymmetric dimeric steroid, was isolated from the marine sponge Crella sp. found on Vanuatu Island. This compound showcases an unprecedented connection through its side chains. Crellastatin A has been shown to possess significant in vitro cytotoxic activity against NSCLC-N6 cells, with an IC50 value of 0.5 μM [201]. This discovery underscores the potential of marine organisms to produce novel bioactive compounds with cytotoxic properties.
A similar dimeric steroid derivative, shishicrellastatin A (226), has been isolated from the marine sponge Crella (Yvesia) spinulata. This compound functions as a cathepsin B inhibitor, exhibiting an IC50 value of 8 μg/mL [202]. Another dimeric steroid, amaroxocane B (227), was discovered in the Caribbean coral reef sponge Phorbas amaranthus collected off Key Largo, Florida. It has shown effectiveness as an antifeedant [203]. Additionally, the same sterol dimer, hamigerol B (227), has been identified in the extract of the Mediterranean sponge Hamigera hamigera [204].
Furthermore, a sulfated sterol dimer named fibrosterol C (228), obtained from Lissodendoryx (Acanthodoryx) fibrosa collected in the Philippines, has been found to inhibit protein kinase Cζ with an IC50 value of 5.6 μM [205]. The 3D graph illustrating its activity is shown in Figure 27. Cephalostatin 20 (229), a member of the cephalostatin family known for its anticancer properties, has been isolated as a minor component of extracts from the marine worm Cephalodiscus gilchristi [206]. A picture of this sea sponge is displayed in Figure 28.
An undescribed marine sponge from the genus Euryspongia serves as the source of sulfated steroids, namely eurysterols A (230) and B (231). Eurysterol A exhibits cytotoxicity against human colon carcinoma (HCT-116) cells, with an IC50 value of 0.3 μM. It also demonstrates antifungal activity against amphotericin-B-resistant Candida albicans [207]. Two sesterterpenoids, oxaspirosuberitenone (232) and isooxaspirosuberitenone (233), have been isolated from the marine sponge Phorbas areolatus. These compounds exhibit significant growth-inhibitory effects against A549, HepG2, HT-29, and MCF-7 tumor cell lines [208].
Furthermore, 12-dehydroxy-16-deacetoxy-22-hydroxyscalarafuran (234) and its corresponding acetate, 12-dehydroxy-16-deacetoxy-22-acetoxy-scalarafuran (235), were identified in the sponge Smenospongia sp. from Soheuksan Island (Korea). These compounds display antimicrobial activity and strong cytotoxicity against the human chronic myelogenous leukemia K562 cell line [209]. Additionally, coscinafuran (236) was detected in the MeOH fraction of the sponge Coscinoderma mathewsi [210].
Sednolide (237) and sednolide 22-acetate (238) were identified in extracts of the nudibranch Chromodoris sedna, collected in Baja California (Mexico). A picture of this mollusc is displayed in Figure 29. Sednolide (237) demonstrated growth inhibition of the marine bacterium Vibrio anguillarum at a concentration of 100 μg/disk [211].
A rare and unusual aminosteroid called clionamine D (239), with its structure shown in Figure 30 and biological activity presented in Table 10, has been isolated from South African specimens of the sponge Cliona celata. This aminosteroid possesses a unique spiro bis-lactone side chain and exhibits cytotoxicity. It also modulates autophagy [212].
Isomalabaricane-type triterpenoids, namely globostelletins P (240) and Q (241), have been isolated from the marine sponge Rhabdastrella globostellata. These triterpenoids, along with the CH2Cl2 fraction of the sponge, demonstrate inhibitory activities against various human tumor cell lines, including A549 (human lung adenocarcinoma), BGC-823 (human gastric carcinoma), HCT-8 (colonic carcinoma), Bel-7402 (human liver carcinoma), and A2780 (human ovarian carcinoma) [213].
The cyanobacterium Scytonema sp. from Bermuda serves as the source of merosterol A (242) [177]. Halogenated polar steroids, including chlorine-containing steroid sulfate (243), iodinated steroid 244, and topsentiasterol sulfate D (245), have been isolated from the marine sponge Topsentia sp. [214,215]. Steroid (243) has been shown to effectively inhibit endo-1,3-β-D-glucanase from the marine mollusc Spisula sachalinensis [214].
A series of steroids known as sinubrasolides A–F (246251), which belong to the class of withanolide-type steroids, were isolated from cultured specimens of Sinularia brassica from Taiwan [216]. Petrosaspongiolides A (252) and B (253) were the first cheilantane sesterterpene lactones to be isolated from a New Caledonian sponge initially assigned to the genus Dactylospongia but later reclassified as a new genus and species called Petrosaspongia nigra [217,218]. Additionally, other petrosasponiolides, namely M (254) and R (255), isolated from the New Caledonian marine sponge Petrosaspongia nigra, exhibited a γ-hydroxybutenolide moiety and a hemiacetal function [219,220]. The 3D graph illustrating the activity of compound 255 is shown in Figure 31.
Cytotoxic steroids named kiheisterones A (256) and B (257) have been isolated from a sponge of the order Poecilosclerida collected along the coast of the island of Maui (Hawaii). These sterols possess an α,β-disubstituted furan in the sidechain, a cis-fused A/B ring, a monoenolized α-di-ketone in the A ring, and a C-21 carboxyl group. Both steroids exhibit mild cytotoxicity against several human tumor cell lines, including A-549 lung carcinoma, HT-29 colon adenocarcinoma, and the P-388 murine lymphocytic leukemia cell line [220]. Furthermore, cytotoxic sesterterpenoids named inorolide A (258), B (259), and C (260) have been isolated from the Japanese nudibranch Chromodoris inornata (Chromodorididae) [221].
A unique 5β-steroid disulfate 261 with a distinctive 4α,9α-ether bridge has been isolated from the ophiuroid species Ophiomastix annulosa [222]. Furthermore, the marine sponge Strepsichordaia aliena from Indonesia has been found to contain the 20,24-bishomoscalarane sesterterpenes honulactones A (262) and B (263). Both sesterterpenes exhibit cytotoxicity against P-388, A-549, HT-29, and MEL-28 cancer cell lines, with an IC50 value of 0.1 μM [223].
A sarcosterol derivative, 22,25-epoxy-23,24-dimethylcholesta-5,17(20)-dien-3-ol (264), has been discovered in the soft coral Sinularia mayi [224]. Additionally, the soft coral Lobophytum depressum contains two steroids: lobophytosterol (265), which is a 22,28-epoxyergost-5-ene-3,25,28-triol, and 5,6-epoxy-lobophytosterol (266). A picture of this soft coral is displayed in Figure 32 [225,226]. Both steroids possess a double bond at positions 5,6. Lobophytosterol from the soft coral Lobophytum laevigatum has demonstrated cytotoxicity against A549 and HL-60 cell lines, with IC50 values of 4.5 and 5.6 μM, respectively [227].

5. Conclusions

This comprehensive review examines the occurrence of furanosteroids and related isoprenoid lipids in various sources, including fungi, plants, and marine organisms. It provides an in-depth analysis of their biological activity, shedding light on their potential applications and significance in drug discovery. Furanosteroids and related isoprenoid lipids have garnered significant attention due to their diverse biological properties. This review aims to explore the occurrence of these compounds in fungi, plants, and marine organisms, as well as delve into their associated biological activity. By analyzing their medicinal and pharmacological potential, this review highlights the importance of furanosteroids as promising natural products for future research and development. The biological activity of furanosteroids and related isoprenoid lipids constitutes a crucial aspect of their investigation. This review focuses on the diverse range of activities exhibited by these compounds, including but not limited to anti-inflammatory, anticancer, and antimicrobial properties.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares that he has no known competing financial interests or personal relationships that could affect the work described in this article.

References

  1. Peterson, L.A. Reactive metabolites in the biotransformation of molecules containing a furan ring. Chem. Res. Toxicol. 2013, 26, 6–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Saeid, H.; Al Sayed, H.; Bader, M. A review on biological and medicinal significance of furan. Alqalam J. Med. App. Sci. 2023, 6, 44–58. [Google Scholar]
  3. Tian, M.; Peng, Y.; Zheng, J. Metabolic activation, and hepatotoxicity of furan-containing compounds. Drug Metabol. Disposit. 2022, 50, 655–670. [Google Scholar] [CrossRef] [PubMed]
  4. Banerjee, R.; Kumar, H.K.S.; Banerjee, M. Medicinal significance of furan derivatives: A Review. Int. J. Rev. Life Sci. 2012, 2, 7–16. [Google Scholar]
  5. Alizadeh, M.; Jalal, M.; Hamed, K.; Saber, A. Recent updates on anti-inflammatory and antimicrobial effects of furan natural derivatives. J. Inflammat. Res. 2020, 13, 451–463. [Google Scholar] [CrossRef]
  6. Mawlong, I.; Sujith Kumar, M.S.; Singh, D. Furan fatty acids: Their role in plant systems. Phytochem. Rev. 2016, 15, 121–127. [Google Scholar] [CrossRef]
  7. Buckingham, J. Dictionary of Natural Products; CRC Press: Boca Raton, FL, USA, 2022. [Google Scholar]
  8. Senapati, B.K. Recent progress in the synthesis of the furanosteroid family of natural products. Org. Chem. Front. 2021, 8, 2608–2642. [Google Scholar] [CrossRef]
  9. Xue, D.; He, H.; Gao, S. Strategies for the total synthesis of the furanosteroids: Wortmannin and viridin. Chem. Lett. 2021, 50, 497–502. [Google Scholar] [CrossRef]
  10. Hu, D.; Gao, Y.H.; Yao, X.S.; Gao, H. Recent advances in dissecting the demethylation reactions in natural product biosynthesis. Curr. Opin. Chem. Biol. 2020, 59, 47–53. [Google Scholar] [CrossRef]
  11. Zhungietu, G.I.; Dorofeenko, G.N. Progress in the field of the chemistry of steroidal heterocycles. Russ. Chem. Rev. 1967, 36, 24–36. [Google Scholar] [CrossRef]
  12. Hatti, K.S.; Muralitharan, R.L.; Kush, H.; Kush, A. Convenient database for neem secondary metabolites. Bioinformation 2014, 10, 314–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Ui, M.; Okada, T.; Hazeki, K.; Hazeki, O. Wortmannin as a unique probe for an intracellular signalling protein, phosphoinositide 3-kinase. Trends Biochem. Sci. 1995, 20, 303–307. [Google Scholar] [CrossRef] [PubMed]
  14. Hanson, J.R. The viridin family of steroidal antibiotics. Nat. Prod. Rep. 1995, 12, 381–384. [Google Scholar] [CrossRef] [PubMed]
  15. Brian, P.W.; Curtis, P.J.; Hemming, H.G. The production of viridin by pigment-forming strains of Trichoderma viride. Ann. Appl. Biol. 1946, 33, 190–200. [Google Scholar] [CrossRef]
  16. Brian, P.W.; McGowan, J.C. Viridin: A highly fungistatic substance produced by Trichoderma viride. Nature 1945, 156, 144–145. [Google Scholar] [CrossRef]
  17. Brian, P.W.; Curtis, P.J.; Hemming, H.G.; Norris, G.L.F. Wortmannin, an antibiotic produced by Penicillium wortmanni. Transact. Br. Mycol. Soc. 1957, 40, 365–368. [Google Scholar] [CrossRef]
  18. Viswanathan, K.; Ononye, S.N.; Cooper, H.D.; Kyle Hadden, M.; Anderson, A.C.; Wright, D.L. Viridin analogs derived from steroidal building blocks. Bioorg. Med. Chem. Lett. 2012, 22, 6919–6922. [Google Scholar] [CrossRef]
  19. Jones, R.W.; Hancock, J.G. Conversion of viridin to viridiol by viridin-producing fungi. Can. J. Microbiol. 1987, 33, 963–966. [Google Scholar] [CrossRef]
  20. Aldridge, D.C.; Turner, W.B.; Geddes, A.J.; Sheldrick, B. Demethoxyviridin and demethoxyviridiol: New fungal metabolites. J. Chem. Soc. Perkin Trans. 1 1975, 10, 943–945. [Google Scholar] [CrossRef]
  21. Andersson, P.F.; Bengtsson, S.; Cleary, M.; Stenlid, J.; Broberg, A. Viridin-like steroids from Hymenoscyphus pseudoalbidus. Phytochemistry 2013, 86, 195–200. [Google Scholar] [CrossRef]
  22. Zhai, M.M.; Li, J.; Jiang, C.X.; Shi, J.Y.P.; Di, D.L.; Crews, P. The bioactive secondary metabolites from Talaromyces species. Nat. Prod. Bioprospect. 2016, 6, 1. [Google Scholar] [CrossRef] [Green Version]
  23. Noinart, J. Bioactive Secondary Metabolites from the Culture of the Marine Sponge-Associated Fungus Talaromyces stipitatus KUFA. Ph.D Thesis, University of Porto, Porto, Portugal, 2017. [Google Scholar]
  24. Ding, H.-E.; Yang, Z.-D.; Sheng, L.; Zhou, S.-Y.; Li, S.; Yao, X.-J.; Zhi, K.-K.; Wang, Y.-G.; Zhang, F. Secovironolide, a novel furanosteroid scaffold with a fivemembered B ring from the endophytic fungus Talaromyces wortmannii LGT-4. Tetrahedron Lett. 2015, 56, 6754–6757. [Google Scholar] [CrossRef]
  25. Phuwapraisirisan, P.; Rangsan, J.; Siripong, P.; Tip-Pyang, S. 9- epi -Viridiol, a novel cytotoxic furanosteroid from soil fungus Trichoderma virens. Nat. Prod. Res. 2006, 20, 14. [Google Scholar] [CrossRef] [PubMed]
  26. Abbas, H.K.; Mirocha, C.J. Isolation and purification of a hemorrhagic factor (wortmannin) from Fusarium oxysporum (N17B). Appl. Environ. Microbiol. 1988, 54, 1268–1274. [Google Scholar] [CrossRef] [Green Version]
  27. Singh, V.; Praveen, V.; Tripathi, D.; Haque, S.; Somvanshi, P. Isolation, characterization and antifungal docking studies of wortmannin isolated from Penicillium radicum. Sci. Rep. 2015, 5, 11948. [Google Scholar] [CrossRef] [Green Version]
  28. Wipf, P.; Halter, R.J. Chemistry and biology of wortmannin. Org. Biomol. Chem. 2005, 3, 2053–2061. [Google Scholar] [CrossRef] [PubMed]
  29. Smith, R.A.; Yuan, H.; Weissleder, R.; Cantley, L.C.; Josephson, L. A wortmannin-cetuximab as a double drug. Bioconjug. Chem. 2009, 20, 2185–2189. [Google Scholar] [CrossRef]
  30. Kong, D.; Yamori, T. Phosphatidylinositol 3-kinase inhibitors: Promising drug candidates for cancer therapy. Cancer Sci. 2008, 99, 1734–1740. [Google Scholar] [CrossRef] [PubMed]
  31. Morales-Oyervidesa, L.; Ruiz-Sánchez, J.P.; Oliveira, J.C. Biotechnological approaches for the production of natural colorants by Talaromyces/Penicillium: A review. Biotechnol. Adv. 2020, 43, 107601. [Google Scholar] [CrossRef] [PubMed]
  32. Nagwa, E.; Kassem, A.H.; Hamed, M.A.; El-Fekya, A.M.; Elnaggard, M.A.A.; Mahmouda, K.; Alie, M.A. Isolation and characterization of the bioactive metabolites from the soil derived fungus Trichoderma viride. Mycology 2018, 9, 70–80. [Google Scholar]
  33. Ghisalberti, E.L.; Sivasithamparam, K. Antifungal antibiotics produced by Trichoderma spp. Soil Biol. Biochem. 1991, 23, 1011–1020. [Google Scholar] [CrossRef]
  34. Mukherjee, P.K.; Horwitz, B.A.; Kenerley, C.M. Secondary metabolism in Trichoderma—A genomic perspective. Microbiology 2012, 158, 35–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Song, Y.-P.; Shi, Z.-Z.; Miao, F.-P.; Fang, S.-T.; Yin, X.-L.; Ji, N.-Y. Tricholumin A, a highly transformed ergosterol derivative from the alga endophytic fungus Trichoderma asperellum. Org. Lett. 2018, 20, 6306–6309. [Google Scholar] [CrossRef] [PubMed]
  36. Hill, R.A.; Makin, H.L.J.; Kirk, D.N.; Murphy, G.M. Dictionary of Steroids, 1st ed.; Chapman and Hall: Boca Raton, FL, USA, 1991. [Google Scholar]
  37. Chen, H.P.; Zhao, Z.Z.; Li, Z.H.; Huang, Y.; Zhang, S.B.; Tang, Y.; Yao, J.N.; Chen, L.; Isaka, M.; Feng, T.; et al. Anti-proliferative and anti-Inflammatory lanostane triterpenoids from the Polish edible mushroom Macrolepiota procera. J. Agric. Food Chem. 2018, 66, 3146–3154. [Google Scholar] [CrossRef]
  38. Guzmán-Guzmán, P.; Kumar, A.; de los Santos-Villalobos, S.; Parra-Cota, F.I.; del Orozco-Mosqueda, C.; Fadiji, A.E.; Hyder, S.; Babalola, O.O.; Santoyo, G. Trichoderma species: Our best fungal allies in the biocontrol of plant diseases—A Review. Plants 2023, 12, 432. [Google Scholar] [CrossRef]
  39. Sood, M.; Kapoor, D.; Kumar, V.; Sheteiwy, M.S.; Ramakrishnan, M.; Landi, M.; Araniti, F.; Sharma, A. Trichoderma: The “Secrets” of a multitalented biocontrol agent. Plants 2020, 9, 762. [Google Scholar] [CrossRef]
  40. Gomdola, D.; Bhunjun, C.S.; Hyde, K.D.; Jeewon, R.; Pem, D.; Jayawardena, R.S. Ten important forest fungal pathogens: A review on their emergence and biology. Mycosphere 2022, 13, 612–671. [Google Scholar] [CrossRef]
  41. Lorenc, F.; Samek, M. Pathogens threatening Czech Republic Forest ecosystems—A review. Sylwan 2021, 165, 853–871. [Google Scholar]
  42. Benjamin, C.R. Ascocarps of Aspergillus and Penicillium. Mycologia 1955, 47, 669–687. [Google Scholar] [CrossRef]
  43. Frisvad, J.C. Taxonomy, chemodiversity, and chemoconsistency of Aspergillus, Penicillium, and Talaromyces species. Front. Microbiol. 2015, 5, 773. [Google Scholar] [CrossRef] [Green Version]
  44. Houbraken, J.; de Vries, R.P.; Samson, R.A. Modern taxonomy of biotechnologically important Aspergillus and Penicillium species. Adv. Appl. Microbiol. 2014, 86, 199–249. [Google Scholar] [PubMed]
  45. Ramani, G.; Meera, B.; Vanitha, C.; Rao, M. Production, purification, and characterization of a β-glucosidase of Penicillium funiculosum NCL1. Appl. Biochem. Biotechnol. 2012, 167, 959–972. [Google Scholar] [CrossRef] [PubMed]
  46. Fravel, D.; Olivain, C.; Alabouvette, C. Fusarium oxysporum and its biocontrol. New Phytol. 2003, 157, 493–502. [Google Scholar] [CrossRef]
  47. Gordon, T.R.; Martyn, R.D. The evolutionary biology of Fusarium oxysporum. Annu. Rev. Phytopathol. 1997, 35, 111–128. [Google Scholar] [CrossRef] [Green Version]
  48. Ekiz, G.; Yılmaz, S.; Yusufoglu, H.; Kırmızıbayrak, P.B.; Bedir, E. Microbial transformation of cycloastragenol and astragenol by endophytic fungi isolated from Astragalus species. J. Nat. Prod. 2019, 82, 2979–2985. [Google Scholar] [CrossRef]
  49. Küçüksolak, M.; Üner, G.; Kırmızıbayrak, P.B.; Bedir, E. Neuroprotective metabolites via fungal biotransformation of a novel sapogenin, cyclocephagenol. Sci. Rep. 2022, 12, 18481. [Google Scholar] [CrossRef] [PubMed]
  50. Galappaththi, M.C.A.; Patabendige, N.M.; Premarathne, B.M.; Hapuarachchi, K.K.; Tibpromma, S.; Dai, D.-Q.; Suwannarach, N.; Rapior, S.; Karunarathna, S.C. A Review of Ganoderma triterpenoids and their bioactivities. Biomolecules 2023, 13, 24. [Google Scholar] [CrossRef] [PubMed]
  51. Su, H.-G.; Wang, Q.; Zhou, L.; Peng, X.-R.; Xiong, W.-Y.; Qiu, M.-H. Functional triterpenoids from medicinal fungi Ganoderma applanatum: A continuous search for antiadipogenic agents. Bioorg. Chem. 2021, 112, 104977. [Google Scholar] [CrossRef] [PubMed]
  52. Li, L.; Peng, X.R.; Dong, J.R.; Lu, S.Y.; Li, X.N.; Zhou, L. Rearranged lanostane-type triterpenoids with anti-hepatic fibrosis activities from Ganoderma applanatum. RSC Adv. 2018, 8, 31287–31295. [Google Scholar] [CrossRef]
  53. Peng, X.-R.; Liu, J.-Q.; Wang, C.-F.; Li, X.-Y.; Shu, Y.; Zhou, L.; Qiu, M.-H. Hepatoprotective effects of triterpenoids from Ganoderma cochlear. J. Nat. Prod. 2014, 77, 737–743. [Google Scholar] [CrossRef]
  54. Niu, X.; Qiu, M.; Li, Z.; Lu, Y.; Cao, P.; Zheng, Q. Two novel 3,4-seco-trinorlanostane triterpenoids isolated from Ganoderma fornicatum. Tetrahedron Lett. 2004, 45, 2989–2993. [Google Scholar] [CrossRef]
  55. Hill, R.A.; Connolly, J.D. Triterpenoids. Nat. Prod. Rep. 2020, 37, 962. [Google Scholar] [CrossRef] [Green Version]
  56. Ma, K.; Li, L.; Bao, L.; He, L.; Sun, C.; Zhou, B.; Si, S.; Liu, H. Six new 3, 4-seco-27-norlanostane triterpenes from the medicinal mushroom Ganoderma boninense and their antiplasmodial activity and agonistic activity to LXRβ. Tetrahedron 2015, 71, 1808–1814. [Google Scholar] [CrossRef]
  57. Zhang, S.-S.; Wang, Y.-G.; Ma, Q.-Y.; Huang, S.-Z.; Hu, L.-L.; Dai, H.-F.; Yu, Z.-F.; Zhao, Y.-X. Three new lanostanoids from the mushroom Ganoderma tropicum. Molecules 2015, 20, 3281–3289. [Google Scholar] [CrossRef] [Green Version]
  58. Peng, X.; Liu, J.; Xia, J.; Wang, C.; Li, X.; Deng, Y.; Bao, N.; Zhang, Z.; Qiu, M. Lanostane triterpenoids from Ganoderma hainanense. Phytochemistry 2015, 114, 137–145. [Google Scholar] [CrossRef]
  59. Li, Q.; Chen, C.; Cheng, L. Emeridones A–F, a series of 3,5-demethylorsellinic acid-based meroterpenoids with rearranged skeletons from an endophytic fungus Emericella sp. TJ29. J. Org. Chem. 2019, 84, 1534–1541. [Google Scholar] [CrossRef]
  60. Liang, X.-R.; Miao, F.-P.; Song, Y.-P.; Guo, Z.-Y.; Ji, N.-Y. Trichocitrin, a new fusicoccane diterpene from the marine brown alga-endophytic fungus Trichoderma citrinoviride cf-27. Nat. Prod. Res. 2016, 30, 1605–1610. [Google Scholar] [CrossRef]
  61. Liu, H.B.; Edrada-Ebel, R.A.; Ebel, R.; Wang, Y.; Schulz, B.; Draeger, S. Ophiobolin sesterterpenoids and pyrrolidine alkaloids from the sponge derived fungus Aspergillus ustus. Helvetica Chim. Acta 2011, 94, 623–631. [Google Scholar] [CrossRef]
  62. Li, Y.L.; Xue, L.J.; Li, J.; Chen, L.M.; Wu, J.J.; Xu, Z.N.; Chen, Y.; Tiana, Y.P.; Yang, X.W. Abinukitrine A, a unique 17,18-cyclolanostane triterpenoid from Abies nukiangensis. Org. Biomol. Chem. 2019, 17, 2107–2109. [Google Scholar]
  63. Sugawara, F.; Strobel, G.; Strange, R.N.; Siedow, J.N.; Van Duyne, G.D.; Clardy, J. Phytotoxins from the pathogenic fungi Drechslera maydis and Drechslera sorghicola. Proc. Natl. Acad. Sci. USA 1987, 84, 3081–3085. [Google Scholar] [CrossRef]
  64. Li, E.; Clark, A.M.; Rotella, D.P.; Hufford, C.D. Microbial metabolites of ophiobolin A and antimicrobial evaluation of ophiobolins. J. Nat. Prod. 1995, 58, 74–81. [Google Scholar] [CrossRef]
  65. Ebel, R. Terpenes from marine-derived fungi. Mar. Drugs 2010, 8, 2340–2368. [Google Scholar] [CrossRef] [Green Version]
  66. Elissawy, A.M.; El-Shazly, M.; Ebada, S.S.; Singab, A.B.; Proksch, P. Bioactive terpenes from marine-derived fungi. Mar. Drugs 2015, 13, 1966–1992. [Google Scholar] [CrossRef] [Green Version]
  67. Zhang, M.; Liu, J.; Chen, R.; Zhao, J.; Xie, K.; Chen, D.; Feng, K.; Dai, J. Two furanharzianones with 4/7/5/6/5 ring system from microbial transformation of harzianone. Org. Lett. 2017, 19, 1168–1171. [Google Scholar] [CrossRef]
  68. Zhang, M.; Liu, J.; Chen, R.; Zhao, J.; Xie, K.; Chen, D.; Feng, K.; Dai, J. Microbial oxidation of harzianone by Bacillus sp. IMM-006. Tetrahedron 2017, 73, 7195–7199. [Google Scholar] [CrossRef]
  69. Dong, H.; Zhou, C.; Li, X.; Gu, H.; Zhang, H.E.; Zhou, F.; Zhao, Z.; Fan, T. Ultraperformance liquid chromatography-quadrupole time-of-flight mass spectrometry based untargeted metabolomics to reveal the characteristics of Dictyophora rubrovolvata from different drying methods. Front. Nutr. 2022, 9, 1056598. [Google Scholar] [CrossRef]
  70. Sato, M.; Kakisawa, H. Structures of three new C16 terpenoids from an Acrostalagmus fungus. J. Chem. Soc. Perkin Trans. I 1976, 22, 2407–2413. [Google Scholar] [CrossRef]
  71. Kakisawa, H.; Sato, M.; Ruo, T.I.; Hayashi, T. Biosynthesis of a C16-terpenoid lactone, a plant growth regulator. J. Chem. Soc. Chem. Commun. 1973, 9, 802–803. [Google Scholar] [CrossRef]
  72. Xia, X.; Zhang, J.; Zhang, Y.; Wei, F.; Liu, X. Pimarane diterpenes from the fungus Epicoccum sp. HS-1 associated with Apostichopus japonicus. Bioorg. Med. Chem. Lett. 2012, 22, 3017–3019. [Google Scholar] [CrossRef]
  73. Li, Y.; Lu, C.; Hu, Z.; Huang, Y.; Shen, Y. Secondary metabolites of Tubercularia sp. TF5, an endophytic fungal strain of Taxus mairei. Nat. Prod. Res. 2009, 23, 70–76. [Google Scholar]
  74. Shiono, Y.; Motoki, S.; Koseki, T.; Murayama, T.; Tojima, M.; Kimura, K.I. Isopimarane diterpene glycosides, apoptosis inducers, obtained from fruiting bodies of the ascomycete Xylaria polymorpha. Phytochemistry 2009, 70, 935–939. [Google Scholar] [CrossRef]
  75. Li, Y.Y.; Hu, Z.Y.; Lu, C.H.; Shen, Y.M. Four new terpenoids from Xylaria sp. 101. Helv. Chim. Acta 2010, 93, 796–802. [Google Scholar] [CrossRef]
  76. Gu, B.B.; Wu, W.; Jiao, F.R.; Jiao, W.J.; Li, L.; Sun, F.; Wang, S.P.; Yang, F.; Lin, H.W. Asperflotone, an 8(14→15)-abeo-ergostane from the sponge derived fungus Aspergillus flocculosus 16D-1. J. Org. Chem. 2019, 84, 300–306. [Google Scholar] [CrossRef]
  77. Gu, B.-B.; Wu, W.; Jiao, F.-R.; Jiao, W.-h.; Li, L.; Sun, F.; Wang, S.-P.; Yang, F.; Lin, H.-W. Aspersecosteroids A and B, two 11(9→10)-abeo-5,10-secosteroids with a dioxatetraheterocyclic ring system from Aspergillus flocculosus 16D-1. Org. Lett. 2018, 20, 7957–7960. [Google Scholar] [CrossRef]
  78. Dong, G.A.N.; Li, C.; Shu, Y.; Wu, J. Steroids and dihydroisocoumarin glycosides from Xylaria sp. by the one strain many compounds strategy and their bioactivities. Chin. J. Nat. Med. 2023, 21, 154–160. [Google Scholar]
  79. Liangsakul, J.; Srisurichan, S.; Pornpakakul, S. Anthraquinone–steroids, evanthrasterol A and B, and a meroterpenoid, emericellic acid, from endophytic fungus, Emericella variecolor. Steroids 2016, 106, 78–85. [Google Scholar] [CrossRef]
  80. Wu, J.; Suzuki, T.; Choi, J.H.; Yasuda, N.; Noguchi, K.; Hirai, H.; Kawagishi, H. An unusual sterol from the mushroom Stropharia rugosoannulata. Tetrahedron Lett. 2013, 54, 4900–4902. [Google Scholar] [CrossRef] [Green Version]
  81. Lee, S.; Kim, C.S.; Yu, J.S.; Kang, H.; Yoo, M.J.; Youn, U.J.; Ryoo, R.; Bae, H.K.; Kim, K.H. Ergopyrone, a styrylpyrone-fused steroid with a hexacyclic 6/5/6/6/6/5 skeleton from a mushroom Gymnopilus orientispectabilis. Org. Lett. 2021, 23, 3315–3319. [Google Scholar] [CrossRef]
  82. Kuo, P.C.; Kuo, T.H.; Damu, A.G.; Su, C.R.; Lee, E.J.; Wu, T.S.; Shu, R.; Chen, C.M.; Kenneth, F.; Bastow, T.-H.; et al. Physanolide A, a novel skeleton steroid, and other cytotoxic principles from Physalis angulate. Org. Lett. 2006, 8, 2953–2956. [Google Scholar] [CrossRef]
  83. Sun, C.P.; Qiu, C.Y.; Zhao, F. Physalins V-IX, 16,24-cyclo-13,14-seco withanolides from Physalis angulata and their antiproliferative and anti-inflammatory activities. Sci. Rep. 2017, 7, 4057. [Google Scholar] [CrossRef] [Green Version]
  84. Arsenault, G.P.; Biemann, K.; Barksdale, A.W.; McMorris, T.C. The structure of antheridiol, a sex hormone in Achyla bisexualis. J. Am. Chem. Soc. 1968, 90, 5635–5636. [Google Scholar] [CrossRef]
  85. Maldonado, E.; Hurtado, N.E.; Pérez-Castorena, A.L.; Martínez, M. Cytotoxic 20,24-epoxy-withanolides from Physalis angulata. Steroids 2015, 104, 72–78. [Google Scholar] [CrossRef] [PubMed]
  86. Meng, Q.; Fan, J.; Liu, Z.; Li, X.; Zhang, F.; Zhang, Y.; Sun, Y.; Li, L.; Xia Liu, X.; Hua, E. Cytotoxic withanolides from the whole herb of Physalis angulata L. Molecules 2019, 24, 1608. [Google Scholar] [CrossRef] [Green Version]
  87. Sun, C.P.; Qiu, C.Y.; Yuan, T.; Nie, X.F.; Sun, H.X.; Zhang, Q.; Li, H.X.; Ding, L.Q.; Zhao, F.; Chen, L.X.; et al. Antiproliferative and anti-inflammatory withanolides from Physalis angulata. J. Nat. Prod. 2016, 79, 1586–1597. [Google Scholar] [CrossRef]
  88. Hirotani, M.; Sai, K.; Hirotani, S.; Yoshikawa, T. Blazeispirols B, C, E and F, des-A-ergostane-type compounds, from the cultured mycelia of the fungus Agaricus blazei. Phytochemistry 2002, 59, 571–577. [Google Scholar] [CrossRef] [PubMed]
  89. Hirotani, M.; Hirotani, S.; Takayanagi, H.; Komiyama, K.; Yoshikawa, T. Agariblazeispirols A and B, an unprecedented skeleton from the cultured mycelia of the fungus, Agaricus blazei. Tetrahedron Lett. 2003, 44, 7975–7979. [Google Scholar] [CrossRef]
  90. Hirotani, M.; Hirotani, S.; Takayanagi, H.; Yoshikawa, T. Blazeispirol A, an unprecedented skeleton from the cultured mycelia of the fungus Agaricus blazei. Tetrahedron Lett. 1999, 40, 329–332. [Google Scholar] [CrossRef]
  91. Tang, Y.; Zhao, Z.Z.; Hu, K.; Feng, T.; Li, Z.H.; Chen, H.P.; Liu, J.K. Irpexolidal Represents a class of triterpenoid from the fruiting bodies of the medicinal fungus Irpex lacteus. J. Org. Chem. 2019, 84, 1845–1852. [Google Scholar] [CrossRef]
  92. Clericuzio, M.; Tabasso, S.; Bianco, M.A.; Pratesi, G.; Beretta, G.; Tinelli, S.; Zunino, F.; Vidari, G. Cucurbitane triterpenes from the fruiting bodies and cultivated mycelia of Leucopaxillus gentianeus. J. Nat. Prod. 2006, 69, 1796–1799. [Google Scholar] [CrossRef]
  93. Cheng, X.; Liang, X.; Yao, F.H.; Liu, X.B.; Qi, S.H. Fusidane-type antibiotics from the marine-derived fungus Simplicillium sp. SCSIO 41513. J. Nat. Prod. 2021, 84, 2945–2952. [Google Scholar] [CrossRef]
  94. Macías-Rubalcava, M.L.; Sánchez-Fernández, R.E. Secondary metabolites of endophytic Xylaria species with potential applications in medicine and agriculture. World J. Microbiol. Biotechnol. 2017, 33, 15. [Google Scholar] [CrossRef] [PubMed]
  95. Huang, K.; El-Seedi, H.R.; Xu, B. Critical review on chemical compositions and health-promoting effects of mushroom Agaricus blazei Murill. Curr. Res. Food Sci. 2022, 5, 2190–2203. [Google Scholar] [CrossRef] [PubMed]
  96. Bertollo, A.G.; Mingoti, M.E.D.; Plissari, M.E.; Betti, G.; Junior, W.A.R.; Luzardo, A.R.; Ignácio, Z.M. Agaricus blazei Murrill mushroom: A review on the prevention and treatment of cancer. Pharmacol. Res. Mod. Chin. Med. 2022, 2, 100032. [Google Scholar] [CrossRef]
  97. Dong, X.M.; Song, X.H.; Liu, K.B.; Dong, C.H. Prospect and current research status of medicinal fungus Irpex lacteus. Mycosystema 2017, 36, 28–34. [Google Scholar]
  98. Mezule, L.; Civzele, A. Bioprospecting white-rot basidiomycete Irpex lacteus for improved extraction of lignocellulose-degrading enzymes and their further application. J. Fungi 2020, 6, 256. [Google Scholar] [CrossRef]
  99. Pachapurkar, R.V.; Kornule, P.M.; Narayana, C.R. A new hexacyclic tetranortriterpenoid. Chem. Lett. 1974, 3, 357–358. [Google Scholar] [CrossRef] [Green Version]
  100. Rogers, L.L.; Zeng, L.; Kozlowski, J.F.; Shimada, H.; Alali, F.Q.; Johnson, H.A.; McLaughlin, J.L. New bioactive triterpenoids from Melia volkensii. J. Nat. Prod. 1998, 61, 64. [Google Scholar] [CrossRef]
  101. Connolly, J.D.; Labbé, C.; Rycroft, D.S.; Taylor, D.A.H. Tetranortriterpenoids and related compounds. Part 22. New apotirucailol derivatives and tetranortriterpenoids from the wood and seeds of Chisocheton paniculatus (Meliaceae). J. Chem. Soc. Perkin Trans. 1979, 21, 2959–2964. [Google Scholar] [CrossRef]
  102. Kraus, W.; Cramer, R.; Sawitzki, G. Tetranortriterpenoids from the seeds of Azadirachta indica. Phytochemistry 1981, 20, 117–120. [Google Scholar] [CrossRef]
  103. Mulholland, D.A.; Monkhe, T.V.; Coombes, P.H.; Rajab, M.S. Limonoids from Turraea holstii and Turraea floribunda. Phytochemistry 1998, 49, 2585. [Google Scholar] [CrossRef]
  104. Kumar, C.S.S.R.; Srinivas, M.; Yakkundi, S. Limonoids from the seeds of Azadirachta indica. Phytochemistry 1996, 43, 451. [Google Scholar] [CrossRef]
  105. Coombesa, P.H.; Mulhollanda, D.A.; Randrianarivelojosia, M. Vilasinin limonoids from Malleastrum antsingyense J.F. Leroy (Meliodeae: Meliaceae). Biochem. Syst. Ecol. 2008, 36, 74–76. [Google Scholar] [CrossRef]
  106. Kowa, T.K.; Jansen, O.; Ledoux, A.; Mamede, L.; Wabo, H.K.; Tchinda, A.T.; Genta-Jouve, G.; Frédérich, M. Bioassay-guided isolation of vilasinin–type limonoids and phenyl alkene from the leaves of Trichilia gilgiana and their antiplasmodial activities. Nat. Prod. Res. 2022, 36, 5039–5047. [Google Scholar] [CrossRef] [PubMed]
  107. Yadav, R.; Pednekar, A.; Avalaskar, A.; Rathi, M. A comprehensive review on Meliaceae family. World J. Pharm. Sci. 2015, 3, 1572–1577. [Google Scholar]
  108. Islas, J.F.; Acosta, E.; Buentello, Z.G.; Delgado-Gallegos, J.L.; Moreno-Treviño, M.G.; Escalante, B.; Moreno-Cueva, J.E. An overview of Neem (Azadirachta indica) and its potential impact on health. J. Function. Foods 2020, 74, 104171. [Google Scholar] [CrossRef]
  109. Ma, C.; Chen, Y.; Chen, J.; Li, X.; Chen, Y. A Review on Annona squamosa L.: Phytochemicals and biological activities. Am. J. Chin. Med. 2017, 45, 1–32. [Google Scholar] [CrossRef] [PubMed]
  110. Ragasa, C.Y.; Nacpil, Z.D.; Natividad, G.M.; Tada, M.; Coll, J.C.; Rideout, J.A. Tetranortriterpenoids from Azadirachta indica. Phytochemistry 1997, 46, 555–558. [Google Scholar] [CrossRef]
  111. Fu, X.; Li, X.C.; Smillie, T.J.; Carvalho, P.; Mabusela, W.; Syce, J.; Johnson, Q.; Folk, W.; Avery, M.A.; Khan, I.A. Cycloartane glycosides from Sutherlandia frutescens. J. Nat. Prod. 2008, 71, 1749–1753. [Google Scholar] [CrossRef] [Green Version]
  112. Yang, M.H.; Wang, J.S.; Luo, J.G.; Wang, X.B.; Kong, L.Y. Tetranortriterpenoids from Chisocheton paniculatus. J. Nat. Prod. 2009, 72, 2014–2018. [Google Scholar] [CrossRef]
  113. Siddiqui, B.S.; Afshana, F.; Ghiasuddina, F.; Faizi, S.; Naqvi, S.N.H.; Tariq, R.M. Two insecticidal tetranortriterpenoids from Azadirachta indica. Phytochemistry 2000, 53, 371–376. [Google Scholar] [CrossRef]
  114. Trinh, B.T.D.; Nguyen, H.D.; Nguyen, H.T.; Pham, P.D.; Ngo, N.T.N. Cytotoxic limonoids from the bark of Walsura cochinchinensis. Fitoterapia 2019, 133, 75–79. [Google Scholar] [CrossRef] [PubMed]
  115. Munehiro, N.; Bo, Z.J.; Kenjiro, T.; Hideo, N. Spirosendan, a novel spiro limonoid from Melia toosendan. Chem. Lett. 1998, 27, 1279–1280. [Google Scholar]
  116. Veitch, N.C.; Wright, G.A.; Stevenson, P.C. Four new tetranortriterpenoids from Cedrela odorata associated with leaf rejection by Exopthalmus jekelianus. J. Nat. Prod. 1999, 62, 1260–1263. [Google Scholar] [CrossRef] [PubMed]
  117. Chen, J.L.; Kernan, M.R.; Jolad, S.D.; Stoddart, C.A.; Bogan, M.; Cooper, R. Dysoxylins A–D, tetranortriterpenoids with potent anti-RSV activity from Dysoxylum gaudichaudianum. J. Nat. Prod. 2007, 70, 312–315. [Google Scholar] [CrossRef]
  118. Waratchareeyakul, W.; Hellemann, E.; Gil, R.R.; Chantrapromma, K.; Langat, M.K.; Mulholland, D.A. Application of residual dipolar couplings and selective quantitative NOE to establish the structures of tetranortriterpenoids from Xylocarpus rumphii. J. Nat. Prod. 2017, 80, 391–402. [Google Scholar] [CrossRef]
  119. Tchimene, M.K.; Tane, P.; Ngamga, D.; Connolly, J.D.; Farrugia, L.J. Four tetranortriterpenoids from the stem bark of Khaya anthotheca. Phytochemistry 2005, 66, 1088–1093. [Google Scholar] [CrossRef]
  120. Fauzi, F.M.; Meilanie, S.R.; Zulfikar; Farabi, K.; Herlina, T.; Al Anshori, J.; Mayanti, T. Kokosanolide D: A New tetranortriterpenoid from fruit peels of Lansium domesticum Corr. cv Kokossan. Molbank 2021, 2021, M1232. [Google Scholar] [CrossRef]
  121. Murphy, B.T. Isolation, and Structure Elucidation of Antiproliferative Natural Products from Madagascar. Ph.D. Thesis, Virginia Polytechnic Institute and State University, Blacksburg, VA, USA, 2007. [Google Scholar]
  122. Murphy, B.T.; Brodie, P.; Slebodnick, C.; Miller, J.S.; Birkinshaw, C.; Randrianjanaka, L.M.; Andriantsiferana, R.; Rasamison, V.E.; Dyke, K.T.; Suh, E.M.; et al. Antiproliferative limonoids of a Malleastrum sp. from the Madagascar rainforest. J. Nat. Prod. 2008, 71, 325–329. [Google Scholar] [CrossRef] [Green Version]
  123. Krief, S.; Martin, M.T.; Grellier, P.; Kasenene, J.; Sevenet, T. Novel antimalarial compounds isolated in a survey of self-medicative behavior of wild chimpanzees in Uganda. Antimicrob. Agents Chemother. 2004, 48, 3196–3199. [Google Scholar] [CrossRef] [Green Version]
  124. Batista, R.; de Jesus Silva Júnior, A.; Braga de Oliveira, A. Plant-derived antimalarial agents: New leads and efficient phytomedicines. Part II. Non-alkaloidal natural products. Molecules 2009, 14, 3037–3072. [Google Scholar] [CrossRef] [Green Version]
  125. Curcino Vieira, I.J.; da Silva Terra, W.; dos Santos Gonçalves, M.; Braz-Filho, R. Secondary metabolites of the genus Trichilia: Contribution to the chemistry of Meliaceae family. Am. J. Anal. Chem. 2014, 5, 42542. [Google Scholar] [CrossRef] [Green Version]
  126. Siddiqui, S.; Ghiasuddin, S.; Siddiqui, B.S.; Faizi, S. Tetranortriterpenoids and steroidal glycosides from the seeds of Azadirachta indica A. Juss. Pak. J. Sci. Indust. Res. 1989, 32, 435. [Google Scholar]
  127. Schwikkard, S.L. Extractives from the Meliaceae and Simaroubaceae of Madagascar. Ph.D Thesis, University of Natal, Durban, South Africa, 1997. [Google Scholar]
  128. Tan, Q.G.; Luo, X.D. Meliaceous limonoids: Chemistry and biological activities. Chem. Rev. 2011, 111, 7437–7522. [Google Scholar] [CrossRef]
  129. Banerji, R.; Misra, G.; Nigam, S.K. On the triterpenes of Azadirachta indica (Melia azadirachta). Fitoterapia 1977, 48, 166. [Google Scholar]
  130. Banerji, B.; Nigam, S.K. Wood constituents of Meliaceae: A review. Fitoterapia 1984, 55, 3. [Google Scholar]
  131. Akhila, A.; Rani, K. Chemistry of the neem tree (Azadirachta indica A. Juss.). In Fortschritte der Chemie Organischer Naturstoffe/Progress in the Chemistry of Organic Natural Products; Herz, W., Falk, H., Kirby, G.W., Moore, R.E., Tamm, C., Eds.; Springer: Vienna, Austria, 1999. [Google Scholar]
  132. Gao, Q.; Sun, J.; Xun, H.; Yao, X.; Wang, J.; Tang, F. A new azadirachta from the crude extracts of neem (Azadirachta indica A. Juss) seeds. Nat. Prod. Res. 2017, 31, 1739–1746. [Google Scholar] [CrossRef]
  133. Ou-Yang, D.-W.; Wu, L.; Li, Y.-L.; Yang, P.-M.; Kong, D.-Y.; Yang, X.-W.; Zhang, W.-D. Miscellaneous terpenoid constituents of Abies nephrolepis and their moderate cytotoxic activities. Phytochemistry 2011, 72, 2197–2204. [Google Scholar] [CrossRef]
  134. Yang, S.P.; Yue, J.M. Discovery of structurally diverse and bioactive compounds from plant resources in China. Acta Pharmacol. Sin. 2012, 33, 1147–1158. [Google Scholar] [CrossRef] [Green Version]
  135. Son, N.T. The genus Walsura: A rich resource of bioactive limonoids, triterpenoids, and other types of compounds. In Progress in the Chemistry of Organic Natural Products 118; Kinghorn, A.D., Falk, H., Gibbons, S., Asakawa, Y., Liu, J.K., Dirsch, V.M., Eds.; Springer: Cham, Switzerland, 2022. [Google Scholar]
  136. Li, Y.; Lin, W.; Huang, J. Anti-cancer effects of Gynostemma pentaphyllum (Thunb.) Makino (Jiaogulan). Chin. Med. 2016, 11, 43. [Google Scholar] [CrossRef] [Green Version]
  137. Razmovski-Naumovski, V.; Huang, T.H.W.; Tran, V.H.; Li, G.Q.; Duke, C.C.; Roufogalis, B.D. Chemistry and pharmacology of Gynostemma pentaphyllum. Phytochem. Rev. 2005, 4, 197–219. [Google Scholar] [CrossRef]
  138. Grover, J.K.; Yadav, S.P. Pharmacological actions and potential uses of Momordica charantia: A review. J. Ethnopharmacol. 2004, 93, 123–132. [Google Scholar] [CrossRef]
  139. Jia, S.; Shen, M.; Zhang, F.; Xie, J. Recent Advances in Momordica charantia: Functional components and biological activities. Int. J. Mol. Sci. 2017, 18, 2555. [Google Scholar] [CrossRef] [Green Version]
  140. Han, M.-L.; Zhang, H.; Yang, S.-P.; Yue, J.-M. Walsucochinoids A and B: New rearranged limonoids from Walsura cochinchinensis. Org. Lett. 2012, 14, 486–489. [Google Scholar] [CrossRef]
  141. Simo Mpetga, J.D.; He, H.-P.; Hao, X.-J.; Leng, Y.; Tane, P. Further cycloartane and friedelane triterpenoids from the leaves of Caloncoba glauca. Phytochem. Lett. 2014, 7, 52–56. [Google Scholar] [CrossRef]
  142. Chen, J.-C.; Liu, W.-Q.; Lu, L.; Qiu, M.H.; Zheng, Y.T.; Yang, L.M. Kuguacins F-S, cucurbitane triterpenoids from Momordica charantia. Phytochemistry 2009, 70, 133–140. [Google Scholar] [CrossRef]
  143. Zhang, X.-S.; Cao, J.-Q.; Zhao, C.; Wang, X.-d.; Wu, X.-j.; Zhao, Y.-Q. Novel dammarane-type triterpenes isolated from hydrolyzate of total Gynostemma pentaphyllum saponins Bioorg. Med. Chem. Lett. 2015, 25, 3095–3099. [Google Scholar] [CrossRef]
  144. Cao, J.-Q.; Zhang, Y.; Cui, J.-M.; Zhao, Y.-Q. Two new cucurbitane triterpenoids from Momordica charantia L. Chin. Chem. Lett. 2011, 22, 583–586. [Google Scholar] [CrossRef]
  145. Wang, W.-H.; Nian, Y.; He, Y.-J.; Wan, L.-S.; Bao, N.-M.; Zhu, G.-L.; Wang, F.; Qiu, M.-H. New cycloartane triterpenes from the aerial parts of Cimicifuga heracleifolia. Tetrahedron 2015, 71, 8018–8025. [Google Scholar] [CrossRef]
  146. Huang, D.; Qing, S.; Zeng, G.; Wang, Y.; Guo, H.; Tan, J.; Zhou, Y. Lipophilic components from fructus Viticis Negundo and their antitumor activities. Fitoterapia 2013, 86, 144–148. [Google Scholar] [CrossRef]
  147. Zhou, Y.; Yang, J.; Peng, L.; Li, Y.; Chen, W. Two novel saponins of 20, 26-epoxy derivatives of pseudojujubogenin from the seeds of Hovenia trichocarpa. Fitoterapia 2013, 87, 65–68. [Google Scholar] [CrossRef]
  148. Wong, C.P.; Shimada, M.; Nugroho, A.E.; Hirasawa, Y.; Kaneda, T.; Hadi, A.H.A.; Osamu, S.; Morita, H. Ceramicines J–L, new limonoids from Chisocheton ceramicus. J. Nat. Med. 2012, 66, 566–570. [Google Scholar] [CrossRef]
  149. Renkonen, O.; Schindler, O.; Reichstein, T. Die Konstitution von Sinogenin. Glykoside und Aglykone. 181. Mitteilung. Croatica Chem. Acta 1957, 29, 239–245. [Google Scholar]
  150. Liu, J.-Q.; Wang, C.-F.; Li, Y.; Chen, J.-C.; Zhou, L.; Qiu, M.-H. Limonoids from the leaves of Toona ciliata var. yunnanensis. Phytochemistry 2012, 76, 141–149. [Google Scholar] [CrossRef]
  151. Xu, D.-M.; Huang, E.-X.; Wang, S.-Q.; Wen, X.-G.; Wu, X.-Y. Studies on the chemical constituents of Fritillaria pallidiflora Schrenk. J. Integr. Plant Biol. 1990, 32, 789–793. [Google Scholar]
  152. Li, Y.; Yili, A.; Li, J.; Muhamat, A.; Aisa, H.A. New isosteroidal alkaloids with tracheal relaxant effect from the bulbs of Fritillaria pallidiflora Schrenk. Bioorg. Med. Chem. Lett. 2016, 26, 1983–1987. [Google Scholar] [CrossRef]
  153. Li, H.J.; Jiang, Y.; Ping Li, W.-C. Puqienine F, a novel veratramine alkaloid from the bulbs of Fritillaria puqiensis. Chem. Pharm. Bull. 2006, 54, 722–724. [Google Scholar] [CrossRef] [Green Version]
  154. Babar, Z.U.; Ata, A.; Meshkatalsadat, M.H. New bioactive steroidal alkaloids from Buxus hyrcana. Steroids 2006, 71, 1045–1051. [Google Scholar] [CrossRef]
  155. Niño, J.; Correa, Y.M.; Mosquera, O.M. Biological activities of steroidal alkaloids isolated from Solanum leucocarpum. Pharm. Biol. 2009, 47, 255–259. [Google Scholar] [CrossRef]
  156. Zhao, Q.-Q.; Song, Q.-Y.; Jiang, K.; Li, G.-D.; Wei, W.-J.; Li, Y.; Gao, K. Spirochensilides A and B, two new rearranged triterpenoids from Abies chensiensis. Org. Lett. 2015, 17, 2760–2763. [Google Scholar] [CrossRef]
  157. Wang, G.-W.; Lv, C.; Fang, X.; Tian, X.-H.; Ye, J.; Li, H.-L.; Shan, L.; Shen, Y.-H.; Zhang, W.-D. Eight pairs of epimeric triterpenoids involving a characteristic spiro-E/F ring from Abies faxoniana. J. Nat. Prod. 2015, 78, 50–60. [Google Scholar] [CrossRef]
  158. Meng, X.H.; Chai, T.; Shi, Y.P.; Yang, J.L. Bungsteroid A: One Unusual C34 Pentacyclic Steroid Analogue from Zanthoxylum bungeanum Maxim. J. Org. Chem. 2020, 85, 10806–10812. [Google Scholar] [CrossRef]
  159. Okuzumi, K.; Hara, N.; Uekusa, H.; Fujimoto, Y. Structure elucidation of cyasterone stereoisomers isolated from Cyathula officinalis. Org. Biomol. Chem. 2005, 3, 1227–1232. [Google Scholar] [CrossRef]
  160. Yi, J.; Luo, Y.; Li, B.; Zhang, G. Phytoecdysteroids and glycoceramides from Eriophyton wallchii. Steroids 2004, 69, 809–815. [Google Scholar] [CrossRef]
  161. Snogan, E.; Vahirua-Lechat, I.; Ho, R.; Bertho, G.; Girault, J.P.; Ortiga, S.; Maria, A.; Lafont, R. Ecdysteroids from the medicinal fern Microsorum scolopendria (Burm. f.). Phytochem. Anal. 2007, 18, 441–450. [Google Scholar] [CrossRef]
  162. Hao, D.-C.; Gu, X.-J.; Xiao, P.-G.; Peng, Y. Phytochemical, and biological research of Fritillaria medicinal resources. Chin. J. Nat. Med. 2013, 11, 0330–0344. [Google Scholar] [CrossRef]
  163. Wang, Y.; Hou, H.; Ren, Q. Natural drug sources for respiratory diseases from Fritillaria: Chemical and biological analyses. Chin. Med. 2021, 16, 40. [Google Scholar] [CrossRef] [PubMed]
  164. Rushforth, K.D. Notes on Chinese silver firs 2. Notes R. Bot. Gard. Edinb. 1984, 41, 535–540. [Google Scholar]
  165. Dai, J.; Han, R.; Xu, Y.; Li, N.; Wang, J.; Dan, W. Recent progress of antibacterial natural products: Future antibiotics candidates. Bioorg. Chem. 2020, 101, 103922. [Google Scholar] [CrossRef]
  166. Zhang, M.; Wang, J.; Zhu, L.; Li, T.; Jiang, W.; Zhou, J.; Peng, W.; Wu, C. Zanthoxylum bungeanum Maxim. (Rutaceae): A systematic review of its traditional uses, botany, phytochemistry, pharmacology, pharmacokinetics, and toxicology. Int. J. Mol. Sci. 2017, 18, 2172. [Google Scholar] [CrossRef] [Green Version]
  167. Bao, Y.; Yang, L.; Fu, Q.; Fu, Y.; Tian, Q.; Wang, C.; Huang, Q. The current situation of Zanthoxylum bungeanum industry and the research and application prospect. A review. Fitoterapia 2023, 164, 105380. [Google Scholar] [CrossRef]
  168. Chawla, A.; Dev, S. A new class of triterpenoids from Ailanthus malabarica DC derivatives of malabaricane. Tetrahedron Lett. 1967, 48, 4837–4843. [Google Scholar] [CrossRef]
  169. Srinivas, P.V.; Rao, R.R.; Rao, J.M. Two new tetracyclic triterpenes from the heartwood of Ailanthus excelsa Roxb. Chem. Biodivers. 2006, 3, 930–934. [Google Scholar] [CrossRef]
  170. Ebada, S.S.; Lin, W.; Proksc, P. Bioactive sesterterpenes and triterpenes from marine sponges: Occurrence and pharmacological significance. Mar. Drugs 2010, 8, 313–346. [Google Scholar] [CrossRef] [Green Version]
  171. Khare, C.P. Indian Medicinal Plants; Springer: Berlin/Heidelberg, Germany, 2007; ISBN 978-0-387-70637-5. [Google Scholar]
  172. Thongnest, S.; Boonsombat, J.; Prawat, H.; Mahidol, C.; Ruchirawat, S. Ailanthusins A-G and nor-lupane triterpenoids from Ailanthus triphysa. Phytochemistry 2017, 134, 98–105. [Google Scholar] [CrossRef] [PubMed]
  173. Ziegler, H.L.; Christensen, J.; Olsen, C.E.; Sittie, A.A.; Jaroszewski, J.W. New dammarane and malabaricane triterpenes from Caloncoba echinata. J. Nat. Prod. 2002, 65, 1764–1768. [Google Scholar] [CrossRef]
  174. Liaw, C.C.; Lo, I.W.; Lin, Y.C.; Huang, H.T.; Zhang, L.J.; Hsiao, P.C. Four cucurbitane glycosides taimordisins A–D with novel furopyranone skeletons isolated from the fruits of Momordica charantia. Food Chem. 2022, 14, 100286. [Google Scholar] [CrossRef] [PubMed]
  175. Rashid, I.; Yaqoob, U. Traditional uses, phytochemistry and pharmacology of genus Fritillaria—A review. Bull. Natl. Res. Cent. 2021, 45, 124. [Google Scholar] [CrossRef]
  176. Xu, Y.-H.; Dai, G.-M.; Tang, S.H.; Yang, H.-J.; Sun, Z.-G. Changes of anti-tuberculosis herbs formula during past three decades in contrast to ancient ones. Chin. J. Integr. Med. 2021, 27, 388–393. [Google Scholar] [CrossRef]
  177. Moosmann, P.; Ueoka, R.; Grauso, L.; Mangoni, A.; Morinaka, B.I.; Gugger, M.; Piel, J. Cyanobacterial ent-sterol-like natural products from a deviated ubiquinone pathway. Angew. Chem. Int. Ed. 2017, 56, 4987–4990. [Google Scholar] [CrossRef]
  178. Govindam, S.V.S.; Choi, B.-K.; Yoshioka, Y.; Kanamoto, A.; Fujiwara, T.; Okamoto, T.; Ojika, M. Novel cytotoxic polyoxygenated steroids from an Okinawan sponge Dysidea sp. Biosci. Biotechnol. Biochem. 2012, 76, 999–1002. [Google Scholar] [CrossRef] [Green Version]
  179. Braekman, J.C.; Daloze, D.; Moussiaux, B.; Vandervyver, G.; Riccio, R. Cholest-6-EN-11β,19-epoxy-3β,5α,8α,9α-tetrol, a novel polyoxygenated steroid from the sponge Dysidea tupha. Bull. Soc. Chim. Belg. 1988, 97, 293. [Google Scholar] [CrossRef]
  180. Corgiat, J.M.; Scheuer, P.J.; Rios Steiner, J.L.; Clardy, J. Three pregnane-10, 2-carbolactones from a sponge, Strongylophora sp. Tetrahedron 1993, 49, 1557. [Google Scholar] [CrossRef]
  181. Teruya, T.; Nakagawa, S.; Koyama, T.; Suenaga, K.; Kita, M.; Uemura, D. Nakiterpiosin, a novel cytotoxic C-nor-D-homosteroid from the Okinawan sponge Terpios hoshinota. Tetrahedron Lett. 2003, 44, 5171. [Google Scholar] [CrossRef]
  182. Afiyatullov, S.S.h.; Kalinovsky, A.I.; Antonov, A.S.; Ponomarenko, L.P.; Dmitrenok, P.S.; Aminin, D.L.; Krasokhin, V.B.; Nosova, V.M.; Kisin, A.V. Isolation and structures of erylosides from the Carribean sponge Erylus goffrilleri. J. Nat. Prod. 2007, 70, 1871–1877. [Google Scholar] [CrossRef]
  183. Kolesnikova, S.A.; Lyakhova, E.G.; Kalinovsky, A.I.; Pushilin, M.A.; Afiyatullov, S.S.; Yurchenko, E.A.; Dyshlovoy, S.A.; Minh, C.V.; Stonik, V.A. Isolation, structures, and biological activities of triterpenoids from a marine sponge Penares sp. J. Nat. Prod. 2013, 76, 1746–1752. [Google Scholar] [CrossRef]
  184. Anjaneyulu, A.S.R.; Rao, V.L.; Sastry, V.G. A new spiroketal steroid from Gorgonella umbraculum. Nat. Prod. Res. 2003, 17, 149. [Google Scholar] [CrossRef]
  185. Tomono, Y.; Hirota, H.; Fusetani, N. Isogosterones A–D, antifouling 13,17-secosteroids from an octocoral Dendronephthya sp. J. Org. Chem. 1999, 64, 2272. [Google Scholar] [CrossRef]
  186. Rao, C.B.; Ramana, K.V.; Rao, D.V.; Fahy, E.; Faulkner, D.J. Metabolites of the gorgonian Isis hippuris from India. J. Nat. Prod. 1988, 51, 954. [Google Scholar] [CrossRef]
  187. Rao, C.B.; Kalidindi, R.S.H.S.N.; Trimurtulu, G.; Rao, D.V. Metabolites of porifera, part III. New 24-methylscalaranes from Phyllospongia dendyi of the Indian ocean. J. Nat. Prod. 1991, 54, 364. [Google Scholar] [CrossRef]
  188. Alvi, K.A.; Crews, P. Homoscalarane sesterterpenes from Lendenfeldia frondosa. J. Nat. Prod. 1992, 55, 859. [Google Scholar] [CrossRef]
  189. Crews, P.; Bescansa, P. Sesterterpenes from a common marine sponge, Hyrtios erecta. J. Nat. Prod. 1986, 49, 1041. [Google Scholar] [CrossRef] [PubMed]
  190. Zeng, L.; Fu, X.; Su, J.; Pordesimo, E.O.; Traeger, S.C.; Schmitz, F.J. Novel bishomoscalarane sesterterpenes from the sponge Phyllospongia foliascens. J. Nat. Prod. 1991, 54, 421. [Google Scholar] [CrossRef]
  191. Chang, Y.-C.; Tseng, S.-W.; Liu, L.-L.; Chou, Y.; Ho, Y.-S.; Lu, M.-C.; Su, J.-H. Cytotoxic sesterterpenoids from a sponge Hippospongia sp. Mar. Drugs 2012, 10, 987–997. [Google Scholar] [CrossRef] [Green Version]
  192. Youssef, D.T.; Yamaki, R.K.; Kelly, M.; Scheuer, P.J. Salmahyrtisol A, a novel cytotoxic sesterterpene from the Red Sea sponge Hyrtios erecta. J. Nat. Prod. 2002, 65, 2–6. [Google Scholar] [CrossRef]
  193. Aoki, S.; Watanabe, Y.; Sanagawa, M.; Setiawan, A.; Kotoku, N.; Kobayashi, M. Cortistatins A, B, C, and D, anti-angiogenic steroidal alkaloids, from the marine sponge Corticium simplex. J. Am. Chem. Soc. 2006, 128, 3148–3149. [Google Scholar] [CrossRef] [PubMed]
  194. Roll, D.M.; Scheuer, P.J.; Matsumoto, G.K.; Clardy, J. Halenaquinone, a pentacyclic polyketide from a marine sponge. J. Am. Chem. Soc. 1983, 105, 6177–6178. [Google Scholar] [CrossRef]
  195. Cimino, G.; Madaio, A.; Trivellone, E. Minor triterpenoids from the Mediterranean sponge, Raspaciona aculeata. J. Nat. Prod. 1994, 57, 784–790. [Google Scholar] [CrossRef]
  196. Cimino, G.; Crispino, A.; Madaio, A.; Trivellone, E. Raspacionin B, a further triterpenoid from the mediterranean sponge Raspaciona aculeata. J. Nat. Prod. 1993, 56, 534–538. [Google Scholar] [CrossRef]
  197. Cimino, G.; Epifanio, R.D.A.; Madaio, A.; Puliti, R.; Trivellone, E. Absolute stereochemistry of raspacionin, the main triterpenoid from the marine sponge Raspaciona aculeata. J. Nat. Prod. 1993, 56, 1622–1626. [Google Scholar] [CrossRef]
  198. Chao, C.H.; Huang, L.F.; Yang, Y.L.; Su, J.H.; Wang, G.H.; Chiang, M.Y.; Wu, Y.C.; Dai, C.F.; Sheu, J.H. Polyoxygenated steroids from the Gorgonian Isis hippuris. J. Nat. Prod. 2005, 68, 880–885. [Google Scholar] [CrossRef]
  199. Anjaneyulu, A.S.R.; Krishna Murthy, M.V.R.; Gowri, P.M. Novel epoxy steroids from the Indian ocean soft coral Sarcophyton crassocaule. J. Nat. Prod. 2000, 63, 112–118. [Google Scholar] [CrossRef] [PubMed]
  200. Huang, C.Y.; Su, J.S.; Liaw, C.C.; Sung, P.J.; Chiang, P.L.; Hwang, T.L.; Chang-Feng Dai, J.-H.; Sheu, H. Bioactive steroids with methyl ester group in the side chain from a reef soft coral Sinularia brassica cultured in a tank. Mar. Drugs 2017, 15, 280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. D’Auria, M.V.; Giannini, C.; Zampella, A.; Minale, L.; Debitus, C.; Roussakis, C. Crellastatin A:  A cytotoxic bis-steroid sulfate from the Vanuatu marine sponge Crella sp. J. Org. Chem. 1998, 63, 7382–7388. [Google Scholar] [CrossRef] [PubMed]
  202. Murayama, S.; Imae, Y.; Takada, K.; Kikuchi, J.; Nakao, Y.; van Soest, R.W.M.; Okada, S.; Matsunaga, S. Shishicrellastatins, inhibitors of cathepsin B, from the marine sponge Crella (Yvesia) spinulata. Bioorg. Med. Chem. 2011, 19, 6594–6598. [Google Scholar] [CrossRef] [PubMed]
  203. Morinaka, B.I.; Pawlik, J.R.; Molinski, T.F. Amaroxocanes A and B: Sulfated dimeric sterols defend the Caribbean coral reef sponge Phorbas amaranthus from fish predators. J. Nat. Prod. 2009, 72, 259–264. [Google Scholar] [CrossRef]
  204. Cheng, J.F.; Lee, J.S.; Sun, F.; Jares-Erijman, E.A.; Cross, S.; Rinehart, K.L. Hamigerols A and B, unprecedented polysulfate sterol dimers from the Mediterranean sponge Hamigera hamigera. J. Nat. Prod. 2007, 70, 1195–1199. [Google Scholar] [CrossRef]
  205. Whitson, E.L.; Bugni, T.S.; Chockalingam, P.S.; Conception, G.P.; Feng, X.; Jin, G. Fibrosterol sulfates from the Philippine sponge Lissodendoryx (Acanthodoryx) fibrosa: Sterol dimers that inhibit PKC. J. Org. Chem. 2009, 74, 5902–5908. [Google Scholar] [CrossRef] [Green Version]
  206. Pettit, G.R.; Xu, J.-P.; Chapuis, J.-C.; Melody, N. The cephalostatins. Isolation, structure, and cancer cell growth inhibition of cephalostatin 20. J. Nat. Prod. 2015, 78, 1446–1450. [Google Scholar] [CrossRef]
  207. Boonlarppradab, C.; Faulkner, D.J. Eurysterols A and B, cytotoxic and antifungal steroidal sulfates from a marine sponge of the genus Euryspongia. J. Nat. Prod. 2007, 70, 846–848. [Google Scholar] [CrossRef]
  208. Solanki, H.; Angulo-Preckler, C.; Calabro, K.; Kaur, N.; Lasserre, P.; Cautain, B.; de la Cruz, M.; Reyes, F.; Avila, C.; Thomas, O.P. Suberitane sesterterpenoids from the Antarctic sponge Phorbas areolatus (Thiele, 1905). Tetrahedron Lett. 2018, 59, 3353–3356. [Google Scholar] [CrossRef]
  209. Song, J.; Jeong, W.; Wang, N.; Lee, H.-S.; Sim, C.J.; Oh, K.-B.; Shin, J. Scalarane sesterterpenes from the sponge Smenospongia sp. J. Nat. Prod. 2008, 71, 1866–1871. [Google Scholar] [CrossRef] [PubMed]
  210. Kimura, J.; Hyosu, M. Two new sesterterpenes from the marine sponge, Coscinoderma mathewsi. Chem. Lett. 1999, 52, 61–62. [Google Scholar] [CrossRef]
  211. Hochlowski, J.E.; Faulkner, D.J.; Bass, L.S.; Clardy, J. Metabolites of the dorid nudibranch Chromodoris sedna. J. Org. Chem. 1983, 48, 1738–1740. [Google Scholar] [CrossRef]
  212. Keyzers, R.A.; Daoust, J.; Davies-Coleman, M.T.; Van Soest, R.; Balgi, A.; Donohue, E.; Roberge, M.; Andersen, R.J. Autophagy-modulating aminosteroids isolated from the sponge Cliona celata. Org. Lett. 2008, 10, 259–262. [Google Scholar] [CrossRef]
  213. Li, J.; Zhu, H.; Ren, J.; Deng, Z.; de Voogd, N.J.; Proksch, P.; Lin, W. Globostelletins JeS, isomalabaricanes with unusual cyclopentane sidechains from the marine sponge Rhabdastrella globostellata. Tetrahedron 2012, 68, 559–565. [Google Scholar] [CrossRef]
  214. Guzii, A.G.; Makarieva, T.N.; Denisenko, V.A.; Dmitrenok, P.S.; Burtseva, Y.V.; Krasokhin, V.B.; Stonik, V.A. Topsentiasterol sulfates with novel iodinated and chlorinated side chains from the marine sponge Topsentia sp. Tetrahedron Lett. 2008, 49, 7191–7193. [Google Scholar] [CrossRef]
  215. Fusetani, N.; Takahashi, M.; Matsunaga, S. Topsentiasterol sulfates, antimicrobial sterol sulfates possessing novel side chains, from a marine sponge, Topsentia sp. Tetrahedron 1994, 50, 7765–7770. [Google Scholar] [CrossRef]
  216. Huang, C.-Y.; Liaw, C.-C.; Chen, B.-W.; Chen, P.-C.; Su, J.-H. Withanolide-based steroids from the cultured soft coral Sinularia brassica. J. Nat. Prod. 2013, 76, 1902–1908. [Google Scholar] [CrossRef]
  217. Lal, A.R.; Cambie, R.C.; Rickard, C.E.F.; Bergquist, P.R. Sesterterpene lactones from a sponge species of the genus Dactylospongia. Tetrahedron Lett. 1994, 35, 2603–2606. [Google Scholar] [CrossRef]
  218. Cambie, C.R.; Lal, A.R.; Rickard, C.E.F. A sesterterpene lactone from Petrosaspongia nigra sp. nov. Acta Cryst. 1996, C52, 709–711. [Google Scholar] [CrossRef]
  219. Randazzo, A.; Debitus, C.; Minale, L.; Pastor, P.G.; Alcaraz, M.J.; Payá, M.; Gomez-Paloma, L. Petrosaspongiolodes M–R: New potent and selective phosphoplipase A2 inhibitors from the New Caledonian marine sponge Petrosaspongia nigra. J. Nat. Prod. 1998, 61, 571–575. [Google Scholar] [CrossRef] [PubMed]
  220. Carney, J.R.; Yoshida, W.Y.; Scheuer, P.J. Kiheisterones, new cytotoxic steroids from a Maui sponge. J. Org. Chem. 1992, 57, 6637–6640. [Google Scholar] [CrossRef]
  221. Miyamoto, T.; Sakamoto, K.; Amano, H.; Higuchi, R.; Komori, T.; Sasaki, T. Three new cytotoxic sesterterpenoids, inorolide A, B, and C from the nudibranch Chromodoris inornata. Tetrahedron Lett. 1992, 33, 5811–5814. [Google Scholar] [CrossRef]
  222. D’Auria, M.V.; Paloma, L.G.; Minale, L.; Riccio, R.; Zampella, A. Isolation, structure characterization and conformational analysis of a unique 4α,9α-epoxysteroid sulphate from the okinawan ophiuroid Ophiomastix annulosa. Tetrahedron Lett. 1992, 33, 4641–4644. [Google Scholar] [CrossRef]
  223. Jiménez, J.I.; Yoshida, W.Y.; Scheuer, P.J.; Lobkovsky, E.; Clardy, J.; Kelly, M. Honulactones:  New bishomoscalarane sesterterpenes from the Indonesian sponge Strepsichordaia aliena. J. Org. Chem. 2000, 65, 6837–6840. [Google Scholar] [CrossRef] [PubMed]
  224. Kobayashi, M. Marine terpenes and terpenoids. IV. Isolation of new cembranoid and secocembranoid lactones from the soft coral Sinularia mayi. Chem. Pharm. Bull. 1988, 36, 488–494. [Google Scholar] [CrossRef] [Green Version]
  225. Carmely, S.; Kashman, Y. Isolation and structure elucidation of lobophytosterol, depresosterol and three other closely related sterols: Five new C28 polyoxygenated sterols from the red sea soft coral Lobophytum depressum. Tetrahedron 1981, 37, 2397–2403. [Google Scholar] [CrossRef]
  226. Kashman, V.; Carmely, S. Four novel C28 sterols from Lobophytum depressum. Tetrahedron Lett. 1980, 21, 4939–4942. [Google Scholar] [CrossRef]
  227. Kim, S.K.; van Ta, Q. Bioactive sterols from marine resources and their potential benefits for human health. Adv. Food Nutrit. Res. 2012, 65, 261–268. [Google Scholar]
  228. Donets, M.M.; Tsygankov, V.Y. Trace elements in commercial marine organisms from the Russian part of the Northwest Pacific (2010–2018). Environ. Chem. Lett. 2019, 17, 1727–1740. [Google Scholar] [CrossRef]
Figure 1. Furanosteroids produced by fungal species.
Figure 1. Furanosteroids produced by fungal species.
Molecules 28 05669 g001
Figure 2. Three-dimensional structure of furanosteroid named viridin (2) and showing a wide range of antifungal and other biological activities. A distinctive characteristic of natural furanosteroids, as was believed in the 1950s of the last century, is the presence of an additional aromatic furan ring in the structure of the steroid. The aromatic furane ring is highlighted in yellow. The percentage of biological activities is shown in Figure 3. Viridin, an antibiotic produced by a pigment-forming strain of the common soil fungus Trichoderma viride, was first described by Brian and McGowan, and Brian, Curtis, Hemming and McGowan. Viridin is not antibacterial but is highly antifungal. Its activity against certain fungi is remarkably high. Germination of the spores of, for example, Botrytis allii is prevented by a concentration of 0.019 p.p.m. of α-viridin or 0.156 p.p.m. of β-viridin. Among synthetic fungicides, only the organo-mercurials are of the same order of activity. gray is carbon, white is hydrogen, and red is oxygen.
Figure 2. Three-dimensional structure of furanosteroid named viridin (2) and showing a wide range of antifungal and other biological activities. A distinctive characteristic of natural furanosteroids, as was believed in the 1950s of the last century, is the presence of an additional aromatic furan ring in the structure of the steroid. The aromatic furane ring is highlighted in yellow. The percentage of biological activities is shown in Figure 3. Viridin, an antibiotic produced by a pigment-forming strain of the common soil fungus Trichoderma viride, was first described by Brian and McGowan, and Brian, Curtis, Hemming and McGowan. Viridin is not antibacterial but is highly antifungal. Its activity against certain fungi is remarkably high. Germination of the spores of, for example, Botrytis allii is prevented by a concentration of 0.019 p.p.m. of α-viridin or 0.156 p.p.m. of β-viridin. Among synthetic fungicides, only the organo-mercurials are of the same order of activity. gray is carbon, white is hydrogen, and red is oxygen.
Molecules 28 05669 g002
Figure 7. Furanosteroids produced by fungi and fungal endophytes.
Figure 7. Furanosteroids produced by fungi and fungal endophytes.
Molecules 28 05669 g007
Figure 8. Three-dimensional graph illustrates the predicted and calculated antiprotozoal activity of steroids (36, 37, 38, 39, and 40) with a confidence level exceeding 95%. These steroids are produced by the medicinal mushroom Ganoderma tropicum.
Figure 8. Three-dimensional graph illustrates the predicted and calculated antiprotozoal activity of steroids (36, 37, 38, 39, and 40) with a confidence level exceeding 95%. These steroids are produced by the medicinal mushroom Ganoderma tropicum.
Molecules 28 05669 g008
Figure 9. Steroids and isoprenoid lipids produced by fungi.
Figure 9. Steroids and isoprenoid lipids produced by fungi.
Molecules 28 05669 g009
Figure 10. Three-dimensional graph illustrates the predicted and calculated activity of steroids (49, 50, 54, and 55) as nitric oxide inhibitors, with a confidence level exceeding 90%.
Figure 10. Three-dimensional graph illustrates the predicted and calculated activity of steroids (49, 50, 54, and 55) as nitric oxide inhibitors, with a confidence level exceeding 90%.
Molecules 28 05669 g010
Figure 11. Steroids and meroterpenoids derived from fungal species.
Figure 11. Steroids and meroterpenoids derived from fungal species.
Molecules 28 05669 g011
Figure 18. (a) Walsura cochinchinensis: This plant, belonging to the Meliaceae family, contains bioactive limonoids known as walsucochinoids A (140) and B (141). Medicinal plants from the Walsura genus, found in tropical areas of several Asian countries, are widely used in traditional medicine systems. Studies have identified over 200 compounds from ten species within this genus, including sesquiterpenoids, flavonoids, sterols, lignans, xanthones, and anthraquinones. Many of these compounds exhibit diverse properties such as cancer cell cytotoxicity, antimicrobial activity, antidiabetic effects, anti-inflammatory effects, antioxidant properties, antifeedant properties, antifertility effects, ichthyotoxic effects, and neuroprotective effects [134,135]. (b) Gynostemma pentaphyllum: Also known as Jiaogulan, Gynostemma pentaphyllum is a dioecious climbing vine from the Cucurbitaceae family. It is widely distributed in South and East Asia, as well as New Guinea. In Chinese medicine, it is commonly used to treat various conditions, including hepatitis, diabetes, and cardiovascular diseases. Extracts from G. pentaphyllum contain sterols, flavonoids, and polysaccharides that exhibit inhibitory activity against cancer cell proliferation. These extracts have demonstrated effects such as cell cycle arrest, apoptosis induction, inhibition of invasion and metastasis, inhibition of glycolysis, and immunomodulatory activity [136,137]. (c) Momordica charantia: The fruit of M. charantia, which contains anticancer steroids (147 and 148), has been consumed as food and used as medicine since ancient times. M. charantia, commonly known as bitter melon, holds a significant place in various systems of traditional medicine. It has been used to treat a wide range of conditions, including diabetes, abortive purposes, anthelmintic effects, contraception, dysmenorrhea, eczema, emmenagogue properties, antimalarial activity, lactagogue effects, gout, jaundice, abdominal pain, renal issues (stones), laxative effects, leprosy, leucorrhea, hemorrhoids, pneumonia, psoriasis, rheumatism, fever, and scabies [138,139].
Figure 18. (a) Walsura cochinchinensis: This plant, belonging to the Meliaceae family, contains bioactive limonoids known as walsucochinoids A (140) and B (141). Medicinal plants from the Walsura genus, found in tropical areas of several Asian countries, are widely used in traditional medicine systems. Studies have identified over 200 compounds from ten species within this genus, including sesquiterpenoids, flavonoids, sterols, lignans, xanthones, and anthraquinones. Many of these compounds exhibit diverse properties such as cancer cell cytotoxicity, antimicrobial activity, antidiabetic effects, anti-inflammatory effects, antioxidant properties, antifeedant properties, antifertility effects, ichthyotoxic effects, and neuroprotective effects [134,135]. (b) Gynostemma pentaphyllum: Also known as Jiaogulan, Gynostemma pentaphyllum is a dioecious climbing vine from the Cucurbitaceae family. It is widely distributed in South and East Asia, as well as New Guinea. In Chinese medicine, it is commonly used to treat various conditions, including hepatitis, diabetes, and cardiovascular diseases. Extracts from G. pentaphyllum contain sterols, flavonoids, and polysaccharides that exhibit inhibitory activity against cancer cell proliferation. These extracts have demonstrated effects such as cell cycle arrest, apoptosis induction, inhibition of invasion and metastasis, inhibition of glycolysis, and immunomodulatory activity [136,137]. (c) Momordica charantia: The fruit of M. charantia, which contains anticancer steroids (147 and 148), has been consumed as food and used as medicine since ancient times. M. charantia, commonly known as bitter melon, holds a significant place in various systems of traditional medicine. It has been used to treat a wide range of conditions, including diabetes, abortive purposes, anthelmintic effects, contraception, dysmenorrhea, eczema, emmenagogue properties, antimalarial activity, lactagogue effects, gout, jaundice, abdominal pain, renal issues (stones), laxative effects, leprosy, leucorrhea, hemorrhoids, pneumonia, psoriasis, rheumatism, fever, and scabies [138,139].
Molecules 28 05669 g018
Figure 19. Three-dimensional graph illustrating the predicted and calculated activity of steroids (140, 141, and 142) as cortisone reductase inhibitors with over 90% confidence. These steroids include the limonoids with an aromatic ring D (140 and 141) found in the leaves of the Walsura cochinchinensis plant, as well as the triterpenoid (142) found in the leaves of Caloncoba glauca.
Figure 19. Three-dimensional graph illustrating the predicted and calculated activity of steroids (140, 141, and 142) as cortisone reductase inhibitors with over 90% confidence. These steroids include the limonoids with an aromatic ring D (140 and 141) found in the leaves of the Walsura cochinchinensis plant, as well as the triterpenoid (142) found in the leaves of Caloncoba glauca.
Molecules 28 05669 g019
Figure 23. (a) The marine sponge belonging to the genus Dysidea is the origin of dysideasterols F (188) and G (189). (b) The Okinawan marine sponge Terpios hoshinota produces the norsteroid nakiterpiosin (193). (c) The marine sponge Lendenfeldia frondosa, found in the Solomon Islands, contains epihomoscalaralactone IIA (204). (d) The Vietnamese marine sponge Penares sp. serves as the source of lanostanes (197 and 198). These diverse marine organisms contribute to the production of biologically active compounds with potential pharmacological applications.
Figure 23. (a) The marine sponge belonging to the genus Dysidea is the origin of dysideasterols F (188) and G (189). (b) The Okinawan marine sponge Terpios hoshinota produces the norsteroid nakiterpiosin (193). (c) The marine sponge Lendenfeldia frondosa, found in the Solomon Islands, contains epihomoscalaralactone IIA (204). (d) The Vietnamese marine sponge Penares sp. serves as the source of lanostanes (197 and 198). These diverse marine organisms contribute to the production of biologically active compounds with potential pharmacological applications.
Molecules 28 05669 g023
Figure 24. Furanosteroids and isoprenoid lipids derived from marine sources.
Figure 24. Furanosteroids and isoprenoid lipids derived from marine sources.
Molecules 28 05669 g024
Figure 25. Three-dimensional graph illustrates the predicted and calculated antineoplastic activity of steroids (210, 218, 219, 220, and 221) with a confidence level exceeding 92%. These cytotoxic, highly polyoxygenated steroids are isolated from marine organisms inhabiting the waters surrounding Taiwan. Specifically, a metabolite (210) is found in the marine sponge Hippospongia sp., while the bamboo coral Isis hippuris contains compounds (218221). These discoveries highlight the potential of marine organisms to provide valuable bioactive compounds with antineoplastic properties.
Figure 25. Three-dimensional graph illustrates the predicted and calculated antineoplastic activity of steroids (210, 218, 219, 220, and 221) with a confidence level exceeding 92%. These cytotoxic, highly polyoxygenated steroids are isolated from marine organisms inhabiting the waters surrounding Taiwan. Specifically, a metabolite (210) is found in the marine sponge Hippospongia sp., while the bamboo coral Isis hippuris contains compounds (218221). These discoveries highlight the potential of marine organisms to provide valuable bioactive compounds with antineoplastic properties.
Molecules 28 05669 g025
Figure 26. Furanosteroids, dimeric and isoprenoid lipids derived from marine sources.
Figure 26. Furanosteroids, dimeric and isoprenoid lipids derived from marine sources.
Molecules 28 05669 g026
Figure 27. Three-dimensional graph presents the predicted and calculated activity of the steroid (228) with a confidence level exceeding 92%. This steroid is recognized as a protein kinase C inhibitor and exhibits potential as an anticancer and antiviral agent. It is a sulfated sterol dimer found in the marine sponge Lissodendoryx (Acanthodoryx) fibrosa, which belongs to the family Coelosphaeridae. This marine sponge is commonly found in the oceanic waters of the Philippines. The discovery of this compound highlights its promising bioactive properties derived from marine sources.
Figure 27. Three-dimensional graph presents the predicted and calculated activity of the steroid (228) with a confidence level exceeding 92%. This steroid is recognized as a protein kinase C inhibitor and exhibits potential as an anticancer and antiviral agent. It is a sulfated sterol dimer found in the marine sponge Lissodendoryx (Acanthodoryx) fibrosa, which belongs to the family Coelosphaeridae. This marine sponge is commonly found in the oceanic waters of the Philippines. The discovery of this compound highlights its promising bioactive properties derived from marine sources.
Molecules 28 05669 g027
Figure 28. (a) The octocoral Dendronephthya sp. from the Sea of Japan produces secosteroids, namely isogosterones A (200) and B (201). (b) The marine sponge Isis hippuris serves as a source of steroids (202) and (203). (c) The marine sponge Corticium simplex produces steroidal alkaloids (212215). (d) The marine worm Cephalodiscus gilchristi synthesizes a unique steroidal alkaloid (229). These diverse marine organisms contribute to the production of bioactive compounds with potential pharmacological applications.
Figure 28. (a) The octocoral Dendronephthya sp. from the Sea of Japan produces secosteroids, namely isogosterones A (200) and B (201). (b) The marine sponge Isis hippuris serves as a source of steroids (202) and (203). (c) The marine sponge Corticium simplex produces steroidal alkaloids (212215). (d) The marine worm Cephalodiscus gilchristi synthesizes a unique steroidal alkaloid (229). These diverse marine organisms contribute to the production of bioactive compounds with potential pharmacological applications.
Molecules 28 05669 g028
Figure 29. (a) The reef soft coral Sinularia brassica is a producer of the cytotoxic steroid sinubrasone B (224). (b) The Caribbean coral reef sponge Phorbas amaranthus contains the dimeric steroid amaroxocane B (227). (c) The Mediterranean sponge Hamigera hamigera produces the same sterol dimer, hamigerol B (227). (d) The nudibranch Chromodoris sedna contains two sesterterpenoids, namely sednolide (237) and sednolide 22-acetate (238). These marine organisms contribute to the production of bioactive compounds with diverse pharmacological properties.
Figure 29. (a) The reef soft coral Sinularia brassica is a producer of the cytotoxic steroid sinubrasone B (224). (b) The Caribbean coral reef sponge Phorbas amaranthus contains the dimeric steroid amaroxocane B (227). (c) The Mediterranean sponge Hamigera hamigera produces the same sterol dimer, hamigerol B (227). (d) The nudibranch Chromodoris sedna contains two sesterterpenoids, namely sednolide (237) and sednolide 22-acetate (238). These marine organisms contribute to the production of bioactive compounds with diverse pharmacological properties.
Molecules 28 05669 g029
Figure 30. Furanosteroids and triterpenoids derived from marine sources.
Figure 30. Furanosteroids and triterpenoids derived from marine sources.
Molecules 28 05669 g030
Figure 31. Three-dimensional graph illustrating the predicted and calculated activity of steroids (254 and 255) with a confidence level exceeding 90%. These steroids are recognized as phospholipase A2 inhibitors, showcasing their potential pharmacological activity.
Figure 31. Three-dimensional graph illustrating the predicted and calculated activity of steroids (254 and 255) with a confidence level exceeding 90%. These steroids are recognized as phospholipase A2 inhibitors, showcasing their potential pharmacological activity.
Molecules 28 05669 g031
Figure 32. (a) The ophiuroid species Ophiomastix annulosa serves as a holder of the steroid (261). (b) The soft coral Lobophytum depressum is a producer of steroids (265) and (266). (c) The marine sponge Rhabdastrella globostellata contains triterpenoids (240 and 241). (d) The marine mollusc Spisula sachalinensis contains the steroid (243). S. sachalinensis is a bivalve mollusc found in the Sea of Japan and the Sea of Okhotsk, inhabiting medium- to fine-grained sands at depths ranging from 0.5 to 10 m. It is the largest of the molluscs, reaching a length of approximately 130 mm and weighing around 500 g. This mollusc is considered a delicacy and is actively fished in Japan and Primorsky Krai (Russia) [228].
Figure 32. (a) The ophiuroid species Ophiomastix annulosa serves as a holder of the steroid (261). (b) The soft coral Lobophytum depressum is a producer of steroids (265) and (266). (c) The marine sponge Rhabdastrella globostellata contains triterpenoids (240 and 241). (d) The marine mollusc Spisula sachalinensis contains the steroid (243). S. sachalinensis is a bivalve mollusc found in the Sea of Japan and the Sea of Okhotsk, inhabiting medium- to fine-grained sands at depths ranging from 0.5 to 10 m. It is the largest of the molluscs, reaching a length of approximately 130 mm and weighing around 500 g. This mollusc is considered a delicacy and is actively fished in Japan and Primorsky Krai (Russia) [228].
Molecules 28 05669 g032
Table 1. Predicted biological activities of steroids derived from fungi (225).
Table 1. Predicted biological activities of steroids derived from fungi (225).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
2Antifungal (0.912)
Anti-inflammatory (0.912)
Antiprotozoal (Plasmodium) (0.794)
Antibacterial (0.783)
Antineoplastic (0.909)
Apoptosis agonist (0.828)
Prostate disorders treatment (0.600)
Proliferative diseases treatment (0.578)
3Antineoplastic (0.802)
Apoptosis agonist (0.647)
Chemopreventive (0.573)
Anti-inflammatory (0.629)
Antibacterial (0.547)
Cytochrome P450 inhibitor (0.503)
4Antineoplastic (0.779)
Apoptosis agonist (0.766)
Antibacterial (0.718)
Antifungal (0.709)
5Antineoplastic (0.837)
Apoptosis agonist (0.704)
Antibacterial (0.745)
6Anti-infertility, female (0.789)
Lipid metabolism regulator (0.789)
Antineoplastic (0.756)
Apoptosis agonist (0.749)
7Antineoplastic (0.913)
Apoptosis agonist (0.907)
Inositol-3-kinase inhibitor (0.898)
Lipid metabolism regulator (0.700)
Hypolipemic (0.693)
Antibacterial (0.642)
8Inositol-4-kinase inhibitor (0.945)
Apoptosis agonist (0.865)
Inositol-3-kinase inhibitor (0.848)
Antifungal (0.876)
Antibacterial (0.733)
9Antineoplastic (0.934)
Apoptosis agonist (0.865)
Antifungal (0.881)
Antibacterial (0.766)
10Inositol-4-kinase inhibitor (0.949)
Inositol-3-kinase inhibitor (0.843)
Monoamine oxidase inhibitor (0.837)
Apoptosis agonist (0.899)
Antineoplastic (0.879)
11Antineoplastic (0.924)
Apoptosis agonist (0.833)
Antifungal (0.876)
Antibacterial (0.733)
12Antineoplastic (0.903)
Apoptosis agonist (0.807)
Prostate disorders treatment (0.800)
Lipid metabolism regulator (0.700)
Hypolipemic (0.693)
13Antineoplastic (0.934)
Apoptosis agonist (0.907)
Prostate disorders treatment (0.866)
Lipid metabolism regulator (0.897)
Antifungal (0.881)
Hypolipemic (0.793)
14Antiproliferative (0.938)
Antineoplastic (0.902)
Apoptosis agonist (0.889)
Prostate disorders treatment (0.900)
Proliferative diseases treatment (0.878)
15Antiproliferative (0.929)
Antineoplastic (0.922)
Prostate disorders treatment (0.876)
Proliferative diseases treatment (0.844)
16Antiproliferative (0.915)
Apoptosis agonist (0.882)
Prostate disorders treatment (0.854)
Proliferative diseases treatment (0.823)
17Antineoplastic (0.877)
Apoptosis agonist (0.856)
Ovulation inhibitor (0.698)
Anti-inflammatory (0.682)
18Antineoplastic (0.879)
Apoptosis agonist (0.876)
Ovulation inhibitor (0.728)
Anti-inflammatory (0.717)
19Antineoplastic (0.865)
Antiprotozoal (Plasmodium) (0.621)
Antifungal (0.779)
Antibacterial (0.747)
20Antineoplastic (0.888)Antifungal (0.843)
21Antineoplastic (0.910)Apoptosis agonist (0.877)
22Nitric oxide inhibitor (0.921)Apoptosis agonist (0.879)
23Nitric oxide inhibitor (0.914)
Antifungal (0.862)
Apoptosis agonist (0.867)
Antineoplastic (0.844)
24Antineoplastic (0.937)Apoptosis agonist (0.892)
25Antineoplastic (0.929)Apoptosis agonist (0.887)
* Only activities with Pa > 0.5 are shown.
Table 2. Predicted biological activities of steroids derived from fungi (2647).
Table 2. Predicted biological activities of steroids derived from fungi (2647).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
26Antineoplastic (0.922)
Apoptosis agonist (0.844)
Antifungal (0.887)
Antibacterial (0.754)
27Antineoplastic (0.929)
Apoptosis agonist (0.839)
Antifungal (0.832)
Antibacterial (0.782)
28Antineoplastic (0.921)
Apoptosis agonist (0.841)
Antifungal (0.833)
Antibacterial (0.755)
29Antineoplastic (0.889)Antibacterial (0.766)
30Antineoplastic (0.833)Antibacterial (0.733)
31Hepatoprotectant (0.860)
Kidney function stimulant (0.719)
Muscle relaxant (0.599)
Spasmolytic (0.587)
32Hepatoprotectant (0.875)
Kidney function stimulant (0.686)
Muscle relaxant (0.612)
Spasmolytic (0.595)
33Hepatoprotectant (0.839)
Kidney function stimulant (0.719)
Muscle relaxant (0.782)
34Antiprotozoal (Plasmodium) (0.718)Anti-inflammatory (0.778)
35Antiprotozoal (0.657)Anti-inflammatory (0.766)
36Antiprotozoal (Plasmodium) (0.947)
Antiprotozoal (Leishmania) (0.888)
Antiprotozoal (0.879)
Antineoplastic (0.866)
Apoptosis agonist (0.795)
Chemopreventive (0.722)
37Antiprotozoal (Plasmodium) (0.949)
Antiprotozoal (Leishmania) (0.868)
Antiprotozoal (0.823)
Antineoplastic (0.859)
Apoptosis agonist (0.791)
Chemopreventive (0.717)
38Antiprotozoal (Plasmodium) (0.927)
Antiprotozoal (Leishmania) (0.914)
Antiprotozoal (0.884)
Antineoplastic (0.843)
Apoptosis agonist (0.747)
Chemopreventive (0.698)
39Antiprotozoal (Plasmodium) (0.912)
Antiprotozoal (Leishmania) (0.886)
Antiprotozoal (0.778)
Antineoplastic (0.876)
Chemopreventive (0.773)
40Antiprotozoal (Plasmodium) (0.955)
Antiprotozoal (Leishmania) (0.948)
Antiprotozoal (0.903)
Chemopreventive (0.903)
Antineoplastic (0.822)
Apoptosis agonist (0.747)
41Antiprotozoal (0.822)Anti-inflammatory (0.712)
42Antibacterial (0.883)
Antifungal (0.876)
Laxative (0.735)
Anti-eczematic (0.717)
43Antibacterial (0.907)
Antifungal (0.865)
Laxative (0.722)
Anti-eczematic (0.686)
44Antibacterial (0.885)
Antifungal (0.876)
Anti-inflammatory (0.712)
45Antibacterial (0.912)
Antifungal (0.875)
Anti-eczematic (0.638)
Anti-inflammatory (0.612)
46Antibacterial (0.901)
Antifungal (0.867)
Anti-eczematic (0.689)
47Antibacterial (0.889)
Antifungal (0.882)
Anti-eczematic (0.654)
* Only activities with Pa > 0.5 are shown.
Table 3. Predicted biological activities of steroids derived from fungi (4872).
Table 3. Predicted biological activities of steroids derived from fungi (4872).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
48Antibacterial (0.856)
Antifungal (0.811)
Antiviral (0.578)
49Nitric oxide inhibitor (0.906)
Apoptosis agonist (0.667)
Antineoplastic (0.764)
50Nitric oxide inhibitor (0.874)
Apoptosis agonist (0.685)
Antineoplastic (0.785)
51Plant growth inhibitor (0.876)Antifungal (0.732)
52Plant growth inhibitor (0.855)Antifungal (0.689)
53Antibacterial (0.769)
Antifungal (0.716)
Antiviral (0.577)
54Nitric oxide inhibitor (0.779)Antibacterial (0.686)
55Nitric oxide inhibitor (0.812)Antibacterial (0.677)
56Antineoplastic (0.755)Anti-inflammatory (0.612)
57Antineoplastic (0.734)Antifungal (0.521)
58Antibacterial (0.772)
Antifungal (0.631)
Antiviral (0.589)
59Antibacterial (0.856)Antiviral (0.618)
60Antifungal (0.789)
Antibacterial (0.634)
Antiviral (0.611)
61Antineoplastic (0.785)Anti-inflammatory (0.622)
62Cytoprotectant (0.758)
63Antibacterial (0.882)Antifungal (0.668)
64Antibacterial (0.872)Antifungal (0.675)
65Antibacterial (0.868)Antifungal (0.712)
66Antibacterial (0.856)Antifungal (0.682)
67Antimicrobial (0.902)
Antibacterial (0.856)
Antifungal (0.754)
68Antibacterial (0.911)Antimicrobial (0.773)
69Antiprotozoal (0.886)
Antibacterial (0.847)
Antimicrobial (0.638)
70Antineoplastic (0.772)Antifungal (0.722)
71Antineoplastic (0.882)
Chemopreventive (0.791)
Anti-inflammatory (0.634)
72Antineoplastic (0.913)
Chemopreventive (0.882)
Apoptosis agonist (0.798)
Antiviral (0.832)
Anti-inflammatory (0.711)
* Only activities with Pa > 0.5 are shown.
Table 4. Predicted biological activities of steroids derived from fungi (73101).
Table 4. Predicted biological activities of steroids derived from fungi (73101).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
73Antineoplastic (0.884)Antifungal (0.621)
74Antineoplastic (0.975)
Apoptosis agonist (0.889)
Chemopreventive (0.887)
Lipid metabolism regulator (0.858)
Hypolipemic (0.677)
75Antineoplastic (0.968)
Chemopreventive (0.882)
Apoptosis agonist (0.798)
Lipid metabolism regulator (0.897)
Hypolipemic (0.693)
76Antineoplastic (0.973)
Apoptosis agonist (0.798)
Chemopreventive (0.882)
Lipid metabolism regulator (0.872)
Hypolipemic (0.654)
77Antineoplastic (0.865)Antifungal (0.633)
78Antineoplastic (0.834)Antifungal (0.611)
79Antineoplastic (0.844)Anti-inflammatory (0.666)
80Antineoplastic (0.851)Antifungal (0.597)
81Antiproliferative (0.834)Cytotoxic (0.658)
82Antiproliferative (0.812)Anti-inflammatory (0.645)
83Antiproliferative (0.876)
Antineoplastic (0.875)
Cytotoxic (0.745)
Antifungal (0.677)
84Antiproliferative (0.876)Anti-inflammatory (0.668)
85Antiproliferative (0.924)
Cytotoxic (0.892)
Cytotoxic (0.821)
Antifungal (0.734)
86Antiproliferative (0.902)
Cytotoxic (0.883)
Cytotoxic (0.833)
Antifungal (0.680)
87Antineoplastic (0.823)Anti-inflammatory (0.619)
88Antiproliferative (0.831)Anti-inflammatory (0.644)
89Antiproliferative (0.849)Hypolipemic (0.623)
90Antineoplastic (0.736)Antifungal (0.567)
91Antineoplastic (0.729)Antifungal (0.564)
92Antineoplastic (0.751)Antifungal (0.592)
93Antineoplastic (0.722)Antimicrobial (0.612)
94Antineoplastic (0.733)Antimicrobial (0.622)
95Antineoplastic (0.728)Antimicrobial (0.641)
96Antineoplastic (0.802)Antimicrobial (0.632)
97Allergic conjunctivitis treatment (0.687)Antifungal (0.566)
98Allergic conjunctivitis treatment (0.705)Antifungal (0.587)
99Antineoplastic (0.902)
Antiproliferative (0.883)
Apoptosis agonist (0.870)
Cytotoxic (0.844)
Anti-inflammatory (0.634)
100Antibacterial (0.903)Antimicrobial (0.734)
101Antibacterial (0.868)Antimicrobial (0.698)
* Only activities with Pa > 0.5 are shown.
Table 5. Predicted biological activities of steroids derived from plants (102127).
Table 5. Predicted biological activities of steroids derived from plants (102127).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
102Antineoplastic (0.879)
Chemopreventive (0.683)
Apoptosis agonist (0.641)
Anti-inflammatory (0.840)
Antifungal (0.671)
103Antineoplastic (0.851)
Chemopreventive (0.662)
Apoptosis agonist (0.637)
Anti-inflammatory (0.790)
Antifungal (0.653)
104Analgesic (0.761)
Antitussive (0.652)
Respiratory analeptic (0.744)
Oxygen scavenger (0.651)
105Antiprotozoal (Plasmodium) (0.869)
Antiprotozoal (0.832)
Antineoplastic (0.798)
Apoptosis agonist (0.795)
106Antiprotozoal (Plasmodium) (0.939)
Antiprotozoal (Leishmania) (0.891)
Antiprotozoal (0.883)
Antineoplastic (0.874)
Apoptosis agonist (0.765)
Chemopreventive (0.703)
107Antiprotozoal (Plasmodium) (0.954)
Antiprotozoal (Leishmania) (0.912)
Antiprotozoal (0.889)
Antineoplastic (0.911)
Apoptosis agonist (0.893)
Chemopreventive (0.791)
108Antiprotozoal (Plasmodium) (0.941)
Antiprotozoal (Leishmania) (0.866)
Antiprotozoal (0.836)
Antineoplastic (0.876)
Apoptosis agonist (0.795)
Chemopreventive (0.722)
109Antiprotozoal (Plasmodium) (0.940)
Antiprotozoal (Leishmania) (0.848)
Antiprotozoal (0.842)
Antineoplastic (0.868)
Apoptosis agonist (0.794)
Chemopreventive (0.713)
110Antineoplastic (0.955)
Apoptosis agonist (0.802)
Antiviral (0.818)
Anti-inflammatory (0.721)
111Antineoplastic (0.946)
Apoptosis agonist (0.766)
Antiviral (0.832)
Anti-inflammatory (0.734)
112Antineoplastic (0.932)
Apoptosis agonist (0.778)
Antiviral (0.822)
113Antineoplastic (0.829)
Cytochrome P450 inhibitor (0.691)
Anti-inflammatory (0.790)
Antimitotic (0.592)
114Antineoplastic (0.867)
Cytochrome P450 inhibitor (0.748)
Anti-inflammatory (0.801)
Antimitotic (0.622)
115Antineoplastic (0.793)Anti-inflammatory (0.715)
116Antiprotozoal (Plasmodium) (0.707)
Antiprotozoal (0.697)
Antineoplastic (0.707)
Apoptosis agonist (0.570)
117Antineoplastic (0.872)
Apoptosis agonist (0.712)
Antiprotozoal (Plasmodium) (0.746)
Antiprotozoal (0.703)
118Antineoplastic (0.854)
Apoptosis agonist (0.743)
Antiprotozoal (Plasmodium) (0.721)
Antiprotozoal (0.693)
119Antineoplastic (0.850)
Apoptosis agonist (0.691)
Cytostatic (0.683)
Hepatoprotectant (0.767)
Immunosuppressant (0.712)
Anti-hypercholesterolemic (0.566)
120Antineoplastic (0.855)
Apoptosis agonist (0.697)
Hepatoprotectant (0.768)
Immunosuppressant (0.718)
121Antineoplastic (0.858)
Apoptosis agonist (0.690)
Hepatoprotectant (0.753)
Anti-hypercholesterolemic (0.584)
122Antineoplastic (0.860)
Apoptosis agonist (0.699)
Hepatoprotectant (0.755)
Anti-hypercholesterolemic (0.616)
123Antiviral (0.832)Anti-inflammatory (0.721)
124Antiviral (0.856)Anti-inflammatory (0.734)
125Antiviral (0.887)
Antiviral (Arbovirus) (0.790)
Antiprotozoal (0.693)
126Antiviral (0.877)Anti-inflammatory (0.715)
127Antiviral (0.839)Anti-inflammatory (0.698)
* Only activities with Pa > 0.5 are shown.
Table 6. Biological activities of steroids derived from plants (128154).
Table 6. Biological activities of steroids derived from plants (128154).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
128Hepatoprotectant (0.784)Immunosuppressant (0.734)
129Hepatoprotectant (0.772)Immunosuppressant (0.728)
130Hepatoprotectant (0.777)Immunosuppressant (0.729)
131Antimicrobial (0.841)Antibacterial (0.823)
132Antineoplastic (0.823)Antiprotozoal (0.821)
133Antineoplastic (0.894)
Chemopreventive (0.763)
Antiprotozoal (0.865)
134Antineoplastic (0.902)
Chemopreventive (0.883)
Antiprotozoal (0.872)
135Antifungal (0.833)Anti-inflammatory (0.790)
136Antifungal (0.797)Antimicrobial (0.671)
137Antifungal (0.803)Antimicrobial (0.666)
138Antimicrobial (0.854)Antifungal (0.687)
139Antimicrobial (0.798)Antifungal (0.693)
140Cortisone reductase inhibitor (0.875)Antineoplastic (0.825)
141Cortisone reductase inhibitor (0.902)Antineoplastic (0.887)
142Cortisone reductase inhibitor (0.898)Antineoplastic (0.879)
143Anti-HIV-1 (0.894)
Antiviral (0.839)
Anti-inflammatory (0.768)
144Anti-HIV-1 (0.904)
Antiviral (0.885)
Anti-inflammatory (0.754)
145Prostate cancer treatment (0.928)
Antineoplastic (0.911)
Antimicrobial (0.723)
146Prostate cancer treatment (0.914)
Antineoplastic (0.896)
Antimicrobial (0.741)
147Antineoplastic (0.907)
Prostate cancer treatment (0.894)
Antimicrobial (0.654)
148Antineoplastic (0.897)
Prostate cancer treatment (0.887)
Antimicrobial (0.678)
149Antineoplastic (0.838)Antimicrobial (0.611)
150Antineoplastic (0.864)
Apoptosis agonist (0.723)
Antileukemic (0.822)
151Antineoplastic (0.843)
Apoptosis agonist (0.711)
Antifungal (0.733)
152Antineoplastic (0.929)
Antimetastatic (0.834)
Antileukemic (0.814)
153Antineoplastic (0.924)
Antimetastatic (0.876)
Antileukemic (0.754)
Antimicrobial (0.629)
154Antineoplastic (0.932)
Antimetastatic (0.839)
Antileukemic (0.724)
Antimicrobial (0.623)
* Only activities with Pa > 0.5 are shown.
Table 7. Predicted biological activities of steroids derived from plants (155186).
Table 7. Predicted biological activities of steroids derived from plants (155186).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
155Antifungal (0.778)Anti-inflammatory (0.709)
156Antineoplastic (0.918)
Prostate cancer treatment (0.915)
Antimetastatic (0.734)
Antifungal (0.734)
Antimicrobial (0.652)
157Antineoplastic (0.923)
Prostate cancer treatment (0.922)
Antimetastatic (0.834)
Antifungal (0.736)
Antimicrobial (0.655)
158Antineoplastic (0.888)
Apoptosis agonist (0.793)
Antifungal (0.713)
Antimicrobial (0.654)
159Antineoplastic (0.854)
Apoptosis agonist (0.721)
Antifungal (0.743)
Antimicrobial (0.664)
160Anti-tuberculosis treatment (0.896)Antibacterial (0.809)
161Anti-tuberculosis treatment (0.903)
Antimicrobial (0.745)
Antibacterial (0.811)
162Anti-tuberculosis treatment (0.937)
Antimicrobial (0.772)
Antibacterial (0.854)
163Anti-tuberculosis treatment (0.875)Antibacterial (0.754)
164Anti-tuberculosis treatment (0.862)Antibacterial (0.761)
165Acetyl-cholinesterase inhibitor (0.931)Antimetastatic (0.829)
166Acetyl-cholinesterase inhibitor (0.942)Antimetastatic (0.842)
167DNA topoisomerase inhibitor (0.928)
Antineoplastic (0.916)
Antimetastatic (0.863)
Apoptosis agonist (0.834)
168DNA topoisomerase inhibitor (0.917)
Antineoplastic (0.922)
Antimetastatic (0.821)
Apoptosis agonist (0.871)
169Anti-inflammatory (0.909)
Antiviral (0.844)
Antimicrobial (0.688)
170Anti-inflammatory (0.902)
Antiviral (0.829)
Antimicrobial (0.679)
171Antineoplastic (0.886)Apoptosis agonist (0.756)
172Antineoplastic (0.874)Apoptosis agonist (0.771)
173Antineoplastic (0.859)Apoptosis agonist (0.754)
174Antineoplastic (0.851)Apoptosis agonist (0.764)
175Antiproliferative (0.889)Anti-inflammatory (0.654)
176Antiproliferative (0.882)Anti-inflammatory (0.659)
177Antiproliferative (0.895)Anti-inflammatory (0.663)
178Anti-hepatitis C virus (0.912)
Antiviral (0.898)
Antimicrobial (0.687)
179Antineoplastic (0.873)Apoptosis agonist (0.775)
180Antineoplastic (0.869)Apoptosis agonist (0.768)
181Hepatocellular carcinoma inhibitor (0.897)Hepatoprotectant (0.860)
182Hepatocellular carcinoma inhibitor (0.906)Hepatoprotectant (0.873)
183Hepatocellular carcinoma inhibitor (0.911)Hepatoprotectant (0.869)
184Hepatocellular carcinoma inhibitor (0.889)Hepatoprotectant (0.856)
185Anti-inflammatory (0.832)
Antiviral (0.754)
Antibacterial (0.785)
186Anti-inflammatory (0.834)
Antiviral (0.765)
Antibacterial (0.744)
* Only activities with Pa > 0.5 are shown.
Table 8. Predicted biological activities of steroids derived from marine sources (187221).
Table 8. Predicted biological activities of steroids derived from marine sources (187221).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
187Antineoplastic (0.913)
Antimetastatic (0.786)
Apoptosis agonist (0.619)
188Antineoplastic (0.908)
Antimetastatic (0.804)
Apoptosis agonist (0.634)
189Antineoplastic (0.902)
Antimetastatic (0.798)
Apoptosis agonist (0.624)
190Cytoprotectant (0.768)
Antineoplastic (0.723)
Antimicrobial (0.776)
191Cytoprotectant (0.787)Antimicrobial (0.685)
192Cytoprotectant (0.766)Antimicrobial (0.692)
193Antineoplastic (0.900)
Apoptosis agonist (0.836)
Antimetastatic (0.698)
194Antineoplastic (0.889)
Apoptosis agonist (0.821)
Antimetastatic (0.682)
195Antineoplastic (0.872)Antiprotozoal (0.765)
196Antineoplastic (0.868)Antiprotozoal (0.756)
197Antiviral (0.776)Anti-inflammatory (0.682)
198Antiviral (0.749)Anti-inflammatory (0.678)
199Antineoplastic (0.773)Apoptosis agonist (0.653)
200Antibacterial (0.747)Antifungal (0.683)
201Antibacterial (0.736)Antifungal (0.651)
202Apoptosis agonist (0.833)Antineoplastic (0.711)
203Apoptosis agonist (0.824)Antineoplastic (0.683)
204Antineoplastic (0.744)Apoptosis agonist (0.621)
205Antineoplastic (0.738)Apoptosis agonist (0.619)
206Antiallergic (0.728)Anti-asthmatic (0.671)
207Antiallergic (0.734)Anti-asthmatic (0.698)
208Antiallergic (0.742)Anti-asthmatic (0.661)
209Antiallergic (0.741)Anti-asthmatic (0.676)
210Antineoplastic (0.926)Antimetastatic (0.855)
211Antineoplastic (0.914)Antimetastatic (0.836)
212Antiproliferative (0.818)Antineoplastic (0.654)
213Antiproliferative (0.828)Antineoplastic (0.662)
214Antiproliferative (0.833)Antineoplastic (0.659)
215Antiproliferative (0.867)Antineoplastic (0.678)
216Apoptosis agonist (0.818)
Antineoplastic (0.762)
Lipid metabolism regulator (0.638)
Anti-inflammatory (0.691)
217Antineoplastic (0.917)Antimetastatic (0.743)
218Antineoplastic (0.919)Antimetastatic (0.767)
219Antineoplastic (0.928)
Apoptosis agonist (0.854)
Antimetastatic (0.711)
220Antineoplastic (0.931)
Apoptosis agonist (0.871)
Antimetastatic (0.721)
221Antineoplastic (0.924)Antimetastatic (0.734)
* Only activities with Pa > 0.5 are shown.
Table 9. Predicted biological activities of steroids derived from marine sources (222238).
Table 9. Predicted biological activities of steroids derived from marine sources (222238).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
222Antiviral (0.788)
Antiviral (Arbovirus) (0.750)
Anti-inflammatory (0.682)
223Allergic conjunctivitis treatment (0.705)Antifungal (0.637)
224Antineoplastic (0.898)
Apoptosis agonist (0.834)
Antimetastatic (0.723)
225Antineoplastic (0.918)
Apoptosis agonist (0.856)
Antimetastatic (0.743)
226Cathepsin B inhibitor (0.889)Apoptosis agonist (0.698)
227Antifeedant (0.867)Antifungal (0.703)
228Protein kinase C inhibitor (0.921)
Antineoplastic (0.834)
Antimetastatic (0.748)
Antiviral (0.829)
Autoimmune disorders treatment (0.619)
229Antineoplastic (0.924)
Apoptosis agonist (0.745)
Antifungal (0.698)
230Antineoplastic (0.931)
Apoptosis agonist (0.805)
Anti-inflammatory (0.654)
231Antineoplastic (0.918)
Apoptosis agonist (0.764)
Anti-inflammatory (0.662)
232Antineoplastic (0.932)
Chemopreventive (0.738)
Apoptosis agonist (0.768)
233Antineoplastic (0.929)
Chemopreventive (0.787)
Apoptosis agonist (0.729)
234Antineoplastic (0.943)
Chemopreventive (0.818)
Apoptosis agonist (0.815)
235Antimicrobial (0.886)Antifungal (0.687)
236Antineoplastic (0.943)
Antimicrobial (0.922)
Apoptosis agonist (0.898)
Antifungal (0.847)
237Antibacterial (0.921)Antifungal (0.727)
238Antibacterial (0.765)
* Only activities with Pa > 0.5 are shown.
Table 10. Predicted biological activities of steroids derived from marine sources (239266).
Table 10. Predicted biological activities of steroids derived from marine sources (239266).
No.Dominated Biological Activity (Pa) *Additional Predicted Activities (Pa) *
239Antineoplastic (0.913)
Apoptosis agonist (0.886)
Antimetastatic (0.829)
240Antineoplastic (0.922)
Chemopreventive (0.731)
Anticarcinogenic (0.698)
Apoptosis agonist (0.876)
241Antineoplastic (0.929)
Chemopreventive (0.720)
Anticarcinogenic (0.678)
Apoptosis agonist (0.854)
242Antineoplastic (0.889)
Anticarcinogenic (0.718)
Chemopreventive (0.683)
243Antibacterial (0.881)Antifungal (0.683)
244Antibacterial (0.879)Antifungal (0.627)
245Antibacterial (0.902)Antifungal (0.638)
246Antineoplastic (0.938)
Apoptosis agonist (0.834)
Chemopreventive (0.712)
247Antineoplastic (0.886)Apoptosis agonist (0.726)
248Antineoplastic (0.887)Apoptosis agonist (0.734)
249Antineoplastic (0.889)Apoptosis agonist (0.722)
250Antineoplastic (0.900)Apoptosis agonist (0.783)
251Antineoplastic (0.911)Apoptosis agonist (0.803)
252Antineoplastic (0.935)
Apoptosis agonist (0.814)
Chemopreventive (0.756)
Antimetastatic (0.820)
253Antineoplastic (0.929)
Apoptosis agonist (0.821)
Chemopreventive (0.744)
Antimetastatic (0.822)
254Phospholipase A2 inhibitor (0.923)
Antibacterial (0.883)
Anti-inflammatory (0.689)
255Phospholipase A2 inhibitor (0.827)
Antibacterial (0.757)
Anti-inflammatory (0.683)
256Antineoplastic (0.884)
Apoptosis agonist (0.756)
Antimetastatic (0.688)
257Antineoplastic (0.891)
Apoptosis agonist (0.785)
Antimetastatic (0.679)
258Antineoplastic (0.814)Anti-inflammatory (0.661)
259Antineoplastic (0.828)Anti-inflammatory (0.658)
260Antineoplastic (0.832)Anti-inflammatory (0.650)
261Antineoplastic (0.915)
Antimetastatic (0.839)
Chemopreventive (0.672)
262Antineoplastic (0.912)
Antimetastatic (0.821)
Chemopreventive (0.669)
263Antineoplastic (0.910)
Apoptosis agonist (0.752)
Antimetastatic (0.712)
264Antineoplastic (0.921)
Apoptosis agonist (0.774)
Antimetastatic (0.729)
265Antineoplastic (0.905)
Apoptosis agonist (0.739)
Antimetastatic (0.704)
266Antineoplastic (0.918)
Apoptosis agonist (0.699)
Antimetastatic (0.687)
* Only activities with Pa > 0.5 are shown.
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

Dembitsky, V.M. Fascinating Furanosteroids and Their Pharmacological Profile. Molecules 2023, 28, 5669. https://doi.org/10.3390/molecules28155669

AMA Style

Dembitsky VM. Fascinating Furanosteroids and Their Pharmacological Profile. Molecules. 2023; 28(15):5669. https://doi.org/10.3390/molecules28155669

Chicago/Turabian Style

Dembitsky, Valery M. 2023. "Fascinating Furanosteroids and Their Pharmacological Profile" Molecules 28, no. 15: 5669. https://doi.org/10.3390/molecules28155669

Article Metrics

Back to TopTop