Next Article in Journal
Carrageenophyte Kappaphycus malesianus Inhibits Microglia-Mediated Neuroinflammation via Suppression of AKT/NF-κB and ERK Signaling Pathways
Next Article in Special Issue
Epigenetic Manipulation Induced Production of Immunosuppressive Chromones and Cytochalasins from the Mangrove Endophytic Fungus Phomopsis asparagi DHS-48
Previous Article in Journal
Structural and Biochemical Analysis Reveals Catalytic Mechanism of Fucoidan Lyase from Flavobacterium sp. SA-0082
Previous Article in Special Issue
Immunosuppressive Cytochalasins from the Mangrove Endophytic Fungus Phomopsis asparagi DHS-48
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unusual Secondary Metabolites from the Mangrove Ecosystems: Structures, Bioactivities, Chemical, and Bio-Syntheses

1
State Key Laboratory of Drug Research, Shanghai Institute of Materia Medica, Chinese Academy of Sciences, 555 Zu Chong Zhi Road, Shanghai 201203, China
2
Collaborative Innovation Center of Yangtze River Delta Region Green Pharmaceuticals and College of Pharmaceutical Science, Zhejiang University of Technology, Hangzhou 310014, China
3
Shandong Laboratory of Yantai Drug Discovery, Bohai Rim Advanced Research Institute for Drug Discovery, Yantai 264117, China
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2022, 20(8), 535; https://doi.org/10.3390/md20080535
Submission received: 28 July 2022 / Revised: 17 August 2022 / Accepted: 18 August 2022 / Published: 20 August 2022
(This article belongs to the Special Issue Bio-Active Products from Mangrove Ecosystems)

Abstract

:
Mangrove ecosystems are widely distributed in the intertidal zone of tropical and subtropical estuaries or coasts, containing abundant biological communities, for example, mangrove plants and diverse groups of microorganisms, featuring various bioactive secondary metabolites. We surveyed the literature from 2010 to 2022, resulting in a collection of 134 secondary metabolites, and classified them into two major families in terms of the biological sources and 15 subfamilies according to the chemical structures. To highlight the structural diversity and bioactivities of the mangrove ecosystem-associated secondary metabolites, we presented the chemical structures, bioactivities, biosynthesis, and chemical syntheses.

Graphical Abstract

1. Introduction

Identifying lead compounds is one of the biggest challenges in drug discovery. Natural products (NPs) and their intricate molecular frameworks have a long tradition as valuable starting points for drug development (for example, artemisinins, taxol, camptothecin, and penicillins). To date, NPs remain a significant source of new compounds, providing a wide range of structural diversities with multiple privileged scaffolds for drug discovery either directly, semi-synthetically, or as a source of inspiration [1,2,3,4,5]. However, discovering new bioactive NPs is generally time-consuming and laborious. Only a few new NP drug pharmacophores have been found over the past 20 years, representing critical issues for NPs-driven lead discovery campaigns.
Mangrove forests are complex ecosystems widely distributed in tropical and subtropical estuaries or coastal intertidal zones. These forests contain diverse biological communities, including mangrove plants and numerous groups of microorganisms. The environment of the mangrove system harbors unique traits, for instance, high salinity, low oxygen, tidal gradients, high temperature, and excessively intense light, resulting in an active ecosystem with various microorganisms [1,2,3,4,5]. Mangrove-associated microorganisms have been demonstrated to be a reliable source of bioactive metabolites and have likewise drawn the attention of NP researchers [6,7,8,9,10,11,12]. A large number of structurally unusual and bioactive NPs have been discovered from the mangrove-associated microorganisms, including fungal and bacterial endophytes isolated from the mangrove plants’ leaves, branches, and roots [13,14]. In addition, Mangrove sediments-derived microbes, a rich reservoir of NP diversity, could be utilized to explore new drugs [15].
The bioactive NPs solely from the true mangrove and semi-mangrove floras worldwide have been summarized in several review papers in 2010 [16,17]. However, the investigation of the mangrove ecosystem is mainly focused on the mangrove-associated microorganism but less on the mangrove flora in recent years. Three comprehensive reviews have recently focused on the NPs from mangrove-associated fungi and mangrove sediments-derived microbes [6,14,15]. However, to our knowledge, no review has been published on the secondary metabolites with unusual skeletons from the mangrove ecosystem. They might merit the attention of chemists and biologists and could be a source of fresh pharmacophores with biological activity for the creation of drugs based on natural products.
In this review, we focus on the mangrove ecosystem-associated NPs featuring new carbon scaffolds, unique ring systems, or unusual functional moieties covering from 2010 to 2022. The structures, biological activities, biosynthesis, and total chemical syntheses of exampled unique compounds were included.
The references were searched using the following keywords as the subject search: natural products/secondary metabolites, mangrove, via Web of Science, Chemical Abstracts, and PubMed databases covering from 2010 to 2022, resulting in a collection of 134 unusual secondary metabolites. We classified them into two major families in terms of biological sources.

2. Unusual Natural Products from Mangrove Flora

2.1. Limonoids

Limonoids are natural tetranortriterpenoids containing a four-ring structure with a 17β-furyl ring mainly distributed in the Meliaceae and Rutaceae families [18]. In the mangrove flora, they are especially abundant and structurally diversified in the genus Xylocarpus moluccensis and X. grantum (family Meliaceae). Up to 2021, approximately 2700 limonoids have been identified. Moreover, owing to their widespread distribution and substantial content in Meliaceae plants, and active biosynthetic pathways, more than 1600, including 30 types of unique rearrangement skeletons, have been isolated and identified in the last 10 years [19]. Among them, nearly 233 new limonoids with 14 kinds of novel skeletons were isolated from mangrove Xylocarpus.
Thaixylomolin A (1), isolated from the seeds of a Thai mangrove Xylocarpus moluccensis collected at the Trang province, was obtained similar to the cleavage of C-6/C-7 by Baeyer–Villiger (BV) oxidation [20], and then the oxidized C-6 formed an unusual 6-oxabicyclo[3.2.1]octan-3-one motif with C-1 [21]. In 2016, the same research group isolated another analogue from X. moluccensis, thaixylomolin R (2) [22], whose C-8 is decarboxylated compared with 1 (Figure 1).
Xylomexicanin F (3) [23], hainangrantums I and J (4 and 5) [24] (Figure 2) are the second examples of a limonoid with an unusual 9, 10-seco and C-9-C-30 linkage, isolated from the seeds of the Chinese mangrove X. granatum. Among them, 3 showed moderate activity against the A549 and RERF cell lines with IC50 values of 18.83 μM and 15.83 μM, respectively. However, the first reported analogue, xylogranatin D, was concluded as an artifact [25,26].
Chemical investigation of the seeds from a Trang (Thailand) mangrove X. moluccensis yielded five structurally intriguing limonoids, namely, trangmolins A–E (610) [27] (Figure 3). Notably, compounds 68 consist of unprecedented ring A/B-fused bicyclic moieties, and compound 10 represents the first example of the oxidative cleavage of the C2-C3 bond among limonoids. In 2021, a trangmolin A derivative krishnolide J (11) was isolated from seeds of the India Krishna mangrove X. moluccensis [28]. The biosynthetic origins of 611 could be traced back to a proposed andirobin-type limonoid with 1,2-bisketone groups [18]. Taking andirobin as the starting point, scientists from the Wu group proposed a biosynthetic pathway characterized by a highly divergent biosynthetic assembly line (Scheme 1) [27]. The three forks of the biosynthetic pathway obtain C-1/C-30 linkage (68, 11), C-3/C-30 linkage (9), and C-2/C-30 linkage (10) based on the main mechanisms of electro- and nucleophilic enzymatic cascade reactions. The diverse cyclization patterns of 611 reveal the remarkable structural plasticity of rings A and B in limonoid biosynthesis.
Andhraxylocarpins A–E (1216) (Figure 4) were isolated and identified as three new types of limonoids from the seeds of two true mangroves, X. granatum (collected at the estuary of Krishna, India) and X. moluccensis (collected in the estuary of Godavari, India), respectively [29]. Among them, andhraxylocarpins A and B (12 and 13) contain an unprecedented 9-oxa-tricyclo-[3.3.2.17,10]undecan-2-ene motif, andhraxylocarpins C-D (14 and 15) harbor a rare (Z)-bicyclo[5.2.1]dec-3-en-8-one substructure, and andhraxylocarpin E (16) possesses a unique tricyclo[3.3.1.13,6]decan-9-one scaffold. In 2016, trangmolin F (17), which shared the same A/B fused carbobicyclic with 16, was obtained from X. moluccensis by the same group [27] Wu et al. suggested a mexicanolide with a Δ8,30 double bond, derived from an andirobin by C-2/C-30 linkage and previously discovered among the genus Xylocarpus, may be the precursor of 1216 [29] (Scheme 1). The presence of bridging rings (C10–C1–C2) in mexicanolide-type limonoids makes C-3 and C-30 close to each other in space, which leads to their coupling.
Krishnadimer A (18) (Figure 5) is the first dimeric limonoid isolated from the seeds of X. moluccensis with an unprecedented axial chirality architecture, with the C2-symmetric architecture, with a P-configured central axis at the C15, C15′-positions of the monomeric units, represents a milestone during decades of work on natural limonoids [30]. It could be obtained by the intermolecular oxidative coupling of the phargmalin-class limonoid, which can be derived from andirobin through C-2/C-30 and C-1/C-29 linkage. (Scheme 1) The semisynthesis of the dimer was successfully conducted. Subsequently, eight new limonoid dimers of four skeletons (two symmetric and two nonsymmetric) were designed and synthesized by oxidative carbon-carbon radical coupling [31].
Two unprecedented limonoids, thaixylomolins B and C (19 and 20) (Figure 5), co-isolates with 1 [21], are limonoids containing a unique pentasubstituted pyridine scaffold that might be generated by aromatization into a pyridine ring from a phargmalin-class limonoid. (Scheme 1) Thaixylomolin B (19) exhibited inhibitory activity against nitric oxide production in lipopolysaccharide and IFN-γ-induced RAW264.7 murine macrophages with an IC50 value of 84.3 μM.
Two pyridine-containing limonoids, xylogranatopyridines A (21) and B (22) (Figure 6), were isolated from the twigs and leaves of X. granatum, collected from the seashore of Dongzhai, Hainan Province [32]. Compared to 21, xylogranatopyridine B (22) possesses an unprecedented rearranged A-ring. Prexylogranatopyridine, a co-occurrence of limonoid with an unusual C-8-C-30 linkage, could be the common biosynthetic precursor of 21 and 22 (Scheme 2). Xylogranatopyridine A (21) exhibited significant inhibitory activity against protein tyrosine phosphatase 1B (PTP1B) with an IC50 value of 22.9 μM.
An unusual tetranortriterpenoid, xylomexicanins E (23) (Figure 7), which is the first example of limonoid with azaspiro skeleton between B (pyrrolidine) and C rings, was isolated from the seeds of the Chinese mangrove, X. granatum [23]. The plausible biosynthetic routes are proposed, as shown in Scheme 2, starting from the limonoid prexylogranatopyridine.
Further investigation of the seeds from the X. granatum led to the isolation of two new tetranortriterpenoids, xylomexicanins I and J (24 and 25) [33] (Figure 7). Notably, 24 represents an unprecedented limonoid with a bridged skeleton between the B- and C-rings, contrasting with analogues possessing bridged A- and B-rings (25). Wu et al. proposed that 24 was obtained from the same natural precursor as 22 after an enolate addition to the allylic alcohol moiety between C-3 and C-11 (Scheme 2).
Three new limonoids, entitled xylomolones A–C (2628, respectively, Figure 8) were discovered from the seeds of the Thai mangrove X. moluccensis, as well as a vital biosynthetic precursor, xylomolone D (a new C11-terpenic acid methyl ester) [34]. Compared to 26, compound 27 is the first 9,10-seco limonoid with a 3,4-dihydro-2H-pyran motif and possesses the reversed alignment of ring A. For the biosynthesis of xylomolone C (28), a five-membered A-ring could be built through a benzylic acid-like rearrangement, forming an unusual 3-oxabicyclo[3.2.1]octan-2,7-dione motif; the C-2 is excluded from the A-ring in the rearrangement process. Wu et al. proposed a novel convergent strategy for limonoid biosynthesis (Scheme 3).

2.2. Diterpenoids

Two new ent-isopimarane-type diterpenoids agallochaexcoerins D and E (29 and 30) [35] (Figure 9), possessing an unusual seven-membered lactone moiety, were isolated from the wood of mangrove Excoecaria agallocha. Notably, 29 is the first report of ent-3,4-secoisopimaratriene diterpenoid with a rare seven-membered ring.
Decandrinin (31) (Figure 9), an unprecedented C-9-spiro-fused 7,8-seco-ent-abietane, was obtained from the bark of an Indian mangrove Ceriops decandra (collected in the estuary of Godavari, Andhra Pradesh) [36]. The biosynthetic precursor might be the naturally more prevalent occurring 7,13-ent-abietadien-3β-ol and the plausible biosynthetic was proposed (Scheme 4). The spiro ring could be formed by oxidative cleavage and lactonization.

3. Unusual Natural Products from Mangrove-Associated Microorganisms

Previous chemical investigations of mangrove microbes especially mangrove-associated fungi resulted in the discovery of various bioactive secondary metabolites, including polyketides, terpenes, alkaloids, and peptides with diverse structural features.

3.1. Polyketides

Polyketides (PKs) are a large family of secondary metabolites with prominent structural diversity and various biological activities, isolated from diverse organisms. Polyketide synthases (PKSs) catalyze the sequential decarboxylative condensations of acyl-CoA thioesters to form structurally diverse PKs [37]. We direct readers to the fantastic reviews for more information on PKSs [38,39,40,41]. An even-increasing number of PKs from the mangrove-associated microorganisms are being reported.

3.1.1. Coumarins and Isocoumarins

Naturally occurring coumarins and isocoumarins are an essential class of benzopyrene derivatives and are present in remarkable amounts in plants, while only a few are found in microorganisms and animal sources.
Up to now, 12 new coumarin and 102 new isocoumarin derivatives have been obtained from mangrove-associated fungi [42,43]. Among them, Peniisocoumarin A and B (32 and 33, Figure 10), a pair of unusual dimeric isocoumarin-type diastereoisomers containing a symmetric four-membered core at C-9/9′ and C-10/10′ were obtained from the fermentation of Penicillium commune QQF-3 (isolated from fresh fruit of the mangrove plant Kandelia candel) [44]. The structures of 32 and 33 were unanimously defined by X-ray diffraction analysis using Cu Kα radiation. In 2015, Darsih et al., discovered penicilliumolide A (34) (Figure 10), an unusual tetracyclic isocoumarin containing a γ-lactone ring fused with a unique spiro framework, from the mangrove endophytic fungus Penicillium chermesinum HLit-ROR2 [45].

3.1.2. Chromones

The chromone and its derivatives have been identified as the central backbone in several functional organic compounds, with strategic importance in many research and industrial domains. Until now, 85 new chromone derivatives have been identified from the mangrove-associated fungal species.
In 2019, two new polycyclic chromones, penixanthones C (35) and D (36) (Figure 11), possessing an unprecedented 6/6/6/5 polycyclic skeleton with a signature C2 bridge, were isolated from the mangrove sediment-derived fungus Penicillium sp. SCSIO041218 [46]. However, 35 and 36 only showed weak cytotoxicity. Furthermore, the configurations of 35 and 36 remain elusive.

3.1.3. Azaphilones

Azaphilones or azaphilonoids are a structurally variable family of fungal polyketide metabolites harboring a highly oxygenated pyranoquinone bicyclic core [47]. In recent years, about 31 azaphilones with unusual structures and remarkable bioactivities were reported from mangrove-associated fungi, including genera Aspergillus, Diaporthe, Penicillium, and Talaromyces.
Two new citrinin derivatives, penicitol A (37) and penicitol B (38) (Figure 12) were identified from Penicillium chrysogenum ML226 obtained from the rhizosphere soil of the mangrove plant Acanthus ilicifolius [48]. The citrinin derivatives are a family of azaphilones, with the first one, namely citrinin, isolated from a P. citrinum strain in 1931 [49]. Penicitol A (37) is the first reported citrinin derivative with an unusual tetracyclic skeleton, and 38 is the first citrinin dimer with a single oxygen bridging center. 37 and 38 exhibited potent cytotoxic activities against HeLa, BEL-7402, HEK-293, HCT-116, and A-549 cell lines with IC50 values of 4.6–10.5 and 3.4–9.6 μM, respectively. In 2011, Hosokawa et al. reported the first total synthesis of penicitol A (37), achieved by acetalization [50].

3.1.4. Benzophenones Derivatives

Benzophenones (BPs) are widely distributed NPs possessing a diphenyl ketone moiety [51]. Given the presence of the chemically active carbonyl group, which can efficiently react with a variety of functional groups, diverse novel skeletons such as isobenzofuran, isoindolinone, and 3-dihydro-1H-indene, etc., could be formed. There are 12 members of BPs that are discovered from mangrove-associated fungi.
Four unusual 2,3-dihydro-1H-indene isomers, diaporindenes A–D (3942) (Figure 13), and an unusual isoprenylisobenzofuran A (43) were isolated from Diaporthe sp. SYSU-HQ3, a fungus obtained from the branches of the mangrove plant Excoecaria agallocha collected from Zhuhai in Guangdong province, China [52]. Compounds 3942 feature a 2,3-dihydro-1H-indene ring and a 1,4-benzodioxan moiety. Isoprenylisobenzofuran A (43) represented the first example of an isoprenylisobenzofuran nucleus possessing a rare 1,4-benzodioxan moiety. Biosynthetically, compounds 3943 could be derived from co-occurrence benzophenone type metabolites such as tenellone B, which is formed by acetyl-CoA and malonyl-CoA through the catalysis of PKSs [53] (Scheme 5). In a bioassay, compounds 3943 were found to exhibit significant inhibitory effects against nitric oxide production with IC50 values from 4.2–9.0 μM and selective index (SI) values from 3.5 to 6.9.

3.1.5. Macrolides

Macrolides, especially those possessing 10- to 19-membered ring systems, have diversified structural features and constitute a prominent group of active secondary metabolites. Since the discovery of well-known progenitor macrolide antibiotic pikromycin in 1950 and the second generation of macrolides such as azithromycin and clarithromycin, naturally occurring macrolides have been found today due to their diverse structures and promising biological properties [54]. A total of 63 macrolides have been isolated from the mangrove-associated fungi.
Sumalarins A−C (4446) (Figure 14) were identified from the cytotoxic extract of Penicillium. sumatrense MA-92 from the rhizosphere of the mangrove Lumnitzera racemose. Notably, they were the unusual and rare examples of sulfur-containing curvularin derivatives isolated for the first time from natural sources [55]. Compounds 4446 displayed cytotoxic activities against Du145, HeLa, Huh 7, MCF-7, NCI-H460, SGC-7901, and SW1990 cell lines with IC50 values ranging from 3.8 μM to 10 μM. Compound 44 is likely formed via Michael’s addition of 3-mercaptolactate to the double bond Δ10,11 of dehydrocurvularin. Esterification or acylation of 45 probably leads to the biosynthesis of 44 and 46 [55].
Ansamycins are characterized by an aromatic nucleus connected with a polyketide chain back to a nonadjacent position through an amide bond. Hertweck et al. isolated four unusual ansa macrolides, compounds 4750 (Figure 15), from Streptomyces sp. HKI0576, a bacterial endophyte separated from the stem of mangrove Bruguiera gymnorrhiza [56]. This was the first report on discovering ansamycins from a plant endophyte. In addition, the degree of “in-built diversification” of these four compounds is unprecedented for complex polyketides. Among them, divergolide A (47) represents an unusual type of ansa macrolide with an unusual branched side chain and a disrupted polyketide backbone. Furthermore, the tricyclic chromophore is unprecedented for macrolides, and related O-heterocyclic substructures are only known from aromatic polyketides, such as the nogalamycin aglycone [57] and chaetoxanthone [58]. Divergolide B (48) represents another unusual type of ansa macrolide featuring a novel benzopyran/chromene core as the first congener of 47. In addition, compounds 49 and 50 share substructures with 47 and 48 but feature structurally intriguing tetracyclic scaffolds. Furthermore, the ansa macrolides display significant antimicrobial and cytotoxic activities, probably regulating the immunity of the mangrove tree. Compounds 4750 are biosynthesized from a common linear polyketide using 3-amino-5-hydroxybenzoic acid (AHBA) as a primer unit. Various reactions, including an optional acyl migration, generate the diverse multicyclic structures [56,59] (Scheme 6).
In 2014, Shen et al., cloned the biosynthetic gene cluster involved in the biosynthesis of the divergolides from the endophytic Streptomyces sp. W112 isolated from Camptotheca acuminata. Following gene disruption, gene overexpression, and bioinformatics analysis, they laid the foundation for further elucidation of the biosynthetic pathway as well as titer improvement [60]. In addition, Zhong et al. [61] conducted genome sequencing, bioinformatics analysis, and further isolations of four new divergolide congeners with a similar endophytic bacteria, Streptomyces sp. from Bruguiera gymnorrhiza. They showed that specialized acyltransferase domains are for selecting extender units, and the branched isobutylmalonyl-CoA is involved.
The total synthesis of divergolide A using the ring-closing metathesis (RCM) approach was published by Dai et al., in 2012 [62]. Subsequently, Rasapalli et al. synthesized the western section of divergolides C (49) and D (50) and demonstrated the robustness of C4-C5 as an appropriate approach for the further total synthesis of divergolides C and D in 2013 [63]. This chemical method was also conducted for divergolides A and B. Studies on the total synthesis of divergolides A-D using inexpensive, readily available starting materials and simple operations have also been constantly reported in recent years [64,65,66].
A macrocyclic polyketide with an unusual carbon skeleton, namely hainanmycin A (51) (Figure 15), was isolated from Streptomyces sp. 219807 (from mangrove soil collected in Sanya) [67]. Compound 51 featured an unprecedented structural skeleton of a 17-membered carbocyclic framework. The cyclo-heptadeca framework containing a cyclopentenone ring substituted with a naturally occurring bridgehead enol motif is unique among NPs. It represents a new subgroup, a minor family of carbocyclic polyketide macrolides. Hong et al. [67] proposed a plausible biosynthetic pathway for 51 based on the biosynthesis of akaeolide [68], an analogue of 51. Shortly, the PKS condenses acetyl-CoA and other building units (e.g., methylmalonyl-CoA and malonyl-CoA) to a linearized polyketide backbone. A thioesterase (TE) then releases the backbone with the formation of a δ-lactone ring. Further construction of the C-C bonds of C-16/C-12 and C-18/C-2 generate the structural core. Notably, a C-18 aldehyde intermediate (S1) might be involved in the C-18/C-12 carbon bond formation (Scheme 7).

3.1.6. Others

Eight new compounds, streptoglycerides A–H (5259) (Figure 16) possessing a unique ring system, were obtained from Streptomyces sp. isolated from a mangrove sample collected on Kosrae Island [69,70]. This is the first report to describe a rare 6/5/5 tricyclic ring system consisting of a glycerol moiety from marine organisms. Streptoglyceride C (54) showed a weak inhibitory effect on nitric oxide production in BV-2 microglia cells. Compounds 5659 showed significant anti-inflammatory activity by inhibiting lipopolysaccharide (LPS)-induced nitric oxide (NO) production in Raw 264.7 cells with IC50 values ranging from 3.5 to 10.9 μM. It should be noted that 57 suppressed the transcription of iNOS and IL-6 without cytotoxicity.
Upon further investigation of the unusual strain, four new compounds, miharadienes A–D (6063), possessing unique ring systems and a rare diene side chain, were isolated [71]. A plausible biosynthetic pathway of miharadienes and related compounds, streptoglycerides is proposed in the literature (Scheme 8a) [71]. However, the formation of 5255 from 63 by attacking the nucleophilic hydroxy on the electron-rich furan ring seems inapplicable. Therefore, we proposed an optional pathway for 5255 (Scheme 8b). In short, the starting lauryl alcohol derivative appears to react with dihydroxyacetone, an oxidation product of glycerol, to form the intermediate int i through Aldol type reaction of the C-4 active methylene of lauryl alcohol derivative with the carbonyl of the dihydroxyacetone. Then the hemiketal formation gives the tetrahydrofuran ring, and ether formation forms the other tetrahydrofuran ring. Afterward, the ketal formation by the interaction of the terminal hydroxyethylene with the hemiketal provides the pyran ring and affords the intermediate int ii, which could be further converted into 5255, possessing a rare 6/5/5 ring system.

3.2. Terpenoids

The new terpenoids from mangrove fungi can be divided into seven groups based on their chemical structures and biosynthetic pathways: monoterpenes, sesquiterpenes, diterpenes, sesterterpenes, triterpenes, and meroterpenes. Sesquiterpenes (138), sesterterpenes (36), and meroterpenes (72) comprise the most significant proportions of new terpenes from mangrove fungi. However, monoterpenes, diterpenes, and triterpenes were rarely isolated from mangrove fungi, and no new skeleton was discovered.

3.2.1. Sesquiterpenoids

Sesquiterpenoids are the largest group of known terpenoids [72]. The mangrove fungi-derived sesquiterpenoids possess a variety of carbon skeletons, including monocyclic, bicyclic, and tricyclic types [73].
One tricyclic and three spirobicyclic norsesquiterpenoids (6467) (Figure 17) were isolated from the endophytic fungus Pseudolagarobasidium acaciicola (from the mangrove Bruguiera gymnorrhiza) [74,75]. Among them, acaciicolin A (64) possesses a previously unknown skeleton with a uniquely connected 6/5/5 ring system and three consecutive oxygenated sp3 quaternary carbons at C-7, C-8, and C-8a. The norsesquiterpene skeleton of 64 was named “acaciicolane”, and was different from the three known sesquiterpene skeletons with 6/5/5 ring systems: cedrane, prezizaane, and zizaane (Figure 18). Spiroacaciicolides A–C (6567) has a hitherto unobserved 5/6 fused spirobicyclic ring system. The absolute configurations of the new compounds 6466 were determined by single-crystal X-ray analysis (Cu-Kα radiation). 6467 could originate from chamigrane endoperoxide A [76] (Scheme 9).
Penicibilaenes A (68) and B (69) (Figure 17), two sesquiterpenes possessing a tricyclo[6.3.1.01,5]dodecane skeleton constituted by [3.3.1]-bridged and [4.3.0]-fused junctions, were characterized from Penicillium bilaiae MA-267, a fungus obtained from the rhizospheric soil of the mangrove plant Lumnitzera racemosa [77]. An X-ray crystallographic study determined the structure and configuration. The hypothetical biosynthetic pathway starting from cis-farnesyl pyrophosphate (FPP) was proposed (Scheme 10). Notably, Compounds 68 and 69 exhibited selective activity against the plant pathogenic fungus Colletotrichum gloeosporioides (MIC = 1.0 and 0.125 μg/mL, respectively).
The first total synthesis of 68 and 69 in their racemic forms was reported by Dong et al. in 2021 [78]. The approach featured a rhodium-catalyzed deconstructive formation of a tricyclic skeleton by C–C activation of cyclobutanone derivatives, generating (±)-68 and (±)-69 in 13 and 14 steps with 0.56% and 0.49% overall yields, respectively. In the same year, K Sugita described another more efficient synthetic pathway for the total practical synthesis of (±)-68 and (±)-69 from commercially available 3-ethoxycyclohex-2-en-1-one with 4.0% overall yields for both compounds [79].

3.2.2. Sesterterpenoids

Sesterterpenoids are a relatively small and rare group of terpenoids found in widespread sources. They always possess interesting carbon skeletons, including linear, monocyclic, polycyclic, and miscellaneous. In addition, they exhibit diverse biological activities such as antimicrobial, cytotoxicity, anti-inflammatory, and protein tyrosine phosphatase B inhibition.
The group of She has been dedicated to the search for structurally unique and biologically active compounds from the mangrove plant-derived fungal endophytes. Five sesterterpenoids of three kinds of carbon skeletons, asperterpenoid A (70) (Figure 19), asperterpenols A and B (71 and 72), and aspterpenacids A and B (73 and 74), have been isolated and characterized from Aspergillus sp. Among them, asperterpenoid A (70), possessing a new 5/7/(3)6/5 pentacyclic carbon skeleton, exhibited potent inhibitory activity against Mycobacterium tuberculosis protein tyrosine phosphatase B (mPTPB) with an IC50 value of 2.2 μM [80].
In addition, asperterpenol A (71) and asperterpenol B (72), two sesterterpenoids with an unusual 5/8/6/6 tetracyclic ring skeleton, showed inhibitory effects on acetylcholinesterase (AChE) with IC50 values of 2.3 μM and 3.0 μM, respectively [81]. Furthermore, the two unusual pentacarbocyclic sesterterpenoids, aspterpenacids A (73) and B (74), with an unusual carbocyclic skeleton containing an unprecedented 5/3/7/6/5 ring system, showed no antibacterial and cytotoxic activities [82]. The structures of 6973 were elucidated based on spectroscopic methods, and the absolute configurations of 7073 were determined by single-crystal X-ray diffraction analysis. She et al. proposed a hypothetical biosynthetic pathway for 7074 [80,81,82]. In brief, they are derived from geranylfarnesyl pyrophosphate (GFPP), followed by cyclization rearrangement and redox reactions (Scheme 11).

3.2.3. Meroterpenoids

Meroterpenoids are secondary metabolites with structures consisting of at least two parts: a terpenoid fragment (mainly mevalonate pathway) and a nonterpenoid fragment [83]. The different nonterpenoid moiety based on the biosynthetic pathway, various terpenoid (the length of the terpenoid chain and its cyclization mode), and the tailoring reactions make the chemical diversity of meroterpenoids.
Chermebilaene A (75) (Figure 20), an unprecedented acorane-type sesquiterpene hybridized with an octadecadienoic acid skeleton, together with an unusual orthoester meroterpenoid, chermebilaene B (76) were isolated from the co-culture extract of P. bilaiae MA-267 (from the rhizosphere of the mangrove Lumnitzera racemosa) and P. chermesinum EN-480 (from the fresh tissue of marine red algal Pterocladiella tenuis) [84]. Compound 75 showed potent inhibitory activities against Ceratobasidium cornigerum and Edwardsiella tarda, and may prove helpful as an antibiotic against aquatic or plant pathogens.
Simpterpenoid A (77) (Figure 21), an unconventional meroterpenoid containing a highly functionalized cyclohexadiene moiety with gem-propane-1,2-dione and methylformate groups, was isolated from the fungal strain Penicillium simplicissimum MA-332, obtained from the rhizospheric soil of the mangrove plant Bruguiera sexangular var. rhynchopetala [85]. The intricate polycyclic skeleton is unique in natural sources. Compound 77 exhibited potent inhibitory activity against influenza neuraminidase with an IC50 value of 8.1 nM.
Two new meroterpenoids, penicianstinoids A and B (78 and 79, Figure 21), were obtained from the mangrove-derived fungus Penicillium sp. TGM112 isolated from the mangrove Bruguiera sexangula var. rhynchopetala [86]. Compared with 79, compound 78 represents an austinoid-like meroterpenoid that is reported for the second time [87], in which a carbon−carbon double bond at C-1′−C-2′ was oxidized to a carbonyl group at C-1′−C-2′. Compounds 78 and 79 showed growth inhibition activity against newly hatched larvae of Helicoverpa armigera (Hubner) with IC50 values of 200 μg/mL. In addition, 78 and 79 displayed insecticidal activity against Caenorhabditis elegans with EC50 values of 9.4 (±1.0) and 9.9 (±0.0) μg/mL, respectively. Biogenetically, compounds 7679 are derived from the same intermediate S2, which is produced by the combination of a polyketide intermediate 3,5-dimethylorsellinicacid (DMOA) and the terpenoid precursor farnesyl pyrophosphate (FPP), following by a series of further modifications to generate a profile of meroterpenoids with diverse skeletons bearing polycyclic cores. DMOA-based meroterpenoids exhibit diverse structures due to the cyclization of the terpenoid moiety, divergence of post-cyclization modification reactions, and various tailoring reactions (Scheme 12) [83].
Two hybrid sesquiterpene-cyclopaldic acid metabolites with an unusual carbon scaffold, namely pestalotiopens A and B (80 and 81) (Figure 22), were obtained from the endophytic fungus Pestalotiopsis sp. (from the leaves of the Chinese mangrove Rhizophora mucronate), together with the known phytotoxin altiloxin B [88]. A plausible biosynthetic pathway of 80 and 81 is proposed (Scheme 13). The cyclopaldic acid and altiloxin B were deduced as precursors.
Indole terpenoids are structurally diverse meroterpenoids containing an indole ring from tryptophan and cyclic sesquiterpenes or diterpene backbone moiety [83].
Three indole sesquiterpenes, indotertine A (82) [89] and indotertine B (83a/83b) [90] (Figure 23) were discovered from actinomycete Streptomyces sp. CHQ-64 (derived from the rhizosphere soil of reeds). They possess an unusual skeleton with a condensed ring system made up of a tryptophan-derived indole moiety and a sesquiterpene unit, which represents a new subgroup of indole terpenoids combining amino acid and mevalonate pathways. Indotertine B (83a/83b) exists as a pair of rotamers about the N−C(O) bond with a 2:1 ratio, inseparable by HPLC because of the dynamic interconversion. The analysis of the NOESY spectrum implied that the formyl−N-1 amide bond was S-trans in 83a and S-cis in 83b. Compound 83 displays cytotoxic activities against HCT-8 and A549 tumor cell lines with IC50 values of 6.96 and 4.88 μM. Further chemical investigation of this fungal strain led to the isolation of drimentine I (84) [91], containing a rare heptacyclic skeleton formed via two bridging linkages. The pentacyclic product indotertine A (82) was hypothetically synthesized by iminium-olefin cyclization. In contrast, tetracyclic product drimentine F could take place from amidic nitrogen by nucleophilic addition to the α-position of the indole moiety (Scheme 14). However, cyclization of 84 happened on indol-NH to afford the linkage between C-14 and N-6 of drimentine F. Compound 84 was found to have weak activity against human cervical carcinoma cell line HeLa, with IC50 values of 16.73 μM.
Secopaxilline A (85) [92] (Figure 24) is the first example of indole diterpenoid derivatives possessing a carbon-nitrogen bond cleavage skeleton, which was isolated from metabolites of the aciduric fungus Penicillium camemberti OUCMDZ-1492 (separated from the soil and mud around the roots of Rhizophora apiculata). A plausible biosynthetic pathway for secopaxilline A (85) from paxilline was postulated, (Scheme 15), and the process has been conducted by chemical reactions with a 45% overall yield. Paxilline was derived from the common indole-diterpenoid precursor 3-geranylgeranylindole (GGI) derived from geranylgeranyl pyrophosphate (GGPP) and indole-3-glycerol phosphate [93] (Scheme 15).
The fungus Mucor irregularis, isolated from the fresh inner tissue of the mangrove Rhizophora stylosa, yields three unusual indole-diterpenes, rhizovarin A–C (8688, Figure 24) [94], which represent the most complex members of the reported indole-diterpenes. Even though the main structural elements resemble those of other reported indole diterpenes, the presence of an unusual acetal linked to a hemiketal (86) or a ketal (87 and 88) unit in an unprecedented 4,6,6,8,5,6,6,6,6-fused indole-diterpene ring system makes them chemically unique. Their structures and absolute configurations were elucidated by spectroscopic analysis, modified Mosher’s method, and chemical calculations. For rhizovarin A (86), the biosynthetic pathway may involve more oxidative steps than penitrem A, a known indole-diterpene derived from a paxilline and two isopentenyl-diphosphate units. (Scheme 15) The biosynthetic pathway has been elucidated by reconstitution of the biosynthetic genes in Aspergillus oryzae [95]. Another unusual indole-diterpene, containing a complex 6,8,6,6,6-fused ring system, rhizovarin D (89), was also obtained in this study. NOESY experiments determined the relative configuration for the stereogenic centers of 89. Each isolated compound was evaluated for antitumor activity against HL-60 and A-549 cell lines. Compounds 86 and 87 showed activities against the human A-549 and HL-60 cancer cell lines (IC50 < 10 μM).
Bioassay-guided fractionation of the bacterial strain Erythrobacter sp. SNB-035 (from mangrove sediments) led to the isolation of erythrazoles A and B (90 and 91) [96] (Figure 25). Structurally, 90 and 91 possess an abenzothiazole moiety, which is rare among NPs. Furthermore, 91 arises from four biosynthetic pathways: NRPS, terpene, shikimate, and polyketide. Although combinations of two of the four pathways are common among NPs, four biosynthetic pathways simultaneously involved are extremely rare (Scheme 16).

3.3. Alkaloids and Other Nitrogen-Containing Metabolites

3.3.1. Diketopiperazines

Diketopiperazines (DKPs) are an essential group of structurally diverse cyclic dipeptides with significant biological properties [97].
Effusin A (92) (Figure 26) is a spirobicyclic N,O-acetal derivative with an unprecedented 3′,3a′,5′,6′-tetrahydrospiro-[piperazine-2,2′-pyrano[2,3,4-de]chromene] ring system. Besides this, a spiro-polyketide-diketopiperazine hybrid dihydrocryptoechinulin D (93) were isolated from a mangrove rhizosphere soil-derived fungus Aspergillus effuses H1-1 [98]. Compounds 92 and 93 occurred as racemates. The enantiomers were separated and characterized by online HPLC-ECD analysis, and their absolute configurations were determined by the TDDFT ECD calculation approach. The spirobicyclic N,O-acetal moiety of 92 could be obtained through a domino ring-closure reaction between the substituted salicylaldehyde moiety in aspergin and the eneamide moiety of the diketopiperazine unit in neoechinulin B [98]. On the contrary, the spirobicycle of 93 is produced by an enzyme-catalyzed regiospecific [4 + 2] Diels Alder reaction (Scheme 17). The cytotoxic effects of 9293 were evaluated, 93 showed potent activity on P388 cells with an IC50 value of 1.83 μM. The target of racemic 93 was also evaluated, and the (12R,28S,31S)-93 enantiomer (93a) showed selectivity against topoisomerase I.
Using the OSMAC (one strain many compounds) approach, a metabolically rich strain of Penicillium brocae MA-231 (isolated from mangrove Avicennia marina) could produce two new diketopiperazines, spirobrocazines A–B (9495) (Figure 27), which had a 6/5/6/5/6 cyclic system with a rare spirocyclic center at C-2 [99]. In addition, a deep-sea sediment-derived fungus Eutypella sp. Also yields three new spirocyclic DKPs, eutypellazines N–P (9698) [100]. Compound 96 was determined as the C-2′ isomer of spirobrocazine A (91). Notably, 97 and 98 are the first compounds isolated from a wild-type fungus to contain a spirocyclic tetrahydrobenzothiophene motif. Furthermore, eight new dioxopiperazines 99106 (penispirozines A-H) were discovered from the mangrove-derived fungus Penicillium janthinellum HDN13-309 [101]. The structures of 99104 were similar to eutypellazines O–P (9788). They were distinguished by not only the existence of a spiro-thiophane or spiro-furan ring system but also the chirality of the pentacyclic moiety. Moreover, penispirozine A (99) had an unusual pyrazino[1,2]oxazadecaline coupled with a thiophane ring system, while penispirozine B (100) possessed a 6/5/6/5/6 pentacyclic ring system with two rare spirocyclic centers. Biosynthetically, the precursor to these structurally diverse penispirozines was considered to be the diketopiperazine cyclo-l-Phe-l-Phe (Scheme 18). In addition, compounds 101 and 102 increased the expression of the two relevant phase-II detoxifying enzymes, SOD2 and HO-1, at 10 μM.
A pair of unusual enantiomeric indole diketopiperazine alkaloid dimers, (±)-asperginulin A (107a/b) (Figure 28), with an unprecedented 6/5/4/5/6 pentacyclic skeleton, were discovered from the mangrove endophytic fungus Aspergillus sp. SK-28, guided by UPLC-HRMS [102]. Chiral-phase HPLC separated the enantiomeric dimers. Their structures, including the absolute configurations, were elucidated by spectroscopic analysis, X-ray diffraction, and quantum chemical calculation. (+)-Asperginulin A (107b) exhibited antifouling activity against the barnacle Balanus reticulatus. 107 was possibly derived, in vivo, from intermolecular [2 + 2] cycloaddition of its monomer precursor by nonenzymatic processes (Scheme 19).
A class of pyrazinopyrimidine-type alkaloids, namely pyrasplorines A–C (108110) (Figure 29) were discovered from the fungus Aspergillus versicolor HDN11-84 [103]. Pyrasplorine A (105) represents the first compound with spiro-cyclopentane in pyrazinopyrimidine-type alkaloids. The cyclopentane moiety is common in terpenes but rare in alkaloids and diketopiperazines, and it is only found in maremycins [104]. The structure is probably constructed by the condensation of anthranilic acid with diketopiperazine and followed by successive steps to yield the key intermediate S3. Then, compound 108 was derived from the S3 via a series of reactions [105] (Scheme 20).

3.3.2. Indole and Isoindole Alkaloids Derivatives

Various mangrove fungi produce indole and isoindole alkaloids with a plethora of biologically active. The indole-terpenes which also belong to meroterpenes have been described in Section 3.2.3.
Cytochalasan alkaloid usually consists of a 10-(indol-3-yl) group, a macrocyclic ring, and a perhydroisoindolone moiety. Chaetoglobosin is one class of cytochalasan alkaloid. The mangrove endophytic fungus Penicillium chrysogenum V11 afforded two unusual new Chaetoglobosins, penochalasin I and K (111 and 112) [106,107] (Figure 30), with an unprecedented six-cyclic 6/5/6/5/6/13 fused ring system formed by the connection of C-5 and C-2′ of the chaetoglobosin class. Additionally, the biomimetic semi-synthesis of 111 and 112 was successfully carried out from the corresponding co-occurrence analogue chaetoglobosin C and chaetoglobosin A, respectively [107]. Compound 112 displayed significant inhibitory activities against Colletotrichum gloeosporioides and Rhizoctonia solani (MICs = 6.13 μM, 12.26 μM, respectively), which was better than those of control carbendazim. It also exhibited potent cytotoxicity against MDA-MB-435, SGC-7901, and A549 cells (IC50 < 10 μM). In addition, compound 111 exhibited significant cytotoxicity against MDA-MB-435 and SGC-7901 cells (IC50 < 10 μM).
The typical paraherquamides (PHQs) are prenylated indole alkaloids with diverse ring systems. PHQs are derived from three building blocks: L-tryptophan, acyclic amino acid (either proline, β-methyl proline, or pipecolic acid), and one or two isoprenyl units. Interestingly, compounds 113115 (Figure 31) (mangrovamides A–C, isolated from the Penicillium sp. Separated from a mangrove sediment sample of the South China Sea) feature a bicyclo [2.2.2] diazaoctane core and contain the first documented examples of isoprene derived dimethyl γ-pyrone and γ-methyl proline, instead of the usual β-methyl proline in the PHQ family [108]. A plausible biosynthetic pathway starting from L-ornithine to account for the formation of the observed γ-methyl proline is outlined (Scheme 21). Moreover, the X-ray data determined the absolute configuration of all chiral centers in 113. In an activity assay, 115 showed a moderate acetylcholinesterase inhibitory effect with an IC50 value of 58.0 μM.
Diaporisoindoles A and B (116 and 117) [109], and D and E (118 and 119) [52] (Figure 32), isolated from the mangrove endophytic fungus Diaporthe sp. SYSU-HQ3 (from a fresh branch of the mangrove plant Excoecaria agallocha) and could be derived from tenellone B, are the first reported examples of isoprenylisoindole alkaloids with a rare 1,4-benzodioxan moiety. In addition, siaporisoindole A (116) showed significant inhibitory activity against Mycobacterium tuberculosis protein tyrosine phosphatase B with IC50 4.2 μM compared to 22.1 μM for the positive control (oleanolic acid,). Furthermore, 116 and 117 exhibited potent inhibitory activity against NO production in RAW 264.7 cells with IC50 values less than 10 μM. Then She et al. continued an extensive study of isolating an unusual diisoprenylisoindole dimer diaporisoindole C (120). It was presumed to be derived from compounds 116 or 117 via addition reaction, dehydration, and aromatization (Scheme 22).
Quinazoline containing indole alkaloids have pyrimidine [2, 1-b] quinazoline and imidazole [1, 2-a] indole groups linked by methylene (and, in some cases, further linked by additional helical Bridges). Two unusual quinazoline-containing indole alkaloids neosartoryadins A and B (121 and 122) (Figure 33) along with fiscalin C (a known compound to be related to biosynthesis) were identified from the mangrove endophytic fungus Neosartorya udagawae HDN13-313 [110]. Compounds 121 and 122 is a quinazoline-containing indole alkaloid featuring a unique 6/6/6/5 quinazoline ring directly linked to the 6/5/5 imidazolinone ring. 121 and 122 differs from conventional fumiquinazoline alkaloids such as fiscalin C by the unprecedented pyrido[2, 1-b]- quinazoline moiety, which binds to a pyridine (C ring) rather than a pyrimidine ring, in addition to the presence of a unique tetrahydrofuran ring (D ring). It is speculated that 121 and 122 are biosynthesized from L-tryptophan, anthranilic acid (ATA), L-valine, and 2-aminoisobutyric acid (Aib). The unprecedented C ring was formed by the key intermediate fiscalin C through further modification by oxidation, hydrolysis, water nucleophilic attack, dehydration, deprotonation, and subsequent aldol reaction (Scheme 23).
Streptocarbazoles, the staurosporine analogues with extraordinary cyclic N-glycosidic connections between 1,3-carbon atoms of the glycosyl moiety and two indole nitrogen atoms of the indolocarbazole core, have also been produced by mangrove actinomycetes.
Streptomyces sp. FMA, isolated from mangrove soil collected in Sanya, Hainan Province of China provided streptocarbazoles A (123) and B (124) [111] (Figure 34). Compound 123 was cytotoxic to HL60, A549, P338, and HeLa cells with IC50 values of 1.4, 5.0, 18.9, and 34.5 μM, respectively, while compound 124 was active against P388 and HeLa cells with IC50 values of 12.8 and 22.5 μM, respectively. In addition, it was demonstrated that streptocarbazoles A arrest the HeLa cells in the G2/M phase at 10 μM. A plausible biogenetic pathway of 123 and 124 was postulated (Scheme 24). The indolocarbazole unit (K252c) was derived from tryptophan, while the glycosyl moiety was probably developed from 2-deoxy-d-pyranoglucose. Subsequently, the first cloning and characterization of an indolocarbazole gene cluster isolated from Streptomyces sanyensis FMA were reported. Indolocarbazole biosynthesis was confirmed by gene inactivation and heterologous expression in Streptomyces coelicolor M1152 [112].

3.3.3. Pyridines

Piericidins feature a 4-pyridinol core linked with a variable methylated polyene side chain. The strain Streptomyces iakyrus SCSIO NS104, isolated from a mangrove sediment sample collected from the Pearl River estuary to the South China Sea, yielded four new piericidin analogues, iakyricidins A–D (125128) [113]. Iakyricidins B–D (126128) represent a new subgroup of piericidin with C-C cyclization and double bond rearrangements in the polyene side chain. In addition, oxidized side chain piericidin analogue iakyricidin A (125) displayed potent antiproliferative activity against human renal carcinoma cell lines ACHN cell with an IC50 value of 20 nM. Compound 125 might be derived by oxidative cleavage between C-13 and C-14 of the precursor. In the plausible biosynthetic pathways of 126128, the most crucial step would be the yet-to-be-identified enzymatic C8-C12 cyclization from the co-occurrence precursor (Scheme 25).
Chemical investigation of the endophytic fungus Campylocarpon sp. HDN13-307, obtained from the root of mangrove plant Sonneratia caseolaris led to the isolation of four new 4-hydroxy-2-pyridone alkaloids, namely campyridones A–D (129132) [114] (Figure 35), which existed as two pairs of diastereoisomers, featuring an additional C ring between the decalin and pyridone units, represented new ring systems for this family of alkaloids. A plausible biosynthetic pathway for 129132 is postulated with the co-occurrence ilicicolin H as a critical intermediate. Ilicicolin H is a typical 4-hydroxy-2-pyridone alkaloid which was considered to be biosynthesized via tetramic acids formed by hybridizing a polyketide unit to a tyrosine (Scheme 26). Compound 132 exhibited activity against Hela cells with IC50 values of 8.8 μM.

3.3.4. Others

A chemical investigation of the fermentation of Penicillium sp. GD6, associated with the Chinese mangrove Bruguiera gymnorrhiza, resulted in the isolation of an unusual pyrrolizidine alkaloid, penibruguieramine A (133) (Figure 36), characterized by an unprecedented 1-alkenyl-2-methyl-8-hydroxymethyl pyrrolizidin-3-one skeleton [115] (Scheme 27).
Talaramide A (134) (Figure 37) is the second example of an alkaloid with a unique oxidized tricyclic system resembling a bird cage, which was obtained from the mangrove endophytic fungus Talaromyces sp [116]. The first example was rubrobramide, obtained from the fungus Cladobotryum ubrobrunnescens [117]. 134 was a PKS-NRPS hybrid metabolite derived from acetyl acid, malonic acid, and L-leucine. A series of polymerizations, cyclizations, rearrangements, and redox reactions finally afforded the unique oxidized tricyclic skeleton of 134 (Scheme 28).

4. Conclusions

In this review, we presented the chemical constituents of the mangrove-associated ecosystem and showcased the diversity of the chemical structures, biological activities, chemical syntheses, and (proposed) biosynthetic pathways.
Structurally diverse secondary metabolites play a crucial role in the discovery campaigns for new NP drug pharmacophores. The mangrove ecosystem is producing various structurally novel compounds that could provide a potent compound library for the identification of lead compounds. Herein, we presented a comprehensive review of 134 mangrove-derived NPs with new carbon skeletons, unique ring systems, or uncommon structural moieties. The majority of them were produced by mangrove-associated microorganisms, and more than 70% were isolated from endophyte fungus, indicating remarkable chemical diversity and interesting bioactivity of the microbial community. The structural novelty and diversity of these metabolites result from the enormous variety of mangrove ecosystems in combination with their potential biosynthetic capabilities. In addition, they display diverse and remarkable biological activities and are frequently reported as antimicrobial and cytotoxic compounds (Table 1 and Table 2), which might attract researchers for further investigations toward chemical synthesis and biosynthesis. Mangrove ecosystems are a rewarding source for producing bioactive substances with novel carbon frameworks and discovering drug lead compounds, attracting pharmaceutical scientists for more in vivo and preclinical studies on these compounds.
In conclusion, through this review, we conveyed that (1) the mangrove-associated ecosystem is still an abundant source of bioactive NPs providing leads for drug development, (2) chemical syntheses of several of the mangrove-associated NPs are completed, but more NPs are to be synthesized and more efficient routes are to be developed, and (3) and the biosynthesis of most of the mangrove-associated NPs remain unclear.

Author Contributions

All authors equally contributed to the review and their specific contributions to the preparation of this review are as follows: investigation and writing—original draft preparation: M.-J.W.; visualization and writing—review and editing: B.X.; conceptualization, review and editing, funding acquisition: Y.-W.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 81991521) and the SKLDR/SIMM Project (No. SIMM2013ZZ-06).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, J.; Feng, Z.; Zhang, W.; Xu, J. Evaluation of the antimicrobial and cytotoxic potential of endophytic fungi extracts from mangrove plants Rhizophora stylosa and R. mucronata. Sci. Rep. 2022, 12, 2733–2744. [Google Scholar] [CrossRef] [PubMed]
  2. Ser, H.L.; Tan, L.T.; Law, J.W.; Chan, K.-G.; Duangjai, A.; Saokaew, S.; Pusparajah, P.; Mutalib, N.-S.; Khan, T.M.; Goh, B.-H.; et al. Focused Review: Cytotoxic and Antioxidant Potentials of Mangrove-Derived Streptomyces. Front. Microbiol. 2017, 8, 2065–2075. [Google Scholar] [CrossRef]
  3. Azman, A.-S.; Othman, I.; Velu, S.S.; Chan, K.-G.; Lee, L.-H. Mangrove rare actinobacteria: Taxonomy, natural compound, and discovery of bioactivity. Front. Microbiol. 2015, 6, 856–870. [Google Scholar] [CrossRef] [PubMed]
  4. Nicoletti, R.; Salvatore, M.M.; Andolfi, A. Secondary Metabolites of Mangrove-Associated Strains of Talaromyces. Mar. Drugs 2018, 16, 12–26. [Google Scholar] [CrossRef] [Green Version]
  5. Xu, J. Bioactive natural products derived from mangrove-associated microbes. RSC Adv. 2015, 5, 841–892. [Google Scholar] [CrossRef]
  6. Chen, S.; Cai, R.; Liu, Z.; Cui, H.; She, Z. Secondary metabolites from mangrove-associated fungi: Source, chemistry and bioactivities. Nat. Prod. Rep. 2022, 39, 560–595. [Google Scholar] [CrossRef]
  7. Ancheeva, E.; Daletos, G.; Proksch, P. Lead Compounds from Mangrove-Associated Microorganisms. Mar. Drugs 2018, 16, 319–349. [Google Scholar] [CrossRef] [Green Version]
  8. Zeng, W.; Huang, G.; Wang, B.; Cai, J.; Zheng, C. Secondary Metabolites and Bioactivities of Penicillium sp. Sourced from Mangrove from 2007 to 2020. Chin. J. Org. Chem. 2021, 41, 4255–4278. [Google Scholar] [CrossRef]
  9. Xu, D.B.; Ye, W.-W.; Han, Y.; Deng, Z.-X.; Hong, K. Natural products from mangrove actinomycetes. Mar. Drugs 2014, 12, 2590–2613. [Google Scholar] [CrossRef] [Green Version]
  10. Deshmukh, S.K.; Gupta, M.K.; Prakash, V.; Reddy, M.S. Mangrove-Associated Fungi: A Novel Source of Potential Anticancer Compounds. J. Fungi 2018, 4, 101–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Wang, X.; Mao, Z.-G.; Song, B.-B.; Chen, C.-H.; Xiao, W.-W.; Hu, B.; Wang, J.-W.; Jiang, X.-B.; Zhu, Y.-H.; Wang, H.-J. Advances in the study of the structures and bioactivities of metabolites isolated from mangrove-derived fungi in the South China Sea. Mar. Drugs 2013, 11, 3601–3616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Manohar, S.M. A Review of the Botany, Phytochemistry and Pharmacology of Mangrove Lumnitzera racemosa Willd. Pharmacogn. Rev. 2021, 15, 107–116. [Google Scholar] [CrossRef]
  13. Wang, K.-W.; Wang, S.-W.; Wu, B.; Wei, J.-G. Bioactive Natural Compounds from the Mangrove Endophytic Fungi. Mini. Rev. Med. Chem. 2014, 14, 370–391. [Google Scholar] [CrossRef] [PubMed]
  14. Cadamuro, R.D.; Bastos, I.M.A.S.; Silva, I.T.; Cruz, A.C.C.; Robl, D.; Sandjo, L.P.; Alves, S.; Lorenzo, J.M.; Rodríguez-Lázaro, D.; Treichel, H.; et al. Bioactive Compounds from Mangrove Endophytic Fungus and Their Uses for Microorganism Control. J. Fungi 2021, 7, 455–473. [Google Scholar] [CrossRef]
  15. Li, K.; Chen, S.; Pang, X.; Cai, J.; Zhang, X.; Liu, Y.; Zhu, Y.; Zhou, X. Natural products from mangrove sediments-derived microbes: Structural diversity, bioactivities, biosynthesis, and total synthesis. Eur. J. Med. Chem. 2022, 230, 114117–114157. [Google Scholar] [CrossRef]
  16. Wu, J.; Xiao, Q.; Xu, J.; Li, M.-Y.; Pan, J.-Y.; Yang, M.-H. Natural products from true mangrove flora: Source, chemistry and bioactivities. Nat. Prod. Rep. 2008, 25, 955–981. [Google Scholar] [CrossRef]
  17. Li, M.Y.; Xiao, Q.; Pan, J.-Y.; Wu, J. Natural products from semi-mangrove flora: Source, chemistry and bioactivities. Nat. Prod. Rep. 2009, 26, 281–298. [Google Scholar] [CrossRef]
  18. Fang, X.; Di, Y.T.; Hao, X.J. The Advances in the Limonoid Chemistry of the Meliaceae Family. Curr. Org. Chem. 2011, 15, 1363–1391. [Google Scholar] [CrossRef]
  19. Luo, J.; Sun, Y.; Kong, L. Research progress of meliaceous limonoids from 2011 to 2021. Nat. Prod. Rep. 2022, 39, 1325–1365. [Google Scholar] [CrossRef]
  20. Tan, Q.-G.; Luo, X.-D. Meliaceous limonoids: Chemistry and biological activities. Chem. Rev. 2011, 111, 7437–7522. [Google Scholar] [CrossRef]
  21. Li, J.; Li, M.-Y.; Bruhn, T.; Katele, F.X.; Xiao, Q.; Pedpradab, P.; Wu, J.; Bringmann, G. Thaixylomolins A-C: Limonoids Featuring Two New Motifs from the Thai Xylocarpus Moluccensis. Org. Lett. 2013, 15, 3682–3685. [Google Scholar] [CrossRef]
  22. Dai, Y.-G.; Li, W.-S.; Pedpradab, P.; Liu, J.-J.; Wu, J.; Shen, L. Thaixylomolins O–R: Four new limonoids from the Trang mangrove, Xylocarpus moluccensis. RSC Adv. 2016, 6, 85978–85984. [Google Scholar] [CrossRef]
  23. Wu, Y.-B.; Qing, X.; Huo, C.-H.; Yan, H.-M.; Shi, Q.-W.; Sauriol, F.; Gu, Y.-C.; Kiyota, H. Xylomexicanins E-H, new limonoids from Xylocarpus granatum. Tetrahedron 2014, 70, 4557–4562. [Google Scholar] [CrossRef]
  24. Pan, J.-Y.; Chen, S.-L.; Li, M.-Y.; Li, J.; Yang, M.-H.; Wu, J. Limonoids from the Seeds of a Hainan Mangrove, Xylocarpus Granatum. J. Nat. Prod. 2010, 73, 1672–1679. [Google Scholar] [CrossRef] [PubMed]
  25. Wu, J.; Zhang, S.; Bruhn, T.; Xiao, Q.; Ding, H.; Bringmann, G. Xylogranatins F-R: Antifeedants from the Chinese mangrove, Xylocarpus granatum, a new biogenetic pathway to tetranortriterpenoids. Chem. Eur. J. 2008, 14, 1129–1144. [Google Scholar] [CrossRef]
  26. Sheng, Y.; Yin, S.; Wang, X.-N.; Lin, L.-P.; Ding, J.; Yue, J.-M. Xylogranatins A−D: Novel Tetranortriterpenoids with an Unusual 9,10-seco Scaffold from Marine Mangrove Xylocarpus granatum. Org. Lett. 2006, 8, 4935–4938. [Google Scholar] [CrossRef]
  27. Li, W.-S.; Shen, L.; Bruhn, T.; Pedpradab, P.; Wu, J.; Bringmann, G. Trangmolins A-F with an Unprecedented Structural Plasticity of the Rings A and B: New Insight into Limonoid Biosynthesis. Chem. Eur. J. 2016, 22, 11719–11727. [Google Scholar] [CrossRef]
  28. He, C.L.; Li, W.-S.; Wu, J.; Shen, L. Krishnolides E-K: New limonoids from the Krishna mangrove Xylocarpus moluccensis. Fitoterapia 2021, 150, 104835–104843. [Google Scholar] [CrossRef]
  29. Li, J.; Li, M.-Y.; Bruhn, T.; Gotz, D.C.G.; Xiao, Q.; Satyanandamurty, T.; Wu, J.; Bringmann, G. Andhraxylocarpins A-E: Structurally intriguing limonoids from the true mangroves Xylocarpus granatum and Xylocarpus moluccensis. Chem. Eur. J. 2012, 18, 14342–14351. [Google Scholar] [CrossRef]
  30. Li, W.-S.; Wu, J.; Li, J.; Satyanandamurty, T.; Shen, L.; Bringmann, G. Krishnadimer A, an Axially Chiral Non-biaryl Natural Product: Discovery and Biomimetic Synthesis. Org. Lett. 2017, 19, 182–185. [Google Scholar] [CrossRef]
  31. Li, W.-S.; Yang, Y.; Liu, J.-J.; Shen, L.; Shi, Z.; Wu, J. Scaffold diversity-oriented synthesis of limonoid dimers: Discovery of an axially chiral agent within vivoanti-breast cancer activity. Org. Chem. Front. 2018, 5, 1079–1091. [Google Scholar] [CrossRef]
  32. Zhou, Z.-F.; Liu, H.-L.; Zhang, W.; Kurtán, T.; Mándi, A.; Bényei, A.; Li, J.; Taglialatela-Scafati, O.; Guo, Y.-W. Bioactive rearranged limonoids from the Chinese mangrove Xylocarpus granatum Koenig. Tetrahedron 2014, 70, 6444–6449. [Google Scholar] [CrossRef]
  33. Wu, Y.-B.; Wang, Y.-Z.; Ni, Z.-Y.; Qing, X.; Shi, Q.-W.; Sauriol, F.; Vavricka, C.J.; Gu, Y.-C.; Kiyota, H. Xylomexicanins I and J: Limonoids with Unusual B/C Rings from Xylocarpus granatum. J. Nat. Prod. 2017, 80, 2547–2550. [Google Scholar] [CrossRef]
  34. Li, W.-S.; Mandi, A.; Liu, J.-J.; Shen, L.; Kurtan, T.; Wu, J. Xylomolones A-D from the Thai Mangrove Xylocarpus moluccensis: Assignment of Absolute Stereostructures and Unveiling a Convergent Strategy for Limonoid Biosynthesis. J. Org. Chem. 2019, 84, 2596–2606. [Google Scholar] [CrossRef] [PubMed]
  35. Ponnapalli, M.G.; Ankireddy, M.; Annam, S.C.V.A.R.; Ravirala, S.; Sukki, S.; Tuniki, V.R. Unusual ent-isopimarane-type diterpenoids from the wood of Excoecaria agallocha. Tetrahedron Lett. 2013, 54, 2942–2945. [Google Scholar] [CrossRef]
  36. Wang, H.; Li, M.-Y.; Katele, F.Z.; Satyanandamurty, T.; Wu, J.; Bringmann, G. Decandrinin, an unprecedented C9-spiro-fused 7,8-seco-ent-abietane from the Godavari mangrove Ceriops decandra. Beilstein J. Org. Chem. 2014, 10, 276–281. [Google Scholar] [CrossRef] [Green Version]
  37. Chooi, Y.-H.; Tang, Y. Navigating the fungal polyketide chemical space: From genes to molecules. J. Org. Chem. 2012, 77, 9933–9953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Vederas, J.C. Explorations of fungal biosynthesis of reduced polyketides—A personal viewpoint. Nat. Prod. Rep. 2014, 31, 1253–1259. [Google Scholar] [CrossRef]
  39. Hang, L.; Liu, N.; Tang, Y. Coordinated and Iterative Enzyme Catalysis in Fungal Polyketide Biosynthesis. ACS Catal. 2016, 6, 5935–5945. [Google Scholar] [CrossRef] [Green Version]
  40. Zhang, X.; Guo, J.; Cheng, F.; Li, S. Cytochrome P450 enzymes in fungal natural product biosynthesis. Nat. Prod. Rep. 2021, 38, 1072–1099. [Google Scholar] [CrossRef]
  41. Little, R.F.; Hertweck, C. Chain release mechanisms in polyketide and non-ribosomal peptide biosynthesis. Nat. Prod. Rep. 2022, 39, 163–205. [Google Scholar] [CrossRef] [PubMed]
  42. Goncalves, G.A.; Spillere, A.R.; Neves, G.M.; Kagami, L.P.; Poser, G.L.; Canto, R.F.S.; Eifler-Lima, V.L. Natural and synthetic coumarins as antileishmanial agents: A review. Eur. J. Med. Chem. 2020, 203, 112514–112533. [Google Scholar] [CrossRef] [PubMed]
  43. Shabir, G.; Saeed, A.; El-Seedi, H.R. Natural isocoumarins: Structural styles and biological activities, the revelations carry on. Phytochemistry 2021, 181, 112568–112590. [Google Scholar] [CrossRef]
  44. Cai, R.; Wu, Y.; Chen, S.; Cui, H.; Liu, Z.; Li, C.; She, Z. Peniisocoumarins A-J: Isocoumarins from Penicillium commune QQF-3, an Endophytic Fungus of the Mangrove Plant Kandelia candel. J. Nat. Prod. 2018, 81, 1376–1383. [Google Scholar] [CrossRef]
  45. Darsih, C.; Prachyawarakorn, V.; Wiyakrutta, S.; Mahidol, C.; Ruchirawat, S.; Kittakoop, P. Cytotoxic metabolites from the endophytic fungus Penicillium chermesinum: Discovery of a cysteine-targeted Michael acceptor as a pharmacophore for fragment-based drug discovery, bioconjugation and click reactions. RSC Adv. 2015, 5, 70595–70603. [Google Scholar] [CrossRef]
  46. Huang, J.; She, J.; Yang, X.; Liu, J.; Zhou, X.; Yang, B. A New Macrodiolide and Two New Polycyclic Chromones from the Fungus Penicillium sp. SCSIO041218. Molecules 2019, 24, 1686–1692. [Google Scholar] [CrossRef] [Green Version]
  47. Gao, J.-M.; Yang, S.-X.; Qin, J.-C. Azaphilones: Chemistry and biology. Chem. Rev. 2013, 113, 4755–4811. [Google Scholar] [CrossRef] [PubMed]
  48. Guo, W.; Li, D.; Peng, J.; Zhu, T.; Gu, Q.; Li, D. Penicitols A-C and penixanacid A from the mangrove-derived Penicillium chrysogenum HDN11-24. J. Nat. Prod. 2015, 78, 306–310. [Google Scholar] [CrossRef]
  49. Blanc, P.J.; Laussac, J.P.; Bars, J.L.; Bars, P.L.; Loret, M.O.; Pareilleux, A.; Prome, D.; Prome, J.C.; Santerre, A.L.; Goma, G. Characterization of monascidin A from Monascus as citrinin. Int. J. Food Microbiol. 1995, 27, 201–213. [Google Scholar] [CrossRef]
  50. Ohashi, T.; Hosokawa, S. Total Syntheses of Stoloniferol B and Penicitol A, and Structural Revision of Fusaraisochromanone. Org. Lett. 2018, 20, 3021–3024. [Google Scholar] [CrossRef]
  51. Wu, S.-B.; Long, C.; Kennelly, E.J. Structural diversity and bioactivities of natural benzophenones. Nat. Prod. Rep. 2014, 31, 1158–1174. [Google Scholar] [CrossRef] [PubMed]
  52. Cui, H.; Liu, Y.; Li, J.; Huang, X.; Yan, T.; Cao, W.; Liu, H.; Long, Y.; She, Z. Diaporindenes A-D: Four Unusual 2,3-Dihydro-1 H-indene Analogues with Anti-inflammatory Activities from the Mangrove Endophytic Fungus Diaporthe sp. SYSU-HQ3. J. Org. Chem. 2018, 83, 11804–11813. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, C.; Ondeyka, J.G.; Herath, K.B.; Guan, Z.; Collado, J.; Platas, G.; Pelaez, F.; Leavitt, P.S.; Gurnett, A.; Nare, B.; et al. Tenellones A and B from a Diaporthe sp.: Two Highly Substituted Benzophenone Inhibitors of Parasite cGMP-Dependent Protein Kinase Activity. J. Nat. Prod. 2005, 68, 611–613. [Google Scholar] [CrossRef] [PubMed]
  54. Dinos, G.P. The macrolide antibiotic renaissance. Br. J. Pharmacol. 2017, 174, 2967–2983. [Google Scholar] [CrossRef]
  55. Meng, L.-H.; Li, X.-M.; Lv, C.-T.; Li, C.-S.; Xu, G.-M.; Huang, C.-G.; Wang, B.-G. Sulfur-containing cytotoxic curvularin macrolides from Penicillium sumatrense MA-92, a fungus obtained from the rhizosphere of the mangrove Lumnitzera racemosa. J. Nat. Prod. 2013, 76, 2145–2149. [Google Scholar] [CrossRef]
  56. Ding, L.; Maier, A.; Fiebig, H.-H.; Gorls, H.; Lin, W.-H.; Peschel, G.; Hertweck, G. Divergolides A-D from a mangrove endophyte reveal an unparalleled plasticity in ansa-macrolide biosynthesis. Angew. Chem. Int. Ed. 2011, 50, 1630–1634. [Google Scholar] [CrossRef]
  57. Wiley, P.F.; Kelly, R.B.; Caron, E.L.; Wiley, V.H.; Johnson, J.H.; MacKellar, F.A.; Mizsak, S.A. Structure of Nogalamycin. J. Am. Chem. Soc. 1977, 99, 542–549. [Google Scholar] [CrossRef]
  58. Pontius, A.; Krick, A.; Kehraus, S.; Brun, R.; König, G.M. Antiprotozoal Activities of Heterocyclic-Substituted Xanthones from the Marine-Derived Fungus Chaetomium sp. J. Nat. Prod. 2008, 71, 1579–1584. [Google Scholar] [CrossRef]
  59. Xu, Z.; Baunach, M.; Ding, L.; Peng, H.; Franke, J.; Hertweck, C. Biosynthetic code for divergolide assembly in a bacterial mangrove endophyte. ChemBioChem 2014, 15, 1274–1279. [Google Scholar] [CrossRef]
  60. Li, S.-R.; Zhao, G.-S.; Sun, M.-W.; He, H.-G.; Wang, H.-X.; Li, Y.-Y.; Lu, C.-H.; Shen, Y.-M. Identification and characterization of the biosynthetic gene cluster of divergolides from Streptomyces sp. W112. Gene 2014, 544, 93–99. [Google Scholar] [CrossRef]
  61. Zhao, G.; Li, S.; Guo, Z.; Sun, M.; Lu, C. Overexpression of div8 increases the production and diversity of divergolides in Streptomyces sp. W112. RSC Adv. 2015, 5, 98209–98214. [Google Scholar] [CrossRef]
  62. Zhao, G.; Wu, J.; Dai, W.-M. Toward a Total Synthesis of Divergolide, A.; Synthesis of the Amido Hydro-quinone Core and the C10–C15 Fragment. Synlett 2012, 23, 2845–2849. [Google Scholar] [CrossRef]
  63. Rasapalli, S.; Jarugumilli, G.; Yarrapothu, G.R.; Golen, J.A.; Rheingold, A.L. Studies toward Total Synthesis of Divergolides C and D. Org. Lett. 2013, 15, 1736–1739. [Google Scholar] [CrossRef] [PubMed]
  64. Nawrat, C.C.; Kitson, R.R.; Moody, C.J. Toward the total synthesis of hygrocin B and divergolide C: Construction of the naphthoquinone-azepinone core. Org. Lett. 2014, 16, 1896–1899. [Google Scholar] [CrossRef]
  65. Trauner, D.; Hager, A.; Kuttruff, C.; Hager, D.; Terwilliger, D. Toward the Total Synthesis of Divergolides C and D. Synlett 2013, 24, 1915–1920. [Google Scholar] [CrossRef]
  66. Rasapalli, S.; Jarugumilli, G.; Yarrapothu, G.R.; Golen, J.A.; Rheingold, A.L. Synthesis of the naphthoquinone core of divergolides (C–D) and model studies for elaboration of the ansabridge. Tetrahedron Lett. 2013, 54, 2615–2618. [Google Scholar] [CrossRef]
  67. Tian, E.-L.; Gu, B.-B.; Han, Y.; Qu, X.-D.; Lin, H.-W.; Deng, Z.-X.; Hong, K. Hainanmycin A, a cyclo-heptadeca macrolide bearing a cyclopentenone moiety from the mangrove-derived Streptomyces sp. 219807. Tetrahedron Lett. 2017, 58, 4348–4351. [Google Scholar] [CrossRef]
  68. Zhou, T.; Komaki, H.; Ichikawa, N.; Hosoyama, A.; Sato, S.; Igarashi, Y. Biosynthesis of akaeolide and lorneic acids and annotation of type I polyketide synthase gene clusters in the genome of Streptomyces sp. NPS554. Mar. Drugs 2015, 13, 581–596. [Google Scholar] [CrossRef] [Green Version]
  69. Choi, B.-K.; Park, S.-Y.; Choi, D.-K.; Shin, B.; Shin, Y.-H.; Oh, D.-C.; Lee, H.-S.; Lee, H.-S.; Lee, Y.-J.; Lee, J.S.; et al. Streptoglycerides A-D with a Rare 6/5/5 Tricyclic Ring Skeleton from a Marine Actinomycete Streptomyces species. Org. Lett. 2018, 20, 6037–6040. [Google Scholar] [CrossRef]
  70. Shin, H.J.; Heo, C.-S.; Anh, C.V.; Yoon, Y.D.; Kang, J.S. Streptoglycerides E-H, Unsaturated Polyketides from the Marine-Derived Bacterium Streptomyces specialis and Their Anti-Inflammatory Activity. Mar. Drugs 2022, 20, 44–55. [Google Scholar] [CrossRef]
  71. Choi, B.-K.; Cho, D.-Y.; Choi, D.-K.; Shin, H.J. Miharadienes A–D with unique cyclic skeletons from a marine-derived Streptomyces miharaensis. Org. Chem. Front. 2021, 8, 4845–4852. [Google Scholar] [CrossRef]
  72. Zhao, W.-Y.; Yi, J.; Chang, Y.-B.; Sun, C.-P.; Ma, X.-C. Recent studies on terpenoids in Aspergillus fungi: Chemical diversity, biosynthesis, and bioactivity. Phytochemistry 2022, 193, 113011–113035. [Google Scholar] [CrossRef]
  73. Jiang, M.; Wu, Z.; Guo, H.; Liu, L.; Chen, S. A Review of Terpenes from Marine-Derived Fungi: 2015–2019. Mar. Drugs 2020, 18, 321–368. [Google Scholar] [CrossRef] [PubMed]
  74. Wibowo, M.; Prachyawarakorn, V.; Aree, T.; Wiyakrutta, S.; Mahidol, C.; Ruchirawat, S.; Kittakoop, P. Tricyclic and Spirobicyclic Norsesquiterpenes from the Endophytic Fungus Pseudolagarobasidium acaciicola. Eur. J. Org. Chem. 2014, 19, 3976–3980. [Google Scholar] [CrossRef]
  75. Wibowo, M.; Prachyawarakorn, V.; Aree, T.; Mahidol, C.; Ruchirawat, S.; Kittakoopet, P. Cytotoxic sesquiterpenes from the endophytic fungus Pseudolagarobasidium acaciicola. Phytochemistry 2016, 122, 126–138. [Google Scholar] [CrossRef]
  76. Liu, D.-Z.; Dong, Z.-J.; Wang, F.; Liu, J.-K. Two novel norsesquiterpene peroxides from basidiomycete Steccherinum ochraceum. Tetrahedron Lett. 2010, 51, 3152–3153. [Google Scholar] [CrossRef]
  77. Meng, L.-H.; Li, X.-M.; Liu, Y.; Wang, B.-G. Penicibilaenes A and B, sesquiterpenes with a tricyclo[6.3.1.01,5]dodecane skeleton from the marine isolate of Penicillium bilaiae MA-267. Org. Lett. 2014, 16, 6052–6255. [Google Scholar] [CrossRef]
  78. Xue, Y.; Dong, G. Total Synthesis of Penicibilaenes via C-C Activation-Enabled Skeleton Deconstruction and Desaturation Relay-Mediated C-H Functionalization. J. Am. Chem. Soc. 2021, 143, 8272–8277. [Google Scholar] [CrossRef]
  79. Matsuo, R.; Watanabe, A.; Kamo, S.; Matsuzawa, A.; Sugita, K. Total syntheses of (±)-penicibilaenes A and B via intramolecular aldol condensation. Org. Chem. Front. 2021, 8, 6063–6066. [Google Scholar] [CrossRef]
  80. Huang, X.; Huang, H.; Li, H.; Sun, X.; Huang, H.; Lu, Y.; Lin, Y.; Long, Y.; She, Z. Asperterpenoid A, a New Sesterterpenoid as an Inhibitor of Mycobacterium tuberculosis Protein Tyrosine Phosphatase B from the Culture of Aspergillus sp. 16-5c. Org. Lett. 2013, 15, 721–723. [Google Scholar] [CrossRef]
  81. Xiao, Z.; Huang, H.; Shao, C.; Xia, X.; Ma, L.; Huang, X.; Lu, Y.; Lin, Y.; Long, Y.; She, Z. Asperterpenols A and B, New Sesterterpenoids Isolated from a Mangrove Endophytic Fungus Aspergillus sp. 085242. Org. Lett. 2013, 15, 2522–2525. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, Z.; Chen, Y.; Chen, S.; Liu, Y.; Lu, Y.; Chen, D.; Lin, Y.; Huang, X.; She, Z. Aspterpenacids A and B, Two Sesterterpenoids from a Mangrove Endophytic Fungus Aspergillus terreus H010. Org. Lett. 2016, 18, 1406–1409. [Google Scholar] [CrossRef] [PubMed]
  83. Jiang, M.; Wu, Z.; Liu, L.; Chen, S. The chemistry and biology of fungal meroterpenoids (2009–2019). Org. Biomol. Chem. 2021, 19, 1644–1704. [Google Scholar] [CrossRef]
  84. Meng, L.-H.; Li, X.-M.; Li, H.-L.; Wang, B.-G. Chermebilaenes A and B, New Bioactive Meroterpenoids from Co-Cultures of Marine-Derived Isolates of Penicillium bilaiae MA-267 and Penicillium chermesinum EN-480. Mar. Drugs 2020, 18, 339–348. [Google Scholar] [CrossRef] [PubMed]
  85. Li, H.-L.; Xu, R.; Li, X.-M.; Yang, S.-Q.; Meng, L.-H.; Wang, B.-G. Simpterpenoid A, a Meroterpenoid with a Highly Functionalized Cyclohexadiene Moiety Featuring gem-Propane-1,2-dione and Methylformate Groups, from the Mangrove-Derived Penicillium simplicissimum MA-332. Org. Lett. 2018, 20, 1465–1468. [Google Scholar] [CrossRef]
  86. Bai, M.; Zheng, C.-J.; Huang, G.-L.; Mei, R.-Q.; Wang, B.; Luo, Y.-P.; Zheng, C.; Niu, Z.-G.; Chen, G.-Y. Bioactive Meroterpenoids and Isocoumarins from the Mangrove-Derived Fungus Penicillium sp. TGM112. J. Nat. Prod. 2019, 82, 1155–1164. [Google Scholar] [CrossRef] [PubMed]
  87. Horikoshi, A.; Tsuchida, M.; Tsujiuchi, T.; Oyama, K.; Mitomi, M. Substance, Its Production Method, Production Strain, and Agricultural and Horticultural Insecticide Containing the Same as an Active Ingredient. JP Patent JP-2010018586-A, 10 July 2008. [Google Scholar]
  88. Hemberger, Y.; Xu, J.; Wray, V.; Proksch, P.; Wu, J.; Bringmann, G. Pestalotiopens A and B: Stereochemically Challenging Flexible Sesquiterpene-Cyclopaldic Acid Hybrids from Pestalotiopsis sp. Chem. Eur. J. 2013, 19, 15556–15564. [Google Scholar] [CrossRef] [PubMed]
  89. Che, Q.; Zhu, T.; Qi, X.; Mándi, A.; Kurtán, T.; Mo, X.; Li, J.; Gu, Q.; Li, D. Hybrid Isoprenoids from a Reeds Rhizosphere Soil Derived Actinomycete Streptomyces sp. CHQ-64. Org. Lett. 2012, 14, 3438–3441. [Google Scholar] [CrossRef]
  90. Che, Q.; Zhu, T.; Keyzers, R.A.; Liu, X.; Li, J.; Gu, Q.; Li, D. Polycyclic hybrid isoprenoids from a reed rhizosphere soil derived Streptomyces sp. CHQ-64. J. Nat. Prod. 2013, 76, 759–763. [Google Scholar] [CrossRef]
  91. Che, Q.; Li, J.; Li, D.; Gu, Q.; Zhu, T. Structure and absolute configuration of drimentine I, an alkaloid from Streptomyces sp. CHQ-64. J. Antibiot. 2016, 69, 467–469. [Google Scholar] [CrossRef] [PubMed]
  92. Fan, Y.; Wang, Y.; Fu, P.; Chairoungdua, A.; Piyachaturawat, P.; Zhu, W. Secopaxilline A, an indole-diterpenoid derivative from an aciduric Penicillium fungus, its identification and semisynthesis. Org. Chem. Front. 2018, 5, 2835–2839. [Google Scholar] [CrossRef]
  93. Byrne, K.M.; Smith, S.K.; Ondeyka, J.G. Biosynthesis of Nodulisporic Acid A: Precursor Studies. J. Am. Chem. Soc. 2002, 124, 7055–7060. [Google Scholar] [CrossRef]
  94. Gao, S.-S.; Li, X.-M.; Williams, K.; Proksch, P.; Ji, N.-Y.; Wang, B.-G. Rhizovarins A-F, Indole-Diterpenes from the Mangrove-Derived Endophytic Fungus Mucor irregularis QEN-189. J. Nat. Prod. 2016, 79, 2066–2074. [Google Scholar] [CrossRef] [PubMed]
  95. Jesus, A.D.; Gorst-Allman, C.P.; Steyn, P.S.; Heerden, F.V.; Vleggaar, R.; Wessels, P.L.; Hull, W.E. Tremorgenic Mycotoxins from Penicillium crustosum. Biosynthesis of Penitrem A. J. Chem. Soc. 1983, 1863. [Google Scholar] [CrossRef]
  96. Hu, Y.; MacMillan, J.B. Erythrazoles A-B, Cytotoxic Benzothiazoles from a Marine-Derived Erythrobacter sp. Org. Lett. 2011, 13, 6580–6583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Huang, R.M.; Yi, X.-X.; Zhou, Y.; Su, X.; Peng, Y.; Gao, C.-H. An update on 2,5-diketopiperazines from marine organisms. Mar. Drugs 2014, 12, 6213–6235. [Google Scholar] [CrossRef]
  98. Gao, H.; Liu, W.; Zhu, T.; Mo, X.; Mándi, A.; Kurtán, T.; Li, J.; Ai, J.; Gu, Q.; Li, D. Diketopiperazine alkaloids from a mangrove rhizosphere soil derived fungus Aspergillus effuses H1-1. Org. Biomol. Chem. 2012, 10, 9501–9506. [Google Scholar] [CrossRef] [Green Version]
  99. Meng, L.-H.; Wang, C.Y.; Mándi, A.; Li, X.-M.; Hu, X.-Y.; Kassack, M.U.; Kurtán, T.; Wang, B.-G. Three Diketopiperazine Alkaloids with Spirocyclic Skeletons and One Bisthiodiketopiperazine Derivative from the Mangrove-Derived Endophytic Fungus Penicillium brocae MA-231. Org. Lett. 2016, 18, 5304–5307. [Google Scholar] [CrossRef]
  100. Niu, S.; Liu, D.; Shao, Z.; Proksch, P.; Lin, W. Eutypellazines N−S, new thiodiketopiperazines from a deep sea sediment derived fungus Eutypella sp. with anti-VRE activities. Tetrahedron Lett. 2017, 58, 3695–3699. [Google Scholar] [CrossRef]
  101. Zhu, M.; Yang, Z.; Wang, H.; Gan, Q.; Zhang, G.; Che, Q.; Zhu, T.; Gu, Q.; Han, B.; Li, D. Penispirozines A-H, Three Classes of Dioxopiperazine Alkaloids with Spirocyclic Skeletons Isolated from the Mangrove-Derived Penicillium janthinellum. J. Nat. Prod. 2020, 83, 2647–2654. [Google Scholar] [CrossRef] [PubMed]
  102. Cai, R.; Jiang, H.; Xiao, Z.; Cao, W.; Yan, T.; Liu, Z.; Lin, S.; Long, Y.; She, Z. (−)- and (+)-Asperginulin A, a Pair of Indole Diketopiperazine Alkaloid Dimers with a 6/5/4/5/6 Pentacyclic Skeleton from the Mangrove Endophytic Fungus Aspergillus sp. SK-28. Org. Lett. 2019, 21, 9633–9636. [Google Scholar] [CrossRef] [PubMed]
  103. Li, F.; Sun, C.; Che, Q.; Zhu, T.; Gu, Q.; Guan, H.; Zhang, G.; Li, D. Pyrazinopyrimidine alkaloids from a mangrove-derived fungus Aspergillus versicolor HDN11-84. Phytochemistry 2021, 188, 112817–112823. [Google Scholar] [CrossRef]
  104. Duan, Y.; Liu, Y.; Huang, T.; Zou, Y.; Huang, T.; Hu, K.; Deng, Z.; Lin, S. Divergent biosynthesis of indole alkaloids FR900452 and spiro-maremycins. Org. Biomol. Chem. 2018, 16, 5446–5451. [Google Scholar] [CrossRef]
  105. Luo, X.; Chen, C.; Tao, H.; Lin, X.; Yang, B.; Zhou, X.; Liu, Y. Structurally diverse diketopiperazine alkaloids from the marine-derived fungus Aspergillus versicolor SCSIO 41016. Org. Chem. Front. 2019, 6, 736–740. [Google Scholar] [CrossRef]
  106. Huang, S.; Chen, H.; Li, W.; Zhu, X.; Ding, W.; Li, C. Bioactive Chaetoglobosins from the Mangrove Endophytic Fungus Penicillium chrysogenum. Mar. Drugs 2016, 14, 172–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Zhu, X.; Zhou, D.; Liang, F.; Wu, Z.; She, Z.; Li, C. Penochalasin K, a new unusual chaetoglobosin from the mangrove endophytic fungus Penicillium chrysogenum V11 and its effective semi-synthesis. Fitoterapia 2017, 123, 23–28. [Google Scholar] [CrossRef]
  108. Yang, B.; Dong, J.; Lin, X.; Zhou, X.; Zhang, Y.; Liu, Y. New prenylated indole alkaloids from fungus Penicillium sp. derived of mangrove soil sample. Tetrahedron 2014, 70, 3859–3863. [Google Scholar] [CrossRef]
  109. Cui, H.; Lin, Y.; Luo, M.; Lu, Y.; Huang, X.; She, Z. Diaporisoindoles A-C: Three Isoprenylisoindole Alkaloid Derivatives from the Mangrove Endophytic Fungus Diaporthe sp. SYSU-HQ3. Org. Lett. 2017, 19, 5621–5624. [Google Scholar] [CrossRef]
  110. Yu, G.; Zhou, G.; Zhu, M.; Wang, W.; Zhu, T.; Gu, Q.; Li, D. Neosartoryadins A and B, Fumiquinazoline Alkaloids from a Mangrove-Derived Fungus Neosartorya udagawae HDN13-313. Org. Lett. 2016, 18, 244–247. [Google Scholar] [CrossRef]
  111. Fu, P.; Yang, C.; Wang, Y.; Liu, P.; Ma, Y.; Xu, L.; Su, M.; Hong, K.; Zhu, W. Streptocarbazoles A and B, Two Novel Indolocarbazoles from the Marine-Derived Actinomycete Strain Streptomyces sp. FMA. Org. Lett. 2012, 14, 2422–2425. [Google Scholar] [CrossRef] [PubMed]
  112. Li, T.; Du, Y.; Cui, Q.; Zhang, J.; Zhu, W.; Hong, K.; Li, W. Cloning, characterization and heterologous expression of the indolocarbazole biosynthetic gene cluster from marine-derived Streptomyces sanyensis FMA. Mar. Drugs 2013, 11, 466–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Li, K.; Liang, Z.; Chen, W.; Luo, X.; Fang, W.; Liao, S.; Lin, X.; Yang, B.; Wang, J.; Tang, L.; et al. Iakyricidins A-D, antiproliferative piericidin analogs bearing carbonyl group or cyclic skeleton from Streptomyces iakyrus SCSIO NS104. J. Org. Chem. 2019, 84, 12626–12631. [Google Scholar] [CrossRef] [PubMed]
  114. Zhu, M.; Zhang, X.; Feng, H.; Che, Q.; Zhu, T.; Gu, Q.; Li, D. Campyridones A–D, pyridone alkaloids from a mangrove endophytic fungus Campylocarpon sp. HDN13-307. Tetrahedron 2016, 72, 5679–5683. [Google Scholar] [CrossRef]
  115. Zhou, Z.-F.; Kurtán, T.; Yang, X.-H.; Mándi, A.; Geng, M.-Y.; Ye, B.-P.; Taglialatela-Scafati, O.; Guo, Y.-W. Penibruguieramine A, a novel pyrrolizidine alkaloid from the endophytic fungus Penicillium sp. GD6 associated with Chinese mangrove Bruguiera gymnorrhiza. Org. Lett. 2014, 16, 1390–1393. [Google Scholar] [CrossRef]
  116. Chen, S.; He, L.; Chen, D.; Cai, R.; Long, Y.; Lu, Y.; She, Z. Talaramide A, an Unusual Alkaloid from the Mangrove Endophytic Fungus Talaromyces sp. (HZ-YX1) as Inhibitor of Mycobacterial PknG. New J. Chem. 2017, 4273–4277. [Google Scholar] [CrossRef]
  117. Gressler, M.; Zaehle, C.; Scherlach, K.; Hertweck, C.; Brock, M. Multifactorial induction of an orphan PKS-NRPS gene cluster in Aspergillus terreus. Chem. Biol. 2011, 18, 198–209. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structures of compounds 1 and 2.
Figure 1. Structures of compounds 1 and 2.
Marinedrugs 20 00535 g001
Figure 2. Structures of compounds 35.
Figure 2. Structures of compounds 35.
Marinedrugs 20 00535 g002
Scheme 1. Proposed biosynthetic pathway for compounds 620 [21,27,29,30].
Scheme 1. Proposed biosynthetic pathway for compounds 620 [21,27,29,30].
Marinedrugs 20 00535 sch001
Figure 3. Structures of compounds 611.
Figure 3. Structures of compounds 611.
Marinedrugs 20 00535 g003
Figure 4. Structures of compounds 1217.
Figure 4. Structures of compounds 1217.
Marinedrugs 20 00535 g004
Figure 5. Structures of compounds 1820.
Figure 5. Structures of compounds 1820.
Marinedrugs 20 00535 g005
Figure 6. Structures of compounds 21 and 22.
Figure 6. Structures of compounds 21 and 22.
Marinedrugs 20 00535 g006
Scheme 2. Proposed biosynthetic pathway for compounds 2124 [23,32,33].
Scheme 2. Proposed biosynthetic pathway for compounds 2124 [23,32,33].
Marinedrugs 20 00535 sch002
Figure 7. Structures of compounds 2325.
Figure 7. Structures of compounds 2325.
Marinedrugs 20 00535 g007
Figure 8. Structures of compounds 2628.
Figure 8. Structures of compounds 2628.
Marinedrugs 20 00535 g008
Scheme 3. Proposed biosynthetic pathway for compounds 2628 [34].
Scheme 3. Proposed biosynthetic pathway for compounds 2628 [34].
Marinedrugs 20 00535 sch003
Figure 9. Structures of compounds 2931.
Figure 9. Structures of compounds 2931.
Marinedrugs 20 00535 g009
Scheme 4. Proposed biosynthetic pathway for compound 31 [36].
Scheme 4. Proposed biosynthetic pathway for compound 31 [36].
Marinedrugs 20 00535 sch004
Figure 10. Structures of compounds 3234.
Figure 10. Structures of compounds 3234.
Marinedrugs 20 00535 g010
Figure 11. Structures of compounds 35 and 36.
Figure 11. Structures of compounds 35 and 36.
Marinedrugs 20 00535 g011
Figure 12. Structures of compounds 37 and 38.
Figure 12. Structures of compounds 37 and 38.
Marinedrugs 20 00535 g012
Figure 13. Structures of compounds 3943.
Figure 13. Structures of compounds 3943.
Marinedrugs 20 00535 g013
Scheme 5. Proposed biosynthetic pathway for compounds 3943 [53].
Scheme 5. Proposed biosynthetic pathway for compounds 3943 [53].
Marinedrugs 20 00535 sch005
Figure 14. Structures of compounds 4446 and dehydrocurvularin.
Figure 14. Structures of compounds 4446 and dehydrocurvularin.
Marinedrugs 20 00535 g014
Figure 15. Structures of compounds 4751.
Figure 15. Structures of compounds 4751.
Marinedrugs 20 00535 g015
Scheme 6. Proposed biosynthetic pathway for compounds 4649 [56].
Scheme 6. Proposed biosynthetic pathway for compounds 4649 [56].
Marinedrugs 20 00535 sch006
Scheme 7. Proposed biosynthetic pathway for compound 51 [67].
Scheme 7. Proposed biosynthetic pathway for compound 51 [67].
Marinedrugs 20 00535 sch007
Figure 16. Structures of compounds 5263.
Figure 16. Structures of compounds 5263.
Marinedrugs 20 00535 g016
Scheme 8. (a) Proposed biosynthetic pathway for compounds 5255 and 6063 in the literature [71]; (b) proposed biosynthetic pathway for 5255 in this review.
Scheme 8. (a) Proposed biosynthetic pathway for compounds 5255 and 6063 in the literature [71]; (b) proposed biosynthetic pathway for 5255 in this review.
Marinedrugs 20 00535 sch008
Figure 17. Structures of compounds 6469.
Figure 17. Structures of compounds 6469.
Marinedrugs 20 00535 g017
Figure 18. Structures of five sesquiterpene skeletons.
Figure 18. Structures of five sesquiterpene skeletons.
Marinedrugs 20 00535 g018
Scheme 9. Proposed biosynthetic pathway for compounds 6366 [74].
Scheme 9. Proposed biosynthetic pathway for compounds 6366 [74].
Marinedrugs 20 00535 sch009
Scheme 10. Proposed biosynthetic pathway for compounds 68 and 69 [77].
Scheme 10. Proposed biosynthetic pathway for compounds 68 and 69 [77].
Marinedrugs 20 00535 sch010
Figure 19. Structures of compounds 7074.
Figure 19. Structures of compounds 7074.
Marinedrugs 20 00535 g019
Scheme 11. (a) Proposed biosynthetic pathway for compounds 7072 [80,81]; (b) Proposed biosynthetic pathway for compounds 7374 [82].
Scheme 11. (a) Proposed biosynthetic pathway for compounds 7072 [80,81]; (b) Proposed biosynthetic pathway for compounds 7374 [82].
Marinedrugs 20 00535 sch011
Figure 20. Structures of compounds 75 and 76.
Figure 20. Structures of compounds 75 and 76.
Marinedrugs 20 00535 g020
Figure 21. Structures of compounds 7679.
Figure 21. Structures of compounds 7679.
Marinedrugs 20 00535 g021
Scheme 12. Proposed biosynthetic pathway for compounds 7679 [83].
Scheme 12. Proposed biosynthetic pathway for compounds 7679 [83].
Marinedrugs 20 00535 sch012
Figure 22. Structures of compounds 80 and 81.
Figure 22. Structures of compounds 80 and 81.
Marinedrugs 20 00535 g022
Scheme 13. Proposed biosynthetic pathway for compounds 80 and 81 [88].
Scheme 13. Proposed biosynthetic pathway for compounds 80 and 81 [88].
Marinedrugs 20 00535 sch013
Figure 23. Structures of compounds 8284.
Figure 23. Structures of compounds 8284.
Marinedrugs 20 00535 g023
Scheme 14. Proposed biosynthetic pathway for compounds 8284 [89].
Scheme 14. Proposed biosynthetic pathway for compounds 8284 [89].
Marinedrugs 20 00535 sch014
Figure 24. Structures of compounds 8589.
Figure 24. Structures of compounds 8589.
Marinedrugs 20 00535 g024
Scheme 15. Proposed biosynthetic pathway for compounds 8689 [93].
Scheme 15. Proposed biosynthetic pathway for compounds 8689 [93].
Marinedrugs 20 00535 sch015
Figure 25. Structures of compounds 90 and 91.
Figure 25. Structures of compounds 90 and 91.
Marinedrugs 20 00535 g025
Scheme 16. Proposed biosynthetic pathway for compounds 90 and 91 [96].
Scheme 16. Proposed biosynthetic pathway for compounds 90 and 91 [96].
Marinedrugs 20 00535 sch016
Figure 26. Structures of compounds 92 and 93.
Figure 26. Structures of compounds 92 and 93.
Marinedrugs 20 00535 g026
Scheme 17. Proposed biosynthetic pathway for compounds 92 and 93 [98].
Scheme 17. Proposed biosynthetic pathway for compounds 92 and 93 [98].
Marinedrugs 20 00535 sch017
Figure 27. Structures of compounds 94106.
Figure 27. Structures of compounds 94106.
Marinedrugs 20 00535 g027
Scheme 18. Proposed biosynthetic pathway for compounds 9498 [99,100].
Scheme 18. Proposed biosynthetic pathway for compounds 9498 [99,100].
Marinedrugs 20 00535 sch018
Figure 28. Structures of compounds 107a/b.
Figure 28. Structures of compounds 107a/b.
Marinedrugs 20 00535 g028
Scheme 19. Proposed biosynthetic pathway for compounds 107a/b [102].
Scheme 19. Proposed biosynthetic pathway for compounds 107a/b [102].
Marinedrugs 20 00535 sch019
Figure 29. Structures of compounds 108110.
Figure 29. Structures of compounds 108110.
Marinedrugs 20 00535 g029
Scheme 20. Proposed biosynthetic pathway for compound 108 [105].
Scheme 20. Proposed biosynthetic pathway for compound 108 [105].
Marinedrugs 20 00535 sch020
Figure 30. Structures of compounds 111 and 112, chaetoglobosin A and C.
Figure 30. Structures of compounds 111 and 112, chaetoglobosin A and C.
Marinedrugs 20 00535 g030
Figure 31. Structures of compounds 113115.
Figure 31. Structures of compounds 113115.
Marinedrugs 20 00535 g031
Scheme 21. Proposed biosynthetic pathway for compounds 113115 [108].
Scheme 21. Proposed biosynthetic pathway for compounds 113115 [108].
Marinedrugs 20 00535 sch021
Figure 32. Structures of compounds 116120.
Figure 32. Structures of compounds 116120.
Marinedrugs 20 00535 g032
Scheme 22. Proposed biosynthetic pathway for compounds 116120 [109].
Scheme 22. Proposed biosynthetic pathway for compounds 116120 [109].
Marinedrugs 20 00535 sch022
Figure 33. Structures of compounds 121 and 122.
Figure 33. Structures of compounds 121 and 122.
Marinedrugs 20 00535 g033
Scheme 23. Proposed biosynthetic pathway for compounds 121 and 122 [110].
Scheme 23. Proposed biosynthetic pathway for compounds 121 and 122 [110].
Marinedrugs 20 00535 sch023
Figure 34. Structures of compounds 123 and 124.
Figure 34. Structures of compounds 123 and 124.
Marinedrugs 20 00535 g034
Scheme 24. Proposed biosynthetic pathway for compounds 123 and 124 [111].
Scheme 24. Proposed biosynthetic pathway for compounds 123 and 124 [111].
Marinedrugs 20 00535 sch024
Scheme 25. Proposed biosynthetic pathway for compounds 125128 [113].
Scheme 25. Proposed biosynthetic pathway for compounds 125128 [113].
Marinedrugs 20 00535 sch025
Figure 35. Structures of compounds 129132.
Figure 35. Structures of compounds 129132.
Marinedrugs 20 00535 g035
Scheme 26. Proposed biosynthetic pathway for compounds 129132 [114].
Scheme 26. Proposed biosynthetic pathway for compounds 129132 [114].
Marinedrugs 20 00535 sch026
Figure 36. Structure of compound 133.
Figure 36. Structure of compound 133.
Marinedrugs 20 00535 g036
Scheme 27. Proposed biosynthetic pathway for compound 133 [115].
Scheme 27. Proposed biosynthetic pathway for compound 133 [115].
Marinedrugs 20 00535 sch027
Figure 37. Structures of compound 134 and rubrobramide.
Figure 37. Structures of compound 134 and rubrobramide.
Marinedrugs 20 00535 g037
Scheme 28. Proposed biosynthetic pathway for compound 134 [116].
Scheme 28. Proposed biosynthetic pathway for compound 134 [116].
Marinedrugs 20 00535 sch028
Table 1. Structurally unusual secondary metabolites from mangrove flora.
Table 1. Structurally unusual secondary metabolites from mangrove flora.
SourceSecondary MetabolitesReported ActivitiesRef
X. moluccensisThaixylomolin A–C (1, 19 and 20)Anti-inflammatory[21]
X. moluccensisThaixylomolin R (2)NR a[22]
X. granatumXylomexicanins E and F (23 and 3)Cytotoxicity[23]
X. granatumHainangrantum I and J (4 and 5)NR[24]
X. moluccensisTrangmolins A–F (610, 17)NR[27]
X. moluccensisKrishnolide J (11)NR[28]
X. moluccensisAndhraxylocarpins A and C (12, 14)NR[29]
X. granatumAndhraxylocarpins A–B and D–E (1213 and 1516)NR[29]
X. moluccensisKrishnadimer A (18)NR[30]
X. granatumXylogranatopyridines A and B (21 and 22)PTP1B inhibitory[32]
X. granatumXylomexicanins I and J (24 and 25)NR[33]
X. moluccensisXylomolones A–C (2628)NR[34]
Excoecaria agallochaAgallochaexceorins D–E (2930)NR[35]
Ceriops decandraDecandrinin (31)NR[36]
a NR: not reported in references.
Table 2. Structurally unusual secondary metabolites from mangrove-associated microorganisms.
Table 2. Structurally unusual secondary metabolites from mangrove-associated microorganisms.
MicroorganismsHostsSecondary MetabolitesReported ActivitiesRef
Penicillium communeKandelia candelPeniisocoumarins A and B (32 and 33)NR a[44]
Penicillium chermesinumMangrove forestPenicilliumolide A (34)NR[45]
Penicillium sp.Mangrove sedimentPenixanthones C and D (35 and 36)NR[46]
Penicillium chrysogenumAcanthus ilicifoliusPenicitols A and B (37-38)Cytotoxicity[48]
Diaporthe sp.Excoecaria agallochaDiaporindenes A–D (39-42)Anti-inflammatory[52]
Diaporthe sp.Excoecaria agallochaIsoprenylisobenzofuran A (43)Anti-inflammatory[52]
Penicillium sumatrenseLumnitzera racemoseSumalarins A−C (4446)Cytotoxicity[55]
Streptomyces sp.Bruguiera gymnorrhizaDivergolide A-D (4750)Antimicrobial,
cytotoxicity
[56]
Streptomyces sp.Mangrove soilHainanmycin A (51)NR[67]
Streptomyces sp.Mangrove sampleStreptoglycerides A–H (5259)Anti-inflammatory[69,70]
Streptomyces sp.Mangrove sampleMiharadiened A–D (6063)NR[71]
Pseudolagarobasidium acaciicolaBruguiera gymnorrhizaAcaciicolin A (64), Spiroacaciicolides A–C (6567)NR[74,75]
Penicillium bilaiaeLumnitzera racemosaPenicibilaenes A–B (6869)Anti-fungal[77]
Aspergillus sp.Mangrove endophyticAsperterpenoid A (70)Antituberculosis[80]
Aspergillus sp.Mangrove endophyticAsperterpenols A–B (7172)Acetylcholinesterase inhibition[81]
Aspergillus terreusKandelia obovateAspterpenacids A–B (7374)NR[82]
Penicillium bilaiaeLumnitzera racemosaChermebilaenes A–B (7576)Antibiotic[84]
Penicillium simplicissimumBruguiera sexangulaSimpterpenoid A (77)Antiviral[85]
Penicillium sp.Bruguiera sexangulaPenicianstinoids A–B (7879)Insecticidal activity[86]
Pestalotiopsis sp.Rhizophora mucronataPestalotiopens A–B (8081)NR[88]
Streptomyces sp.Mangrove rhizosphere soilIndotertine A (82), indotertine B (83a/83b)Antitumor[89,90]
Streptomyces sp.Mangrove rhizosphere soilDrimentine I (84)Antitumor[91]
Penicillium camembertiRhizosphere soil of Rhizophora apiculataSecopaxilline A (85)NR[92]
Mucor irregularisRhizophora stylosaRhizovarin A–D (8689)Antitumor[94]
Erythrobacter sp.Mangrove sedimentErythrazoles A–B (9091)NR[96]
Aspergillus effusesMangrove rhizosphere soilEffusin A (92), dihydrocryptoechinulin D (93)Cytotoxicity[98]
Penicillium brocaeAvicennia marinaSpirobrocazines A–B (9495)NR[99]
Eutypella sp.Deep sea sedimentEutypellazines N–P (9698)NR[100]
Penicillium janthinellumMangrovePenispirozines A–H (99106)Antioxidant[101]
Aspergillus sp.Mangrove endophytic(±)-Asperginulin A (107a/b)Antifouling[102]
Aspergillus versicolorRhizosphere soil of Thespesia populneaPyrasplorines A–C (108110)NR[103]
Penicillium chrysogenumMangrove endophyticPenochalasin I and K (111112)Antibiotic, cytotoxicity[106,107]
penicillium sp.Mangrove sedimentMangrovamides A–C (113115)Acetylcholinesterase inhibition[108]
Diaporthe sp.Excoecaria agallochaDiaporisoindoles A–B (116117), diaporisoindoles D–E (118119), diaporisoindole C (120)Antituberculosis, anti-inflammatory[52,109]
Neosartorya udagawaeMangrove endophyticNeosartoryadins A–B (121122)NR[110]
Streptomyces sp.Mangrove soilStreptocarbazoles A–B (123124)Antitumor[111]
Streptomyces iakyrusMangrove sedimentIakyricidins A–D (125128)NR[113]
Campylocarpon sp.Sonneratia caseolarisCampyridones A–D (129132)Antitumor[114]
Penicillium sp.Bruguiera gymnorrhizaPenibruguieramine A (133)NR[115]
Talaromyces sp.Mangrove endophyticTalaramide A (134)NR[116]
a NR: not reported in references.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, M.-J.; Xu, B.; Guo, Y.-W. Unusual Secondary Metabolites from the Mangrove Ecosystems: Structures, Bioactivities, Chemical, and Bio-Syntheses. Mar. Drugs 2022, 20, 535. https://doi.org/10.3390/md20080535

AMA Style

Wu M-J, Xu B, Guo Y-W. Unusual Secondary Metabolites from the Mangrove Ecosystems: Structures, Bioactivities, Chemical, and Bio-Syntheses. Marine Drugs. 2022; 20(8):535. https://doi.org/10.3390/md20080535

Chicago/Turabian Style

Wu, Meng-Jun, Baofu Xu, and Yue-Wei Guo. 2022. "Unusual Secondary Metabolites from the Mangrove Ecosystems: Structures, Bioactivities, Chemical, and Bio-Syntheses" Marine Drugs 20, no. 8: 535. https://doi.org/10.3390/md20080535

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