Next Article in Journal
Deciphering the Association among Pathogenicity, Production and Polymorphisms of Capsule/Melanin in Clinical Isolates of Cryptococcus neoformans var. grubii VNI
Next Article in Special Issue
Biosynthesis of Fungal Natural Products Involving Two Separate Pathway Crosstalk
Previous Article in Journal
Taxonomic Reappraisal of Periconiaceae with the Description of Three New Periconia Species from China
Previous Article in Special Issue
Polyketide Derivatives from the Endophytic Fungus Phaeosphaeria sp. LF5 Isolated from Huperzia serrata and Their Acetylcholinesterase Inhibitory Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diterpenes Specially Produced by Fungi: Structures, Biological Activities, and Biosynthesis (2010–2020)

School of Pharmaceutical Sciences, South-Central University for Nationalities, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
J. Fungi 2022, 8(3), 244; https://doi.org/10.3390/jof8030244
Submission received: 25 January 2022 / Revised: 25 February 2022 / Accepted: 27 February 2022 / Published: 28 February 2022

Abstract

:
Fungi have traditionally been a very rewarding source of biologically active natural products, while diterpenoids from fungi, such as the cyathane-type diterpenoids from Cyathus and Hericium sp., the fusicoccane-type diterpenoids from Fusicoccum and Alternaria sp., the guanacastane-type diterpenoids from Coprinus and Cercospora sp., and the harziene-type diterpenoids from Trichoderma sp., often represent unique carbon skeletons as well as diverse biological functions. The abundances of novel skeletons, biological activities, and biosynthetic pathways present new opportunities for drug discovery, genome mining, and enzymology. In addition, diterpenoids peculiar to fungi also reveal the possibility of differing biological evolution, although they have similar biosynthetic pathways. In this review, we provide an overview about the structures, biological activities, evolution, organic synthesis, and biosynthesis of diterpenoids that have been specially produced by fungi from 2010 to 2020. We hope this review provides timely illumination and beneficial guidance for future research works of scholars who are interested in this area.

1. Introduction

Fungi are widely distributed in terrestrial environments, freshwater, and marine habitats; more than one million distinctive fungal species exist, but only approximately 100,000 of these have been classified [1]. These eukaryotic microbes produce specialized metabolites that participate in a variety of ecological functions, such as quorum sensing, chemical defense, allelopathy, and maintenance of symbiotic interactions [2]. There are more than 40,000 terpenoid compounds in nature, which compose the largest family of natural products [3]. Terpenoids exist in all domains of life, but are particularly prevalent in plants, fungi, and marine invertebrates, and are essential constituents of secondary metabolism [3,4].
Diterpenoids are a class of C20 compounds derived from isoprenoid precursor geranylgeranyl diphosphate (GGPP) under the catalysis of diterpene synthases (DTSs) [5,6,7,8,9,10,11]. Prenyltransferase (PT) and terpene synthase (TPS) are key enzymes in the formation of the basic carbon skeletons of terpenoids [8,12]. The PT enzymes determine the prenyl carbon chain length, whereas the TPS enzymes generate the structural complexity of the molecular scaffolds, forming various ring structures [8]. Fungi are among the most important microbial resources for drug discovery, owing to their capability to produce structurally diverse and biologically important secondary metabolites [13,14]. It is also well known that fungi possess extraordinary biosynthetic gene clusters that may encode highly diverse biosynthetic pathways of natural products [15,16,17,18].
Between 2010 and 2020, about 400 fungal-specific diterpenes have been reported. In addition to 172 cyathane diterpenes reviewed by Bailly et al. [19] and Gao et al. [20], a total of 232 diterpenes were collected in this review (Chart 1). These diterpenoids are mainly tricyclic or tetracyclic skeletal structures such as cyathane-type, fusicoccane-type, guanacastane-type, and harziene-type diterpenoids (Chart 1). Judging from the distribution of fungal diterpenoid resources, the diterpenes from the genera Trichoderma, Penicillium, Cyathus, Hericium, and Crinipellis account for 60% of the total (Chart 2). In addition, systematic studies on the chemical constituents of fungi have shown that a large number of fungal diterpenoids exhibited significant biological functions such as anti-inflammatory, cytotoxic, antimicrobial, and antiviral activities (Chart 3). For instance, the semi-synthetic pleuromutilin analogues tiamulin 193 and valnemulin 194 have been used for over three decades as antibiotics to treat economically important infections in swine and poultry [21,22,23,24,25].
Consequently, a wealth of novel skeletons, biosynthetic pathways, and bioactivities have provided new opportunities for drug discovery, genome mining, enzymology, and chemical synthesis. During the period covered in this review, there have been several more specialized reviews of fungal metabolites [26,27,28], including benzene carbaldehydes [29], trichothecenes [30,31], protoilludane sesquiterpenoids [32], meroterpenoids [33,34,35], meroterpenoid cyclases [36], terpenoids [37], and natural product biosynthetic genes and enzymes of fungi [17,18,38,39]. In addition, the isolation and chemistry of diterpenoids from terrestrial sources have been summarized [40]. In this review, we provide an overview of diterpenoids that were specially produced by fungi during the period from 2010 to 2020, and focus on their structures, biological activities, and biosynthesis, and we also conduct an evolutionary analysis.
In particular, literature investigation of known databases such as PubMed and Web of Science was conducted from 2010 to July 2020 using the keywords “diterpenes/diterpenoids” paired with “fungi”, “fungal diterpenoids” paired with “structure elucidation”, or “fungal diterpenoids” paired with “biosynthesis”. There were no language restrictions imposed. The references were further scrutinized and, finally, 210 references were selected. The data inclusion criteria included: (1) diterpenes/diterpenoids isolated from fungi, (2) carbon skeleton obtained only from fungi or rarely from other sources, (3) studies on the biological activities of diterpenes/diterpenoids and their derivatives that had been carried out in vitro or in vivo, (4) studies on the biosynthesis of diterpenes/diterpenoids and their derivatives. The data exclusion criteria included: (1) carbon skeleton of diterpenes/diterpenoids obtained in abundance from other sources, such as plants, bacteria and so on, (2) duplication of data and titles and/or abstracts not meeting the inclusion criteria.

2. Cyathane

Jof 08 00244 i001
Cyathane diterpenes are a group of natural products that possess unusual, angularly fused 5/6/7 tricyclic cores, and they are characteristic of certain basidiomycete species including Cyathus, Hericium, and Sarcodon (Figure 1). For example, there have been more than 170 compounds isolated from fungi such as Cyathus africanus and Hericium erinaceus [19,20,41]. These compounds have a common biosynthetic precursor and can be produced via biosynthesis, hemi-synthesis, or total synthesis [42,43,44,45,46,47]. The cyathane diterpenoids include the classes of cyathins, striatins, sarcodonins, scabronines, and erinancines, according to their origins. Among them, the striatals, striatins, and erinacines, called cyathane-xylosides, which represent an unusual group of cyathane diterpenoids attached to a modified pentose (D-xylose) moiety, have been isolated from cultures of Cyathus and Hericium species [20]. The cyathane structure is different from the homoverrucosane, mulinane, and valparane diterpenoids which also possess a 5/6/7 tricarbocyclic system [48,49]. The cyathanes are most similar to cyanthiwigins and can be differentiated by the orientation of the angular methyl groups, mainly present in some sponges [50,51,52,53,54,55,56,57,58,59]. These compounds display a diverse range of biological activities, including anticancer, antimicrobial, anti-MRSA (methicillin-resistant Staphylococcus aureus), anti-inflammatory, anti-proliferative, and nerve growth factor (NGF)-like properties [19,20,60,61]. An overview of cyathane-type diterpenes including isolation, structure diversity, synthesis, and bioactivity has been reviewed by Bailly et al. [19] and Gao et al. [20]. Therefore, in this review, we no longer summarize the details of cyathane diterpenoids.
To understand the source genera of cyathane diterpenoids, we performed a phylogenetic analysis by using the maximum likelihood method and the general time reversible model [62,63,64] for all the species involved in the reviews by Bailly et al. [19] and Gao et al. [20]. The results show that source genera are grouped based on their regiospecificity, i.e., genera Cyathus, Hericium, and Sarcodon were clustered into different clades (Figure 1). Taxonomically, Cyathus africanus, C. hookeri, C. gansuensis, C. subglobisporus, C. stercoreus, and C. striatus all belonged to the genus Cyathus. They were close to each other, and first, they gathered into one branch, then, they gathered into one branch with Strobilurus tenacellus of the genus Strobilurus, and finally gathered into one branch with other genera (Figure 1). C. earlei and C. helenae also belonged to the genus Cyathus, they were close to each other, and first, they gathered into one branch, then, they gathered into one branch with Gerronema albidum of the genus Gerronema. Similarly, Hericium erinaceus, H. flagellum, and Hericium sp. WBSP8, Sarcodon scabrosus, S. glaucopus, and other species were close to each other. Existing studies have shown that most fungal metabolites are encoded by biosynthetic gene clusters (BGCs) [17]. The natural product BGCs of species in the same genus tend to be highly homologous, and BGC functional divergence gives rise to the evolution of new secondary metabolites, indicating that species-level sampling in these three genera for natural products mining will yield significant returns for cyathane diterpenoids discovery.

3. Cyclopiane

Jof 08 00244 i002
Cyclopiane diterpenoids comprise a class of tetracyclic diterpenes with unique scaffolds. They are characterized by a highly fused 6/5/5/5 ring system. The structural variations of cyclopiane diterpenoids are mainly owing to oxidation occurring at various sites to generate hydroxy groups [65]. In general, cyclopiane diterpenoids have mainly been isolated from different species of the genus Penicillium (Figure 2) and have been classified into two groups according to the functionality at C-1, i.e., conidiogenols and conidiogenones. The former featured with a hydroxy group at C-1, while the later possessed a carbonyl group at C-1 [66]. Specifically, Penicillium commune MCCC 3A00940, P. sp. F23-2, P. sp. YPGA11, P. cyclopium, P. roqueforti IFM 48062, P. sp. TJ403-2, P. chrysogenum QEN-24S, and Leptosphaeria sp. XL026 have been reported to produce conidiogenol-type diterpenoids, while P. commune MCCC 3A00940, P. chrysogenum MT-12, P. sp. YPGA11, and P. cyclopium have been reported to produce conidiogenone-type diterpenoids (Figure 2). Structurally, cyclopiane diterpenoids differ from the aberrarane-type diterpenoid aberrarone, which has shown in vitro antimalarial activity against a chloroquine-resistant strain of the protozoan parasite Plasmodium falciparum isolated from the Caribbean sea whip Pseudopterogorgia elisabethae [67]. The molecular structure of aberrarone was established by spectral analysis and subsequently confirmed by X-ray crystallographic analysis. Some cyclopiane compounds exhibited pronounced biological activities, such as conidiation induction, cytotoxic, anti-inflammatory, antimicrobial, and antiallergic effects.
Jof 08 00244 i003

3.1. Conidiogenol Type

Conidiogenol 1 is a potent and selective inducer of conidiogenesis in the liquid culture of Penicillium cyclopium under non-nutrient limiting conditions [66]. Conidiogenol B 2 has been obtained from the deep-sea derived fungus P. commune MCCC 3A00940 [68]. Conidiogenols C 3 and D 4 have been isolated from a deep-sea derived fungus P. sp. YPGA11 [65].
The absolute structure of cyclopiane diterpenoids was first confirmed by Abe and co-workers, in 2018, with the aid of the crystal sponge method [69]. Using the genome-mining approach, a chimeric enzyme of prenyltransferase-diterpene synthase (PT-TS) discovered from P. chrysogenum MT-12 was designated as P. chrysogenum cyclopiane-type diterpene synthase (PcCS). The new diterpene alcohol metabolite 5 was produced after the gene heterologously expressed in Aspergillus oryzae, and the crystalline sponge method also revealed the absolute configuration of 5 [69]. The PT domain of PcCS first generated geranylgeranyl diphosphate (GGPP) from dimethylallyl pyrophosphate (DMAPP) and isopentenyl pyrophosphate (IPP) (Scheme 1A). Then, GGPP was converted into 5 by a cyclization reaction catalyzed by the TS domain of PcCS (Scheme 1B).

3.2. Conidiogenone Type

Conidiogenone 6, first isolated from Penicillium cyclopium, was also an inducer of conidiation [66,70]. The biosynthetic pathway of (–)-conidiogenone 6 has been fully elucidated by the heterologous expression of biosynthetic genes in Aspergillus oryzae and by in vitro enzyme assay with 13C-labeled substrates [71]. After construction of deoxyconidiogenol by the action of bifunctional terpene synthases (PchDS gene obtained from Penicillium chrysogenum, and PrDS gene identified from Penicillium roqueforti showed significant homology to PchDS), one cytochrome P450 catalyzed two rounds of oxidation to furnish conidiogenone 6. The cyclization mechanism catalyzed by terpene synthase, involving successive 1,2-alkyl shifts, was fully elucidated using 13C-labeled geranylgeranyl pyrophosphate (GGPP) as a substrate (Scheme 2).
A series of new conidiogenone-type diterpenoids have been obtained from several Penicillium species including conidiogenones B–G 712 from the fungus P. sp. F23-2 [72], conidiogenones H 13 and I 14 from the endophytic fungus P. chrysogenum QEN-24S [73], conidiogenones J 16 and K 15 from the fungus P. commune [68], and conidiogenone L 17 from P. sp. YPGA11 [65]. Conidiogenone B 7 showed potent activity against methicillin-resistant Staphylococcus aureus (MRSA), Pseudomonas fluorescens, P. aeruginosa, and Staphylococcus epidermidis (each with a MIC value of 8 μg/mL) [73]. Conidiogenone C 8 showed potent cytotoxicity against HL-60 and BEL-7402 cell lines with IC50 values of 0.038 and 0.97 μM, and conidiogenone G 12 showed potent cytotoxicity against HL-60 cell line with an IC50 value of 1.1 μM [72]. Provoked by the novelty of structures and potent bioactivities, total syntheses of 1, 6, and 7 were achieved, which led to further determination of their absolute configurations [74].
Three new cyclopiane diterpenes 13β-hydroxy conidiogenone C 18 and 12β-hydroxy conidiogenones C 19 and D 20 have been isolated and identified from a sea sediment-derived fungus Penicillium sp. TJ403-2 [75]. Their absolute configurations were further established by X-ray crystallography experiment. Compounds 1820 were evaluated for their anti-inflammatory activity against LPS-induced NO production, and compound 18 showed notable inhibitory potency with an IC50 value of 2.19 μM, which was three-fold lower than the positive control indomethacin (IC50 8.76 μM). Further Western blot and immunofluorescence experiments demonstrated that 18 inhibited the NF-κB-activated pathway.
Leptosphin C 21 has been isolated from the solid cultures of an endophytic fungus Leptosphaeria sp. XL026 [76]. Its structure was elucidated by extensive spectroscopic methods and single-crystal X-ray diffraction.

4. Fusicoccane

4.1. Structural and Biological Diversity

Fusicoccane diterpenoids, characterized by 5/8/5, 5/8/6, 5/9/4, and 5/9/5 fused carbocyclic ring systems, include the fusicoccins, cotylenins, brassicicenes, heterodimers, and homodimers [77,78,79,80]. They were first isolated as glycosides from the phytopathogenic fungus Fusicoccum amygdali, in 1964 [81]. Substances exhibiting this structural motif have been isolated from a variety of sources including fungi such as Talaromyces purpureogenus, Alternaria brassicicola XXC, and Trichoderma citrinoviride cf-27 (Figure 3), and rarely from liverworts, algae, ferns, streptomycetes, and higher plants, some of which showed remarkable biological effects relevant for drug discovery, such as antibacterial, antitumor, anti-inflammatory, and antifungal activities [82,83,84,85,86,87,88,89].
Jof 08 00244 i004
Talaronoids A–D 2225, four diterpenoids with an unexpected tricyclic 5/8/6 carbon skeleton isolated from Talaromyces stipitatus, represent a new class of fusicoccane diterpenoids with a benzo[a]cyclopenta[d]cyclooctane skeleton [80]. Plausible biosynthetic pathways of talaronoids A–D 2225 have been proposed starting from geranylgeranyl diphosphate with a Wagner–Meerwein rearrangement as the key step (Scheme 3). Talaronoids A–D 2225 have shown moderate butyrylcholinesterase (BCHe) inhibitory activity with IC50 values of 14.71, 26.47, 31.51, and 11.37 μM, respectively. A new diterpenoid roussoellol C 26 that exhibited moderate antiproliferative activities against human breast adenocarcinoma (MCF-7) cell line with an IC50 value of 6.5 μM has been isolated from an extract of laboratory cultures of the marine-derived fungus Talaromyces purpurogenus [90]. Roussoellols A 27 and B 28 have been isolated from the plant-inhabiting ascomycetous fungus Roussoella hysterioides KT1651 [91]. Compound 28 inhibited the hyphal growth of the phytopathogen Cochliobolus miyabeanus at 10 μg/mL.
Six new fusicoccane-type diterpenoids, 14-hydroxycyclooctatin 30, 12α-hydroxycyclooctatin 31, 12β-hydroxycyclooctatin 32, fusicomycin A 33, fusicomycin B 34, and isofusicomycin A 35, along with a known compound, cyclooctatin 29 [82,92], have been isolated from the fermentation broth of Streptomyces violascens [93]. Compounds 3335 have demonstrated cytotoxicity against five human cancer cell lines (BGC823, H460, HCT116, HeLa, and SMMC7721), with IC50 values ranging from 3.5 to 14.1 μM. Cell adhesion, migration, and invasion assays have shown that fusicomycin B 34 inhibited the migration and invasion of human hepatocellular carcinoma SMMC7721 cells in a dose-dependent manner. Through further investigation, it was revealed that 34 inhibited the enzymatic activity of matrix metalloproteinase-2 (MMP-2) and matrix metalloproteinase-9 (MMP-9), in addition to downregulating the expressions of MMP-2 and MMP-9 at both the protein and mRNA levels to influence the migration and invasion of cancer cells.
Jof 08 00244 i005
Between 1999 and 2014, eleven new fusicoccane-like diterpenoids were isolated from the phytopathogenic fungus Alternaria brassicicola [94,95,96]. With the aid of computational predictions, experimental validation, and biosynthetic logic-based strategies, Zhang and co-workers first rectified the conclusion that all brassicicenes were originally proposed to have a 5/8/5 fused skeleton and, thus, reassigned brassicicenes C–H 3641, J 42, and K 43 to have a unique bridgehead double-bond-containing 5/9/5 fused skeleton [97]. Meanwhile, brassicicenes L–N 4446 were three highly modified fusicoccanes also isolated from the fungus Alternaria brassicicola [97]. Afterward, alterbrassicene A 47 [78] and alterbrassicicene A 48 [98], two unprecedented fusicoccane-derived diterpenoids featuring a 5/9/4-fused carbocyclic skeleton and a newly transformed monocyclic carbon skeleton (Scheme 4), respectively, were obtained from the same fungal strain and found to function on different targets in the NF-κB signaling pathway of anti-inflammatory activity. Later, the biogenetically related intermediates, brassicicenes O 49 and P 50, were also discovered [78].
Brassicicenes Q–X 5158 have been isolated from the phytopathogenic fungus Alternaria brassicicola [99]. Brassicicene S 53 was found to show significant anti-inflammatory activity against the production of NO, TNF-α, and IL-1β at 10 μM. Further Western blot and immunofluorescence experiments found the mechanism of 53 inhibiting the NF-κB-activated pathway.
Jof 08 00244 i006
Seven new modified fusicoccane-type diterpenoids 5965, together with two known congeners, have been obtained from A. brassicicola [100]. Alterbrassicicenes B 60 and C 62 represented the first examples of fusicoccane-type diterpenoids featuring two previously undescribed tetracyclic 5/6/6/5 ring systems, while 1β,2β-epoxybrassicicene I 63 featured a previously undescribed tetracyclic 5/8/5/3 ring system. Alterbrassicicene E 65 showed moderate anti-inflammatory activity against NO production in lipopolysaccharide (LPS)-induced RAW264.7 cell with an IC50 value of 24.3 μM. In addition, alterbrassicicene B 60, 3-ketobrassicicene W 61, 1β,2β-epoxybrassicicene I 63, and alterbrassicicene E 65 exerted weak cytotoxicity against certain human tumor cell lines (OCVAR, MDA-MB-231, HeLa, HT-29, and Hep3B cells) with IC50 values ranging from 25.0 to 38.2 μM.
Jof 08 00244 i007
Alterbrassinoids A–D 6669, the first examples of fusicoccane-derived diterpene dimers furnished by forming an undescribed C-12–C-18′ linkage, have been isolated from modified cultures of Alternaria brassicicola [79]. Alterbrassinoids A 66 and B 67 represented unprecedented heterodimers, whereas alterbrassinoids C 68 and D 69 represented unprecedented homodimers, and alterbrassinoid D 69 also featured an undescribed anhydride motif. Alterbrassinoids A–D 6669 showed moderate activities against five cancer cells (including OCVAR, MDAMB-231, HeLa, HT-29, and Hep3B). Afterward, three rearranged fusicoccane diterpenoids bearing a rare bridgehead double-bond-containing tricyclo[9.2.1.03,7]tetradecane (5/9/5 ring system) core skeleton, namely alterbrassicenes B–D 7072, were obtained from the same fungus A. brassicicola [101]. Their structures were assigned via spectroscopic methods, ECD calculations, and single-crystal X-ray diffraction. Compounds 7072 showed moderate cytotoxicity against several human tumor cell lines with IC50 values ranging from 15.87 to 36.85 μM.
Five new diterpenoid glycosides, dongtingnoids A–E 7377, two new diterpenoid aglycones, dongtingnoids F 78 and G 79, and two known analogues, cotylenins E and J, belonging to the fusicoccane family, have been isolated from the fungus Penicillium sp. DT10 [102]. Dongtingnoids A 73, D 76, and E 77 showed comparable seed-germination-promoting activities to the growth regulator cotylenin E [103,104]. Such diterpene glucosides have been used for the production of an intermediate compound suitable for semi-synthesis by a mutant constructed by disruption of a specific gene by homologous recombination [105,106].
Trichocitrin 80, representing the first Trichoderma-derived and furan-bearing fusicoccane diterpene, has been isolated from the culture of marine brown alga-endophytic Trichoderma citrinoviride [107]. A new class of fusicoccane-type diterpenoid alkaloids with an unusual 5/5/8/5 tetracyclic system, i.e., pericolactines A–C 8183, have been isolated from Periconia sp. [108].

4.2. Biosynthesis of Fusicoccane Diterpenes

A unique chimeric enzyme PaFS, possessing both a geranylgeranyl diphosphate (GGDP) synthase domain and a diterpene cyclase domain, has been identified from Phomopsis amygdali [109]. A biosynthetic gene cluster of brassicicene C 36, a fusicoccadiene synthase (AbFS) containing 11 genes (orf1 to orf11, Scheme 5A), has been identified in Alternaria brassicicola ATCC 96836 from genome database search [110,111]. In vivo and in vitro studies have clearly revealed the function of Orf8 and Orf6 as a fusicoccadiene synthase similar to PaFS and methyltransferase, respectively. In this gene cluster, five genes (orf1, orf2, orf5, orf7, and orf11) encoded cytochrome P450s. Orf9 was a key dioxygenase to determine the aglycon structures of fusicoccin and brassicicene [112].
Other fusicoccane-type diterpene synthases have been identified from bacteria or fungus, such as CotB2 from bacteria responsible for the biosynthesis of cyclooctat-9-en-7-ol 84 [113], and SdnA from fungus responsible for the biosynthesis of cycloaraneosene 85 [114]. The same 5/8/5 tricyclic skeleton occurred in the sesterterpene ophiobolin F for which the terpene synthase AcOS has been reported from Aspergillus clavatus [115]. Oikawa and co-workers applied the Aspergillus oryzae heterologous expression system to functionally characterize cryptic bifunctional terpene synthase genes found in fungal genomes and identified the sesterfisherol (contains a characteristic 5/6/8/5 tetracyclic system) synthase gene (NfSS) from Neosartorya fischeri [116].
A unique P450 enzyme bscF has been identified in the phytopathogen Pseudocercospora fijiensis that generated two structurally different products from the single substrate. In addition to the heterologous expression of the eight genes, bscA-bscH elucidated the biosynthetic pathway for brassicicenes (Scheme 5B) [117].
A new fusicoccane-type diterpene synthase MgMS has been identified from the fungus Myrothecium graminearum by the genome mining method, which catalyzed the formation of the new diterpene alcohol myrothec-15(17)-en-7-ol 86 with all the seven stereocenters being introduced in the cyclization steps and conserved in the structure of the product. Based on this, its novel cyclization mode was unambiguously assigned (Scheme 6) [118].

5. Guanacastane

The discovery of 5/7/6 ring-fused guanacastane diterpenoids has been limited to several fungal species in the different genera Cercospora, Cortinarius, Coprinellus, Coprinus, Psathyrella, and Verticillium (Figure 4). Coprinellus heptemerus and C. radians M65 belong to the same genus, and first, they gather into one branch. Psathyrella candolleana and Cercospora sp. gather into one branch although they come from different genera. They are all able to produce guanacastane diterpenoids, indicating that highly homologous BGCs may also exist in fungi of different genera. The first member guanacastepene A 88, a new diterpene antibiotic against methicillin-resistant Staphylococcus aureus (MRSA) and vancomycin-resistant Enterococcus faecalis (VREF), has been isolated from an unidentified endophytic fungus [119]. Meanwhile, fourteen new analogues guanacastepenes B–O 89102 have been isolated from the same resource [120]. The novel skeleton has attracted great interests for organic synthesis [121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136]. The biological activities of guanacastanes have mainly been reported to possess cytotoxicity and antimicrobial effects.
Jof 08 00244 i008
Heptemerones A–G 103109 have been isolated from cultures of Coprinus heptemerus [137,138]. Radianspenes A–M 110122 have been obtained from Coprinus radians [139]. Among the biological activities of these isolates, the inhibition of fungal germination was the most potent, and depended highly on the composition of the assay medium [137]. Radianspene C 112 showed inhibitory activity against human breast carcinoma (MDA-MB-435) cell with an IC50 value of 0.91 μM [139]. Investigation of secondary metabolites from the fungal Coprinus plicatilis led to the discovery of several new guanacastane-type diterpenoids, named plicatilisins A–D 123126 [140] and E–H 127130 [141]. In vitro cytotoxic activities against the human cancer cell lines (HepG2, HeLa, MDA-MB-231, BGC-823, HCT 116, and U2OS) showed that plicatilisin A 123 exhibited significant cytotoxicity with IC50 values ranging from 1.2 to 6.0 μM [140].
Guanacastepenes P–T 131135 have been isolated from cultures of the fungus Psathyrella candolleana [142]. Guanacastepene R 133 exhibited inhibitory activity against both human and mouse isozymes of 11β-hydroxysteroid dehydrogenase (11β-HSD1) with IC50 values of 6.2 and 13.9 μM, respectively. Cercosporenes A–F 136141, including two homodimers 140 and 141, have been isolated from the fungus Cercospora sp. [143]. Cercosporene F 141 was cytotoxic to five human tumor cell lines (HeLa, A549, MCF-7, HCT116, and T24) with IC50 values of 8.16–46.1 μM, and induced autophagy in HCT116 cell.
Jof 08 00244 i009
Eleven new guanacastane-type diterpenoids dahlianes A–K 142152 have been obtained from the fungus Verticillium dahlia that was isolated from the gut of insect Coridius chinensis [144,145]. In the cytotoxicity evaluation against human tumor cell lines, dahlianes B 143 and C 144 exhibited significant cytotoxicity against human breast cancer cell MCF-7 with IC50 values of 3.35 and 4.72 μM, respectively [144]. In addition, the isolates were evaluated for their cytotoxicity toward drug-sensitive and DOX resistant MCF-7 cells by MTT assay. As a result, dahliane G 148 showed an 80-fold potentiation effect on the sensitization of doxorubicin at the concentration of 15 μM when screening the reversal activity on doxorubicin-resistant human breast cancer cell (MCF-7/DOX) [145].
Pyromyxones A–D 153156 have been isolated from fruiting bodies of Cortinarius pyromyxa, which possessed an undescribed nor-guanacastane skeleton of a 5/7/6 tricyclic system [146]. Pyromyxones A 153, B 154, and D 156 exhibited weak activity against Gram-positive Bacillus subtilis and Gram-negative Aliivibrio fischeri, as well as the phytopathogenic fungi Botrytis cinerea, Septoria tritici, and Phytophthora infestans [146].
Jof 08 00244 i010

6. Harziene

Harziene is a small group of diterpenoids that have a unique 4/7/5/6 tetracyclic scaffold. They have mainly been obtained from different Trichoderma species and rarely from liverworts [147]. Harziandione 157 was the first harziene diterpenoid isolated from the liquid culture of T. harzianum, in 1992 [148]. Harzianone 158, a new harziene diterpene, has been isolated from an alga-endophytic isolate of T. longibrachiatum [149]. The structure with absolute configuration of 158 was unambiguously identified by NMR and mass spectrometric methods as well as quantum chemical calculations. In addition, the absolute configuration of harziandione 157 was supported by optical rotation calculation, and the structure of isoharziandione isolated from culture filtrate of a strain of Trichoderma viride [150] was revised to harziandione 157 on the basis of 13C NMR data comparison and calculation.
Jof 08 00244 i011
The terpene cyclization mechanism of harzianone 158 has been studied by feeding experiments using selectively 13C- and 2H-labeled synthetic mevalonolactone isotopologues, followed by the analysis of the incorporation patterns of 13C NMR spectroscopy and GC/MS, and the structure of harzianone 158 was further supported from a 13C-13C COSY experiment of the in vivo generated fully 13C-labeled diterpenoid (Scheme 7) [151].
Four new harziene-related compounds 159162 have been isolated from an endophytic fungus Trichoderma atroviridae UB-LMA [152]. Among them, 159 is a potential derivative of geranylgeranyl diphosphate and may represent the biosynthetic precursor of this scarce family of compounds (Scheme 7). Recently, the first total synthesis of nominal harziene diterpenoid 160 has been achieved; stereochemical analysis and subsequent synthesis of the epimeric tertiary alcohol led to the reassignment of configuration for compound 160 as shown for harzianol I 180 [153].
(9R,10R)-Dihydro-harzianone 163 and harzianelactone 164 have been isolated from the endophytic fungus Trichoderma sp. Xy24 [154]. Compound 163 was the reductive product of harzianone 158 while 164 possessed a 6/5/7/5-fused ring core containing a lactone. The latter was the Baeyer–Villiger monooxygenase catalyzed oxidation product of harzianone 158. Compound 163 exhibited cytotoxicity against HeLa and MCF-7 cell lines with IC50 values of 30.1 and 30.7 μM, respectively.
3R-Hydroxy-9R,10R-dihydroharzianone 165 has been isolated from an endophytic fungus Trichoderma harzianum X-5 [155]. 11-Hydroxy-9-harzien-3-one 166, isolated from T. asperellum cf44-2, showed inhibitory activity against pathogenic bacteria Vibrio parahaemolyticus with a 6.2 mm zone [156]. 3S-Hydroxyharzianone 167, isolated from T. asperellum A-YMD-9-2, could highly inhibit four marine phytoplankton species (Chattonella marina, Heterosigma akashiwo, Karlodinium veneficum, and Prorocentrum donghaiense) with the IC50 values ranging from 3.1 to 7.7 μg/mL [157]. Deoxytrichodermaerin 168, a harziene lactone possessing potent inhibition against the four phytoplankton species (C. marina, H. akashiwo, K. veneficum, and P. donghaiense), has been obtained from an endophyte Trichoderma longibrachiatum A-WH-20-2 [158].
Two new harziene diterpene lactones, i.e., harzianelactones A 169 and B 170, and five new ones, i.e., harzianones A–D 171174 and harziane 175, have been isolated from the soft coral-derived fungus Trichoderma harzianum XS-20090075 [159]. These compounds exhibited potent phytotoxicity against seedling growth of amaranth and lettuce. Harzianone E 176, which exhibited weak antibacterial activity against Photobacterium angustum, has been obtained from the culture of coral-derived fungus T. harzianum treated with 10 µM sodium butyrate [160]. Harzianols F–J 177181 and three known derivatives have been obtained from the liquid fermentation of an endophytic fungus T. atroviride B7 [161]. Among them, compound 180 exhibited significant antibacterial effect against Staphylococcus aureus, Bacillus subtilis, and Micrococcus luteus with EC50 values of 7.7, 7.7, and 9.9 μg/mL, respectively. Meanwhile, cytotoxic activity of 180 against three cancer cell lines was also observed [161].
Furanharzianones A 182 and B 183 are two new harziene-type diterpenoids with an unusual 4/7/5/6/5 ring system, while harzianols A–E 184188 and harziane acid 189 are six new oxidized derivatives of harzianone [162,163]. These compounds have all been obtained from microbial transformation by the bacterial strain Bacillus sp. IMM-006.
Jof 08 00244 i012

7. Phomopsene

Structure determination of the novel diterpene hydrocarbon phomopsene 190 has been provided by enzymatic synthesis with the recombinant terpene synthase PaPS from Phomopsis amygdali, and screening of fungal broth extracts regarding characteristic NMR signals of phomopsene 190 resulted in the isolation of a new diterpene, methyl phomopsenonate 191 (Scheme 8) [164].
The cyclization mechanism of tetracyclic diterpene phomopsene 190 with phomopsene synthase (PaPS) has been examined through systematically deuterium-labeled geranylgeranyl diphosphate (GGPP), starting from site-specific deuterium-labeled isopentenyl diphosphates (IPPs) using IPP isomerase and three prenyltransferases (Scheme 9) [165].
Otherwise, other phomopsene synthases have been identified from actinomycetes such as Allokutzneria albata (PmS), Nocardia testacea (NtPS), and Nocardia rhamnosiphila (NrPS) [166,167]. All enzymes were subjected to in-depth mechanistic studies involving isotopic labeling experiments, metal-cofactor variation, and site-directed mutagenesis.
Jof 08 00244 i013

8. Pleuromutilin

Pleuromutilin 192 is a diterpene with a tricyclic skeleton possessing antimicrobial properties. It was first discovered from two basidiomycete fungal species including Pleurotus mutilis (synonymous to Clitopilus scyphoides f. mutilus) and Pleurotus passeckerianus (synonymous to Clitopilus passeckerianus) [168], and then produced by a number of other related species [169]. Its chemical structure and cyclisation mechanism has been elucidated by independent works [170,171,172], while total synthesis has been achieved by [173,174]. The semi-synthetic pleuromutilin analogues tiamulin 193 and valnemulin 194 have been used for over three decades to treat economically important infections in swine and poultry without showing any significant development of resistance in their target bacteria [21,22,23,24,25]. In recent years, extensive research including structure–activity relationship studies have been conducted to generate new orally available pleuromutilin derivatives having been used systemically in human medicine to treat acute bacterial skin and skin structure infections, as well as multidrug-resistant tuberculosis [175,176,177,178].
The gene cluster for the antibiotic pleuromutilin 192 has been isolated in Clitopilus passeckerianus [179]. Total de novo biosynthesis of pleuromutilin 192 was achieved through the expression of the entire gene cluster in the secondary host Aspergillus oryzae, proving that the seven genes isolated were sufficient for biosynthesis of the diterpene antibiotic. Heterologous expression of genes from the pleuromutilin gene cluster in A. oryzae revealed the biosynthesis of the antibiotic pleuromutilin 192 (Scheme 10) and generated bioactive semi-synthetic derivatives [180].
Jof 08 00244 i014

9. Sordaricin

Sordarin 195, an antifungal antibiotic possessing a unique 5/6/5/5-fused ring system, was discovered in 1971 as a metabolite of Sordaria araneosa [181]. A number of related semisynthetic sordarin derivatives have also been reported and some have been developed as antifungal agents such as zofimarin 196, hypoxysordarin (FR231956) 197, and FR290581 198 [182,183,184,185,186]. Sordarin 195 and related compounds have been shown to inhibit protein synthesis by a mechanism involving selective binding to the elongation factor 2 (EF-2) and ribosome complex in fungi [187,188,189].
Sordarins C–F 199202, possessing a unique 5/6/5/5 or 5/6/5/5/3 ring system varied at the C-11 position and the branch attached to C-12 of the sordaricin-type diterpene skeleton, have been isolated from the fungus Xylotumulus gibbisporus [190]. Genome mining of the sordarin biosynthetic gene cluster from Sordaria araneosa has been carried out, and the results suggest that the identified sdn gene cluster is responsible for the biosynthesis of sordarin 195 and hypoxysordarin 197 (Scheme 11) [114].

10. Tetraquinane

Several antibiotic crinipellin-related diterpenoids containing a 5/5/5/5 tetraquinane skeleton have been obtained from the basidiomycetous fungus Crinipellis stipitaria [191,192]. Up to now, the total synthesis of (±)-crinipellin B 203 and (–)-crinipellin A 204 have been reported [186,193,194,195].
Four novel diterpenoids, namely (4β)-4,4-O-dihydrocrinipellin A 205, (4β,8α)-4,4-O,8,8-O-tetrahydrocrinipellin B 206, crinipellins C 207 and D 208, along with three known diterpenoids have been isolated from the fungus Crinipellis sp. 113 [196]. Antitumor assays demonstrated that the compounds possess moderate activities against HeLa cell.
Four new tetraquinane diterpenoids crinipellins E–H 209212 have been isolated from fermentations of a Crinipellis species [197]. Crinipellins E–G 209211 inhibited the LPS/IFN-γ induced CXCL10 promoter activity in transiently transfected human MonoMac6 cell in a dose-dependent manner with IC50 values of 15, 1.5, and 3.15 μM, respectively. Moreover, crinipellins E–G 209211 reduced mRNA level and synthesis of proinflammatory mediators such as cytokines and chemokines in LPS/IFN-γ stimulated MonoMac6 cell.
A new crinipellin derivative crinipellin I 213 together with the known crinipellin A 204 have been obtained from the fungus Crinipellis rhizomaticola [198]. Crinipellin A 204 exhibited a wide range of antifungal activity in vitro against Colletotrichum coccodes, Magnaporthe oryzae, Botrytis cinerea, and Phytophthora infestans (MICs of 1, 8, 31, and 31 µg/mL, respectively).
Jof 08 00244 i015

11. Others

11.1. Spirograterpene

A novel spiro-tetracyclic diterpene featuring a 5/5/5/5 spirocyclic carbon skeleton, i.e., spirograterpene A 214, has been isolated from the deep-sea-derived fungus Penicillium granulatum [199]. Spiroviolene 215, bearing the same carbon skeleton to that of 214, has been obtained from a bacterial terpene synthase [200]. Spirograterpene A 214 showed an antiallergic effect on immunoglobulin E (IgE)-mediated rat mast RBL-2H3 cell with 18% inhibition as comparedwitho 35% inhibition for the positive control (loratadine) at the same concentration of 20 μg/mL [199].
Jof 08 00244 i016

11.2. Psathyrin

Two skeletally novel tetracyclic diterpenoids that possess a novel 5/5/4/6 tetracyclic system, psathyrins A 216 and B 217, have been characterized from cultures of the basidiomycete Psathyrella candolleana. They displayed weak antibacterial activities against Staphylococcus aureus and Salmonella enterica. The biosynthetic pathway of 216 and 217 was proposed to start from GGPP and the final products were obtained through a series of reactions (Scheme 12) [201].

11.3. Coicenal

Coicenals A–D 218221, possessing a previously undescribed 6/6 fused carbon skeleton, have been isolated from the solid culture of the plant pathogenic fungus Bipolaris coicis [202]. Coicenals A 218 and B 219 could be transformed into 221 and compound 222 by treatment with acetyl chloride, respectively. Coicenals A–D 218221 showed moderate inhibitory activity against NO release with IC50 values of 16.34, 23.55, 10.82, and 54.20 µM, respectively.
Jof 08 00244 i017

11.4. Eryngiolide

Eryngiolide A 223 has been isolated from the solid culture of the edible mushroom Pleurotus eryngii [203]. It is the first member of diterpenoids with a novel skeleton deriving from a cyclododecane core fused with two γ-lactone units [203]. It has exhibited moderate cytotoxicity against two human cancer cell lines (Hela and HepG2) in vitro. Biogenetically, eryngiolide A 223 could be the first diterpene not synthesized from GGPP unit, which indicated a completely new route for diterpene biosynthesis in nature (Scheme 13).

11.5. Trichodermanin

Trichodermanin A 224, a structurally unique diterpenoid with skeletal carbons arranged compactly in a 6/5/6/6 ring system, has been isolated from cultures of Trichoderma atroviride [204]. Its absolute configuration was elucidated by single crystal X-ray diffraction. Wickerols A 225 and B 226 were two novel diterpenoids produced by Trichoderma atroviride and the absolute configuration of 226 was confirmed by X-ray crystallographic analysis [205,206]. Wickerol A 225 showed potent antiviral activity against the A/H1N1 flu virus (A/PR/8/34 and A/WSN/33 strains) with an IC50 value of 0.07 μg/mL, but not active against the A/H3N2 virus. Wickerol B 226 also showed anti-influenza virus activity against A/PR/8/34 virus with an IC50 value of 5.0 μg/mL [206].
Jof 08 00244 i018
The new skeleton of wickerols A 225 and B 226 was revealed by the feeding experiments of [1-13C]-, [2-13C]-, and [1,2-13C2]-acetates, respectively [206]. The cyclization mechanism of wickerol B 226 was predicted, as shown in Scheme 14. First, pyrophosphate was ejected from the terminus of the boat-like transition state of GGPP, forming a verticillen-12-yl cation intermediate, the same as the first step of phomactatriene and taxadiene biosynthesis [207]. 1,2-Rearrangements of β-methyl and α-hydride occurred at the six-membered ring part, then, the ring inversion and cyclization progressed to form the 6/5/9 ring intermediate. A rearrangement proceeded to expand the ring from five to six membered, and the last step resulted in the formation of the 6/5/6/6 ring skeleton. The C-8 position of wickerol A 225 was oxidized by a cytochrome P450 to give wickerol B 226.
Trichodermanins C–H 227232 are new diterpenes with a 6/5/6/6 tetracyclic system that have been isolated from the marine sponge-derived fungus Trichoderma harzianum [208,209]. Trichodermanin C 227 potently inhibited the growth of murine P388 leukemia, human HL-60 leukemia, and murine L1210 leukemia cell lines with IC50 values of 7.9, 6.8, and 7.6 μM, respectively [208].

12. Conclusions and Future Prospects

Diterpenoids show huge potential for drug discovery and development due to their extensive biological functions and structural diversity. Fungal diterpenoids are a diverse family of hybrid natural products with potent bioactivities and intriguing structural architectures. A large number of fungal diterpenoids have exhibited significant anti-inflammatory, cytotoxic, anti-MRSA, antimicrobial, antiviral, antihypertensive, and platelet aggregation-inhibitory activities. Consequently, these bioactive diterpenoids are always hot trending topics for the synthesis community [173,174,186]. Nevertheless, the structural complexity and limited availability of natural products remain obstacles to synthesizing a large collection of natural products and their structural analogues in sufficient amounts. Thus, a synthetic biology method based on the combination of heterologous biosynthesis and genome mining is a promising approach to translate enormous amounts of biosynthetic gene information to richly diverse natural products. Interestingly, while fungi have evolved their systems to create terpenoid diversity, they have also biosynthesized some of the same classes of terpenoids found in plants, bacteria, and other organisms. These relationships provide accessible and renewable prokaryotic systems for eukaryotic natural product biosynthesis and enzymology. In conclusion, we hope it is evident from this review that most of the fungal diterpenoids are biologically active with a few key scaffolds paving a path towards potential drug discovery and development.

Author Contributions

Conceptualization, T.F.; discussion of the contents, F.-L.Z., and T.F.; writing—original draft preparation, F.-L.Z.; writing—review and editing, and editing, T.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of China (grant number 81872762).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blackwell, M. The fungi: 1, 2, 3 … 5.1 million species? Am. J. Bot. 2011, 98, 426–438. [Google Scholar] [CrossRef]
  2. Spiteller, P. Chemical ecology of fungi. Nat. Prod. Rep. 2015, 32, 971–993. [Google Scholar] [CrossRef] [PubMed]
  3. Dictionary of Natural Products. 2020. Available online: http://dnp.chemnetbase.com (accessed on 26 August 2021).
  4. Buckingham, J.; Cooper, C.M.; Purchase, R. Natural Products Desk Reference; CRC Press: Boca Raton, FL, USA, 2016; pp. 1–219. [Google Scholar]
  5. Christianson, D.W. Structural and chemical biology of terpenoid cyclases. Chem. Rev. 2017, 117, 11570–11648. [Google Scholar] [CrossRef] [Green Version]
  6. Quin, M.B.; Flynn, C.M.; Schmidt-Dannert, C. Traversing the fungal terpenome. Nat. Prod. Rep. 2014, 31, 1449–1473. [Google Scholar] [CrossRef]
  7. Minami, A.; Ozaki, T.; Liu, C.; Oikawa, H. Cyclopentane-forming di/sesterterpene synthases: Widely distributed enzymes in bacteria, fungi, and plants. Nat. Prod. Rep. 2018, 35, 1330–1346. [Google Scholar] [CrossRef] [PubMed]
  8. Mitsuhashi, T.; Abe, I. Chimeric terpene synthases possessing both terpene cyclization and prenyltransfer activities. ChemBioChem 2018, 19, 1106–1114. [Google Scholar] [CrossRef] [PubMed]
  9. Dickschat, J.S. Bacterial diterpene biosynthesis. Angew. Chem. Int. Ed. 2019, 58, 15964–15976. [Google Scholar] [CrossRef] [PubMed]
  10. Mafu, S.; Zerbe, P. Plant diterpenoid metabolism for manufacturing the biopharmaceuticals of tomorrow: Prospects and challenges. Phytochem. Rev. 2018, 17, 113–130. [Google Scholar] [CrossRef]
  11. Rudolf, J.D.; Alsup, T.A.; Xu, B.; Li, Z. Bacterial terpenome. Nat. Prod. Rep. 2020, 38, 905–980. [Google Scholar] [CrossRef]
  12. Pemberton, T.A.; Chen, M.; Harris, G.G.; Chou, W.K.; Duan, L.; Koksal, M.; Genshaft, A.S.; Cane, D.E.; Christianson, D.W. Exploring the influence of domain architecture on the catalytic function of diterpene synthases. Biochemistry 2017, 56, 2010–2023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Keller, N.P.; Turner, G.; Bennett, J.W. Fungal secondary metabolism—From biochemistry to genomics. Nat. Rev. Microbiol. 2005, 3, 937–947. [Google Scholar] [CrossRef] [PubMed]
  14. Bills, G.F.; Gloer, J.B. Biologically active secondary metabolites from the fungi. Microbiol. Spectr. 2016, 4, 1087–1119. [Google Scholar] [CrossRef] [PubMed]
  15. Sanchez, J.F.; Somoza, A.D.; Keller, N.P.; Wang, C.C.C. Advances in Aspergillus secondary metabolite research in the post-genomic era. Nat. Prod. Rep. 2012, 29, 351–371. [Google Scholar] [CrossRef] [Green Version]
  16. Fischer, M.J.; Rustenhloz, C.; Leh-Louis, V.; Perriere, G. Molecular and functional evolution of the fungal diterpene synthase genes. BMC Microbiol. 2015, 15, 221. [Google Scholar] [CrossRef] [Green Version]
  17. Rokas, A.; Mead, M.E.; Steenwyk, J.L.; Raja, H.A.; Oberlies, N.H. Biosynthetic gene clusters and the evolution of fungal chemodiversity. Nat. Prod. Rep. 2020, 37, 868–878. [Google Scholar] [CrossRef]
  18. Gressler, M.; Löhr, N.A.; Schäfer, T.; Lawrinowitz, S.; Seibold, P.S.; Hoffmeister, D. Mind the mushroom: Natural product biosynthetic genes and enzymes of Basidiomycota. Nat. Prod. Rep. 2021, 38, 702–722. [Google Scholar] [CrossRef]
  19. Bailly, C.; Gao, J.M. Erinacine A and related cyathane diterpenoids: Molecular diversity and mechanisms underlying their neuroprotection and anticancer activities. Pharmacol. Res. 2020, 159, 104953. [Google Scholar] [CrossRef] [PubMed]
  20. Tang, H.Y.; Yin, X.; Zhang, C.C.; Jia, Q.; Gao, J.M. Structure diversity, synthesis, and biological activity of cyathane diterpenoids in higher fungi. Curr. Med. Chem. 2015, 22, 2375–2391. [Google Scholar] [CrossRef]
  21. Laber, G.; Schütze, E. In Vivo efficacy of 81.723 hfu, a new pleuromutilin derivative against experimentally induced airsacculitis in chicks and turkey poults. Antimicrob. Agents Chemother. 1975, 7, 517–521. [Google Scholar] [CrossRef] [Green Version]
  22. Burch, D.G.; Jones, G.T.; Heard, T.W.; Tuck, R.E. The synergistic activity of tiamulin and chlortetracycline: In-feed treatment of bacterially complicated enzootic pneumonia in fattening pigs. Vet. Rec. 1986, 119, 108–112. [Google Scholar] [CrossRef]
  23. Stipkovits, L.; Ripley, P.H.; Tenk, M.; Glávits, R.; Molnár, T.; Fodor, L. The efficacy of valnemulin (Econor®) in the control of disease caused by experimental infection of calves with Mycoplasma bovis. Res. Vet. Sci. 2005, 78, 207–215. [Google Scholar] [CrossRef] [PubMed]
  24. Rittenhouse, S.; Biswas, S.; Broskey, J.; McCloskey, L.; Moore, T.; Vasey, S.; West, J.; Zalacain, M.; Zonis, R.; Payne, D. Selection of retapamulin, a novel pleuromutilin for topical use. Antimicrob. Agents Chemother. 2006, 50, 3882–3885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zhao, D.H.; Zhang, Z.; Zhang, C.Y.; Liu, Z.C.; Deng, H.; Yu, J.J.; Guo, J.P.; Liu, Y.H. Population pharmacokinetics of valnemulin in swine. J. Vet. Pharmacol. Ther. 2014, 37, 59–65. [Google Scholar] [CrossRef]
  26. Zhang, L.; Fasoyin, O.E.; Molnár, I.; Xu, Y. Secondary metabolites from hypocrealean entomopathogenic fungi: Novel bioactive compounds. Nat. Prod. Rep. 2020, 37, 1181–1206. [Google Scholar] [CrossRef] [PubMed]
  27. Kuang, Y.; Li, B.; Wang, Z.; Qiao, X.; Ye, M. Terpenoids from the medicinal mushroom Antrodia camphorata: Chemistry and medicinal potential. Nat. Prod. Rep. 2021, 38, 83–102. [Google Scholar] [CrossRef] [PubMed]
  28. Jiang, M.H.; Wu, Z.E.; Guo, H.; Liu, L.; Chen, S.H. A review of terpenes from marine-derived fungi: 2015–2019. Mar. Drugs 2020, 18, 321. [Google Scholar] [CrossRef]
  29. Ran, H.; Li, S.M. Fungal benzene carbaldehydes: Occurrence, structural diversity, activities and biosynthesis. Nat. Prod. Rep. 2021, 38, 240–263. [Google Scholar] [CrossRef]
  30. Zhu, M.Z.; Cen, Y.F.; Ye, W.; Li, S.N.; Zhang, W.M. Recent advances on macrocyclic trichothecenes, their bioactivities and biosynthetic pathway. Toxins 2020, 12, 417. [Google Scholar] [CrossRef]
  31. Proctor, R.H.; McCormick, S.P.; Gutierrez, S. Genetic bases for variation in structure and biological activity of trichothecene toxins produced by diverse fungi. Appl. Microbiol. Biotechnol. 2020, 104, 5185–5199. [Google Scholar] [CrossRef]
  32. Cadelis, M.M.; Copp, B.R.; Wiles, S. A review of fungal protoilludane sesquiterpenoid natural products. Antibiotics 2020, 9, 928. [Google Scholar] [CrossRef]
  33. El-Demerdash, A.; Kumla, D.; Kijjoa, A. Chemical diversity and biological activities of meroterpenoids from marine derived-fungi: A comprehensive update. Mar. Drugs 2020, 18, 317. [Google Scholar] [CrossRef]
  34. Zhao, M.; Tang, Y.; Xie, J.; Zhao, Z.; Cui, H. Meroterpenoids produced by fungi: Occurrence, structural diversity, biological activities, and their molecular targets. Eur. J. Med. Chem. 2021, 209, 112860. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, M.; Wu, Z.; Liu, L.; Chen, S. The chemistry and biology of fungal meroterpenoids (2009–2019). Org. Biomol. Chem. 2020, 19, 1644–1704. [Google Scholar] [CrossRef] [PubMed]
  36. Barra, L.; Abe, I. Chemistry of fungal meroterpenoid cyclases. Nat. Prod. Rep. 2020, 38, 566–585. [Google Scholar] [CrossRef] [PubMed]
  37. Zhao, Y.; Cui, J.; Liu, M.Y.J.; Zhao, L. Progress on terpenoids with biological activities produced by plant endophytic fungi in China between 2017 and 2019. Nat. Prod. Commun. 2020, 15, 1934578X20937204. [Google Scholar] [CrossRef]
  38. Zhang, L.; Yue, Q.; Wang, C.; Xu, Y.; Molnár, I. Secondary metabolites from hypocrealean entomopathogenic fungi: Genomics as a tool to elucidate the encoded parvome. Nat. Prod. Rep. 2020, 37, 1164–1180. [Google Scholar] [CrossRef]
  39. Lyu, H.N.; Liu, H.W.; Keller, N.P.; Yin, W.B. Harnessing diverse transcriptional regulators for natural product discovery in fungi. Nat. Prod. Rep. 2020, 37, 6–16. [Google Scholar] [CrossRef]
  40. Hanson, J.R.; Nichols, T.; Mukhrish, Y.; Bagley, M.C. Diterpenoids of terrestrial origin. Nat. Prod. Rep. 2019, 36, 1499–1512. [Google Scholar] [CrossRef]
  41. Friedman, M. Chemistry, nutrition, and health-promoting properties of Hericium erinaceus (Lion’s Mane) mushroom fruiting bodies and mycelia and their bioactive compounds. J. Agric. Food Chem. 2015, 63, 7108–7123. [Google Scholar] [CrossRef]
  42. Enquist, J.A.; Stoltz, B.M. Synthetic efforts toward cyathane diterpenoid natural products. Nat. Prod. Rep. 2009, 26, 661–680. [Google Scholar] [CrossRef]
  43. Kim, K.; Cha, J.K. Total synthesis of cyathin A3 and cyathin B2. Angew. Chem. Int. Ed. 2009, 48, 5334–5336. [Google Scholar] [CrossRef] [PubMed]
  44. Kanoh, N.; Sakanishi, K.; Iimori, E.; Nishimura, K.; Iwabuchi, Y. Asymmetric total synthesis of (−)-scabronine G via intramolecular double Michael reaction and Prins cyclization. Org. Lett. 2011, 13, 2864–2867. [Google Scholar] [CrossRef]
  45. Kobayakawa, Y.; Nakada, M. Enantioselective total synthesis of (−)-cyathin B2. J. Antibiot. 2014, 67, 483–485. [Google Scholar] [CrossRef]
  46. Nakada, M. Enantioselective total syntheses of cyathane diterpenoids. Chem. Rec. 2014, 14, 641–662. [Google Scholar] [CrossRef]
  47. Wu, G.J.; Zhang, Y.H.; Tan, D.X.; He, L.; Cao, B.C.; He, Y.P.; Han, F.S. Synthetic studies on enantioselective total synthesis of cyathane diterpenoids: Cyrneines A and B, glaucopine C, and (+)-allocyathin B2. J. Org. Chem. 2019, 84, 3223–3238. [Google Scholar] [CrossRef] [PubMed]
  48. Marcos, I.S.; Moro, R.F.; Gil-Mesón, A.; Díez, D. Chapter 5. 7-6-5 Tricarbocyclic diterpenes: Valparanes, mulinanes, cyathanes, homoverrucosanes, and related ones. In Studies in Natural Products Chemistry; ur Rahman, A., Ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2016; Volume 48, pp. 137–207. [Google Scholar]
  49. De Jesus Dzul-Beh, A.; Uc-Cachon, A.H.; Borquez, J.; Loyola, L.A.; Pena-Rodriguez, L.M.; Molina-Salinas, G.M. Mulinane- and azorellane-type diterpenoids: A systematic review of their biosynthesis, chemistry, and pharmacology. Biomolecules 2020, 10, 1333. [Google Scholar] [CrossRef] [PubMed]
  50. Sennett, S.H.; Pompeni, S.A.; Wright, A.E. Diterpene metabolites from two chemotypes of the marine sponge Myrmekioderma styx. J. Nat. Prod. 1992, 55, 1421–1429. [Google Scholar] [CrossRef] [PubMed]
  51. Green, D.; Goldberg, I.; Stein, Z.; Ilan, M.; Kashman, Y. Cyanthiwigin A–D, novel cytotoxic diterpenes from the sponge Epipolasis reiswigi. Nat. Prod. Lett. 1992, 1, 193–199. [Google Scholar] [CrossRef]
  52. Peng, J.; Walsh, K.; Weedman, V.; Bergthold, J.D.; Lynch, J.; Lieu, K.L.; Braude, I.A.; Kelly, M.; Hamann, M.T. The new bioactive diterpenes cyanthiwigins E–AA from the Jamaican sponge Myrmekioderma styx. Tetrahedron 2002, 58, 7809–7819. [Google Scholar] [CrossRef]
  53. Pfeiffer, M.W.B.; Phillips, A.J. Conversion of cyanthiwigin U to related cyanthiwigins: Total syntheses of cyanthiwigin W and cyanthiwigin Z. Tetrahedron Lett. 2008, 49, 6860–6861. [Google Scholar] [CrossRef] [Green Version]
  54. Miller, L.C.; Ndungu, J.M.; Sarpong, R. Parallel kinetic resolution approach to the cyathane and cyanthiwigin diterpenes using a cyclopropanation/cope rearrangement. Angew. Chem. Int. Ed. 2009, 48, 2398–2402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Enquist, J.A., Jr.; Virgil, S.C.; Stoltz, B.M. Total syntheses of cyanthiwigins B, F, and G. Chem. Eur. J. 2011, 17, 9957–9969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Wang, C.; Wang, D.; Gao, S. Total synthesis of cyanthiwigins A, C, G, and H. Org. Lett. 2013, 15, 4402–4405. [Google Scholar] [CrossRef]
  57. Kim, K.E.; Stoltz, B.M. A second-generation synthesis of the cyanthiwigin natural product core. Org. Lett. 2016, 18, 5720–5723. [Google Scholar] [CrossRef] [Green Version]
  58. Chang, Y.; Shi, L.; Huang, J.; Shi, L.; Zhang, Z.; Hao, H.D.; Gong, J.; Yang, Z. Stereoselective total synthesis of (±)-5-epi-cyanthiwigin I via an intramolecular Pauson–Khand reaction as the key step. Org. Lett. 2018, 20, 2876–2879. [Google Scholar] [CrossRef]
  59. Kim, K.E.; Adams, A.M.; Chiappini, N.D.; Du Bois, J.; Stoltz, B.M. Cyanthiwigin natural product core as a complex molecular scaffold for comparative late-stage C–H functionalization studies. J. Org. Chem. 2018, 83, 3023–3033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Cao, C.Y.; Zhang, C.C.; Shi, X.W.; Li, D.; Cao, W.; Yin, X.; Gao, J.M. Sarcodonin G derivatives exhibit distinctive effects on neurite outgrowth by modulating NGF signaling in PC12 cells. ACS Chem. Neurosci. 2018, 9, 1607–1615. [Google Scholar] [CrossRef]
  61. Dixon, E.; Schweibenz, T.; Hight, A.; Kang, B.; Dailey, A.; Kim, S.; Chen, M.Y.; Kim, Y.; Neale, S.; Groth, A.; et al. Bacteria-induced static batch fungal fermentation of the diterpenoid cyathin A3, a small-molecule inducer of nerve growth factor. J. Ind. Microbiol. Biotechnol. 2011, 38, 607–615. [Google Scholar] [CrossRef]
  62. Nei, M.; Kumar, S. Molecular Evolution and Phylogenetics; Oxford University Press: New York, NY, USA, 2000. [Google Scholar]
  63. Felsenstein, J. Confidence limits on phylogenies: An approach using the bootstrap. Evolution 1985, 39, 783–791. [Google Scholar] [CrossRef]
  64. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  65. Cheng, Z.; Li, Y.; Xu, W.; Liu, W.; Liu, L.; Zhu, D.; Kang, Y.; Luo, Z.; Li, Q. Three new cyclopiane-type diterpenes from a deep-sea derived fungus Penicillium sp. YPGA11 and their effects against human esophageal carcinoma cells. Bioorg. Chem. 2019, 91, 103129. [Google Scholar] [CrossRef] [PubMed]
  66. Roncal, T.; Cordobés, S.; Ugalde, U.; He, Y.; Sterner, O. Novel diterpenes with potent conidiation inducing activity. Tetrahedron Lett. 2002, 43, 6799–6802. [Google Scholar] [CrossRef] [Green Version]
  67. Rodriíguez, I.I.; Rodriíguez, A.D.; Zhao, H. Aberrarone: A gorgonian-derived diterpene from Pseudopterogorgia elisabethae. J. Org. Chem. 2009, 74, 7581–7584. [Google Scholar] [CrossRef] [Green Version]
  68. Niu, S.; Fan, Z.; Tang, X.; Liu, Q.; Shao, Z.; Liu, G.; Yang, X.W. Cyclopiane-type diterpenes from the deep-sea-derived fungus Penicillium commune MCCC 3A00940. Tetrahedron Lett. 2018, 59, 375–378. [Google Scholar] [CrossRef]
  69. Mitsuhashi, T.; Kikuchi, T.; Hoshino, S.; Ozeki, M.; Awakawa, T.; Shi, S.P.; Fujita, M.; Abe, I. Crystalline sponge method enabled the investigation of a prenyltransferase-terpene synthase chimeric enzyme, whose product exhibits broadened NMR signals. Org. Lett. 2018, 20, 5606–5609. [Google Scholar] [CrossRef] [PubMed]
  70. Roncal, T.; Cordobes, S.; Sterner, O.; Ugalde, U. Conidiation in Penicillium cyclopium is induced by conidiogenone, an endogenous diterpene. Eukaryot. Cell 2002, 1, 823–829. [Google Scholar] [CrossRef] [Green Version]
  71. Shiina, T.; Nakagawa, K.; Fujisaki, Y.; Ozaki, T.; Liu, C.; Toyomasu, T.; Hashimoto, M.; Koshino, H.; Minami, A.; Kawaide, H.; et al. Biosynthetic study of conidiation-inducing factor conidiogenone: Heterologous production and cyclization mechanism of a key bifunctional diterpene synthase. Biosci. Biotechnol. Biochem. 2019, 83, 192–201. [Google Scholar] [CrossRef]
  72. Du, L.; Li, D.; Zhu, T.; Cai, S.; Wang, F.; Xiao, X.; Gu, Q. New alkaloids and diterpenes from a deep ocean sediment derived fungus Penicillium sp. Tetrahedron 2009, 65, 1033–1039. [Google Scholar] [CrossRef]
  73. Gao, S.S.; Li, X.M.; Zhang, Y.; Li, C.S.; Wang, B.G. Conidiogenones H and I, two new diterpenes of cyclopiane class from a marine-derived endophytic fungus Penicillium chrysogenum QEN-24S. Chem. Biodivers. 2011, 8, 1748–1753. [Google Scholar] [CrossRef]
  74. Hou, S.H.; Tu, Y.Q.; Wang, S.H.; Xi, C.C.; Zhang, F.M.; Wang, S.H.; Li, Y.T.; Liu, L. Total syntheses of the tetracyclic cyclopiane diterpenes conidiogenone, conidiogenol, and conidiogenone B. Angew. Chem. Int. Ed. 2016, 128, 4532–4536. [Google Scholar] [CrossRef]
  75. Li, F.; Sun, W.; Zhang, S.; Gao, W.; Lin, S.; Yang, B.; Chai, C.; Li, H.; Wang, J.; Hu, Z.; et al. New cyclopiane diterpenes with anti-inflammatory activity from the sea sediment-derived fungus Penicillium sp. TJ403-2. Chin. Chem. Lett. 2020, 31, 197–201. [Google Scholar] [CrossRef]
  76. Chen, H.Y.; Liu, T.K.; Shi, Q.; Yang, X.L. Sesquiterpenoids and diterpenes with antimicrobial activity from Leptosphaeria sp. XL026, an endophytic fungus in Panax notoginseng. Fitoterapia 2019, 137, 104243. [Google Scholar] [CrossRef]
  77. De Boer, A.H.; De Vries-Van Leeuwen, I.J. Fusicoccanes: Diterpenes with surprising biological functions. Trends Plant Sci. 2012, 17, 360–368. [Google Scholar] [CrossRef] [PubMed]
  78. Hu, Z.; Sun, W.; Li, F.; Guan, J.; Lu, Y.; Liu, J.; Tang, Y.; Du, G.; Xue, Y.; Luo, Z.; et al. Fusicoccane-derived diterpenoids from Alternaria brassicicola: Investigation of the structure-stability relationship and discovery of an IKKβ inhibitor. Org. Lett. 2018, 20, 5198–5202. [Google Scholar] [CrossRef]
  79. Li, F.L.; Lin, S.; Zhang, S.T.; Hao, X.C.; Li, X.N.; Yang, B.Y.; Liu, J.J.; Wang, J.P.; Hu, Z.X.; Zhang, Y.H. Alterbrassinoids A–D: Fusicoccane-derived diterpenoid dimers featuring different carbon skeletons from Alternaria brassicicola. Org. Lett. 2019, 21, 8353–8357. [Google Scholar] [CrossRef]
  80. Zhang, M.; Yan, S.; Liang, Y.; Zheng, M.; Wu, Z.; Zang, Y.; Yu, M.; Sun, W.; Liu, J.; Ye, Y.; et al. Talaronoids A–D: Four fusicoccane diterpenoids with an unprecedented tricyclic 5/8/6 ring system from the fungus Talaromyces stipitatus. Org. Chem. Front. 2020, 7, 3486–3492. [Google Scholar] [CrossRef]
  81. Ballio, A.; Chain, E.B.; De Leo, P.; Erlanger, B.F.; Mauri, M.; Tonolo, A. Fusicoccin: A new wilting toxin produced by Fusicoccum amygdali Del. Nature 1964, 203, 297. [Google Scholar] [CrossRef]
  82. Aoyagi, T.; Aoyama, T.; Kojima, F.; Hattori, S.; Honma, Y.; Hamada, M.; Takeuch, T. Cyclooctatin, a new inhibitor of lysophospholipase, produced by Streptomyces melanosporofaciens MI614-43F2. Taxonomy, production, isolation, physico-chemical properties and biological activities. J. Antibiot. 1992, 45, 1587–1591. [Google Scholar] [CrossRef] [Green Version]
  83. Muromtsev, G.S.; Voblikova, V.D.; Kobrina, N.S.; Koreneva, V.M.; Krasnopolskaya, L.M.; Sadovskaya, V.L. Occurrence of fusicoccanes in plants and fungi. J. Plant Growth Regul. 1994, 13, 39–49. [Google Scholar] [CrossRef]
  84. Rasoamiaranjanahary, L.; Marston, A.; Guilet, D.; Schenk, K.; Randimbivololona, F.; Hostettmann, K. Antifungal diterpenes from Hypoestes serpens (Acanthaceae). Phytochemistry 2003, 62, 333–337. [Google Scholar] [CrossRef]
  85. Komala, I.; Ito, T.; Nagashima, F.; Yagi, Y.; Kawahata, M.; Yamaguchi, K.; Asakawa, Y. Zierane sesquiterpene lactone, cembrane and fusicoccane diterpenoids, from the Tahitian liverwort Chandonanthus hirtellus. Phytochemistry 2010, 71, 1387–1394. [Google Scholar] [CrossRef] [PubMed]
  86. Gilabert, M.; Ramos, A.N.; Schiavone, M.A.M.; Arena, M.E.; Bardoón, A. Bioactive sesqui- and diterpenoids from the Argentine liverwort Porella chilensis. J. Nat. Prod. 2011, 74, 574–579. [Google Scholar] [CrossRef] [PubMed]
  87. Kawamura, A.; Iacovidou, M.; Hirokawa, E.; Soll, C.E.; Trujillo, M. 17-Hydroxycyclooctatin, a fused 5–8–5 ring diterpene, from Streptomyces sp. MTE4a. J. Nat. Prod. 2011, 74, 492–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Kenmoku, H.; Tada, H.; Oogushi, M.; Esumi, T.; Takahashi, H.; Noji, M.; Sassa, T.; Toyota, M.; Asakawa, Y. Seed dormancy breaking diterpenoids from the liverwort Plagiochila sciophila and their differentiation inducing activity in human promyelocytic leukemia HL-60 cells. Nat. Prod. Commun. 2014, 9, 915–920. [Google Scholar] [CrossRef] [Green Version]
  89. Takahashi, M.; Kawamura, A.; Kato, N.; Nishi, T.; Hamachi, I.; Ohkanda, J. Phosphopeptide-dependent labeling of 14–3–3ζ proteins by fusicoccin-based fluorescent probes. Angew. Chem. Int. Ed. 2012, 51, 509–512. [Google Scholar] [CrossRef]
  90. Wang, W.; Wan, X.; Liu, J.; Wang, J.; Zhu, H.; Chen, C.; Zhang, Y. Two new terpenoids from Talaromyces purpurogenus. Mar. Drugs 2018, 16, 150. [Google Scholar] [CrossRef] [Green Version]
  91. Takekawa, H.; Tanaka, K.; Fukushi, E.; Matsuo, K.; Nehira, T.; Hashimoto, M. Roussoellols A and B, tetracyclic fusicoccanes from Roussoella hysterioides. J. Nat. Prod. 2013, 76, 1047–1051. [Google Scholar] [CrossRef]
  92. Aoyama, T.; Naganawa, H.; Muraoka, Y.; Aoyagi, T.; Takeuchi, T. The structure of cyclooctatin, a new inhibitor of lysophospholipase. J. Antibiot. 1992, 45, 1703–1704. [Google Scholar] [CrossRef] [Green Version]
  93. Zheng, D.; Han, L.; Qu, X.D.; Chen, X.; Zhong, J.L.; Bi, X.X.; Liu, J.; Jiang, Y.; Jiang, C.L.; Huang, X.S. Cytotoxic fusicoccane-type diterpenoids from Streptomyces violascens isolated from Ailuropoda melanoleuca feces. J. Nat. Prod. 2017, 80, 837–844. [Google Scholar] [CrossRef]
  94. Mackinnon, S. Components from the phytotoxic extract of Alternaria brassicicola, a black spot pathogen of canola. Phytochemistry 1999, 51, 215–221. [Google Scholar] [CrossRef]
  95. Pedras, M.S.; Chumala, P.B.; Jin, W.; Islam, M.S.; Hauck, D.W. The phytopathogenic fungus Alternaria brassicicola: Phytotoxin production and phytoalexin elicitation. Phytochemistry 2009, 70, 394–402. [Google Scholar] [CrossRef] [PubMed]
  96. Kenmoku, H.; Takeue, S.; Oogushi, M.; Yagi, Y.; Sassa, T.; Toyota, M.; Asakawa, Y. Seed dormancy breaking diterpenoids, including novel brassicicenes J and K, from fungus Alternaria brassicicola, and their necrotic/apoptotic activities in HL-60 cells. Nat. Prod. Commun. 2014, 9, 351–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Tang, Y.; Xue, Y.; Du, G.; Wang, J.; Liu, J.; Sun, B.; Li, X.N.; Yao, G.; Luo, Z.; Zhang, Y. Structural revisions of a class of natural products: Scaffolds of aglycon analogues of fusicoccins and cotylenins isolated from fungi. Angew. Chem. Int. Ed. 2016, 55, 4069–4073. [Google Scholar] [CrossRef] [PubMed]
  98. Li, F.; Sun, W.; Guan, J.; Lu, Y.; Zhang, S.; Lin, S.; Liu, J.; Gao, W.; Wang, J.; Hu, Z.; et al. Alterbrassicicene A, a highly transformed fusicoccane-derived diterpenoid with potent PPAR-γ agonistic activity from Alternaria brassicicola. Org. Lett. 2018, 20, 7982–7986. [Google Scholar] [CrossRef] [PubMed]
  99. Li, F.; Sun, W.; Guan, J.; Lu, Y.; Lin, S.; Zhang, S.; Gao, W.; Liu, J.; Du, G.; Wang, J.; et al. Anti-inflammatory fusicoccane-type diterpenoids from the phytopathogenic fungus Alternaria brassicicola. Org. Biomol. Chem. 2018, 16, 8751–8760. [Google Scholar] [CrossRef]
  100. Li, F.; Lin, S.; Zhang, S.; Pan, L.; Chai, C.; Su, J.C.; Yang, B.; Liu, J.; Wang, J.; Hu, Z.; et al. Modified fusicoccane-type diterpenoids from Alternaria brassicicola. J. Nat. Prod. 2020, 83, 1931–1938. [Google Scholar] [CrossRef]
  101. Li, F.L.; Pan, L.F.; Lin, S.; Zhang, S.T.; Li, H.Q.; Yang, B.Y.; Liu, J.J.; Wang, J.P.; Hu, Z.X.; Zhang, Y.H. Fusicoccane-derived diterpenoids with bridgehead double-bond-containing tricyclo[9.2.1.03,7]tetradecane ring systems from Alternaria brassicicola. Bioorg. Chem. 2020, 100, 103887. [Google Scholar] [CrossRef]
  102. Bie, Q.; Chen, C.M.; Yu, M.Y.; Guo, J.R.; Wang, J.P.; Liu, J.J.; Zhou, Y.; Zhu, H.C.; Zhang, Y.H. Dongtingnoids A–G: Fusicoccane diterpenoids from a Penicillium species. J. Nat. Prod. 2019, 82, 80–86. [Google Scholar] [CrossRef]
  103. Harada, J.; Tanaka, T.; Sassa, T. Sprouting of dormant tubers of Sagittaria trifolia, a perennial paddy weed, caused by cotylenin E, a new plant growth regulator. J. Weed Sci. Technol. 1981, 26, 37–39. [Google Scholar] [CrossRef] [Green Version]
  104. Takeuchi, Y.; Sassa, T.; Kawaguchi, S.; Ogasawara, M.; Yoneyama, K.; Konnai, M. Stimulation of germination of Monochoria vaginalis seeds by seed coat puncture and cotylenins. J. Weed Sci. Technol. 1995, 40, 221–224. [Google Scholar] [CrossRef] [Green Version]
  105. Noike, M.; Ono, Y.; Araki, Y.; Tanio, R.; Higuchi, Y.; Nitta, H.; Hamano, Y.; Toyomasu, T.; Sassa, T.; Kato, N.; et al. Molecular breeding of a fungus producing a precursor diterpene suitable for semi-synthesis by dissection of the biosynthetic machinery. PLoS ONE 2012, 7, e42090. [Google Scholar] [CrossRef]
  106. Arens, J.; Engels, B.; Klopries, S.; Jennewein, S.; Ottmann, C.; Schulz, F. Exploration of biosynthetic access to the shared precursor of the fusicoccane diterpenoid family. Chem. Commun. 2013, 49, 4337–4339. [Google Scholar] [CrossRef] [PubMed]
  107. 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]
  108. Wu, Y.H.; Chen, G.D.; He, R.R.; Wang, C.X.; Hu, D.; Wang, G.Q.; Guo, L.D.; Yao, X.S.; Gao, H. Pericolactines A–C, a new class of diterpenoid alkaloids with unusual tetracyclic skeleton. Sci. Rep. 2015, 5, 17082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Toyomasu, T.; Tsukahara, M.; Kaneko, A.; Niida, R.; Mitsuhashi, W.; Dairi, T.; Kato, N.; Sassa, T. Fusicoccins are biosynthesized by an unusual chimera diterpene synthase in fungi. Proc. Natl. Acad. Sci. USA 2007, 104, 3084–3088. [Google Scholar] [CrossRef] [Green Version]
  110. Minami, A.; Tajima, N.; Higuchi, Y.; Toyomasu, T.; Sassa, T.; Kato, N.; Dairi, T. Identification and functional analysis of brassicicene C biosynthetic gene cluster in Alternaria brassicicola. Bioorg. Med. Chem. Lett. 2009, 19, 870–874. [Google Scholar] [CrossRef]
  111. Hashimoto, M.; Higuchi, Y.; Takahashi, S.; Osada, H.; Sakaki, T.; Toyomasu, T.; Sassa, T.; Kato, N.; Dairi, T. Functional analyses of cytochrome P450 genes responsible for the early steps of brassicicene C biosynthesis. Bioorg. Med. Chem. Lett. 2009, 19, 5640–5643. [Google Scholar] [CrossRef] [PubMed]
  112. Ono, Y.; Minami, A.; Noike, M.; Higuchi, Y.; Toyomasu, T.; Sassa, T.; Kato, N.; Dairi, T. Dioxygenases, key enzymes to determine the aglycon structures of fusicoccin and brassicicene, diterpene compounds produced by fungi. J. Am. Chem. Soc. 2011, 133, 2548–2555. [Google Scholar] [CrossRef]
  113. Kim, S.Y.; Zhao, P.; Igarashi, M.; Sawa, R.; Tomita, T.; Nishiyama, M.; Kuzuyama, T. Cloning and heterologous expression of the cyclooctatin biosynthetic gene cluster afford a diterpene cyclase and two P450 hydroxylases. Chem. Biol. 2009, 16, 736–743. [Google Scholar] [CrossRef] [Green Version]
  114. Kudo, F.; Matsuura, Y.; Hayashi, T.; Fukushima, M.; Eguchi, T. Genome mining of the sordarin biosynthetic gene cluster from Sordaria araneosa Cain ATCC 36386: Characterization of cycloaraneosene synthase and GDP-6-deoxyaltrose transferase. J. Antibiot. 2016, 69, 541–548. [Google Scholar] [CrossRef]
  115. Chiba, R.; Minami, A.; Gomi, K.; Oikawa, H. Identification of ophiobolin F synthase by a genome mining approach: A sesterterpene synthase from Aspergillus clavatus. Org. Lett. 2013, 15, 594–597. [Google Scholar] [CrossRef] [PubMed]
  116. Ye, Y.; Minami, A.; Mandi, A.; Liu, C.; Taniguchi, T.; Kuzuyama, T.; Monde, K.; Gomi, K.; Oikawa, H. Genome mining for sesterterpenes using bifunctional terpene synthases reveals a unified intermediate of di/sesterterpenes. J. Am. Chem. Soc. 2015, 137, 11846–11853. [Google Scholar] [CrossRef] [PubMed]
  117. Tazawa, A.; Ye, Y.; Ozaki, T.; Liu, C.; Ogasawara, Y.; Dairi, T.; Higuchi, Y.; Kato, N.; Gomi, K.; Minami, A.; et al. Total biosynthesis of brassicicenes: Identification of a key enzyme for skeletal diversification. Org. Lett. 2018, 20, 6178–6182. [Google Scholar] [CrossRef]
  118. Lin, F.L.; Lauterbach, L.; Zou, J.; Wang, Y.H.; Lv, J.M.; Chen, G.D.; Hu, D.; Gao, H.; Yao, X.S.; Dickschat, J.S. Mechanistic characterization of the fusicoccane-type diterpene synthase for myrothec-15(17)-en-7-ol. ACS Catal. 2020, 10, 4306–4312. [Google Scholar] [CrossRef]
  119. Brady, S.F.; Singh, M.P.; Janso, J.E.; Clardy, J. Guanacastepene, a fungal-derived diterpene antibiotic with a new carbon skeleton. J. Am. Chem. Soc. 2000, 122, 2116–2117. [Google Scholar] [CrossRef]
  120. Brady, S.F.; Bondi, S.M.; Clardy, J. The guanacastepenes: A highly diverse family of secondary metabolites produced by an endophytic fungus. J. Am. Chem. Soc. 2001, 123, 9900–9901. [Google Scholar] [CrossRef]
  121. Dudley, G.B.; Danishefsky, S.J. A four-step synthesis of the hydroazulene core of guanacastepene. Org. Lett. 2001, 3, 2399–2402. [Google Scholar] [CrossRef]
  122. Dudley, G.B.; Tan, D.S.; Kim, G.; Tanski, J.M.; Danishefsky, S.J. Remarkable stereoselectivity in the alkylation of a hydroazulenone: Progress towards the total synthesis of guanacastepene. Tetrahedron Lett. 2001, 42, 6789–6791. [Google Scholar] [CrossRef]
  123. Mehta, G.; Umarye, J.D. Studies toward the total synthesis of diterpene antibiotic guanacastepene A:  construction of the hydroazulenic core. Org. Lett. 2002, 4, 1063–1066. [Google Scholar] [CrossRef]
  124. Lin, S.; Dudley, G.B.; Tan, D.S.; Danishefsky, S.J. A stereoselective route to guanacastepene A through a surprising epoxidation. Angew. Chem. Int. Ed. 2002, 41, 2188–2191. [Google Scholar] [CrossRef]
  125. Brummond, K.M.; Gao, D. Unique strategy for the assembly of the carbon skeleton of guanacastepene A using an allenic Pauson–Khand-type reaction. Org. Lett. 2003, 5, 3491–3494. [Google Scholar] [CrossRef] [PubMed]
  126. Du, X.; Chu, H.V.; Kwon, O. Synthesis of the [5–7–6] tricyclic core of guanacastepene A via an intramolecular Mukaiyama aldol reaction. Org. Lett. 2003, 5, 1923–1926. [Google Scholar] [CrossRef] [PubMed]
  127. Hughes, C.C.; Kennedy-Smith, J.J.; Trauner, D. Synthetic studies toward the guanacastepenes. Org. Lett. 2003, 5, 4113–4115. [Google Scholar] [CrossRef] [PubMed]
  128. Hughes, C.C.; Miller, A.K.; Trauner, D. An electrochemical approach to the guanacastepenes. Org. Lett. 2005, 7, 3425–3428. [Google Scholar] [CrossRef]
  129. Battiste, M.A.; Pelphrey, P.M.; Wright, D.L. The cycloaddition strategy for the synthesis of natural products containing carbocyclic seven-membered rings. Chem. Eur. J. 2006, 12, 3438–3447. [Google Scholar] [CrossRef]
  130. Iimura, S.; Overman, L.E.; Paulini, R.; Zakarian, A. Enantioselective total synthesis of guanacastepene N using an uncommon 7-endo Heck cyclization as a pivotal step. J. Am. Chem. Soc. 2006, 128, 13095–13101. [Google Scholar] [CrossRef] [Green Version]
  131. Li, C.C.; Wang, C.H.; Liang, B.; Zhang, X.H.; Deng, L.J.; Liang, S.; Chen, J.H.; Wu, Y.D.; Yang, Z. Synthetic study of 1,3-butadiene-based IMDA approach to construct a [5–7–6] tricyclic core and its application to the total synthesis of C8-epi-guanacastepene O. J. Org. Chem. 2006, 71, 6892–6897. [Google Scholar] [CrossRef]
  132. McGowan, C.A.; Schmieder, A.K.; Roberts, L.; Greaney, M.F. Synthesis of the guanacastepene A–B hydrazulene ring system through photochemical ring transposition. Org. Biomol. Chem. 2007, 5, 1522–1524. [Google Scholar] [CrossRef]
  133. Michalak, K.; Michalak, M.; Wicha, J. Construction of the tricyclic 5-7-6 scaffold of fungi-derived diterpenoids. Total synthesis of (±)-heptemerone G and an approach to Danishefsky’s intermediate for guanacastepene A synthesis. J. Org. Chem. 2010, 75, 8337–8350. [Google Scholar] [CrossRef]
  134. Oonishi, Y.; Taniuchi, A.; Sato, Y. Rhodium(I)-catalyzed hydroacylation/cycloisomerization cascade reaction: Application to the construction of the tricyclic core of guanacastepenes. Synthesis 2010, 2010, 2884–2892. [Google Scholar] [CrossRef]
  135. Gampe, C.M.; Carreira, E.M. Total syntheses of guanacastepenes N and O. Angew. Chem. Int. Ed. 2011, 50, 2962–2965. [Google Scholar] [CrossRef]
  136. Michalak, K.; Michalak, M.; Wicha, J. A facile construction of the tricyclic 5-7-6 scaffold of fungi-derived diterpenoids. The first total synthesis of (±)-heptemerone G and a new approach to Danishefsky’s intermediate for a guanacastepene A synthesis. Tetrahedron Lett. 2010, 51, 4344–4346. [Google Scholar] [CrossRef]
  137. Kettering, M.; Valdivia, C.; Sterner, O.; Anke, H.; Thines, E. Heptemerones A–G, seven novel diterpenoids from Coprinus heptemerus: Producing organism, fermentation, isolation and biological activities. J. Antibiot. 2005, 58, 390–396. [Google Scholar] [CrossRef] [PubMed]
  138. Valdivia, C.; Kettering, M.; Anke, H.; Thines, E.; Sterner, O. Diterpenoids from Coprinus heptemerus. Tetrahedron 2005, 61, 9527–9532. [Google Scholar] [CrossRef]
  139. Ou, Y.X.; Li, Y.Y.; Qian, X.M.; Shen, Y.M. Guanacastane-type diterpenoids from Coprinus radians. Phytochemistry 2012, 78, 190–196. [Google Scholar] [CrossRef]
  140. Liu, Y.Z.; Li, Y.Y.; Ou, Y.X.; Xiao, S.Y.; Lu, C.H.; Zheng, Z.H.; Shen, Y.M. Guanacastane-type diterpenoids with cytotoxic activity from Coprinus plicatilis. Bioorg. Med. Chem. Lett. 2012, 22, 5059–5062. [Google Scholar] [CrossRef]
  141. Liu, Y.Z.; Lu, C.H.; Shen, Y.M. Guanacastane-type diterpenoids from Coprinus plicatilis. Phytochem. Lett. 2014, 7, 161–164. [Google Scholar] [CrossRef]
  142. Yin, X.; Feng, T.; Li, Z.H.; Leng, Y.; Liu, J.K. Five new guanacastane-type diterpenes from cultures of the fungus Psathyrella candolleana. Nat. Prod. Bioprospect. 2014, 4, 149–155. [Google Scholar] [CrossRef] [Green Version]
  143. Feng, Y.; Ren, F.X.; Niu, S.B.; Wang, L.; Li, L.; Liu, X.Z.; Che, Y.S. Guanacastane diterpenoids from the plant endophytic fungus Cercospora sp. J. Nat. Prod. 2014, 77, 873–881. [Google Scholar] [CrossRef]
  144. Wu, F.B.; Li, T.X.; Yang, M.H.; Kong, L.Y. Guanacastane-type diterpenoids from the insect-associated fungus Verticillium dahliae. J. Asian Nat. Prod. Res. 2016, 18, 117–124. [Google Scholar] [CrossRef]
  145. Zhang, H.; Yang, M.H.; Li, Y.; Cheng, X.B.; Pei, Y.H.; Kong, L.Y. Seven new guanacastane-type diterpenoids from the fungus Verticillium dahliae. Fitoterapia 2019, 133, 219–224. [Google Scholar] [CrossRef]
  146. Lam, Y.T.H.; Palfner, G.; Lima, C.; Porzel, A.; Brandt, W.; Frolov, A.; Sultani, H.; Franke, K.; Wagner, C.; Merzweiler, K.; et al. Nor-guanacastepene pigments from the Chilean mushroom Cortinarius pyromyxa. Phytochemistry 2019, 165, 112048. [Google Scholar] [CrossRef] [PubMed]
  147. Wang, X.; Jin, X.Y.; Zhou, J.C.; Zhu, R.X.; Qiao, Y.N.; Zhang, J.Z.; Li, Y.; Zhang, C.Y.; Chen, W.; Chang, W.Q.; et al. Terpenoids from the Chinese liverwort Heteroscyphus coalitus and their anti-virulence activity against Candida albicans. Phytochemistry 2020, 174, 112324. [Google Scholar] [CrossRef]
  148. Ghisalberti, E.L.; Hockless, D.C.R.; Rowland, C.; White, A.H. Harziandione, a new class of diterpene from Trichoderma harzianum. J. Nat. Prod. 1992, 55, 1690–1694. [Google Scholar] [CrossRef]
  149. Miao, F.P.; Liang, X.R.; Yin, X.L.; Wang, G.; Ji, N.Y. Absolute configurations of unique harziane diterpenes from Trichoderma species. Org. Lett. 2012, 14, 3815–3817. [Google Scholar] [CrossRef]
  150. Mannina, L.; Segre, A.L.; Ritieni, A.; Fogliano, V.; Vinale, F.; Randazzo, G.; Maddau, L.; Bottalico, A. A new fungal growth inhibitor from Trichoderma viride. Tetrahedron 1997, 53, 3135–3144. [Google Scholar] [CrossRef]
  151. Barra, L.; Dickschat, J.S. Harzianone biosynthesis by the biocontrol fungus Trichoderma. ChemBioChem 2017, 18, 2358–2365. [Google Scholar] [CrossRef]
  152. Adelin, E.; Servy, C.; Martin, M.T.; Arcile, G.; Iorga, B.I.; Retailleau, P.; Bonfill, M.; Ouazzani, J. Bicyclic and tetracyclic diterpenes from a Trichoderma symbiont of Taxus baccata. Phytochemistry 2014, 97, 55–61. [Google Scholar] [CrossRef]
  153. Hönig, M.; Carreira, E.M. Total synthesis and structural revision of a harziane diterpenoid. Angew. Chem. Int. Ed. 2020, 59, 1192–1196. [Google Scholar] [CrossRef]
  154. Zhang, M.; Liu, J.M.; Zhao, J.L.; Li, N.; Chen, R.D.; Xie, K.B.; Zhang, W.J.; Feng, K.P.; Yan, Z.; Wang, N.; et al. Two new diterpenoids from the endophytic fungus Trichoderma sp. Xy24 isolated from mangrove plant Xylocarpus granatum. Chin. Chem. Lett. 2016, 27, 957–960. [Google Scholar] [CrossRef]
  155. Song, Y.P.; Fang, S.T.; Miao, F.P.; Yin, X.L.; Ji, N.Y. Diterpenes and sesquiterpenes from the marine algicolous fungus Trichoderma harzianum X-5. J. Nat. Prod. 2018, 81, 2553–2559. [Google Scholar] [CrossRef]
  156. Song, Y.P.; Liu, X.H.; Shi, Z.Z.; Miao, F.P.; Fang, S.T.; Ji, N.Y. Bisabolane, cyclonerane, and harziane derivatives from the marine-alga-endophytic fungus Trichoderma asperellum cf44-2. Phytochemistry 2018, 152, 45–52. [Google Scholar] [CrossRef] [PubMed]
  157. Song, Y.P.; Miao, F.P.; Liang, X.R.; Yin, X.L.; Ji, N.Y. Harziane and cadinane terpenoids from the alga-endophytic fungus Trichoderma asperellum A-YMD-9-2. Phytochem. Lett. 2019, 32, 38–41. [Google Scholar] [CrossRef]
  158. Zou, J.X.; Song, Y.P.; Ji, N.Y. Deoxytrichodermaerin, a harziane lactone from the marine algicolous fungus Trichoderma longibrachiatum A-WH-20-2. Nat. Prod. Res. 2021, 35, 216–221. [Google Scholar] [CrossRef]
  159. Zhao, D.L.; Yang, L.J.; Shi, T.; Wang, C.Y.; Shao, C.L.; Wang, C.Y. Potent phytotoxic harziane diterpenes from a soft coral-derived strain of the fungus Trichoderma harzianum XS-20090075. Sci. Rep. 2019, 9, 13345. [Google Scholar] [CrossRef] [Green Version]
  160. Shi, T.; Shao, C.L.; Liu, Y.; Zhao, D.L.; Cao, F.; Fu, X.M.; Yu, J.Y.; Wu, J.S.; Zhang, Z.K.; Wang, C.Y. Terpenoids From the coral-derived fungus Trichoderma harzianum (XS-20090075) induced by chemical epigenetic manipulation. Front. Microbiol. 2020, 11, 572. [Google Scholar] [CrossRef]
  161. Li, W.Y.; Liu, Y.; Lin, Y.T.; Liu, Y.C.; Guo, K.; Li, X.N.; Luo, S.H.; Li, S.H. Antibacterial harziane diterpenoids from a fungal symbiont Trichoderma atroviride isolated from Colquhounia coccinea var. mollis. Phytochemistry 2020, 170, 112198. [Google Scholar] [CrossRef] [PubMed]
  162. 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] [PubMed]
  163. 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]
  164. Toyomasu, T.; Kaneko, A.; Tokiwano, T.; Kanno, Y.; Kanno, Y.; Niida, R.; Miura, S.; Nishioka, T.; Ikeda, C.; Mitsuhashi, W.; et al. Biosynthetic gene-based secondary metabolite screening: A new diterpene, methyl phomopsenonate, from the fungus Phomopsis amygdali. J. Org. Chem. 2009, 74, 1541–1548. [Google Scholar] [CrossRef]
  165. Shinde, S.S.; Minami, A.; Chen, Z.; Tokiwano, T.; Toyomasu, T.; Kato, N.; Sassa, T.; Oikawa, H. Cyclization mechanism of phomopsene synthase: Mass spectrometry based analysis of various site-specifically labeled terpenes. J. Antibiot. 2017, 70, 632–638. [Google Scholar] [CrossRef] [PubMed]
  166. Lauterbach, L.; Rinkel, J.; Dickschat, J.S. Two bacterial diterpene synthases from Allokutzneria albata produce bonnadiene, phomopsene, and allokutznerene. Angew. Chem. Int. Ed. 2018, 57, 8280–8283. [Google Scholar] [CrossRef] [PubMed]
  167. Rinkel, J.; Steiner, S.T.; Dickschat, J.S. Diterpene biosynthesis in actinomycetes: Studies on cattleyene synthase and phomopsene synthase. Angew. Chem. Int. Ed. 2019, 58, 9230–9233. [Google Scholar] [CrossRef]
  168. Kavanagh, F.; Hervey, A.; Robbins, W.J. Robbins, W.J. Antibiotic substances from Basidiomycetes VIII. Pleurotus Multilus (Fr.) Sacc. and Pleurotus passeckerianus Pilat. Proc. Natl. Acad. Sci. USA 1951, 37, 570. [Google Scholar] [CrossRef] [Green Version]
  169. Hartley, A.J.; De Mattos-Shipley, K.; Collins, C.M.; Kilaru, S.; Foster, G.D.; Bailey, A.M. Investigating pleuromutilin-producing Clitopilus species and related basidiomycetes. FEMS Microbiol. Lett. 2009, 297, 24–30. [Google Scholar] [CrossRef]
  170. Arigoni, D. La struttura di un terpene di nuovo genere. Gazz. Chim. Ital. 1962, 92, 884–901. [Google Scholar]
  171. Birch, A.J.; Holzapfel, C.W.; Rickards, R.W. The structure and some aspects of the biosynthesis of pleuromutilin. Tetrahedron 1966, 22, 359–387. [Google Scholar] [CrossRef]
  172. Arigoni, D. Some studies in the biosynthesis of terpenes and related compounds. Pure Appl. Chem. 1968, 17, 331–348. [Google Scholar] [CrossRef]
  173. Hu, Y.J.; Li, L.X.; Han, J.C.; Min, L.; Li, C.C. Recent advances in the total synthesis of natural products containing eight-membered carbocycles (2009–2019). Chem. Rev. 2020, 120, 5910–5953. [Google Scholar] [CrossRef]
  174. Min, L.; Liu, X.; Li, C.C. Total synthesis of natural products with bridged bicyclo[m.n.1] ring systems via type II [5 + 2] cycloaddition. Acc. Chem. Res. 2020, 53, 703–718. [Google Scholar] [CrossRef] [PubMed]
  175. Ling, C.; Fu, L.; Gao, S.; Chu, W.; Wang, H.; Huang, Y.; Chen, X.; Yang, Y. Design, synthesis, and structure–activity relationship studies of novel thioether pleuromutilin derivatives as potent antibacterial agents. J. Med. Chem. 2014, 57, 4772–4795. [Google Scholar] [CrossRef] [PubMed]
  176. Wang, X.; Ling, Y.; Wang, H.; Yu, J.; Tang, J.; Zheng, H.; Zhao, X.; Wang, D.; Chen, G.; Qiu, W.; et al. Novel pleuromutilin derivatives as antibacterial agents: Synthesis, biological evaluation and molecular docking studies. Bioorg. Med. Chem. Lett. 2012, 22, 6166–6172. [Google Scholar] [CrossRef]
  177. Dong, Y.J.; Meng, Z.H.; Mi, Y.Q.; Zhang, C.; Cui, Z.H.; Wang, P.; Xu, Z.B. Synthesis of novel pleuromutilin derivatives. Part 1: Preliminary studies of antituberculosis activity. Bioorg. Med. Chem. Lett. 2015, 25, 1799–1803. [Google Scholar] [CrossRef] [PubMed]
  178. Novak, R. Are pleuromutilin antibiotics finally fit for human use? Ann. N. Y. Acad. Sci. 2011, 1241, 71–81. [Google Scholar] [CrossRef] [PubMed]
  179. Bailey, A.M.; Alberti, F.; Kilaru, S.; Collins, C.M.; de Mattos-Shipley, K.; Hartley, A.J.; Hayes, P.; Griffin, A.; Lazarus, C.M.; Cox, R.J.; et al. Identification and manipulation of the pleuromutilin gene cluster from Clitopilus passeckerianus for increased rapid antibiotic production. Sci. Rep. 2016, 6, 25202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Alberti, F.; Khairudin, K.; Venegas, E.R.; Davies, J.A.; Hayes, P.M.; Willis, C.L.; Bailey, A.M.; Foster, G.D. Heterologous expression reveals the biosynthesis of the antibiotic pleuromutilin and generates bioactive semi-synthetic derivatives. Nat. Commun. 2017, 8, 1831. [Google Scholar] [CrossRef] [Green Version]
  181. Hauser, D.; Sigg, H.P. Isolation and decomposition of sordarin. Helv. Chim. Acta 1971, 54, 1178–1190. [Google Scholar] [CrossRef]
  182. Ogita, T.; Hayashi, T.; Sato, A.; Furutani, W. Antibiotic Substance Zofimarin. JPN Patent No. JPS 6240292A, 21 February 1987. [Google Scholar]
  183. Michael, D.; Sarah, M.; Timm, A.; Olov, S. Hypoxysordarin, a new sordarin derivative from Hypoxylon croceum. Z. Naturforsch. C 1999, 54, 474–480. [Google Scholar] [CrossRef] [Green Version]
  184. Hori, Y.; Nitta, K.; Kobayashi, M.; Takase, S.; Hino, M. Novel Sordarin Derivative as a Therapeutic Antimicrobial Agent. JPN Patent No. WO/2001/000639, 4 January 2001. [Google Scholar]
  185. Hanadate, T.; Tomishima, M.; Shiraishi, N.; Tanabe, D.; Morikawa, H.; Barrett, D.; Matsumoto, S.; Ohtomo, K.; Maki, K. FR290581, a novel sordarin derivative: Synthesis and antifungal activity. Bioorg. Med. Chem. Lett. 2009, 19, 1465–1468. [Google Scholar] [CrossRef]
  186. Büschleb, M.; Dorich, S.; Hanessian, S.; Tao, D.; Schenthal, K.B.; Overman, L.E. Synthetic strategies toward natural products containing contiguous stereogenic quaternary carbon atoms. Angew. Chem. Int. Ed. 2016, 55, 4156–4186. [Google Scholar] [CrossRef] [Green Version]
  187. Justice, M.C.; Hsu, M.J.; Tse, B.; Ku, T.; Balkovec, J.; Schmatz, D.; Nielsen, J. Elongation factor 2 as a novel target for selective inhibition of fungal protein synthesis. J. Biol. Chem. 1998, 273, 3148–3151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Domiínguez, J.M.; Martiín, J.J. Identification of elongation factor 2 as the essential protein targeted by sordarins in Candida albicans. Antimicrob. Agents Chemother. 1998, 42, 2279–2283. [Google Scholar] [CrossRef] [Green Version]
  189. Domínguez, J.M.; Kelly, V.A.; Kinsman, O.S.; Marriott, M.S.; Gómez de las Heras, F.; Martín, J.J. Sordarins: A new class of antifungals with selective inhibition of the protein synthesis elongation cycle in yeasts. Antimicrob. Agents Chemother. 1998, 42, 2274–2278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Chang, Y.C.; Lu, C.K.; Chiang, Y.R.; Wang, G.J.; Ju, Y.M.; Kuo, Y.H.; Lee, T.H. Diterpene glycosides and polyketides from Xylotumulus gibbisporus. J. Nat. Prod. 2014, 77, 751–757. [Google Scholar] [CrossRef]
  191. Kupka, J.; Anke, T.; Oberwinkler, F.; Schramm, G.; Steglich, W. Antibiotics from basidiomycetes. VII. Crinipellin, a new antibiotic from the basidiomycetous fungus Crinipellis stipitaria (Fr.) Pat. J. Antibiot. 1979, 32, 130–135. [Google Scholar] [CrossRef] [Green Version]
  192. Anke, T.; Heim, J.; Knoch, F.; Mocek, U.; Steffan, B.; Steglich, W. Crinipellins, the first natural products with a tetraquinane skeleton. Angew. Chem. Int. Ed. 1985, 24, 709–711. [Google Scholar] [CrossRef]
  193. Piers, E.; Renaud, J. Total synthesis of the tetraquinane diterpenoid (±)-crinipellin B. J. Org. Chem. 1993, 58, 11–13. [Google Scholar] [CrossRef]
  194. Piers, E. Tetraquinane diterpenoids: Total synthesis of (±)-crinipellin B. Synthesis 1998, 1998, 590–602. [Google Scholar] [CrossRef]
  195. Kang, T.; Song, S.B.; Kim, W.Y.; Kim, B.G.; Lee, H.Y. Total synthesis of (−)-crinipellin A. J. Am. Chem. Soc. 2014, 136, 10274–10276. [Google Scholar] [CrossRef] [PubMed]
  196. Li, Y.Y.; Shen, Y.M. Four novel diterpenoids from Crinipellis sp. 113. Helv. Chim. Acta 2010, 93, 2151–2157. [Google Scholar] [CrossRef]
  197. Rohr, M.; Oleinikov, K.; Jung, M.; Sandjo, L.P.; Opatz, T.; Erkel, G. Anti-inflammatory tetraquinane diterpenoids from a Crinipellis species. Biorg. Med. Chem. 2017, 25, 514–522. [Google Scholar] [CrossRef] [PubMed]
  198. Han, J.W.; Oh, M.; Lee, Y.J.; Choi, J.; Choi, G.J.; Kim, H. Crinipellins A and I, two diterpenoids from the basidiomycete fungus Crinipellis rhizomaticola, as potential natural fungicides. Molecules 2018, 23, 2377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Niu, S.; Fan, Z.W.; Xie, C.L.; Liu, Q.; Luo, Z.H.; Liu, G.; Yang, X.W. Spirograterpene A, a tetracyclic spiro-diterpene with a fused 5/5/5/5 ring system from the deep-sea-derived fungus Penicillium granulatum MCCC 3A00475. J. Nat. Prod. 2017, 80, 2174–2177. [Google Scholar] [CrossRef] [PubMed]
  200. Rabe, P.; Rinkel, J.; Dolja, E.; Schmitz, T.; Nubbemeyer, B.; Luu, T.H.; Dickschat, J.S. Mechanistic investigations of two bacterial diterpene cyclases: Spiroviolene synthase and tsukubadiene synthase. Angew. Chem. Int. Ed. 2017, 56, 2776–2779. [Google Scholar] [CrossRef]
  201. Liu, Y.P.; Dai, Q.; Wang, W.X.; He, J.; Li, Z.H.; Feng, T.; Liu, J.K. Psathyrins: Antibacterial diterpenoids from Psathyrella candolleana. J. Nat. Prod. 2020, 83, 1725–1729. [Google Scholar] [CrossRef]
  202. Wang, Q.X.; Qi, Q.Y.; Wang, K.; Li, L.; Bao, L.; Han, J.J.; Liu, M.M.; Zhang, L.X.; Cai, L.; Liu, H.W. Coicenals A–D, four new diterpenoids with new chemical skeletons from the plant pathogenic fungus Bipolaris coicis. Org. Lett. 2013, 15, 3982–3985. [Google Scholar] [CrossRef]
  203. Wang, S.J.; Li, Y.X.; Bao, L.; Han, J.J.; Yang, X.L.; Li, H.R.; Wang, Y.Q.; Li, S.J.; Liu, H.W. Eryngiolide A, a cytotoxic macrocyclic diterpenoid with an unusual cyclododecane core skeleton produced by the edible mushroom Pleurotus eryngii. Org. Lett. 2012, 14, 3672–3675. [Google Scholar] [CrossRef]
  204. Sun, P.X.; Zheng, C.J.; Li, W.C.; Jin, G.L.; Huang, F.; Qin, L.P. Trichodermanin A, a novel diterpenoid from endophytic fungus culture. J. Nat. Med. 2011, 65, 381–384. [Google Scholar] [CrossRef]
  205. Omura, S.; Shiomi, K.; Masuma, R.; Ui, H.; Nagai, T.; Yamada, H. Wickerol and Process for Production Thereof. JPN Patent No. WO 2009116604, 24 September 2009. [Google Scholar]
  206. Yamamoto, T.; Izumi, N.; Ui, H.; Sueki, A.; Masuma, R.; Nonaka, K.; Hirose, T.; Sunazuka, T.; Nagai, T.; Yamada, H.; et al. Wickerols A and B: Novel anti-influenza virus diterpenes produced by Trichoderma atroviride FKI-3849. Tetrahedron 2012, 68, 9267–9271. [Google Scholar] [CrossRef]
  207. Tokiwano, T.; Fukushi, E.; Endo, T.; Oikawa, H. Biosynthesis of phomactins: Common intermediate phomactatriene and taxadiene. Chem. Commun. 2004, 40, 1324–1325. [Google Scholar] [CrossRef]
  208. Yamada, T.; Suzue, M.; Arai, T.; Kikuchi, T.; Tanaka, R. Trichodermanins C–E, new diterpenes with a fused 6-5-6-6 ring system produced by a marine sponge-derived fungus. Mar. Drugs 2017, 15, 169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Yamada, T.; Fujii, A.; Kikuchi, T. New diterpenes with a fused 6-5-6-6 ring system isolated from the marine sponge-derived fungus Trichoderma harzianum. Mar. Drugs 2019, 17, 480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Chart 1. Fungal diterpenoids (2010–2020) classified by skeleton.
Chart 1. Fungal diterpenoids (2010–2020) classified by skeleton.
Jof 08 00244 ch001
Chart 2. Source genera of fungal diterpenoids (2010–2020).
Chart 2. Source genera of fungal diterpenoids (2010–2020).
Jof 08 00244 ch002
Chart 3. The proportion of one activity as compared with the whole occurrence of activities of bioactive fungal diterpenoids (2010–2020).
Chart 3. The proportion of one activity as compared with the whole occurrence of activities of bioactive fungal diterpenoids (2010–2020).
Jof 08 00244 ch003
Figure 1. The evolutionary analysis tree constructed with selected fungi producing cyathane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the internal transcribed spacer (ITS) sequences as follows: Cyathus africanus (JX103204.1), C. earlei (KY964272.1), C. gansuensis (KC869661.1), C. helenae (DQ463334.1), C. hookeri (KC005989.1), C. stercoreus (MH543350.1), C. striatus (KU865513.1), C. subglobisporus (MH156046.1), Gerronema albidum (MF318924.1), Hericium erinaceus (KU855351.1), H. flagellum (MG649451.1), H. ramosum (U27043.1), H. sp. WBSP8 (MN243091.1), Hydnum repandum (LC377888.1), Laxitextum incrustatum (KT722621.1), Phellodon niger (MH310794.1), Sarcodon glaucopus (MT955152.1), S. scabrosus (MN992643.1), Strobilurus tenacellus (MF063128.1). Since the ITS sequence of Sarcodon cyrneus was not available, Sarcodon sp. (MK049936.1) was selected, since it is in the same family with S. cyrneus. The evolutionary history was inferred by using the maximum likelihood method and the general time reversible model [62]. The bootstrap consensus tree inferred from 1000 replicates is taken to represent the evolutionary history of the taxa analyzed [63]. Branches corresponding to partitions reproduced in less than 50% bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches [63]. Initial tree(s) for the heuristic search were obtained automatically by applying the Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances estimated using the maximum composite likelihood (MCL) approach, and then selecting the topology with superior log likelihood value. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 1.2219)). The rate variation model allowed for some sites to be evolutionarily invariable ([+I], 0.00% sites). This analysis involved 20 nucleotide sequences. Codon positions included were 1st + 2nd + 3rd + noncoding. There were 924 positions in the final dataset. The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Figure 1. The evolutionary analysis tree constructed with selected fungi producing cyathane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the internal transcribed spacer (ITS) sequences as follows: Cyathus africanus (JX103204.1), C. earlei (KY964272.1), C. gansuensis (KC869661.1), C. helenae (DQ463334.1), C. hookeri (KC005989.1), C. stercoreus (MH543350.1), C. striatus (KU865513.1), C. subglobisporus (MH156046.1), Gerronema albidum (MF318924.1), Hericium erinaceus (KU855351.1), H. flagellum (MG649451.1), H. ramosum (U27043.1), H. sp. WBSP8 (MN243091.1), Hydnum repandum (LC377888.1), Laxitextum incrustatum (KT722621.1), Phellodon niger (MH310794.1), Sarcodon glaucopus (MT955152.1), S. scabrosus (MN992643.1), Strobilurus tenacellus (MF063128.1). Since the ITS sequence of Sarcodon cyrneus was not available, Sarcodon sp. (MK049936.1) was selected, since it is in the same family with S. cyrneus. The evolutionary history was inferred by using the maximum likelihood method and the general time reversible model [62]. The bootstrap consensus tree inferred from 1000 replicates is taken to represent the evolutionary history of the taxa analyzed [63]. Branches corresponding to partitions reproduced in less than 50% bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches [63]. Initial tree(s) for the heuristic search were obtained automatically by applying the Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances estimated using the maximum composite likelihood (MCL) approach, and then selecting the topology with superior log likelihood value. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 1.2219)). The rate variation model allowed for some sites to be evolutionarily invariable ([+I], 0.00% sites). This analysis involved 20 nucleotide sequences. Codon positions included were 1st + 2nd + 3rd + noncoding. There were 924 positions in the final dataset. The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Jof 08 00244 g001
Figure 2. The evolutionary analysis tree constructed with selected fungi producing cyclopiane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the ITS sequences as follows: Penicillium commune MCCC 3A00940 (KY978585.1), P. sp. F23-2 (EU770318.1), P. sp. YPGA11 (MG835908.1), P. sp. TJ403-2 (MK613138.1), P. chrysogenum MT-12 (MF765611.1), P. chrysogenum QEN-24S (GU985086.1), P. roqueforti IFM 48062 (AB041202.1), and Leptosphaeria sp. XL026 (MK603060.1). Since the ITS sequence of strain P. cyclopium IMI 229034 was not available, P. cyclopium IFM 41611 (AB041169.1) was selected, since it was in the same family as P. cyclopium. The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Figure 2. The evolutionary analysis tree constructed with selected fungi producing cyclopiane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the ITS sequences as follows: Penicillium commune MCCC 3A00940 (KY978585.1), P. sp. F23-2 (EU770318.1), P. sp. YPGA11 (MG835908.1), P. sp. TJ403-2 (MK613138.1), P. chrysogenum MT-12 (MF765611.1), P. chrysogenum QEN-24S (GU985086.1), P. roqueforti IFM 48062 (AB041202.1), and Leptosphaeria sp. XL026 (MK603060.1). Since the ITS sequence of strain P. cyclopium IMI 229034 was not available, P. cyclopium IFM 41611 (AB041169.1) was selected, since it was in the same family as P. cyclopium. The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Jof 08 00244 g002
Scheme 1. (A) Reaction catalyzed by the prenyltransferase domain of PcCS; (B) reaction catalyzed by the terpene synthase domains of PcCS [69].
Scheme 1. (A) Reaction catalyzed by the prenyltransferase domain of PcCS; (B) reaction catalyzed by the terpene synthase domains of PcCS [69].
Jof 08 00244 sch001
Scheme 2. Proposed cyclization mechanism catalyzed by PchDS/PrDS [71].
Scheme 2. Proposed cyclization mechanism catalyzed by PchDS/PrDS [71].
Jof 08 00244 sch002
Figure 3. The evolutionary analysis tree constructed with selected fungi producing fusicoccane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the ITS sequences as follows: Alternaria brassicicola XXC (KR779774.1), Penicillium sp. DT10 (MH458525.1), Periconia sp. No. 19-4-2-1 (KP873157.1), Roussoella hysterioides KT1651 (KJ474829.1), Talaromyces stipitatus (MH857968.1), Talaromyces purpurogenus (MH120320.1), and Trichoderma citrinoviride cf-27 (KT259441.1). The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Figure 3. The evolutionary analysis tree constructed with selected fungi producing fusicoccane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the ITS sequences as follows: Alternaria brassicicola XXC (KR779774.1), Penicillium sp. DT10 (MH458525.1), Periconia sp. No. 19-4-2-1 (KP873157.1), Roussoella hysterioides KT1651 (KJ474829.1), Talaromyces stipitatus (MH857968.1), Talaromyces purpurogenus (MH120320.1), and Trichoderma citrinoviride cf-27 (KT259441.1). The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Jof 08 00244 g003
Scheme 3. Proposed biosynthetic pathways of talaronoids A–D 2225 [80].
Scheme 3. Proposed biosynthetic pathways of talaronoids A–D 2225 [80].
Jof 08 00244 sch003
Scheme 4. Hypothetical biosynthetic pathways for alterbrassicene A 47 and alterbrassicicene A 48 [78,98].
Scheme 4. Hypothetical biosynthetic pathways for alterbrassicene A 47 and alterbrassicicene A 48 [78,98].
Jof 08 00244 sch004
Scheme 5. (A) Biosynthetic gene clusters of the brassicicenes in P. fijiensis and A. brassicicola; (B) proposed biosynthetic pathway for brassicicenes (dashed arrows are those deduced from expected protein function) [117].
Scheme 5. (A) Biosynthetic gene clusters of the brassicicenes in P. fijiensis and A. brassicicola; (B) proposed biosynthetic pathway for brassicicenes (dashed arrows are those deduced from expected protein function) [117].
Jof 08 00244 sch005
Scheme 6. Mechanistic hypothesis for the cyclization of GGPP to myrothec-15(17)-en-7-ol 86 and myrotheca-7,15(17)-diene 87 [118].
Scheme 6. Mechanistic hypothesis for the cyclization of GGPP to myrothec-15(17)-en-7-ol 86 and myrotheca-7,15(17)-diene 87 [118].
Jof 08 00244 sch006
Figure 4. The evolutionary analysis tree constructed with selected fungi producing guanacastane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the ITS sequences as follows: Coprinus heptemerus D99052 (JN159553.1), Coprinus radians M65 (HM045514.1), Coprinus plicatilis 82 (Parasola plicatilis) (FM163216.1), Psathyrella candolleana (MF401519.1), Cercospora sp. (KF577929.1), and Verticillium dahlia (HQ839784.1). Since the ITS sequence of Cortinarius pyromyxa was not available, Cortinarius misermontii (NR_130230.1) was selected since their ITS sequences were the most similarly. The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Figure 4. The evolutionary analysis tree constructed with selected fungi producing guanacastane diterpenoids. The evolutionary analysis was reconstructed by the maximum likelihood method from the ITS sequences as follows: Coprinus heptemerus D99052 (JN159553.1), Coprinus radians M65 (HM045514.1), Coprinus plicatilis 82 (Parasola plicatilis) (FM163216.1), Psathyrella candolleana (MF401519.1), Cercospora sp. (KF577929.1), and Verticillium dahlia (HQ839784.1). Since the ITS sequence of Cortinarius pyromyxa was not available, Cortinarius misermontii (NR_130230.1) was selected since their ITS sequences were the most similarly. The evolutionary analysis was conducted in MEGA X (version 10.2.2) [64].
Jof 08 00244 g004
Scheme 7. Biosynthetic mechanism to harziene and taxadiene scaffolds [151,152].
Scheme 7. Biosynthetic mechanism to harziene and taxadiene scaffolds [151,152].
Jof 08 00244 sch007
Scheme 8. Proposed biosynthetic pathway of phomopsene 190 and methyl phomopsenonate 191 [164].
Scheme 8. Proposed biosynthetic pathway of phomopsene 190 and methyl phomopsenonate 191 [164].
Jof 08 00244 sch008
Scheme 9. Proposed cyclization mechanism catalyzed by PaPS [165].
Scheme 9. Proposed cyclization mechanism catalyzed by PaPS [165].
Jof 08 00244 sch009
Scheme 10. Proposed biosynthetic pathway to pleuromutilin 192 in Clitopilus passeckerianus [180].
Scheme 10. Proposed biosynthetic pathway to pleuromutilin 192 in Clitopilus passeckerianus [180].
Jof 08 00244 sch010
Scheme 11. The biosynthetic pathway for sordarin 195 and hypoxysordarin 197 [114].
Scheme 11. The biosynthetic pathway for sordarin 195 and hypoxysordarin 197 [114].
Jof 08 00244 sch011
Scheme 12. Proposed biosynthetic pathway for psathyrins A 216 and B 217 [201].
Scheme 12. Proposed biosynthetic pathway for psathyrins A 216 and B 217 [201].
Jof 08 00244 sch012
Scheme 13. Plausible biogenetic origin of eryngiolide A 223 [203].
Scheme 13. Plausible biogenetic origin of eryngiolide A 223 [203].
Jof 08 00244 sch013
Scheme 14. Incorporation patterns of [1-13C]-, [2-13C]-, and [1,2-13C2]-acetates enriched wickerol B 226, and proposed mechanism of cyclization from GGPP to wickerols [206].
Scheme 14. Incorporation patterns of [1-13C]-, [2-13C]-, and [1,2-13C2]-acetates enriched wickerol B 226, and proposed mechanism of cyclization from GGPP to wickerols [206].
Jof 08 00244 sch014
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, F.-L.; Feng, T. Diterpenes Specially Produced by Fungi: Structures, Biological Activities, and Biosynthesis (2010–2020). J. Fungi 2022, 8, 244. https://doi.org/10.3390/jof8030244

AMA Style

Zhang F-L, Feng T. Diterpenes Specially Produced by Fungi: Structures, Biological Activities, and Biosynthesis (2010–2020). Journal of Fungi. 2022; 8(3):244. https://doi.org/10.3390/jof8030244

Chicago/Turabian Style

Zhang, Fa-Lei, and Tao Feng. 2022. "Diterpenes Specially Produced by Fungi: Structures, Biological Activities, and Biosynthesis (2010–2020)" Journal of Fungi 8, no. 3: 244. https://doi.org/10.3390/jof8030244

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop