Bioactive natural products derived from mangrove-associated microbes

Jing Xu *
Key Laboratory of Protection and Development Utilization of Tropical Crop Germplasm Resources, Ministry of Education, College of Material and Chemical Engineering, Hainan University, Haikou 570228, P. R. China. E-mail: happyjing3@163.com; Fax: +86 898 6627 9010; Tel: +86 898 6627 9226

Received 3rd October 2014 , Accepted 19th November 2014

First published on 19th November 2014


Abstract

This review summarizes new findings concerning the sources and characteristics of various natural products that can be extracted from mangrove-associated microbes over the past three years (January 2011–December 2013). The natural products are discussed with a focus on bioactivity, highlighting the unique chemical diversity of these metabolic products.


image file: c4ra11756e-p1.tif

Jing Xu

Jing Xu obtained her BS and MS degrees from Hainan University. In 2006, she began her doctoral research with Professor Huashi Guan at the College of Medicine and Pharmaceutics, Ocean University of China. In 2007, she continued with her doctoral research in a group working with Professor Peter Proksch and received her PhD degree in 2010 at Heinrich-Heine-Universität Düsseldorf, Germany. She joined the faculty at Hainan University after her graduation where she is currently an Associate Professor of Pharmaceutical Chemistry. Simultaneously, she undertook a postdoctoral appointment at Nanjing University, with Professor Renxiang Tan from 2011 to 2014. Her research interests are focused on natural products originating from endophytic microbes and medicinal plants.


1. Introduction

Microorganisms from special ecological niches such as the mangrove endosymbionts are a rich source of diverse and structurally unique bioactive natural products. Microbes inhabiting mangrove ecosystems adapt to frequent and sporadic environmental changes, including high salinity, low oxygen, nutrient limitation, tidal gradients, high temperatures, excessively high light, and drought.1 The warm and damp conditions result in an active microbial community, which can act as an effective selector for metabolic pathway adaptations via generation of unique functional metabolites of pharmaceutical importance.

Mangrove forests are complex ecosystems that harbor diverse groups of microorganisms including actinomycetes, bacteria, fungi, cyanobacteria, microalgae, macroalgae and fungus-like protozoa. In the tropical mangrove microbial community, bacteria and fungi constitute 91% of the total microbial biomass, whereas algae and protozoa represent only 7% and 2%, respectively. Although the mangrove ecosystem has extensive microbial diversity, culture-dependent technologies reveal only a small percentage (<1%) of the microbial community while 99% of the microorganisms remain undiscovered. Among the isolated microbes, less than 5% of species have been presently chemically estimated, and these include the actinomycetes, bacteria and fungi. The advent of modern techniques provides the opportunity to find novel metabolites.2

Mangrove associated microbial natural products have been the subject of several review articles. Advances in the chemistry and bioactivity of fungi from mangroves have been reviewed periodically in Natural Product Reports by Blunt and his co-workers.3,4 Wang et al. reviewed the studies on structure and bioactivity of metabolites isolated from mangrove-derived fungi in the South China Sea.5 Prof. Proksch' group compiled a comprehensive review detailing a decade of studies on potential of mangrove-derived endophytic fungi as promising bioactive natural products.6,7 Our previous review discussed the role of these endophytic fungi as a major source of antimicrobial agents. These fungi produce a wide range of medicinal compounds, including antitumor, neuroprotective, antioxidant, anti-AChE, anti-BKCa channel agents, and 348 molecules were discovered before 2011.8

However, 464 new metabolites have been discovered recently; therefore, we describe here the source, chemistry, and biology of the newly discovered biomolecules from the mangrove-associated microbes. We also summarize the chemical synthesis and the biosynthetic relationship of metabolites. All relevant studies focusing on known secondary metabolites in terms of bioactivity, the source of the microorganism, and the location of collection are examined, and particular emphasis is given to their potential use as drug leads. The running titles are presented in the following order: actinomycetic, bacterial, and fungal producers with structures classified within a biogenetic context as polyketides, terpenes, and nitrogenated compounds. Bioassays showed that the antitumor, anti-influenza A H1N1, anti-enzyme, anti- or activate-subintestinal vessel plexus and brine shrimp lethality activities were the most notable bioactivities of secondary metabolites isolated during 2011–2013. Possible biogenetic relationships between several alkaloids, quinones, phenols, lactones, sesquiterpenes, sesterterpenoids, and meroterpenes are discussed. The biomacromolecules originating from the title microbes were not within the scope of this review. The Metabolites Name Index in conjunction with the Microbial and Host Source Index, the Bioactivity Index and the Leading References on isolation in the accompanying tables, will help understand the fascinating chemistry and biology of naturally occurring mangrove-associated microbial metabolites.

2. Actinomyces-derived molecules

Due to the high rediscovery rate of known compounds from Actinomycetes, there has been a renewed interest in the development of new antimicrobial agents from this species. These compounds can help combat the increasing number of multidrug-resistant human pathogens.9 Mangrove-derived actinomycetes are a potentially vast resource of structurally diverse natural products with unusual biological activity. Streptomyces are Gram-positive bacteria known for the production of an enormous variety of biologically active secondary metabolites, including antibiotics, immunosuppressants, and anticancer agents and have been become the focus of antibiotic research in the past three years.

2.1. Terpenoids

Eudesmene sesquiterpenes are proposed to be originated from further cyclization of the monocyclic germacranes.10 They are most abundant as eukaryotic secondary metabolites and may be referred to as selinanes, commonly metabolized by a variety of plants, such as Inula japonica,11 Blumea balsamifera,12 and Nectandra cissiflora.13 The only examples from prokaryotes are the recently described selina-4(14),7(11)-diene-8,9-diol from a marine Streptomyces sp.;14 and 1,6,11-eudesmanetriol and 11-eudesmene-1,6-diol from an endophytic Streptomyces sp. of Drymaria diandra.15 Five new members of this family, kandenols A–E (1–5), were characterized from a culture filtrate of endophytic Streptomyces sp. HKI0595 isolated from Kandelia candel (L.) (Xiamen, Fujian, China) using an HPLC-MS hyphenated system. The configuration of these eudesmenes was rigorously established by a combination of the Mosher method and comparison of CD spectra with α-rotunol and β-rotunol. Hydroperoxide moieties are relatively rare in compounds 3 and 4. Metabolite 5 is the first actinomycete agarofuran, which belongs to an important group of antibiotics. Unfortunately, none of these sesquiterpenes exhibited any cytotoxicity against the 12 selected human cell lines tested. However, weak to moderate antimicrobial activities were detected against Bacillus subtilis ATCC 6633 and Mycobacterium vaccae IMET 10670, with MIC values ranging from 12.5–50 μg mL−1.16
image file: c4ra11756e-u1.tif

image file: c4ra11756e-u2.tif

2.2. Nitrogenated compounds

Indole terpenoids encompass a highly diverse group of natural products, including infamous psychotropic agents such as lysergic acid derivatives, the aphrodisiac yohimbine, and the potassium channel blockers paxilline and lolitrem. A remarkable fact about this multifarious class is that practically all indole terpene alkaloids have been isolated from plants and fungi. Further work on the aforementioned Streptomyces sp. HKI0595 endophyted in K. candel and another endophyte Streptomyces sp. GT2002/1503 associated in Bruguiera gymnorrhiza revealed five new pentacyclic indolosesquiterpene congeners, xiamycin A (6) and its methyl ester (7), xiamycin B (8), the seco-derivative indosespene (9), and the novel bridged spiro compound sespenine (10) by the same research team.17,18 Their most likely biosynthetic pathway and interrelationship are proposed based on the heterologous gene expression and mutational analysis as shown in Scheme 1.19 A broad bioactivity screen revealed that compound 6 was moderately active against HIV, specifically blocked R5, but had no effects on X4 tropic HIV-1 infection. Compound 7 exhibited potential cytotoxicity compared with 6 in a modified propidium iodide assay, with IC50 values of 10.13 μM. Compound 8–10 showed moderate to strong antimicrobial activities in agar diffusion assays against several bacteria, including methicillin-resistant Staphylococcus aureus and vancomycin-resistant Enterococcus faecalis; however, these compounds did not show cytotoxicity against human tumor cell lines. These rare endophyte metabolites likely play an ecological role in their habitat because their diverse antiviral, antibacterial, and antifungal activities may contribute to the antibiotic reservoir of the mangrove plants.18 It is intriguing that these prokaryotes are endophytes producing typical plant metabolites. A unique diazaphenanthrene alkaloid, named 1-N-methyl-3-methylamino-[N-butanoic acid-3′-(9′-methyl-8′-propen-7′-one)-amide]-benzo[f][1,7]naphthyridine-2-one (11) and a known metabolite N-[[3,4-dihydro-3S-hydroxy-2S-methyl-2-(4′R-methyl-3′S-pentenyl)-2H-1-benzopyran-6-yl]carbonyl]-threonine (12) were isolated from Streptomyces albogriseolus and Streptomyces xiamenensis, respectively. Both of these compounds were obtained from mangrove sediments collected in the national mangrove reserve in Fujian province of China. A bioassay disclosed that the compound 12 inhibited the proliferation of human lung fibroblasts (WI26); blocks adhesion of human acute monocytic leukemia cells (THP-1) to a monolayer of WI26 cells; and reduces the contractile capacity of WI26 cells in three-dimensional free-floating collagen gels.20,21 The total synthesis of 11 was accomplished by coupling the tricyclic heteroaromatic carboxylic acid with the chiral unsaturated amino ketoacid to yield a fused tricyclic heteroaromatic system as shown in Scheme 2 and the absolute configuration of 11 was firstly determined.22 A new benzamide, 3-hydroxyl-2-N-iso-butyryl-anthranilamide (13), together with two known benzamides (14, 15) and three known quinazolines (16–18), was isolated from a mangrove actinomycetes Streptomyces sp. no. 061316 isolated from mangrove soils (Wenchang, Hainan, China). Although the extract from which it was isolated displayed in vitro inhibitory caspase-3 activity, the pure natural products only 14 and 18 were found to be moderately active with IC50 values of 32 and 36 μM, respectively.23 Several studies have demonstrated that Streptomyces is a successful source of bioactive benzoxazole metabolites, such as antitumor agents UK-1 (ref. 24) and nataxazole,25 antibiotics A33853 (ref. 26) and caboxamycin.27 An antitumour benzoxazole derivative AJI956 (19), originally obtained from the actinomycete strain Streptomyces sp., was purified from Streptomyces sp. A1626 in Kandelia candel.28 Micro-broth dilution assay disclosed its additional strength in moderating antifungal activity against fungal strains Candida albicans ATCC90028, Cryptococcus neoformans SLA1017 and Aspergillus fumigatus SIA1524, with IC50/MIC values of 0.78/1.25, 1.5/2.5, and 15/20 μg mL−1. Halogenation was reported to be an important feature for the bioactivity of a large number of distinct natural products. Chlorination was the most frequently found modification, followed by bromination, while iodination and fluorination are rare in nature.29 Using PCR-based genetic screening techniques on mangrove sediments, 163 strains of actinomycetes were isolated and investigated for their potential to produce halogenated metabolites. A previously known antibiotic, enduracidin (20), was identified in the mangrove-derived Streptomyces atrovirens MGR140.30,31
image file: c4ra11756e-s1.tif
Scheme 1 Biosynthetic pathways for the generation of compounds 6–10.

image file: c4ra11756e-s2.tif
Scheme 2 Synthesis of benzonaphthyridine alkaloid 11. Reagents and conditions: (a) triphosgene, Et3N, THF, 0 °C, 2 h, 78%; (b) LiAlH4, THF, reflux, 2 h, 68%; (c) MnO2, THF, rt, 20 h, 77%; (d) diethyl malonate, piperidine, 170 °C, 5 h, 74%; (e) KOH, ethanol, reflux, 0.5 h, then glacial acetic acid, 87%; (f) DPPA, Et3N, t-BuOH, reflux, 5 h, 73%; (g) CH3I, Cs2CO3, DMF, 0 °C to rt, overnight, 96%; (h) SeO2, 70% TBHP, 1,4-dioxane, 50 °C, 5 h, 82%; (j) N,O-dimethylhydroxylaminehydrochloride, ClCO2-i-Bu, NMM, CH2Cl2, −15 °C to rt, 2.5 h, 96%; (k) 2-methyl-1-propenyl magnesium bromide solution (0.5 M in THF), THF, 0 °C, 4 h, 73%; (l) concentrated H2SO4, tert-butyl acetate, 0 °C, 2 h, then saturated NaHCO3, 80%; (m) (i) Ag2O, KOH, H2O, ethanol, rt, 1 h, then filtered through Celite and dried; (ii) HATU, DIPEA, image file: c4ra11756e-u3.tif, DMF, rt, overnight, 45% over two steps; (n) CF3COOH, CH2Cl2, rt, 5 h, 89%.

3. Bacterial metabolites

Information on bacterial strains, hosts, natural products, their biological activities, and relevant articles is included in Table 1. A predominant focus of natural products study of mangrove-associated bacteria has been on species of the genus Bacillus. However other bacteria, such as Gram-negative Erythrobacter have also proven to be an additional source of biologically and chemically interesting natural products. Chemical examination of an extract from an Erythrobacter sp. isolated from mangrove sediments yielded erythrazoles A (21) and B (22). The erythrazoles are of mixed biosynthetic origin containing a tetrasubstituted benzothiazole, an appended diterpene side chain, and a glycine unit. Based on the structural analysis, formation of the benzothiazole moiety was proposed to include a series of well precedented steps found in ubiquinone biosynthesis, i.e. prenylation, decarboxylation, hydroxylation, methylation, oxidation, and cyclisation. Their most likely biosynthetic pathway and interrelationship are shown in Scheme 3. Compound 21 is cytotoxic to a panel of non-small cell lung cancer (NSCLC) cell lines, with IC50 values of 1.5, 2.5, and 6.8 μM against H1325, H2122, and HCC366, respectively.32 This encouraged subsequent efforts to investigate other bacteria that were not necessarily scarce in nature, but were rarely ever brought into culture and study.
Table 1 Actinomycetic and bacterial metabolites
Bacterial species Host(s) Natural product(s) Biological activity Ref.
Erythrobacter sp. Sediments Erythrazole A (21) Cytotoxic 32
Erythrazole B (22)  
Streptomyces albogriseolus Sediments 1-N-Methyl-3-methylamino-[N-butanoic acid-3′-(9′-methyl-8′-propen-7′-one)-amide]-benzo[f][1,7]naphthyridine-2-one (11)   20
Streptomyces xiamenensis Sediments N-[[3,4-Dihydro-3S-hydroxy-2S-methyl-2-(4′R-methyl-3′S-pentenyl)-2H-1-benzopyran-6-yl]carbonyl]-threonine (12) Anti-fibrotic 21
Streptomyces sp. A1626 Kandelia candel AJI956 (19) Antitumour; antifungal 28
Streptomyces sp. GT2002/1503 Bruguiera gymnorrhiza Xiamycin B (8)   18
Indosespene (9)  
Sespenine (10)  
Streptomyces sp. HKI0595 Kandelia candel Kandenol A (1) Moderate antimicrobial 16
Kandenol B (2) Moderate antimicrobial
Kandenol C (3) Moderate antimicrobial
Kandenol D (4) Weak antimicrobial
Kandenol E (5) Moderate antimicrobial
Xiamycin A (6) Selective anti-HIV 17
Xiamycin A methyl ester (7) Cytotoxic
Streptomyces atrovirens MGR140   Enduracidin (20) Antibiotic 30
Streptomyces sp. no. 061316 Soil 3-Hydroxyl-2-N-iso-butyryl-anthranilamide (13)   23
3-Hydroxyl-anthranilamide (14) Caspase-3 catalytic inhibitor
Anthranilamide (15)  
8-Hydroxy-4(3H)-quinazoline (16)  
8-Hydroxy-2-methyl-4(3H)-quinazoline (17)  
8-Hydroxy-2,4-dioxoquinazoline (18) Caspase-3 catalytic inhibitor



image file: c4ra11756e-s3.tif
Scheme 3 Possible biosynthetic pathway of the formation of compounds 21 and 22.

4. Fungal metabolites

Mangrove-associated fungi produce a larger number of secondary metabolites compared to any other mangrove-derived microbes, especially the endophytic fungi.8 Many of these fungi are capable of synthesizing bioactive compounds that may contribute to their adaptation to harsh environmental factors such as predation and microbial infection during long co-evolution. An ever-increasing number of compounds are being reported from fungi associated with mangrove plants and soils. Some of these compounds have been shown to have a broad spectrum of biological activity, and they can be grouped into several categories, including isocoumarins, chromones, flavonoids, quinones, xanthones, phenols and phenolic acids, lactones, terpenoids, nitrogenated compounds, steroids, etc. Table 2 lists the hosts, natural products, biological activities, and articles related to these fungi and their products.
Table 2 Fungal metabolites
Fungal species Host(s) Natural product(s) Biological activity Ref.
Acremonium sp. PSU-MA70 Rhizophora apiculata Acremonone B (34)   40
Acremonone C (35)  
Acremonone D (36)  
Acremonone E (37)  
Acremonone F (38)  
Acremonone G (39)  
Acremonone H (40)  
Acremonide (216)  
Acremonone A (217)  
(+)-Brefeldin A (218) Strong protein secretion inhibitor
5,7-Dimethoxy-3,4-dimethyl-3-hydroxyphthalide (219)  
4-Methyl-1-phenyl-2,3-hexanediol (272)  
(2R,3R)-4-Methyl-1-phenyl-2,3-pentanediol (273)  
8-Deoxytrichothecin (314) Selectively cytotoxic
Trichodermol (315) Antifungal
Guangomide A (440) Weak antibacterial
Guangomide B (441) Weak antibacterial
Sch 54794 (442) PAF inhibitor
Sch 54796 (443)  
Alternaria tenuissima EN-192 Rhizophora stylosa Tricycloalternarene 3a (87) Moderate antibacterial 60
Tricycloalternarene 1b (88)  
Tricycloalternarene 2b (89)  
Paspaline (406)  
Penijanthine A (409)  
Paspalinine (410)  
Penitrem A (411)  
Djalonensone  
Alternariol  
Alternaria sp. ZJ9-6B Aegiceras corniculatum Alterporriol K (70) Moderate cytotoxic 52
Alterporriol L (71) Moderate cytotoxic
Alterporriol M (72)  
Anthraquinones (73–76)  
Aspergillus effuses H1-1 Rhizosphere soil Effusin A (393)   86 and 210
Dihydrocryptoechinulin D (394) Strong cytotoxic; selectivity topoisomerase I inhibitor
Auroglaucin (166)  
Dihydroneochinulin B (395) Weak cytotoxic
Didehydroechinulin B (396)  
Neoechinulin B (397) Strong cytotoxic
Cryptoechinuline D  
Aspergillus nidulans MA-143 Rhizophora stylosa Aniduquinolone A (376)   203 and 226
Aniduquinolone B (377) Brine shrimp lethality
Aniduquinolone C (378) Brine shrimp lethality
6-Deoxyaflaquinolone E (339)  
Isoaflaquinolone E (380)  
14-Hydroxyaflaquinolone F (381)  
Aflaquinolone A (382) Brine shrimp lethality
Aniquinazoline A (429) Strong brine shrimp lethality
Aniquinazoline B (430) Strong brine shrimp lethality
Aniquinazoline C (431) Strong brine shrimp lethality
Aniquinazoline D (432) Strong brine shrimp lethality
Aspergillus niger MA-132 Avicennia marina Nigerapyrone A (187)   103
Nigerapyrone B (188) Selective cytotoxic
Nigerapyrone C (189)  
Nigerapyrone D (190) Weak cytotoxic
Nigerapyrone E (191) Cytotoxic
Nigerapyrone F (192)  
Nigerapyrone G (193)  
Nigerapyrone H (194)  
Asnipyrone A (195)  
Asnipyrone B (196) Selective cytotoxic
Nigerasterol A (456) Strong cytotoxic 232
Nigerasterol B (457)  
Malformin A1 (445) Weak antibacterial
Malformin C (446) Weak antibacterial
Aspergillus taichungensis ZHN-7-07 Acrostichum aureum Prenylterphenyllin A (130) Moderate cytotoxic 76
Prenylterphenyllin B (131)  
Prenylterphenyllin C (132)  
Prenylcandidusin A (133)  
Prenylcandidusin A (134) Moderate cytotoxic
Prenylcandidusin C (135)  
4′′-Dehydro-3-hydroxyterphenyllin (136) Moderate cytotoxic
Prenylterphenyllin (137) Moderate cytotoxic
Terprenin (138)  
Deoxyterhenyllin (139)  
3-Hydroxyterphenyllin (140)  
Terphenyllin (141)  
3,3′-Dihydroxyterphenyllin (142)  
Candidusin A (143)  
Candidusin C (144)  
Aspergillus terreus Gwq-48 Rhizosphere soil Isoaspulvinone E (232) Significant anti-influenza A H1N1; significant H1N1 neuraminidase inhibitor 122
Aspulvinone E (233) Significant anti-influenza A H1N1
Pulvic acid (234) Significant anti-influenza A H1N1
Aspergillus terreus A8-4 Sediment 7′′-Hydroxybutyrolactone III (197) Weak cytotoxic 105
Butyrolactone I (198) Weak cytotoxic; eukaryotic CDK inhibitor
Terretrione A (372)  
Terretrione B (373)  
Terretrione C (374)  
cyclo(Leu-Pro)  
cyclo(Val-Pro)  
cyclo(Ile-Pro)  
cyclo(Phe-Pro)  
Aspergillus terreus (no. GX7-3B) Bruguiera gymnoihiza 8-Hydroxy-2-[1-hydroxyethyl]-5,7-dimethoxynaphtho[2,3-b]thiophene-4,9-dione (90)   61 and 158
Anhydrojavanicin (91) Remarkable AChE inhibitor
8-O-Methyljavanicin (92)  
Botryosphaerone D (93)  
6-Ethyl-5-hydroxy-3,7-dimethoxynaphthoquinone (94)  
Beauvericin (447) Insecticidal
8-O-Methylbostrycoidin (454)  
NGA0187 (458) Neuritogenic
3β,5α-Dihydroxy-(22E,24R)-ergosta-7,22-dien-6-one (459)  
3β,5α,14α-Trihydroxy-(22E,24R)-ergosta-7,22-dien-6-one (460)  
Botryosphaerin F (309) Strong cytotoxic
(13,14,15,16-Tetranorlabd-7-ene-19,6b:12,17-diolide) (310)  
Botryosphaerin B (311)  
LL-Z1271b (312) Significant cytotoxic
Aspergillus tubingensis (GX1-5E) Pongamia pinnata Rubasperone A (47)   46 and 47
Rubasperone B (48)  
Rubasperone C (49)  
Rubasperone D (50) Mild cytotoxic
Rubasperone E (51)  
Rubasperone F (52)  
Rubasperone G (53)  
Rubrofusarin (54) Moderate tyrosinase inhibitor
Rubrofusarin B (55) Mild α-glucosidase inhibitor; mild cytotoxic
TMC 256 A1 (56) Cytotoxic
Fonsecin (57)  
Flavasperone (58) Mild cytotoxic
Aspergillus sp. strain FSY-01 and FSW-02 Avicennia Neoaspergillic acid (348) Significant antibacterial 192
Aspergicin (415) Moderate antibacterial
Ergosterol  
Aspergillus sp. 16-5c Sonneratia apetala Asperterpenoid A (317) Strong mPTPB inhibitor 164
Aspergillus sp. 085241B   Acetoxydehydroaustin B (325)   184
1,2-Dihydro-acetoxydehydroaustin B (326)  
Aspergillus sp. 085242 Acanthus ilicifolius Asperterpenol A (318) Strong AChE inhibitor 174
Asperterpenol B (319) Strong AChE inhibitor
Aspergillus sp. Bruguiera gymnorrhiza Aspergillumarin A (32) Weak antibacterial 39
Aspergillumarin B (33) Weak antibacterial
Bionectria ochroleuca Sonneratia caseolaris Pullularin E (434) Moderate cytotoxic 227
Pullularin F (434)  
Pullularin A (435) Moderate cytotoxic
Pullularin C (436) Moderate cytotoxic
Verticillin D (437) Pronounced cytotoxic
Cladosporium sp. PJX-41 Soil 3-Hydroxyglyantrypine (416)   189
Oxoglyantrypine (417a, 417b) Compound 417b significant anti-H1N1
Cladoquinazoline (418)  
epi-Cladoquinazoline (419)  
Norquinadoline A (420) Significant anti-H1N1
Quinadoline A (421)  
Deoxynortryptoquivaline (422) Significant anti-H1N1
Deoxytryptoquivaline (423) Significant anti-H1N1
Tryptoquivaline (424) Significant anti-H1N1
CS-C (425)  
Quinadoline B (426)  
Prelapatin B (427)  
Glyantrypine (428)  
Corynespora cassiicola Laguncularia racemosa 2,5,7-Trihydroxy-3-methoxynaphthalene-1,4-dione (99)   62 and 106
6-(3′-Hydroxybutyl)-7-O-methylspinochrome B (99)  
Xestodecalactone D (199)  
Xestodecalactone E (200)  
Xestodecalactone F (201)  
Corynesidone C (202)  
Corynesidone A (203)  
Corynesidone B (204) Protein kinase inhibitor
Coryoctalactone A (205)  
Coryoctalactone B (206)  
Coryoctalactone C (207)  
Coryoctalactone D (208)  
Coryoctalactone E (209)  
Deuteromycete sp. Drift wood Deuteromycol A (123)   73
Deuteromycol B (124)  
Diaporthe phaseolorum Laguncularia racemosa 3-Hydroxypropionic acid (3-HPA) (267) Antibacterial 139
Diaporthe sp. Rhizophora stylosa Diaporol A (291)   150
Diaporol B (292)  
Diaporol C (293)  
Diaporol D (294)  
Diaporol E (295)  
Diaporol F (296)  
Diaporol G (297)  
Diaporol H (298)  
Diaporol I (299)  
3β-Hydroxyconfertifolin (300)  
Diplodiatoxin (301)  
Emericella sp. Aegiceras corniculatum Acetoxydehydroaustin B (325)   185
Austin (327) nAChR antagonist
Deacetylaustin (328)  
Dehydroaustin (329) nAChR antagonist
Emerimidine A (349) Moderate anti-H1N1
Emerimidine B (350) Moderate anti-H1N1
Emeriphenolicin A (351)  
Emeriphenolicin B (352)  
Emeriphenolicin C (353)  
Emeriphenolicin D (354)  
Aspernidine A (355)  
Aspernidine B (356)  
Eurotium rubrum Hibiscus tiliaceus Emodic acid (99)   63
9-Dehydroxyeurotinone (210) Weak antibacterial
2-O-Methyl-9-dehydroxyeurotinone (211)  
12-Demethyl-12-oxo-eurotechinulin B (384) Moderate cytotoxic
Variecolorin J (386)  
Eurotechinulin B (387)  
Variecolorin G (388) Moderate cytotoxic
Alkaloid E-7 (389) Moderate cytotoxic
Cryptoechinuline G (390)  
Isoechinulin B (391)  
7-Isopentenylcryptoechinuline D (392)  
Emodin  
Flavodon flavus PSU-MA201 Rhizophora apiculata Flavodonfuran (266) Mild antibacterial and antifungal 138
Tremulenolide A (306) Mild antibacterial and antifungal
Fusarium equiseti AGR12 Rhizophora stylosa Equisetin (357) Phytotoxic; moderate antibacterial 193
epi-Equisetin (358) Phytotoxic; moderate antibacterial
Ergosterol  
Fusarium incarnatum (HKI0504 Aegiceras corniculatum N-2-Methylpropyl-2-methylbutenamide (363)   197
Fusamine (364) Weak cytotoxic
3-(1-Aminoethylidene)-6-methyl-2H-pyran-2,4(3H)-dione (365) Weak cytotoxic
2-Acetyl-1,2,3,4-tetrahydro-β-carboline (412) Weak cytotoxic
Fusarine (452)  
Fusarium oxysporum Rhizophora annamalayana Taxol (316) Clinical antitumor medicine 170
Fusarium sp. Kandelia candel Fusaric acid (455) Antimycobacterial 233
Fusarium sp. ZZF60 Kandelia candel 5-Hydroxy-7-methoxy-4′-O-(3-methylbut-2-enyl)isoflavone (66)   50
Eriodictyol (67)  
3,6,7-Trihydroxy-1-methoxyxanthone (111)  
1,3,6-Trihydroxy-8-methylxanthone (112)  
Vittarin-B (167)  
cyclo(Phe-Tyr)  
Hypocrea virens Rhizophora apiculata 2-Methylimidazo[1,5-b]isoquinoline-1,3,5(2H)-trione (375)   201
Leucostoma persoonii Rhizophora mangle Cytosporone R (150)   79
Cytosporone E (220) Anti-infective; antimicrobial
Cytosporones B  
Cytosporones C  
Nigrospora sp. no. 1403 Kandelia candel Deoxybostrycin (77) Strong antitumor; strong antimicrobial; strong antimycobacteria 56 and 57
Methyl 3-chloro-6-hydroxy-2-(4-hydroxy-2-methoxy-6-methylphenoxy)-4-methoxybenzoate (145)  
(2S,5′R,E)-7-Hydroxy-4,6-dimethoxy-2-(1-methoxy-3-oxo-5-methylhex-1-enyl)-benzofuran-3(2H)-one (251)  
Dechlorogriseofulvin (254)  
Griseofulvin  
Bostrycin Phytotoxic; antibacterial; strong antitumor; strong antimicrobial
Nigrospora sp. no. MA75 Pongamia pinnata 2,3-Didehydro-19α-hydroxy-14-epicochlioquinone B (78) Strong cytotoxic 58
4-Deoxytetrahydrobostrycin (79) Moderate antimicrobial
3,8-Dihydroxy-6-methoxy-1-methylxanthone (108)  
3,6,8-Trihydroxy-1-methylxanthone (109)  
Griseophenone C (110) Antibacterial
6-O-Desmethyldechlorogriseofulvin (252)  
6′-Hydroxygriseofulvin (253)  
Dechlorogriseofulvin (254)  
Griseofulvin  
Tetrahydrobostrycin  
Paecilomyces sp.   Paeciloxocins A (179) Strong cytotoxic; mildly antimicrobial 89
Paeciloxocin B (180)  
Penicillium camemberti OUCMDZ-1492 Mangrove soil and mud (2S,4bR,6aS,12bS,12cS,14aS)-3-Deoxo-4b-deoxypaxilline (398) Significant anti-H1N1 219
(2S,4aR,4bR,6aS,12bS,12cS,14aS)-4a-Demethylpaspaline-4a-carboxylic acid (399) Significant anti-H1N1
(2S,3R,4R,4aS,4bR,6aS,12bS,12cS,14aS)-4a-Demethylpaspaline-3,4,4a-triol (400) Significant anti-H1N1
(2R,4bS,6aS,12bS,12cR,14aS)-2′-Hydroxypaxilline (401)  
(2R,4bS,6aS,12bS,12cR,14aS)-9,10-Diisopentenylpaxilline (402) Significant anti-H1N1
(6S,7R,10E,14E)-16-(1H-Indol-3-yl)-2,6,10,14-tetramethylhexadeca-2,10,14-triene-6,7-diol (403) Significant anti-H1N1
Emindole SB (404) Significant anti-H1N1
21-Isopentenylpaxilline (405) Significant anti-H1N1
Paspaline (406) Significant anti-H1N1
Paxilline (407) Significant anti-H1N1
Dehydroxypaxilline (408)  
Penicillium chermesinum (ZH4-E2) Kandelia candel 6′-O-Desmethylterphenyllin (125) Strong α-glucosidase inhibitor 53
3-Hydroxy-6′-O-desmethylterphenyllin (126) Strong α-glucosidase inhibitor
3′′-Deoxy-6′-O-desmethylcandidusin B (127) Strong α-glucosidase inhibitor
3,3′′-Dihydroxy-6′-O-desmethylterphenyllin (128, 129) Acetylcholinesterase inhibitor
6′-O-Desmethylcandidusin B  
Chermesinone B (185)  
Chermesinone C (186)  
Chermesinone A (259) Mild R-glucosidase inhibitor
Penicillium chrysogenum PXP-55 Rhizophora stylosa Chrysogeside A (341)   189
Chrysogeside B (342) Antimicrobial
Chrysogeside C (343)  
Chrysogeside D (344)  
Chrysogeside E (345)  
Chrysogedone A (346)  
Chrysogedone B (347)  
Penicillium expansum 091006 Excoecaria agallocha Expansol C (151) Weak cytotoxic 80
Expansol D (152)  
Expansol E (153) Weak cytotoxic
ExpansolF (154)  
3-O-Methyldiorcinol (155)  
(+)-(7S)-7-O-Methylsydonic acid (156)  
3,7-Dihydroxy-1,9-dimethyldibenzofuran (157)  
Orcinol (158)  
2,4-Dimethoxyphenol (159)  
4-Hydroxybenzoic acid (160)  
Butyrolactone I (198) Weak cytotoxic; eukaryotic CDK inhibitor
Butyrolactone V (231) Moderate antimalarial
WIN 64821 (444)  
Expansol A  
Expansol B  
Diorcinol  
S-(+)-Sydonic acid  
Penicillium sumatrense MA-92 Lumnitzera racemosa Sumalarin A (240) Strong cytotoxic 130
Sumalarin B (241) Strong cytotoxic
Sumalarin C (242) Strong cytotoxic
Curvularin (239)  
Dehydrocurvularin (243) Strong cytotoxic
Curvularin-7-O-β-D-glucopyranoside (244)  
Penicillium sp. MA-37 Bruguiera gymnorrhiza 7-O-Acetylsecopenicillide C (146)   78
Hydroxytenellic acid B (147)  
6-[2-Hydroxy-6-(hydroxymethyl)-4-methylphenoxy]-2-methoxy-3-(1-methoxy-3-methylbutyl)benzoic acid (148)  
Secopenicillide C (149)  
Δ1′,3′-1′-Dehydroxypenicillide (212)  
Penicillide (213) Brine shrimp lethality
Dehydroisopenicillide (214) Brine shrimp lethality
3′-O-Methyldehydroisopenicillide (215)  
4,25-Dehydrominiolutelide B (330)  
4,25-Dehydro-22-deoxyminiolutelide B (331)  
Isominiolutelide A (332)  
Berkeleyacetals A (333)  
Berkeleyacetal B (334)  
22-Epoxyberkeleydione (335)  
Penicillium sp. sk5GW1L Kandelia candel Arigsugacin I (338) Potential AChE inhibitor 188
Arigsugacins F (339) Potential AChE inhibitor
Territrem B (340) Potential AChE inhibitor
Penicillium sp. ZH16 Avicennia 5-Methyl-8-(3-methylbut-2-enyl)furanocoumarin (23) Modest cytotoxic 33
Bergapten (24)  
Scopoletin (25)  
Umbelliferone (26)  
Sterequinone C (80)  
1,7-Dihydroxyxanthone (113)  
3,5-Dimethoxybiphenyl (168)  
cyclo(6,7-en-Pro-L-Phe)  
Penicillium sp. ZH58 Avicennia 5-Hydroxy-3-hydroxymethyl-7-methoxy-2-methyl-4H-1-benzopyran-4-one (63) Antibiotic 49
4-(Methoxymethyl)-7-methoxy-6-methyl-1(3H)-isobenzofuranone (238)  
Curvularin (239)  
5,5′-Oxy-dimethylene-bis(2-furaldehyde) (281)  
Dilation (313)  
Harman (1-methyl-β-carboline) (413)  
N9-Methyl-1-methyl-β-carboline (414)  
Lumichrome (450)  
Pestalotiopsis clavispora Bruguiera sexangula 3-Hydroxy-4-methoxystyrene (169)   86
3,4-Dihydroxyphenylethanol (170)  
p-Tyrosol (171)  
Diisobutyl phthalate (271)  
Ursolic acid (319)  
3β,22β,24-Trihydroxy-olean-12-ene (320)  
Thymidine (449)  
Ergosta-4,6,8(14),22-tetraen-3-one (461)  
3β-Hydroxy-5α,8α-epidioxyergosta-6,22-diene  
(15α)-15-Hydroxysoyasapogenol B (322)   176
(7β,15α)-7,15-Dihydroxysoyasapogenol B (323)  
(7β)-7,29-Dihydroxysoyasapogenol B (324)  
Pestalotiopsis foedan Bruguiera sexangula (−)-(4S,8S)-Foedanolide (236) Modest cytotoxic 124
(+)-(4R,8R)-Foedanolide (237) Modest cytotoxic
Pestalotiopsis heterocornis (L421) Bruguiera gymnorrhiza 7-Hydroxy-5-methoxy-4,6-dimethyl-7-O-a-L-rhamnosyl-phthalide (182)   90
7-Hydroxy-5-methoxy-4,6-dimethyl-7-O-b-D-glucopyranosyl-phthalide (183)  
7-Hydroxy-5-methoxy-4,6-dimethylphthalide (184)  
Pestalotiopsis virgatul Sonneratia caseolaris Pestalotiopyrone I (245)   131
Pestalotiopyrone J (246)  
Pestalotiopyrone K (247)  
Pestalotiopyrone L (248)  
(6S,10S,20S)-Hydroxypestalotin (249)  
Pestalotiopsis JCM2A4 Rhizophora mucronata n-Hexadecanoic acid (264)   137
Elaidic acid (265)
Altiloxin B (308)   157
Pestalotiopen A (336) Moderate antimicrobial
Pestalotiopen B (337)  
Pestalotiopsis sp. PSU-MA69 Rhizophora apiculata Pestalochromone A (60)   48
Pestalochromone B (61)  
Pestalochromone C (62)  
Asperpentyn (100)  
Siccayne (101)  
Pestaloxanthone (115)  
Pestalotethers A (161)  
Pestalotethers B (162)  
Pestalotethers C (163)  
Pestalotether D (164)  
Pestalolide (228) Weak antifungal
Seiridin (229)  
(S)-Penipratynolene (276)  
Anofinic acid (277) DNA-damaging active
Pestalotiopsis sp. PSU-MA92/Pestalotiopsis sp. PSU-MA119 Rhizophora apiculata/Rhizophora mucronata Pestalotiopyrone A (221)   111
Pestalotiopyrone B (222)  
Pestalotiopyrone C (223)  
Pestalotioprolide A (224)  
Pestalotioprolide B (225)  
Seiricuprolide (226)  
2′-Hydroxy-30,40-didehydropenicillide (227)  
Phoma herbarum VB7   Dibutylphthalate (269)   140
Mono(2-ethylhexyl)phthalate (270)  
Phoma sp. SK3RW1M Avicennia marina 1-Hydroxy-8-(hydroxymethyl)-6-methoxy-3-methyl-9H-xanthen-9-one (105)   66
1-Hydroxy-8-(hydroxymethyl)-3-methoxy-6-methyl-9H-xanthen-9-one (106)  
1,8-Dihydroxy-10-methoxy-3-methyldibenzo[b,e]oxepine-6,11-dione (178)  
Phomopsis sp. IM 41-1 Rhizhopora mucronata Phomoxanthone A (121) Moderate antimicrobial 156
12-O-Deacetyl-phomoxanthone A (122) Moderate antimicrobial
Phomopsis sp. (no. SK7RN3G1) Sediment 2,6-Dihydroxy-3-methyl-9-oxoxanthene-8-carboxylic acid methyl ester (117) Modest cytotoxic 69
Lichenxanthone (118)  
Griseoxanthone C (119)  
1,3,6-Trihydroxy-8-methyl-9H-xanthen-9-one (120)  
Phomopsis sp. (ZH76) Excoecaria agallocha 3-O-(6-O-α-L-Arabinopyranosyl)-β-D-glucopyranosyl-1,4-dimethoxyxanthone (116) Modest cytotoxic 68
Phomapyrone D (235)  
2-Methoxy-3,4-methylenedioxybenzophenone  
cyclo(D-6-Hyp-L-Phe)  
Phomopsis sp. (no. ZH-111) Sediment (3R,4S)-3,4-Dihydro-4,5,8-trihydroxy-3-methylisocoumarin (27) Strong SIV accelerator; weak cytotoxic 35
4-(Hydroxymethyl)-7-methoxy-6-methyl-1(3H)-isobenzofuranone (181) SIV inhibitor
Exumolide A (439) Strong SIV accelerator
Phomopsis sp. PSU-MA214 Rhizophora apiculata (2R,3S)-7-Ethyl-1,2,3,4-tetrahydro-2,3,8-trihydroxy-6-methoxy-3-methyl-9,10-anthracenedione (81) Weak cytotoxic 59
Tetrahydroaltersolanol B (82)  
Tetrahydroaltersolanol C (83)  
Ampelanol (84)  
Macrosporin (85)  
1-Hydroxy-3-methoxy-6-methylanthraquinone (86)  
Phenethyl alcohol hydracrylate (274)  
Butanamide (275)  
Phomonitroester (453)  
Phomopsis sp. ZSU-H26 Excoecaria agallocha 5-Hydroxy-6,8-dimethoxy-2-benzyl-4H-naphtho[2,3-b]-pyran-4-one (43) Cytotoxic 43
5,7-Dihydroxy-2-methylbenzopyran-4-one (44)  
3,5-Dihydroxy-2,7-dimethylbenzopyran-4-one (45)  
cyclo(Tyr-Tyr)  
Phomopsis sp. (#zsu-H76) Excoecaria agallocha Phomopsis-H76 A (46) Strong SIV accelerator 44
Phomopsis-H76 B (177)  
Phomopsis-H76 C (450) SIV inhibitor
Sporothrix sp. (no. 4335) Kandelia candel 7-Chloro-2′,5,6-trimethoxy-6′-methylspiro(benzofuran-2(3H),1′-(2)cyclohexene)-3,4′-dione (250)   134
2-Acetyl-7-methoxybenzofuran (256)  
Diaporthin (42)   41
5-Hydroxy-2-methylchromanone (64)  
5-Methoxy-2-methylchromone (65)  
1,8-Dihydroxy-4-methylanthraquinone (102)  
1,8-Dihydroxy-5-methoxy-3-methyl-9H-xanthen-9-one (114)  
3-Methoxy-6-methyl-1,2-benzenediol (173)  
4-Methoxypyrocatechol (174)  
3,5-Dimethylphenol (175)  
1-Hydroxy-8-methoxynaphthalene (176)  
7-Chloro-2′,5,6-trimethoxy-6′-methylspiro[benzofuran-2(3H),1′-(2)cyclohexene]-3,4′-dione (255)  
1,8-Dimethoxynaphthalene (279)  
Methyl 7-methylbenzofuran-2-carboxylate (280)  
Cytochalasin IV (361)  
Sporothrin A  
Sporothrin B  
Sporothrin C  
1,8-Dihydroxy-5-methoxy-3-methyl-9H-xanthen-9-one  
5-Carboxymellein  
Peroxyergosterol  
cyclo(L-Leu-L-Pro)  
cyclo-L-Phenylalanyl-L-alanine  
2,4-Dihydroxypyrimidine  
Talaromyces flavus Sonneratia apetala Talaperoxide A (287)   149
Talaperoxide B (288) Cytotoxic
Talaperoxide C (289)  
Talaperoxide D (290) Cytotoxic
Merulin A (steperoxide B) (282)  
Talaromyces sp. ZH-154 Kandelia candel 7-Epiaustdiol (257) Moderate cytotoxic; significant antimicrobial 135
8-O-Methylepiaustdiol (258) Moderate cytotoxic
Stemphyperylenol  
Skyrin  
Secalonic acid A  
Emodin  
Norlichexanthone  
Xylaria cubensis PSU-MA34 Bruguiera parviflora (R)-(−)-5-Carboxymellein (29)   38
(R)-(−)-5-Methoxycarbonylmellein (30)  
(R)-(−)-Mellein methyl ester (31)  
2-Chloro-5-methoxy-3-methylcyclohexa-2,5-diene-1,4-dione (95)  
Isosclerone (96)  
Xylacinic acid A (260)  
Xylacinic acid B (261)  
2-Hexylidene-3-methyl succinic acid 4-methyl ester (262)  
Cytochalasin D (359)  
Xylaria sp. BL321 Acanthus ilicifolius Isocoumarin (41)   113 and 153
Lactone (230)  
07H239-A (305) Cytotoxic; concentration dependent α-glucosidase activator and inhibitor
Tricyclic lactone (307)  
Cytochalasin C (359)  
Cytochalasin D (359)  
19,20-Epoxycytochalasin C (360)  
Three new eremophilane sesquiterpenes (302304)  
Unidentified fungus A1 Scyphiphora hydrophyllacea R-3-Hydroxyundecanoic acid methyl ester-3-O-α-L-rhamnopyranoside (263) Modest antimicrobial 136
Unidentified fungus no. AMO 3-2 Avicennia marina Farinomalein C (366)   197
Farinomalein D (367)  
Farinomalein E (368)  
Farinomalein (369)  
Farinomalein methyl ester (370)  
(3R)-5,7-Dihydroxy-3-methylisoindolin-1-one (371)  
Unidentified fungal strains K38 and E33 Kandelia candel/Eucheuma muricatum 8-Hydroxy-3-methyl-9-oxo-9H-xanthene-1-carboxylic acid methyl ether (107) Antifungal 37 and 67
6,8-Dihydroxy-4-acetylisocoumarin (28)  
2,5-Hydroxy-6,8-dimethoxy-2,3-dimethyl-4H-naphtho-[2,3-b]-pyran-4-one (59)  
2-Formyl-3,5-dihydroxy-4-methyl-benzoic acid (172)  
Allitol (268)  
7,22-(E)-Diene-3β,5α,6β-triol-ergosta (463)  
Unidentified fungus XG8D Xylocarpus granatum Merulin A (282) Significant cytotoxic 147
Merulin B (283)  
Merulin C (284) Significant cytotoxic; promising antiangiogenic
Merulin D (285)   148
Steperoxide A (286)  
Unidentified fungus Zh6-B1 Sonneratia apetala 3,4-Dihydro-4,8-dihydroxy-7-(2-hydroxyethyl)-6-methoxy-1(2H)-naphthalen-1-one (68)   51
10-Norparvulenone (69) Anti-influenza
3R,5R-Sonnerlactone Cytotoxic
3R,5S-Sonnerlactone  
Unidentified fungus no. ZSU-H16 Avicennia 3,5,8-Trihydroxy-2,2-dimethyl-3,4,4-trihydro-2H,6H-pyrano[3,2-b]-xanthen-6-one (103)   65
5,8-Dihydroxy-2,2-dimethyl-2H,6H-pyrano[3,2-b]xanthen-6-one (104)  
cyclo-(N-O-Methyl-L-Trp-L-Ile-D-Pip-L-2-amino-8-oxo-decanoyl) (439)  
cyclo-(Phe-Tyr)  
Unidentified fungi nos K38 and E33   Ethyl 5-ethoxy-2-formyl-3-hydroxy-4-methylbenzoate (165) Weak antifungal 84
Unidentified two co-cultured fungi   Marinamide (383) Potent cytotoxic 204
Marinamide methyl ester (384) Potent cytotoxic


4.1. Polyketides

4.1.1 Coumarins. Only one previously unknown furanocoumarin, named 5-methyl-8-(3-methylbut-2-enyl)furanocoumarin (23), together with three known metabolites, bergapten (24), scopoletin (25), and umbelliferone (26), were identified from mangrove sources from 2011–2013. These metabolites were isolated from Penicillium sp. ZH16, which is an endophytic fungus from the mangrove tree Avicennia (Hainan, China). Compound 23 showed modest cytotoxic activity against the human tumor cell line KB and KBV200 with IC50 values of 5 and 10 μg mL−1, respectively.33
image file: c4ra11756e-u4.tif
4.1.2 Isocoumarin derivatives. Naturally occurring isocoumarin and 3,4-dihydroisocoumarin derivatives abound in a wide variety of microbial sources and have been shown to possess an impressive array of biological activities. Many of these are known to have substituents at C-3.34 A new isochroman, (3R,4S)-3,4-dihydro-4,5,8-trihydroxy-3-methylisocoumarin (27), isolated from mangrove sediment was characterized from Phomopsis sp. (no. ZH-111) (ZhuHai, China). This metabolite was tested on zebrafish embryo collections in a primary antibacterial assay wherein it was shown to accelerate the growth of blood vessel markedly. Furthermore, it showed weak cytotoxicity against Hep-2 and HepG2 cells.35 In mixed fermentation, co-cultured microbes may activate the silent genes related to metabolite biosynthesis for new biomolecules, which cannot be otherwise detected in individual strain cultures. This approach can be used to fully exploit the metabolic potential of cultivable microbes.36 Application of mixed fermentation technique on unidentified mangrove fungal strains Kandelia candel endophytic K38 and Eucheuma muricatum endophytic E33 yielded a representative 4-substituted 3-nonsubstituted known isocoumarin, 6,8-dihydroxy-4-acetylisocoumarin (28) (Zhanjiang, South China Sea, China).37 Three previously known mellein derivatives (29–31) were isolated from Xylaria cubensis PSU-MA34, an endophyte obtained from a branch of Bruguiera parviflora (Suratthani, Thailand). Even the crude extract from the culture broth exhibited cytotoxic activity against KB cells derived from the epidermoid carcinoma of the oral cavity, with an IC50 value of 2.79 μg mL−1. Moreover, the crude extract showed antibacterial activity against Staphylococcus aureus ATCC 25923 and methicillin-resistant Staphylococcus aureus (MRSA) with equal MIC values of 200 μg mL−1. None of 29–31 exhibited either cytotoxic or antibacterial activities.38 Two new dihydroisocoumarin derivatives, aspergillumarins A (32) and B (33), were isolated from the liquid culture of Aspergillus sp., an endophytic fungus isolated from the leaf of Chinese mangrove plant Bruguiera gymnorrhiza (South China Sea, China). Both compounds exhibited a weak antibacterial activity against Staphylococcus aureus and Bacillus subtilis at a concentration of 50 μg mL−1.39 Chromatographic analysis of Rhizophora apiculata (Satun, Thailand) endophytic Acremonium sp. PSU-MA70 yielded seven new analogues, acremonones B–H (34–40).40 A previously unknown isocoumarin (41), bearing an exocyclic double bond, was identified from an endophytic fungus Xylaria sp. BL321, present in Acanthus ilicifolius L. (Guangdong, South China Sea, China).153 Another two known analogues, phytotoxic diaporthin (42) and 5-carboxymellein,8 were also found to occur in a fungal endophyte Sporothrix sp. (#4335) of Kandelia candel (South China Sea, China). Compound 42 was found to be toxic to tumor cell line HepG2, giving an IC50 value of 23 μM.41
image file: c4ra11756e-u5.tif
4.1.3 Chromones. Naphthopyrones combining naphthalene and a γ-pyrone moiety exist as naturally occurring scaffolds with broad biological activity ranging from antimicrobial, antiviral, insecticidal and anti-estrogenic activity.42 A new cytotoxic naphtho-γ-pyrone,5-hydroxy-6,8-dimethoxy-2-benzyl-4H-naphtho[2,3-b]-pyran-4-one (43), was characterized from the fungal strain Phomopsis sp. ZSU-H26 residing in Excoecaria agallocha (DongZhai, Hainan, China). Two other well-known metabolites 5,7-dihydroxy-2-methylbenzopyran-4-one (44) and 3,5-dihydroxy-2,7-dimethylbenzopyran-4-one (45) were also discovered from Phomopsis sp. ZSU-H26 alongside the above described compound. Compound 43 showed cytotoxic activity against Hep-2 and HepG2 cells, with IC50 values of 10 and 8 μg mL−1, respectively.43 Chemical investigation of another Excoecaria agallocha (DongZhai, Hainan, China) endophytic fungal strain Phomopsis sp. (#zsu-H76) yielded a unique dimer metabolite, phomopsis-H76 A (46), which could accelerate the growth of subintestinal vessel plexus (SIV) branch.44 “One strain many compounds” (OSMAC) is an effective approach to activate the biosynthetic pathway of microorganisms45 and has been successfully applied towards the discovery of new metabolites from Aspergillus tubingensis (GX1-5E). A. tubingensis (GX1-5E) obtained from the radix of Pongamia pinnata (Guangxi, China), yielded seven dimeric naphtho-γ-pyrones, rubasperone A–G (47–53), which co-exist with their biosynthetically related monomers, rubrofusarin (54), rubrofusarin B (55), TMC 256 A1 (56), fonsecin (57), and flavasperone (58) by solid-substrate fermentation cultures. The absolute configurations of compounds 47 and 48 were determined by X-ray crystallography. Compounds 52/53 were obtained as an inseparable mixture of two slowly equilibrating atropisomers, which could be distinguished by NMR spectroscopy. The hindered internal rotation allows distinguishing these compounds, thus, indicating a non-enzymatic biosynthesis or a particularly facile mode of racemization. Among these metabolites, compound 54 displayed moderate tyrosinase inhibitory activity, with an IC50 value of 65.6 μM, and compound 55 exhibited mild α-glucosidase inhibitory activity, with an IC50 value of 97.3 μM. Compound 56 was found to be toxic towards a small panel of tumor cell lines such as MCF-7, MDA-MB-435, Hep3B, Huh7, SNB19, and U87 MG, with IC50 values ranging from 19.92 to 47.98 μM while compounds 50, 55, and 58 showed mild cytotoxicities.46,47 Further investigation of the aforementioned mixed culture of the two unidentified fungal strains K38 and E33 isolated from mangrove plants Kandelia candel endophytic and Eucheuma muricatum, yielded another previously known naphtho-γ-pyrone,5-hydroxy-6,8-dimethoxy-2,3-dimethyl-4H-naphtho-[2,3-b]-pyran-4-one (59).37 Nonchlorinated chromones are common metabolites encountered in mangrove fungi.8 Three previously unknown chlorinated tetrahydrochromanone derivatives, pestalochromones A–C (60–62), were recently identified in a mangrove plant Rhizophora apiculata-derived fungus Pestalotiopsis sp. PSU-MA69 (Satun, Thailand). The biosynthetic pathway for compounds 60–62 is proposed as shown in Scheme 4.48 Several ubiquitous microbial chromone derivatives were also isolated, including 5-hydroxy-3-hydroxymethyl-7-methoxy-2-methyl-4H-1-benzopyran-4-one (63) from Penicillium sp. ZH58 obtained from the leaves of the mangrove tree Avicennia (Hainan, China);49 and 5-hydroxy-2-methylchromanone (64) and 5-methoxy-2-methylchromanone (65) from Sporothrix sp. (#4335) of Kandelia candel.41
image file: c4ra11756e-u6.tif

image file: c4ra11756e-s4.tif
Scheme 4 Proposed pathways of compounds 100–101 to compounds 60–62.
4.1.4 Flavonoids. An isoflavone, 5-hydroxy-7-methoxy-4′-O-(3-methylbut-2-enyl)isoflavone (66) and a structurally related known vasodilatory agent eriodictyol (67) were identified from the cultures of Fusarium sp. ZZF60 inhabiting the leaves of Kandelia candel (Hainan, China).50 Huang et al. (2012) reported these metabolites for the first time in mangrove microbes.
image file: c4ra11756e-u7.tif
4.1.5 Quinones. The widespread occurrence and function of quinone and semiquinone derivatives in mangrove-derived fungi has been well established. These compounds have anticancer, antibacterial, antimalarial and fungicidal properties. Two previously reported semiquinones, 3,4-dihydro-4,8-dihydroxy-7-(2-hydroxyethyl)-6-methoxy-1(2H)-naphthalen-1-one (68) and an anti-influenza virus antibiotic 10-norparvulenone (69) were isolated from an unidentified plant endophyte Zh6-B1, obtained from the bark of Sonneratia apetala (Guangdong, China).51 Three newly identified bianthraquinone derivatives, alterporriols K–M (70–72), and four co-occurring anthraquinones (73–76) were obtained from extracts of the endophytic fungus Alternaria sp. ZJ9-6B isolated from Aegiceras corniculatum (Guangdong, China). Compounds 70–72 were the first isolated alterporriols featuring a rarely encountered a C-2–C-2′ linkage. Compounds 70 and 71 were found to be moderately toxic towards MDA-MB-435 and MCF-7 cells with IC50 values ranging from 13.1 to 29.1 μM.52 The anticancer mechanism of compound 71 towards breast cancer cells lines was estimated, and it is known to play a vital role in breast cancer cells by destroying the mitochondrial membrane.53 Early studies on Kandelia candel endophytic fungus Nigrospora sp. no. 1403 revealed the presence of five cytotoxic anthracenediones.54 Further investigation of this fungus yielded an additional multi-active substance deoxybostrycin (77), which is a known phytotoxin, also an antibacterial agent bostrycin 1403P-3 and a deoxy-analogue 1403P-2 nigrosporin. Bioassay indicated that 77 could significantly suppress the growth of six human tumor cell lines: A549, Hep-2, HepG2, KB, MCF-7, and MCF-7/Adr, with IC50 values of 2.44, 3.15, 4.41, 3.15, 4.76, and 5.46 μg mL−1, respectively. Moreover, its 21 synthetic derivatives also exhibited strong to moderate cytotoxicity while some displayed remarkable activity against MDA-MB-435, which was comparable with the positive control epirubicin.55 Further tests disclosed its strong antimicrobial activities against Staphylococcus aureus, Escherichia coli, Pseudomonas aeruginosa, Sarcina ventriculi, Bacillus subtilis with an IC50 of 3.13 μg mL−1; it also inhibited Candida albicans with an IC50 of 12.5 μg mL−1.56 Moreover, 77 exhibited remarkable inhibitory effect on mycobacteria, especially Mycobacterium tuberculosis, and this inhibitory effect was much more effective than some of the first line anti-tuberculosis (TB) drugs against clinical multidrug-resistant (MDR) M. tuberculosis strains.57 A new cochlioquinone derivative, 2,3-didehydro-19α-hydroxy-14-epicochlioquinone B (78) and a known hydronaphthalenone derivative 4-deoxytetrahydrobostrycin (79) were characterized from another endophytic Nigrospora sp. strain no. MA75 obtained from the semi-mangrove plant Pongamia pinnata in different culture media. This compound (78) displayed remarkable antibacterial activity toward MRSA, E. coli, P. aeruginosa, P. fluorescens, and S. epidermidis, with MIC values of 8, 4, 4, 0.5, and 0.5 μg mL−1, respectively. However, compound 79 exhibited moderate and somewhat selective inhibition of MRSA, E. coli, and S. epidermidis. Moreover, compound 78 potently inhibited the growth of MCF-7, SW1990, and SMMC7721 tumor cell lines with IC50 value of 5 μg mL−1.58 A known phytochemical sterequinone C (80) was characterized from Penicillium sp. ZH16 from the leaves of Avicennia.33 The leaves of Rhizophora apiculata (Songkhla, Thailand) harbor the endophytic fungus, Phomopsis sp. PSU-MA214 from which a new tetrahydroanthraquinone derivative, (2R,3S)-7-ethyl-1,2,3,4-tetrahydro-2,3,8-trihydroxy-6-methoxy-3-methyl-9,10-anthracenedione (81), and five other known anthraquinones (82–86), were characterized. Compound 81 is a rare ethyltetrahydroanthraquinone that shows weak cytotoxicity against breast cancer MCF-7 cell lines. Furthermore, it shows antibacterial activity against the standard Staphylococcus aureus ATCC 25923 and methicillin-resistant S. aureus SK1. The proposed biosynthetic pathway of these anthraquinones (81–86) starts from an octaketide precursor, which is obtained by the condensation reaction of acetate and malonate units as shown in Scheme 5.59 A chemical investigation of the ethyl acetate extract of the fermentation broth of Alternaria tenuissima EN-192, an endophytic fungus obtained from the stems of Rhizophora stylosa, resulted in the isolation of three known tricycloalternarene metabolites (87–89). In disk diffusion assays, compound 87 was shown have moderate antibacterial activity against the aquaculture pathogenic bacterium Vibrio anguillarum.60 A rare thiophene compound, named 8-hydroxy-2-[1-hydroxyethyl]-5,7-dimethoxynaphtho[2,3-b]thiophene-4,9-dione (90), and four previously described metabolites anhydrojavanicin (91), 8-O-methyljavanicin (92), botryosphaerone D (93), 6-ethyl-5,8-dihydroxy-3,7-dimethoxynaphthoquinone (94) were characterized from the liquid culture of Aspergillus terreus (no. GX7-3B), an endophyte present in a branch of Bruguiera gymnoihiza (Linn.) Savigny (Guangxi, China). Compound 91 displayed remarkable inhibition against α-acetylcholinesterase (AChE) with IC50 values 2.01 μM in bioactivity assays in vitro.61
image file: c4ra11756e-s5.tif
Scheme 5 A plausible biosynthetic pathway for compounds 81–86.

Several known compounds (96–102) were also isolated from other endophytes associated with various plants.38,41,48,62,63 Amongst, compound 98 exhibited significant inhibitory effect against a panel of human protein kinases, such as IGF1-R and VEGF-R2 with IC50 values mostly in the low micromolar range.62

image file: c4ra11756e-u8.tif

4.1.6 Xanthones. Xanthones are usually polysubstituted and occur as either fully aromatized, dihydro-, tetrahydro-, or, more rarely, hexahydroderivatives. This family of compounds appeals to medicinal chemists because of its members' pronounced biological activity within a notably broad spectrum of disease profiles, a result of their interactions with a correspondingly diverse range of target biomolecules. Studies on structure–activity relationships of xanthone derivatives were reviewed comprehensively.64 A new xanthone derivative,3,5,8-trihydroxy-2,2-dimethyl-3,4,4-trihydro-2H,6H-pyrano[3,2-b]-xanthen-6-one (103), together with its analogue,5,8-dihydroxy-2,2-dimethyl-2H,6H-pyrano[3,2-b]xanthen-6-one (104), were found to be produced by unidentified endophytic fungus no. ZSU-H16, which was separated from the leaves of mangrove avicennia from the South China Sea coast. Compound 103 was shown to be moderate cytotoxic against KB and KBV200 cells.65 On the other hand, two new xanthones, named 1-hydroxy-8-(hydroxymethyl)-6-methoxy-3-methyl-9H-xanthen-9-one (105) and 1-hydroxy-8-(hydroxymethyl)-3-methoxy-6-methyl-9H-xanthen-9-one (106), were isolated from endophytic Phoma sp. SK3RW1M associated with the roots of Avicennia marina (Guangxi, China). This is the first report on xanthone derivatives isolated as metabolites from Phoma species.66 The same collection of a two-endophytic fungi association of unidentified mangrove fungal strains Kandelia candel endophytic K38 and Eucheuma muricatum endophytic E33 yielded an additional new xanthone antibiotic, 8-hydroxy-3-methyl-9-oxo-9H-xanthene-1-carboxylic acid methyl ester (107). Biological investigation demonstrated that compound 107 has broad antifungal activity against some plant pathogens Gloeosporium musae, Blumeria graminearum, Fusarium oxysporum, Peronophthora cichoralearum and Colletotrichum glocosporioides.67 Bioassay-guided purification of another endophytic Nigrospora sp. strain no. MA75, obtained from the semi-mangrove plant Pongamia pinnata on various culture media, resulted in the characterization of three previously known xanthone derivatives (108–110). Compound 108 exhibited moderate and somewhat selective inhibition of human tumor cell HepG2, whereas compound 110 was shown to be broad-spectrum antibacterial toward MRSA, E. coli, P. aeruginosa, P. fluorescens, and S. epidermidis, with MIC values of 0.5, 2, 0.5, 0.5, and 16 μg mL−1, respectively.58 Four related known derivatives were obtained from the chemical investigation of Kandelia candel endophytic Fusarium sp. ZZF60 (ref. 50) and Sporothrix sp. (#4335),41 and an endophytic Penicillium sp. ZH16 of a mangrove tree Avicennia.33 These were: 3,6,7-trihydroxy-1-methoxyxanthone (111), 1,3,6-trihydroxy-8-methylxanthone (112) 1,7-dihydroxyxanthone (113), and 1,8-dihydroxy-5-methoxy-3-methyl-9H-xanthen-9-one (114). A new xanthone, pestaloxanthone (115) was characterized from Pestalotiopsis sp. PSU-MA69, an endophytic fungus living in the interior part of Rhizophora apiculata branch.48 A new xanthone O-glycoside, 3-O-(6-O-α-L-arabinopyranosyl)-β-D-glucopyranosyl-1,4-dimethoxyxanthone (116) was discovered in the endophyte Phomopsis sp. (ZH76) isolated from the stem of Excoecaria agallocha (Hainan, China). This compound showed modest cytotoxicity against HEp-2 cells with IC50 of 9 μg mL−1 and against HepG2 cells with IC50 of 16 μg mL−1.68 A new xanthone, named 2,6-dihydroxy-3-methyl-9-oxoxanthene-8-carboxylic acid methyl ester (117), alongside three known compounds lichenxanthone (118), griseoxanthone C (119) and 1,3,6-trihydroxy-8-methyl-9H-xanthen-9-one (120) were characterized from Phomopsis sp. (no. SK7RN3G1), isolated from the mangrove sediment. Compound 117 displayed modest in vitro inhibition of the proliferation of HEp-2 and HepG2 cells lines, giving IC50 values of 8 and 9 μg mL−1, respectively.69 Xanthone dimers possess increasingly complex and interesting structures and in many cases have very specific and selective biological properties. Two dimeric tetrahydroxanthones, a known phomoxanthone A (121) and a structurally related new 12-O-deacetyl-phomoxanthone A (122) were obtained from the rice culture of Phomopsis sp. IM 41-1 separated from a mangrove plant Rhizhopora mucronata (Jakarta, Indonesia). Both of these compounds exhibited moderate antimicrobial activities against Botrytis cinerea, Sclerotinia sclerotiorum, Diaporthe medusaea, and Staphylococcus aureus. Furthermore, they are implicated in the protection of the host plant from degradation and disease caused by pathogens. However, in contrast to previous reports,70,71 the acetylated moiety did not have any significant effect on biological activity. Moreover, the differences in the antimicrobial activities between 121 and 122 were small.72
image file: c4ra11756e-u9.tif
4.1.7 Phenol and phenolic acids. Over 70 natural phenolic acids produced by mangrove fungi were identified during 2011–2013, and since then, their number has steadily increased. Two new benzofuranoids, deuteromycols A and B (123, 124), were isolated from the fermentation broth of Deuteromycete sp., which was isolated by micromanipulation from the mangrove drift wood (Red Sea, El Gouna, Egypt). Although the extract from which it was isolated exhibited in vitro antibacterial activity toward pathogenic strains such as Staphylococcus aureus (ATCC 6538), Bacillus subtilis (ATCC 6051), Escherichia coli (ATCC 11229), Pseudomonas aeruginosa (ATCC 22853) and multiresistant strains Staphylococcus aureus (Northern Germany Epidemic Strain, NGES), Staphylococcus epidermidis (LHI, no. 847) and Staphylococcus haemolyticus (LHI, no. 535), the pure natural product has not yet been evaluated.73

Three new p-terphenyls, 6′-O-desmethylterphenyllin (125), 3-hydroxy-6′-O-desmethylterphenyllin (126), 3′′-deoxy-6′-O-desmethylcandidusin B (127) along with two known p-terphenyls (128, 129) were isolated from Penicillium chermesinum (ZH4-E2), which is endophytic to Kandelia candel (South China Sea, Hainan, China). All of these compounds showed enzyme inhibitory activities. Compounds 125–127 showed strong inhibitory effects against R-glucosidase with IC50 values of 0.9, 4.9, and 2.5 μM, respectively, whereas compounds 128 and 129 showed inhibitory activity toward acetylcholinesterase, with IC50 values of 7.8 and 5.2 μM.53

Polyhydroxy-p-terphenyl-type chemical entities, mainly found in mycomycetes, include scaffolds with a C-18 tricyclic or polycyclic C-18 skeleton and those exhibiting alkylated (methylated or prenylated) side chains.74 So far, less than 20 polyhydroxy-p-terphenyl metabolites have been isolated from lower fungi, with the majority reported from Aspergillus section Candidi.75 Members of this family of natural products include six new prenylated polyhydroxy-p-terphenyl metabolites, named prenylterphenyllins A–C (130–132) and prenylcandidusins A–C (133–135), and one new polyhydroxy-p-terphenyl compound with a simple tricyclic C-18 skeleton, named 4′′-dehydro-3-hydroxyterphenyllin (136). All of these metabolites were obtained together with eight known analogues (137–144) from Aspergillus taichungensis ZHN-7-07, a root soil fungus isolated from Acrostichum aureum. These compounds were evaluated for cytotoxic activity in vitro against HL-60, A-549, and P-388 cell lines. It was found that compounds 130 and 137 exhibited moderate activities against all three cell lines, with IC50 values ranging from 1.53–10.90 μM, whereas compounds 134 and 136 displayed moderate activities only against the P-388 cell line, with IC50 values of 1.57 and 2.70 μM, respectively.76 Simple halogenated aromatic metabolites are very common in mangrove fungi, thus, reflecting the availability of chloride and bromide ions in seawater.77 Methyl 3-chloro-6-hydroxy-2-(4-hydroxy-2-methoxy-6-methylphenoxy)-4-methoxybenzoate (145) was characterized56 from Nigrospora sp. no. 1403, isolated from Kandelia candel.

Two new diphenyl ethers, namely, 7-O-acetylsecopenicillide C (146), and hydroxytenellic acid B (147), along with two related derivatives (148, 149) were characterized from the shaken culture of Penicillium sp. MA-37, which was obtained from the rhizospheric soil of the mangrove plant Bruguiera gymnorrhiza (South China Sea, Hainan, China). The absolute configuration of 147 was determined by the modified Mosher's method. None of these diphenyl ethers exhibited either brine shrimp lethality or antibacterial activity.78 Epigenetic modification was used to stimulate the biological profile of cytosporones, which induce the generation of additional derivatives. A previously undescribed cytosporone R (150) was isolated from the culture medium of Leucostoma persoonii, an endophytic fungus obtained from Rhizophora mangle (Florida Everglades, USA). The medium was supplemented with histone deacetylase inhibitor (HDAC), sodium butyrate, and a DNA methyltransferase (DNMT) inhibitor, 5-azacytidine to obtain the cytosporone.79

Apart from previously reported expansols A and B, diorcinol, and S-(+)-sydonic acid, four newly discovered polyphenols, expansols C–F (151–154), and one new diphenyl ether derivative, 3-O-methyldiorcinol (155), as well as a series of known compounds, (+)-(7S)-7-O-methylsydonic acid (156), 3,7-dihydroxy-1,9-dimethyldibenzofuran (157), orcinol (158), 2,4-dimethoxyphenol (159), and 4-hydroxybenzoic acid (160), were isolated from the refermentation broth of Penicillium expansum 091006 endogenous with Excoecaria agallocha. Compounds 151–154 are a group of natural polyphenols featuring a rarely encountered phenolic bisabolane sesquiterpenoid and diphenyl ether moieties of fungal origin. Except 151 and 153, all compounds showed weak cytotoxicity against HL-60 cell lines, and they showed this activity at IC50 values of 18.2 and 20.8 μM, respectively. The co-isolation of known precursors, expansols A and B, further supports the proposed biogenetic pathway for compounds 151–156 as illustrated in Scheme 6.80 Chlorinated diphenyl ethers are microbially metabolized readily,81 but they rarely occur as fungal metabolites. Culture of Pestalotiopsis sp. PSU-MA69, an endophytic fungus from Rhizophora apiculata, gave four previously unknown diphenyl ethers, pestalotethers A–D (161–164), together with their biosynthetic precursors, pestheic acid, chloroisosulochrin, chloroisosulochrin dehydrate, isosulochrin and isosulochrindehydrate from P. theae, and pestaloxanthone (115).82,83 In addition to ring-generating biosynthetic cyclizations, some oxidation and hydrolysis and subsequent adornment reactions (e.g. esterification/decarboxylation/methanolysis) with xanthone precursors were found to be involved in the biosynthesis of the individual diphenyl ethers 161–164. These processes are shown in Scheme 7.48 A new polysubstituted benzaldehyde derivative, known as ethyl 5-ethoxy-2-formyl-3-hydroxy-4-methylbenzoate (165) was discovered using a mixed fermentation technique on unidentified mangrove endophytic fungal strains nos K38 and E33.84 It was found to exhibit weak antifungal activity against Fusarium graminearum, Gloeosporium musae, Rhizoctonia solani Kuhn, and Phytophthora sojae.57 An antimicrobial metabolite auroglaucin (166), previously characterized from Aspergillus glaucus,85 was isolated again from the mangrove rhizosphere soil derived fungus, Aspergillus effuses H1-1.86


image file: c4ra11756e-s6.tif
Scheme 6 Postulated metabolic relationship between compounds 151–156.

image file: c4ra11756e-s7.tif
Scheme 7 A plausible biosynthetic pathway for the generation of compounds 161–164.

A number of known antifungal phenolic acids, such as vittarin-B (167), 3,5-dimethoxybiphenyl (168), 3-hydroxy-4-methoxystyrene (169), 3,4-dihydroxyphenylethanol (170), p-tyrosol (171), 2-formyl-3,5-dihydroxy-4-methyl-benzoic acid (172), 3-methoxy-6-methyl-1,2-benzenediol (173), 4-methoxypyrocatechol (174), 3,5-dimethylphenol (175), and 1-hydroxy-8-methoxynaphthalene (176), were first reported to be produced by different mangrove endophytes, as Kandelia candel endophytic Sporothrix sp. (no. 4335)41 and Fusarium sp. ZZF60,50 Avicennia endophytic Penicillium sp. ZH16,33 Bruguiera sexangula endophytic Pestalotiopsis clavispora,87 and two unidentified fungal strains K38 and E33 from Kandelia candel and Eucheuma muricatum.37

image file: c4ra11756e-u10.tif

4.1.8 Lactones. Lactones continue to be a prominent feature of mangrove-derived secondary metabolism. A unique δ-lactone dimer, phomopsis-H76 B (177) was obtained from the liquid culture of Phomopsis sp. (#zsu-H76) isolated from Excoecaria agallocha.44 A new dibenzoxepin derivative, 1,8-dihydroxy-10-methoxy-3-methyldibenzo[b,e]oxepine-6,11-dione (178)66 was identified from the fungus Phoma sp. SK3RW1M isolated from Avicennia marina. This dibenzoxepin derivative was of interest as it had an aromatic lactone ring framework, which was structurally related to that of possible biosynthetic intermediates in the synthesis of flavones by Aspergillus variecolor.88 Paeciloxocins A and B (179, 180), the two new depsidones, were isolated from an endophytic fungus Paecilomyces sp. (Taiwan, China). Interestingly, it was reported that compound 179 exhibits strong cytotoxicity effects against the growth of hepG2 cell line at 1 μg mL−1 and mildly antimicrobial activity against Curvularia lunata (Walker) Boedijn and Candida albicans ATCC 10231.89 A known isobenzofuran derivative, 4-(hydroxymethyl)-7-methoxy-6-methyl-1(3H)-isobenzofuranone (181), was found to be produced by Phomopsis sp. (no. ZH-111) isolated from the mangrove sediment, which also inhibited the growth of subintestinal vessel plexus (SIV) branches.35 Two new phthalide glycosides, 7-hydroxy-5-methoxy-4,6-dimethyl-7-O-a-L-rhamnosyl-phthalide (182) and 7-hydroxy-5-methoxy-4,6-dimethyl-7-O-b-D-glucopyranosyl-phthalide (183), were isolated from the extract of fermentation broth of an endophyte Pestalotiopsis heterocornis (L421) found in the stem of Bruguiera gymnorrhiza (Hainan, China). The known related aglycon 7-hydroxy-5-methoxy-4,6-dimethylphthalide (184) was also isolated with the above mentioned compounds 182 and 183.90

Azaphilones are a structurally diverse family of natural products containing a highly oxygenated bicyclic core and a quaternary center. These molecules exhibit a wide range of significant biological activities, such as inhibitions of monoamine oxidase,91 Gp120–CD4 binding,92 Grb2–SH2 interaction,93 MDM2–p53 interaction94 and heat shock protein 90 (Hsp90),95 as well as antimicrobial, antiviral, cytotoxic, anticancer, and anti-inflammatory activities.96,97 Azaphilones have interesting structural features and biological properties; therefore, a number of synthetic efforts concerning these compounds have been reported during the last four decades.98–101 Two new members of this product type, named chermesinones B and C (185, 186), were identified from Penicillium chermesinum (ZH4-E2) obtained from Kandelia candel. Neither of the isolates showed inhibitory effects against α-glucosidase or acetylcholinesterase.102 Eight newly identified α-pyrone derivatives, namely, nigerapyrones A–H (187–194), coexisting with two known congeners, asnipyrones A (195) and B (196), were isolated from Aspergillus niger MA-132, an endophytic fungus obtained from the fresh tissue of Avicennia marina (Hainan, China). All metabolites were examined by cytotoxicity and antimicrobial bioassays. Nigerapyrone E (191) exerted cytotoxicities against SW1990, MDA-MB-231, and A549 cell lines with IC50 values of 38, 48, and 43 μM, respectively, which were stronger than that of the positive control fluorouracil. It also showed weak or moderate activity against MCF-7, HepG2, Du145, NCI-H460, and MDA-MB-231 cell lines with IC50 values of 105, 86, 86, 43, and 48 μM, respectively. Among these, compound 188 showed selective activity against the HepG2 while 196 showed activity against A549, both with an IC50 of 62 μM. Compound 190 showed moderate or weak activity against the MCF-7, HepG2, and A549 cell lines, with IC50 values of 121, 81, and 81 μM, respectively. However, none of the metabolites showed significant antimicrobial activity.103

A new butyrolactone, 7′′-hydroxybutyrolactone III (197) and its desoxidation precursor butyrolactone I (198), isolated from the culture of Aspergillus terreus A8-4 obtained from mangrove-associated sediments, possessed weak cytotoxicity in vitro against HCT-8, Bel-7402, BGC-823, A2780 cell lines. Butyrolactone I 198 was reported as a specific inhibitor of eukaryotic cyclin-dependent kinases (CDK).104 The putative biosynthetic transformations between these metabolites and conversions can be deduced to the epoxidation of the double bond of 198 on the prenyl side chain to give butyrolactone III, followed by the additional hydroxylation to afford compound 197, as shown in Scheme 8.105


image file: c4ra11756e-s8.tif
Scheme 8 A putative biotransformation of precursor compound 198 to 197.

A comprehensive analysis of Corynespora cassiicola endophyte from Laguncularia racemosa, led to the discovery of four new metabolites, including three decalactones, xestodecalactones D–F (199–201), a new depsidone corynesidone C (202), as well as two known analogues, corynesidones A (203) and B (204). Absolute configurations of the optically active compounds 199–201 were determined by TDDFT ECD (Time-Dependent Density Functional Theory, Electronic Circular Dichroism) calculations of their solution conformers. Compound 199 was separated as a mixture of two diastereomers. All compounds were tested against a panel of human protein kinases. Amongst these, compound 204 could inhibit several kinases such as IGF1-R and VEGF-R2 with IC50 values mostly in the low micromolar range. Furthermore, compound 204 inhibited PIM1 with an IC50 value of 3.5 × 10−7 M, suggesting a tenfold higher specificity of this naturally occurring inhibitor against this particular protein kinase in comparison to most of the other kinases investigated.62 Later, a detailed analysis of the minor metabolites obtained from the cultivation of this fungus yielded a series of unusual octalactones, coryoctalactones A–E (205–209). It is interesting to note that the absolute configuration of the side chain in 205–207 and 209 were tentatively deduced based on biogenetic consideration in comparison with xestodecalactones as shown in Scheme 9. All isolated compounds were evaluated for their antimicrobial, cytotoxic, and antitrypanosomal activities, and were active in all of the bioassays performed.106


image file: c4ra11756e-s9.tif
Scheme 9 Proposed biosynthetic pathway for isolated octalactones (205–209).

The endophytic fungus Eurotium rubrum has been isolated from a semi-mangrove plant Hibiscus tiliaceus and produced a new anthraquinone, 9-dehydroxyeurotinone (210) and 2-O-methyl-9-dehydroxyeurotinone (211). Compound 210 showed weak antibacterial activity against Escherichia coli and modest cytotoxic activity against human tumor cell line SW1990.63

A novel diphenyl ether derivative, namely, Δ1′,3′-1′-dehydroxypenicillide (212), was purified from the extracts of Penicillium sp. MA-37, which was separated from the rhizospheric soil of the mangrove plant Bruguiera gymnorrhiza. Furthermore, three related known compounds, penicillide (213), dehydroisopenicillide (214), and 3′-O-methyldehydroisopenicillide (215) were isolated in addition to the compound 212. All isolated compounds were evaluated in the brine shrimp lethality assay and compound 213 and 214 were active, with LD50 values of 135.9 and 160.0 μM, respectively.78

Fermentation of an Acremonium sp. PSU-MA70, an endophyte from a branch of Rhizophora apiculata, yielded a new phthalide derivative acremonide (216) and a new depsidone acremonone A (217), together with two previously known metabolites, (+)-brefeldin A (218) and 5,7-dimethoxy-3,4-dimethyl-3-hydroxyphthalide (219). The antiviral antibiotic brefeldin A (BFA) (218) could strongly inhibit the protein secretion such as the Golgi membrane enzyme that catalyzes the exchange of guanine nucleotide bound to ribosylation factor (ARF).107 BFA was found to display a weak antifungal activity against Candida albicans NCPF3153.40

Octaketides such as cytosporones are common metabolites found to be produced by three endophytic fungal strains: Aegiceras corniculatum endophytic Dothiorella sp. HTF3, Rhizophora mucronata endophytic Pestalotiopsis sp., and Excoecaria agallocha endophytic Phomopsis sp. ZSU-H76.8 Epigenetic tailoring of Rhizophora mangle endophytic to Leucostoma persoonii, led to the reisolation of a strongly antibacterial trihydroxybenzene cytosporone E (220). This compound 220 had previously been isolated by Brady et al. from Cytospora sp. CR200 and Diaporthe sp. CR146 endophytic in plants Conocarpus erecta and Forsteronia spicata.108 This compound displayed anti-infective activity at an IC90 of 13 μM toward Plasmodium falciparum, with A549 cytotoxicity IC90 of 437 μM. Thus, it represented a 90% inhibition therapeutic index (TI90 = IC90A459/IC90 P. falciparum) of 33. Moreover, it was active against MRSA with a minimal inhibitory concentration (MIC) value of 72 μM; the inhibition of MRSA biofilm at roughly half that value, minimum biofilm eradication counts, MBEC90, was found to be 39 μM.79 Its broad-spectrum biological activity has attracted interest to synthesize this compound. Efficient stereoselective synthesis of the natural enantiomer of this compound was performed by different groups. An overall six-step process starting from 3,4,5-trimethoxybenzoic acid with a yield of 41% is summarized in Scheme 10.109 However, in contrast to previous reports, (±)-220, (R)-220 and (S)-220 were all found to have weak antimicrobial activities which were similar in all three compounds.110 Three previously unknown α-pyrones, pestalotiopyrones A–C (221–223), together with two previously unknown seiricuprolides, pestalotioprolides A (224) and B (225), and two known compounds, seiricuprolide (226) and 2′-hydroxy-3′,4′-didehydropenicillide (227), were characterized from two mangrove-derived fungal strains, Pestalotiopsis sp. PSU-MA92 endophytic to Rhizophora apiculata (Trang, Thailand) and Pestalotiopsis sp. PSU-MA119 endophytic to Rhizophora mucronata (Satun, Thailand).111 Compounds 221–223 have different structures but they are all called pestalotiopyrones, as described in a previous report.112 Compound 222 was the only compound obtained in amounts sufficient for antibacterial testing against Staphylococcus aureus ATCC 25923 and the methicillin resistant S. aureus clinical isolate. It was also tested for antifungal activity against Candida albicans NCPF3153, Cryptococcus neoformans ATCC90113, and Microsporum gypseum clinical isolate; however, it showed no antibacterial or antifungal activity.111


image file: c4ra11756e-s10.tif
Scheme 10 Synthesis of (±)-cytosporone E (220). Reagents and conditions: (a) EDC, HOBt; then CH2Cl2, Et3N, 91%; (b) LiBH4, then THF, room temperature (rt), 94%; (c) MsCl, Et3N, then CH2Cl2, rt, 97%; (d) sec-BuLi, THF, 6 h, −78 °C; then (b). CH3(CH2)5CHO, −78 °C to rt; (e) NH4Cl (sat.), 56%; (f) 3N HCl, reflux 30 min, 99%; (g) BBr3, CH2Cl2, −30 °C to rt, 90%.

Continuous study of Pestalotiopsis sp. PSU-MA69 endophytic to Rhizophora apiculata led to the discovery of a new butenolide pestalolide (228) and a known phytotoxic analog seiridin (229). Compound 228 showed weak antifungal activity against Candida albicans and Cryptococcus neoformans, both with MIC values of 128 μg mL−1.48 A previously unknown lactone (230), was also isolated from previously described endophyte Xylaria sp. BL321 residing in Acanthus ilicifolius L.113 Two known butenolides, butyrolactones I (198) and V (231), previously characterized from different Aspergillus terreus strains,114,115 were reisolated from the culture of Penicillium expansum 091006, an endophytic fungus from Excoecaria agallocha.80 Butyrolactone V 231 showed moderate antimalarial activity.104,116,117 Aspulvinones, a set of lactones composed of a butenolide unit and substituted by a phenyl group at C-3 and a benzyl group at C-5, were generally produced as pigments by fungi belonging to the Aspergillus family. This class of natural products was first reported from Aspergillus terreus in 1973.118 So far, more than 35 natural analogues, all exhibiting Z-configuration for the exocyclic double bond have been isolated. Furthermore, these metabolites were demonstrated to display only moderate biological activities, including antibacterial,119 luciferase inhibitory120 and glucose-6-phosphate translocase T1 inhibitory121 effects. Given their intriguing structure and antibacterial activity, they have served as synthetic targets of luciferase inhibitors in recent years.120 A new member of this group, isoaspulvinone E (232), together with two known metabolites aspulvinone E (233) and pulvic acid (234) were purified from the culture broth of Aspergillus terreus Gwq-48 in mangrove rhizosphere soil. Application of bioassay tests revealed that each of these compounds showed significant anti-influenza A H1N1 virus activities, with IC50 values of 32.3, 56.9, and 29.1 μg mL−1 respectively. Moreover, only compound 232 exhibited effective inhibitory activity against H1N1 viral neuraminidase (NA), and docking of two isomers 232 and 233 into the active sites of NA indicated that the E double bond Δ5(10) was essential to achieve this activity.122 On the other hand, phomapyrone D (235) originally characterized from the fungal pathogen Leptosphaeria maculans/Phoma lingam,123 might biogenetically result from the methylation of a tetraketide or decarboxylation of a pentaketide. It was isolated as a metabolite from another Phomopsis sp. (ZH76) living in Excoecaria agallocha.68 A pair of new spiro-γ-lactone enantiomers, (−)-(4S,8S)-foedanolide (236) and (+)-(4R,8R)-foedanolide (237), were isolated from the fermentation broth of the endophytic fungus Pestalotiopsis foedan inhabiting the branches of Bruguiera sexangula (Hainan, China). Both compounds exhibited moderate inhibition of human tumor cells HeLa, A-549, U-251, HepG2, and MCF-7, giving IC50 values between 5.4 and 296 μg mL−1.124 In a survey of Penicillium sp. ZH58 of Avicennia, a new isobenzofuranone, 4-(methoxymethyl)-7-methoxy-6-methyl-1(3H)-isobenzofuranone (238) and a known antibiotic curvularin (239) were reported.49 Macrocyclic polyketides are an important feature of the fungi secondary metabolite chemistry. Curvularin and its derivatives have potential antibacterial activity against Bacillus megaterium,125 inhibitor of multifunctional cytokine TGF-β,126 antitrypanosomal against Trypanosoma brucei,127 and anti-inflammatory through decreasing proinflammatory gene expression in an in vivo model of a chronic inflammatory disease.128 In recent years, they are reported to be produced by some fungal species mainly from Curvularia,125 Aspergillus,129 and Penicillium.49 Sumalarins A–C (240–242), the new and rare examples of sulfur-containing curvularin derivatives, along with three known analogues (239, 243, 244), were identified from the cytotoxic extract of Penicillium sumatrense MA-92, a fungus obtained from the rhizosphere of the mangrove Lumnitzera racemosa. The absolute configuration of compound 240 was established by X-ray crystallographic analysis. Considering the possible biogenetic pathway for these sulfur-containing compounds 240–242, compounds 240 and 241 can be derived from 242 by esterification or esterification and acylation. However, the putative intermediate 242 is likely produced via Michael addition of the cysteine metabolite 3-mercaptolactate to the double bond of dehydrocurvularin 243 as the putative precursor shown in Scheme 11. It is noteworthy that compounds 240–243 showed potent cytotoxicity against some of the tested tumor cell lines. Sulfur substitution at C-11 or a double bond at C-10 significantly increased the cytotoxic activities of the curvularin analogues.130 Chromatographic analysis of mangrove Sonneratia caseolaris (Dongzhai, Hainan, China) endophytic Pestalotiopsis virgatula rice culture extracts yielded four new α-pyrone derivatives, pestalotiopyrones I–L (245–248), and a new (6S,10S,20S)-hydroxypestalotin (249).131

image file: c4ra11756e-u11.tif


image file: c4ra11756e-s11.tif
Scheme 11 Possible pathway for compounds 240–242 from 243.
4.1.9 Miscellaneous polyketides. Griseofulvin is a classic antifungal agent used clinically for the treatment of dermatomycoses. It was originally obtained from the fungal strain Penicillium griseofulvum in 1939; however, it was reisolated from the diverse mangrove-derived fungi. It was first reported from the unidentified mangrove endophytic fungus no. 1403 (ref. 132) and later isolated from Nigrospora sp. no. 1403 endophyte of Kandelia candel56 and Pongamia pinnata endophytic to Nigrospora sp. no. MA75.58 Its distinct activities have attracted a lot of attention and more than a hundred research papers describe its analogues synthesis. Since 1950, more than 400 analogues have been disclosed covering most positions and many have displayed significantly increased activity.133 Related metabolites were also discovered recently. Four new griseofulvin derivatives, 7-chloro-2′,5,6-trimethoxy-6′-methylspiro(benzofuran-2(3H),1′-(2)cyclohexene)-3,4′-dione (250), (2S,5′R,E)-7-hydroxy-4,6-dimethoxy-2-(1-methoxy-3-oxo-5-methylhex-1-enyl)-benzofuran-3(2H)-one (251), 6-O-desmethyldechlorogriseofulvin (252) and 6′-hydroxygriseofulvin (253); two known analogue dechlorogriseofulvin (254) and 7-chloro-2′,5,6-trimethoxy-6′-methylspiro[benzofuran-2(3H),1′-(2)cyclohexene]-3,4′-dione (255), were produced by three endophytic fungal strains: Kandelia candel endophytic to Sporothrix sp. (no. 4335),41,132 Nigrospora sp. no. 1403,56 and Pongamia pinnata endophytic Nigrospora sp. strain no. MA75.58 Compounds 252–254 were devoid of any significant antibacterial activity toward MRSA, E. coli, P. aeruginosa, P. fluorescens, and S. epidermidis.58 This further confirmed the prediction that the activity of griseofulvin should be strictly related to both its planar structure and spatial configuration. A known metabolite 2-acetyl-7-methoxybenzofuran (256) was also isolated from the aforementioned endophytic fungus Sporothrix sp. (no. 4335).134 Bioactivity-directed fractionation of the extract of the endophytic Talaromyces sp. ZH-154 isolated from the stem bark of Kandelia candel (L.) Druce, Rhizophoraceae, afforded two new austdiol analogues, 7-epiaustdiol (257) and 8-O-methylepiaustdiol (258). The absolute configuration of 257 was unequivocally determined by single-crystal X-ray diffraction. Although, both compounds were found to be moderately cytotoxic against KB and KBv200 cells (IC50 = 16.37–37.16 μg mL−1), 257 displayed significant inhibitory activity against Pseudomonas aeruginosa with a MIC value of 6.25 μg mL−1.135 Chermesinone A (259), a new azaphilone confirmed by X-ray analysis, showed a mild anti-α-glucosidase effect with an IC50 value of 24.5 μM, was isolated from culture of Penicillium chermesinum (ZH4-E2) associated with Kandelia candel.74

Exploration of the Bruguiera parviflora endophytic fungus Xylaria cubensis PSU-MA34 provided two new succinic acid derivatives, xylacinic acids A (260) and B (261), along with one known analog, 2-hexylidene-3-methyl succinic acid 4-methyl ester (262).38 A new fatty acid glycoside was obtained from an unidentified endophytic fungus A1 isolated from Scyphiphora hydrophyllacea Gaertn. F., and it was identified as R-3-hydroxyundecanoic acid methyl ester-3-O-α-L-rhamnopyranoside (263). It showed modest inhibitory effect on Staphylococcus aureus and methicillin-resistant S. aureus (MRSA).136 A specific fatty acid methyl esters (FAME) profile was also used in taxonomic studies of the Pestalotiopsis sp. strain. The specific FAME profile of Rhizophora mucronata endophytic Pestalotiopsis JCM2A4 was constructed. Meanwhile, two of these components were isolated and determined to be n-hexadecanoic acid (264) and elaidic acid (265). These compounds (264 and 265) were produced by Pestalotiopsis JCM2A4, an endophytic fungus originally isolated from leaves of the Chinese mangrove plant.137 Flavodonfuran (266), a new difuranylmethane derivative from another Rhizophora apiculata endophytic fungus Flavodon flavus PSU-MA201, is proposed to be biosynthesized from linoleic acid. Considering the chemical reactions, a three-step mechanism was proposed that involves the formation of a furan fatty acid. The proposed mechanism involves the oxidation of an unsaturated unit of linoleic acid that would provide 5-pentyl-2-furaldehyde; and reduction of the furaldehyde followed by coupling of two units of the resulting furfuryl alcohol that would furnish the difuranylmethane as shown in Scheme 12.138 Another known metabolite reported in this category included 3-hydroxypropionic acid (3-HPA) (267), an antibacterial agent from Diaporthe phaseolorum obtained from the branches of Laguncularia racemosa. In bioassays, 3-HPA showed antimicrobial activities against both Staphylococcus aureus and Salmonella typhi.139 Analysis of mixed fermentation of unidentified mangrove fungal strains Kandelia candel endophytic K38 and Eucheuma muricatum endophytic E33 yielded a common metabolite allitol (268).37 Two usual phthalate derivatives, dibutylphthalate (269) and mono (2-ethylhexyl) phthalate (270), were reported from the antibacterial extract of the fungus Phoma herbarum VB7.140 Diisobutyl phthalate (271), a co-occurring metabolite, was reported from the endophytic fungus Pestalotiopsis clavispora isolated from Bruguiera sexangula.87 Acremonium sp. PSU-MA70 endophytic to Rhizophora apiculata was also the source of two known metabolites, 4-methyl-1-phenyl-2,3-hexanediol (272), (2R,3R)-4-methyl-1-phenyl-2,3-pentanediol (273).40 A diverse array of compounds was isolated from the Rhizophora apiculata endophytic fungus Phomopsis sp. PSU-MA214, including a previously known phenethyl alcohol hydracrylate (274) and a known butanamide (275).59 The previously known acetylenic nematicide (S)-penipratynolene (276) and DNA-damaging active anofinic acid (277) were characterized from the endophytic Pestalotiopsis sp. PSU-MA69 of Rhizophora apiculata. Compound 276 showed more nematicidal potential against Pratylenchus penetrans than the positive control, aspyrone. It killed 77% P. penetrans at a concentration of 300 μg mL−1.48,141,142 Compound 278 isolated from an Excoecaria agallocha endophyte Phomopsis sp. (ZH76) was identified as a previously known metabolite, 2-methoxy-3,4-methylenedioxybenzophenone.68 Two known products, 1,8-dimethoxynaphthalene (279) and methyl 7-methylbenzofuran-2-carboxylate (280) were isolated from Sporothrix sp. (#4335) of Kandelia candel.41 A known metabolite 5,5′-oxy-dimethylene-bis(2-furaldehyde) (281) was purified from Avicennia endophytic fungus Penicillium sp. ZH58.49

image file: c4ra11756e-u12.tif


image file: c4ra11756e-s12.tif
Scheme 12 Possible biosynthetic pathway of the formation of compound 266.

4.2. Terpenoids

4.2.1 Sesquiterpenes. Terpene peroxides possess a wide range of biological activities, including antimalarial, antitumor, antimicrobial, and antiviral activities.143–145 A typical example of a sesquiterpene peroxide is artemisinin, which has already been clinically applied as an antimalarial drug.146 A culture of unidentified fungus XG8D, isolated from the leaves of Xylocarpus granatum collected from (Samutsakorn, Thailand), yielded a new nor-chamigrane endoperoxide, merulin A (282), and two new chamigrane endoperoxides, merulins B and C (283, 284). X-ray crystallographic analysis confirmed the structure of 282.147 Reinvestigation of this fungus in a large-scale fermentation experiment led to the isolation of an additional new chamigrane endoperoxide merulin D (285) and one known analogue steperoxide A (286). Of these, compounds 282 and 284 displayed significant cytotoxicity against human breast cancer line (BT474) with IC50 values of 4.98 and 1.57 μg mL−1, respectively. These compounds (282 and 284) also showed cytotoxicity against colon (SW620) cancer cell lines with IC50 values of 4.84 and 4.11 μg mL−1, respectively. Moreover, merulin C 284 displayed promising activity in a rat aortic ring sprouting (ex vivo) and a mouse Matrigel (in vivo) assay. It also exhibited potent antiangiogenic activity mainly by suppression of endothelial cell proliferation and migration in a dose-dependent manner, and its effect is mediated by the reduction in the phosphorylation of Erk1/2.148 Species Talaromyces flavus from Sonneratia apetala was the source of four new norsesquiterpene peroxides, called talaperoxides A–D (287–290), and the known analogue merulin A (steperoxide B) (282). The stereochemistries of 287, 288, and 282 were fully established through single-crystal X-ray diffraction using Cu Kα radiation. With the exception of 288 and 290, all of the sesquiterpenes were cytotoxic against human cancer cell lines MCF-7, MDA-MB-435, HepG2, HeLa, and PC-3 with IC50 values between 0.70 and 2.78 μg mL−1.149 Cultures of Diaporthe sp. from the leaves of Rhizophora stylosa produced a unique novel sesquiterpenoid with a tricyclic lactone framework, named diaporol A (291); eight new drimane sesquiterpenoids, diaporols B–I (292–299); and the known compounds 3β-hydroxyconfertifolin (300) and diplodiatoxin (301). Structures of 291–295 were further confirmed by low-temperature (100 K), single-crystal X-ray diffraction, thus establishing their absolute configurations using Cu Kα radiation. Neither of the isolated metabolites exhibited significant cytotoxicity against the tested cell lines at a concentration of 20 μM.150 Eremophilane-type sesquiterpenes are well-known secondary metabolites found in both fungi and higher plants and they have been proved to possess phytotoxic, mycotoxic and phytohormonic activity.151,152 The mangrove-derived endophytic fungus Xylaria sp. BL321 yielded three new eremophilane sesquiterpenes (302–304) and a known analogue 07H239-A (305). All but 305 showed the activation of α-glucosidase at 0.15 μM (146%), but gradually α-glucosidase showed inhibitory activity as concentrations increased, with an IC50 value of 6.54 μM.113 Metabolite 305 also showed cytotoxic activity against human breast cancer cell lines MCF-7 and MDA-MB-435, with IC50 values of 22.5 and 7.1 μM, respectively.153 Tremulenolide A (306), a representative member of the substituted perhydroazulene tremulanes, previously reported from the liquid cultures of the aspen (Populus tremuloides) rotting fungus Phellinus tremulae,154 was reisolated from Flavodon flavus PSU-MA201 endophytic to Rhizophora apiculata.138 The tricyclic compound 307 was isolated for the first time as a natural product from Acanthus ilicifolius L. endophytic fungus Xylaria sp. BL321. Attempted single-crystal X-ray crystallography resulted in establishing its stereochemistry.153,155 A previously known phytotoxin altiloxin B (308) isolated from culture filtrate of Phoma asparagi Sacc., the causal fungus of stem blight disease on asparagus,156 has also been found in our laboratory from Pestalotiopsis sp., an endophyte from the leaves of the Chinese mangrove plant Rhizophora mucronata.157 Aspergillus terreus (no. GX7-3B) isolated from the branch of Bruguiera gymnoihiza (L.) Savigny (Guangxi, China) was the source of four sesquiterpenes, involving a new compound botryosphaerin F (309), together with three known compounds (13,14,15,16-tetranorlabd-7-ene-19,6b:12,17-diolide) (310), botryosphaerin B (311) and LL-Z1271b (312). Compound 309 was found to have strong inhibitory activity towards MCF-7 (IC50 4.49 μM) and HL-60 (IC50 3.43 μM), and compound 312 exhibited significant activity against HL-60 cell line with an IC50 value of 0.6 μM. The possible biosynthetic approach for these four sesquiterpenes is proposed in Scheme 13.158 A known analogue, dilation (313), was characterized from Penicillium sp. ZH58 on mangrove tree Avicennia (Hainan, China).49 Two previously known mycotoxins, 8-deoxytrichothecin (314) and trichodermol (315), were obtained from Acremonium sp. PSU-MA70 of Rhizophora apiculata.40 It was reported that compound 314 showed selective cytotoxicity against human breast cancer (BC-1) and human small cell lung cancer (NCI-H187) lines with IC50 values of 0.88 and 1.48 μM, respectively.159 Furthermore, 315 was shown to have antifungal activity.160
image file: c4ra11756e-u13.tif

image file: c4ra11756e-s13.tif
Scheme 13 Possible biogenetic relationship between compounds 309–312.
4.2.2 Diterpenes. The world's first billion-dollar anticancer drug, called paclitaxel (taxol), was previously reported only from the inner bark of the Pacific yew Taxus brevifolia,161,162 and its production by bioengineering has still not reached a commercial scale. Paclitaxel has a complex and fascinating structure, which probably involves lengthy syntheses steps resulting in extremely low overall yields that impede the large-scale commercial preparation of this drug.163 Alternative approaches for paclitaxel production involving the use of taxol producing fungi have made significant progress worldwide. Various genera including Taxomyces,164,165 Pestalotiopsis,166 Alternaria,167 Fusarium168 and Periconia169 were found to have the capability to produce taxol and its derivatives. Recently, a mangrove endophytic fungus, Fusarium oxysporum obtained from Rhizophora annamalayana was verified for the production of taxol (316).170 It would be remarkable if mangrove endophytic fungi could become a potential alternative source for the production of taxol on a commercial scale.
image file: c4ra11756e-u14.tif
4.2.3 Sesterterpenoids. Sesterterpenoids (C25) form the smallest class of the terpenoid family; they are rarely encountered in marine-derived fungi and no example has been reported from mangrove microbes to date.171–173 Recently, a novel 5/7/(3)6/5 pentacyclic sesterterpenoid named asperterpenoid A (317) has been isolated from a Sonneratia apetala endophyte, which was identified from Aspergillus sp. 16-5c.174 Another two novel sesterterpenoids with an unusual 5/8/6/6 tetracyclic ring skeleton, named asperterpenols A (318) and B (319) were isolated from Acanthus ilicifolius endophytic Aspergillus sp. 085242. Absolute configuration of these metabolites 317–319 were determined by single crystal X-ray diffraction analysis and was further confirmed from the proposed biosynthetic pathway. These metabolites are derivable from the same precursor geranylfarnesyl diphosphate (GFPP), through a series of cationic cyclizations, rearrangements and oxidation, as shown in Scheme 14. An extensive bioassay disclosed strong inhibitory activity of compound 317 against Mycobacterium tuberculosis protein tyrosine phosphatase B (mPTPB) with an IC50 value of 2.2 μM while compounds 318 and 319 strongly inhibited acetylcholinesterase (AChE) with IC50 values of 2.3 and 3.0 μM, respectively.175
image file: c4ra11756e-s14.tif
Scheme 14 Possible biogenetic pathway for sesterterpenoids 317–319.
4.2.4 Triterpenes. Along with ursolic acid (320), 3β,22β,24-trihydroxy-olean-12-ene (321), three new oleanane-type triterpenoids, (15α)-15-hydroxysoyasapogenol B (322), (7β,15α)-7,15-dihydroxysoyasapogenol B (323), and (7β)-7,29-dihydroxysoyasapogenol B (324) were isolated from the liquid culture of Pestalotiopsis clavispora, an endophytic fungus isolated from the Chinese mangrove plant Bruguiera sexangula (Dongzhai, Hainan, China).87,176
image file: c4ra11756e-u15.tif
4.2.5 Meroterpenes. The term meroterpenoid was first applied by Cornforth, in 1968, to describe natural products of mixed biosynthetic origin, which are partially derived from terpenoids. These natural products are most often isolated from fungi and marine organisms and some have provided valuable access to novel hybrid chemotypes originating from simple compounds. These compounds comprise a prenyl unit linked to the polyketide unit to form unique architectural scaffolds, and are widely distributed in mangrove-derived fungi.177 Austin-like metabolites represent a class of meroterpenoids mainly isolated from Aspergillus and Penicillium genera.178–181 It was previously reported that these derivatives show notable toxicities to insects.182,183 Chemical investigation on the mangrove endophytic fungus Aspergillus sp. 085241B, led to the isolation of two new austin-like hybrid spirolactone meroterpenes, acetoxydehydroaustin B (325) and 1,2-dihydro-acetoxydehydroaustin B (326).184 Culture of Emericella sp., isolated as an endophytic fungus associated with Aegiceras corniculatum resulted in the purification of further members of the series such as austin (327), deacetylaustin (328) and dehydroaustin (329).185 Recently, compounds 325, 327 and 329 were shown to have a blocking action on cockroach nicotinic acetylcholine receptors.186 A static culture of Penicillium sp. MA-37, which was obtained from the rhizospheric soil of the mangrove plant Bruguiera gymnorrhiza led to the identification three new meroterpenoid derivatives, namely 4,25-dehydrominiolutelide B (330), 4,25-dehydro-22-deoxyminiolutelide B (331), and isominiolutelide A (332), together with three known congeners, berkeleyacetals A and B (333, 334), and 22-epoxyberkeleydione (335).78 The genus Pestalotiopsis has been shown to be a good source of sesquiterpenes, some of which are analogues arising from the metabolic hybrid with mero-sesquiterpenoid skeletons.187 Two novel hybrid sesquiterpene-cyclopaldic acid metabolites with an unusual carbon skeleton, named pestalotiopens A and B (336, 337), were obtained from the endophytic fungus Pestalotiopsis sp. JCM2A4 isolated from the leaves of the Chinese mangrove, Rhizophora mucronata. The absolute configurations of the new metabolites were determined by a combination of spectroscopic methods and quantum-chemical optical-rotatory dispersion (ORD) and circular dichroism (CD) calculations. A biosynthetic pathway to pestalotiopens A and B was proposed with altiloxin B (308) as one of the suggested precursors as illustrated in Scheme 15. Both compounds were evaluated for their antimicrobial and antifungal activities against six bacterial strains, namely, Escherichia coli, Enterococcus faecalis, Staphylococcus aureus, Streptococcus pyogenes, Pseudomonas aeruginosa, and Klebsiella pneumonia. All compounds except 336 showed moderate antimicrobial activity against Enterococcus faecalis.157 The search for new acetylcholinesterase inhibitors has led to the characterization of a series of potential inhibitors, 338–340, from Kandelia candel endophytic Penicillium sp. sk5GW1L, with IC50 values of 0.64 ± 0.08 μM, 0.37 ± 0.11 μM, and 7.03 ± 0.20 nM, respectively. Of these, α-pyrone meroterpene arigsugacin I (338) is new. Low temperature (100 K), single crystal X-ray diffraction with Cu Kα radiation led to the determination of the absolute configuration as shown in compounds 338 and 339.188
image file: c4ra11756e-u16.tif

image file: c4ra11756e-s15.tif
Scheme 15 Proposed biogenetic transformation from compound 308 to compound 336, 337.

4.3. Nitrogenated compounds

4.3.1 Amines and amides. A halotolerant fungus, Penicillium chrysogenum PXP-55, was isolated from the root surface of Rhizophora stylosa (Hainan, China) and was later cultured in a hypersaline medium. It provided five new cerebrosides, chrysogesides A–E (341–345), and two new 2-pyridone alkaloids, chrysogedones A and B (346, 347). Among chrysogesides B–D (342–344) were the first cerebrosides that featured an unsaturated C19-fatty acid. Attemped application of both the CD excitation chirality and the modified Mosher method resulted in the determination of 341–347 as stereochemistries, of which 342 was antimicrobial against Enterobacter aerogenes with an MIC value of 1.72 μM.189 Several studies demonstrated that mangrove-derived fungi are capable of producing a number of important bioactive secondary metabolites, hitherto known in trace quantities. This is of particular significance since it highlights the potential of using these fungi as alternative and effective source for these metabolites.190 Different Aspergillus strains are capable of making various hydroxypyrazine derivatives. A known antibiotic, neoaspergillic acid (348), originally obtained from the fungal species Aspergillus sclerotiorum191 was reisolated from the mixed cultured mycelia of two mangrove Avicennia epiphytic Aspergillus sp. fungi strain FSY-01 and FSW-02. Compound 348 showed significant antibacterial activity against three Gram-positive bacteria, namely, Staphylococcus aureus, Staphylococcus epidermidis, and Bacillus subtilis (IC50 0.49–1.95 μg mL−1), and three Gram-negative bacteria, Bacillus dysenteriae, Bacillus proteus, and Escherichia coli, (IC50 7.80–15.62 μg mL−1).192 Bioassay-guided fractionation of extracts from the endophytic fungus Emericella sp. (HK-ZJ) isolated from Aegiceras corniculatum led to the isolation of eight isoindolone derivatives, termed as emerimidines A and B (349, 350) and emeriphenolicins A–D (351–354), and two previously reported compounds, aspernidines A and B (355, 356). A close biogenetic relationship probably existed among these compounds; thus, a mixed biosynthetic route involving polyketide and mevalonate pathways is proposed in Scheme 16. All isolated metabolites were tested for in vitro activity against H1N1 replication in MDCK cells. However, only compounds 349 and 350 showed moderate inhibitory effects with IC50 values of 42.07 and 62.05 μg mL−1, respectively.185 Two previously known phytotoxins, equisetin (357) and epi-equisetin (358) were isolated from Rhizophora stylosa (Hainan, China) endophytic Fusarium equiseti AGR12, and showed moderate antibacterial activity.193,194 Equisetin was a classical antibiotic agent and showed selectivity against several strains of Gram-positive bacteria.195 Surprisingly, it was reported to inhibit the 2,4-dinitrophenol (DNP) stimulated ATPase activity of rat liver mitochondria and mitoplasts in a concentration-dependent manner; effectively 50% inhibition was caused by about 8 nmol equisetin per mg protein.196 Cytochalasans are a group of substances whose structures include the presence of a hydrogenated and highly substituted bicyclic isoindolone moiety fused to a marcrocyclic ring and several have been reported to originate from mangrove fungi.8 Studies are currently being conducted into the active constituents of the endophytic fungus Xylaria sp. BL321 of Acanthus ilicifolius L. Some of these studies have already led to the characterization of three known amides, cytochalasins C–D (359, 360), and 19,20-epoxycytochalasin C (361).153 Cytochalasin IV (362) was found to be present in a fungal endophyte Sporothrix sp. (#4335) of Kandelia candel.41 Compounds 359–361 exhibited cytotoxic activity against MCF-7 and MDA-MB-435 cell lines, with IC50 values of 33.0, 26.3, and 13.6 μM and 9.3, 11.1, and 6.9 μM, respectively.153 A number of unusual alkaloids were reported from the culture broth of endophytic Fusarium incarnatum (HKI0504), which was isolated from the fruit of Aegiceras corniculatum, including N-2-methylpropyl-2-methylbutenamide (363), fusamine (364), and 3-(1-aminoethylidene)-6-methyl-2H-pyran-2,4(3H)-dione (365). Of these, the compounds 364 and 365 exhibited weak antiproliferative activity and were cytotoxic against HUVEC, K-562, and HeLa human cell lines, with IC50 values ranging from 9.0 to 103.6 μM.197 The mycelium extract of a cultured, unidentified endophytic fungus (no. AMO 3-2) originated from the inner tissues of the healthy leaves from the mangrove plant Avicennia marina, from Oman. This afforded five farinomalein derivatives, including three new farinomaleins C–E (366–368), and one new isoindoline congener (369), together with the known farinomalein (370) and its methyl ester (371), with only 367 being moderately cytotoxic to mouse lymphoma cells L5178Y with an IC50 value of 4.4 μg mL−1. To the best of our knowledge, only two such farinomalein derivatives are known to exert this functionality: farinomalein from entomopathogenic Paecilomyces farinosus,198 and pestalotiopsoid from Rhizophora mucronata endophytic Pestalotiopsis sp.199 A plausible biogenetic pathway of the isolated farinomaleins is proposed as shown in Scheme 17.200 Terretrione A–C (372–374) were new cycloheptanetriones derived probably from the condensation of amino acids with hydrophobic side chain and phenylalanine analogue, were isolated from the culture of Aspergillus terreus A8-4 (sediment). Unlike the production of diketopiperazines, an additional carbonyl group was randomly inserted between amino-nitrogen and α-carbon of phenylalanine followed by the cyclization reaction forming the so-called cycloheptanetriones skeleton, which rarely occurred in the metabolites isolated from microorganisms.105
image file: c4ra11756e-u17.tif

image file: c4ra11756e-s16.tif
Scheme 16 Hypothetic biosynthesis pathway of 349–356.

image file: c4ra11756e-s17.tif
Scheme 17 Proposed biogenetic pathway of farinomaleins 366–368, 370, 371.
4.3.2 Isoquinolines. During 2011–2013, only one previously unknown isoquinoline derivative, 2-methylimidazo[1,5-b]isoquinoline-1,3,5(2H)-trione (375), was isolated from the culture of Hypocrea virens obtained from the mangrove flora Rhizophora apiculata (Guangxi, China).201
image file: c4ra11756e-u18.tif
4.3.3 Quinolines. Quinoline alkaloids have been widely isolated from plants; however, very few examples of such compounds have been described from filamentous fungi.202 Six new members of these compounds were isolated and identified from the cultures of Aspergillus nidulans MA-143, an endophytic fungus obtained from the fresh leaves of Rhizophora stylosa. These compounds incorporate a 4-phenyl-3,4-dihydroquinolin-2-one moiety, and are: aniduquinolones A–C (376–378), 6-deoxyaflaquinolone E (379), isoaflaquinolone E (380), and 14-hydroxyaflaquinolone F (381), as well as one related known compound, aflaquinolone A (382). X-ray analysis confirmed and established the absolute configuration of compound 376. In bioscreening experiments, none of the isolated compounds showed potent antibacterial or cytotoxic activity. However, compounds 377, 378, and 382 exhibited lethality against brine shrimp (Artemia salina), with LD50 values of 7.1, 4.5, and 5.5 μM, respectively.203 Co-cultures of two unidentified mangrove-derived endophytic fungi from the coast of the South China Sea provided marinamide (383) and its methyl ester (384), which originally had been reported as pyrrolyl 1-isoquinolone alkaloids. Their structures were alternatively revised as pyrrolyl 4-quinolone, based on the recrystallization of 384 from pyridine; this yielded the known pesticide quinolactacidewhich was treated with methyl iodide to afford N-methyl quinolactacide and was then identified using X-ray crystallography. Potential cytotoxicity against HepG2, 95-D, MGC832, and HeLa tumour cell lines reaching nM degree were observed for both 383 and 384, ranging from 0.4 nM to 2.52 μM.204
image file: c4ra11756e-u19.tif
4.3.4 Indole derivatives. Isolation and fermentation of the fungal strain Eurotium rubrum, an endophyte inhabiting the inner stem tissues of Hibiscus tiliaceus, previously yielded several dioxopiperazines.205 Further research on the same fungus yielded another new dioxopiperazine alkaloid, 12-demethyl-12-oxo-eurotechinulin B (385), together with seven related known derivatives, including variecolorin J (386), eurotechinulin B (387), variecolorin G (388), alkaloid E-7 (389), cryptoechinuline G (390), isoechinulin B (391) and 7-isopentenylcryptoechinuline D (392). All isolates were evaluated for their cytotoxic activity against seven tumor cell lines of MCF-7, SW1990, HepG2, NCI-H460, SMMC-7721, HeLa, and Du145 and only compounds 385, 388, and 389 displayed moderate cytotoxic activities against one or two of these cell lines.63 Effusin A (393), 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, and a spiro-polyketide-diketopiperazine hybrid dihydrocryptoechinulin D (394) were isolated from a mangrove rhizosphere soil derived fungus, Aspergillus effuses H1-1. Both isolated compounds occurred as racemates, the enantiomers of which were separated and characterized by online HPLC-ECD analysis and their absolute configurations were determined by the solution TDDFT ECD calculation approach. Compound 393 could be obtained by a domino ring-closure reaction between the substituted salicylaldehyde moiety in aspergin and the enamide moiety of the diketopiperazine unit in neoechinulin B. In contrast, an enzyme-catalyzed regiospecific [4 + 2] Diels–Alder reaction produces the spirobicycle of 394 as shown in Scheme 18. A similar Diels–Alder biosynthetic reaction was suggested for a few recent examples, such as yaoshanenolides and lanceolatins.206 Notably, all of the key intermediates including neoechinulin B,207 isodihydroauroglaucin208 and aspergin209 were isolated from this fungus Aspergillus sp., which supported our biosynthetic hypothesis. Preliminarily cytotoxic effects were evaluated and compound 394 showed potent activity on P388 cells with an IC50 value of 1.83 μM. The target of racemic 394 was also investigated and the (12R,28S,31S)-394 enantiomer showed selectivity against topoisomerase I.210 Another new prenylated indole diketopiperazine, named dihydroneochinulin B (395), alongside three co-occurring known metabolites didehydroechinulin B (396), neoechinulin B (397), and spiropolyketide-diketopiperazine hybrid cryptoechinuline D were reisolated from the mangrove rhizosphere soil derived fungus, Aspergillus effuses H1-1. The cytotoxic effects of these compounds were preliminarily evaluated against P388, HL-60, BEL-7402 and A-549. Compound 395 showed weak activity against BEL-7402 and A-549 cell lines while compound 397 showed significant inhibitory activity against A-549 and BEL-7402. The IC50 values of A-549 and BEL-7402 were 1.43 and 4.20 μM, respectively.86 Indole-diterpenoids are known as a large and structurally diverse group of fungal secondary metabolites211 that possess a common cyclic diterpene backbone derived from geranylgeranyl diphosphate and an indole group derived from indole-3-glycerol phosphate.212 They are known as tremorgenic mycotoxins213,214 and display anti-insectan215,216 and antibiotic activities.217,218 Six new indole-diterpenoid alkaloids (398–403), as well as five known analogues, emindole SB (404), 21-isopentenylpaxilline (405), paspaline (406), paxilline (407), and dehydroxypaxilline (408) were isolated and identified from an aciduric fungal strain, Penicillium camemberti OUCMDZ-1492 obtained from an acidic mangrove soil and mud niche around the roots of Rhizophora apiculata. Compared with ribavirin (IC50 113.1 μM), compounds 398–400 and 402–408 showed significant activity in vitro against the H1N1 virus with IC50 values of 28.3, 38.9, 32.2, 73.3, 34.1, 26.2, 6.6, 77.9, and 17.7 μM, respectively.219 Another four known structure resembling congeners, penijanthine A (409), paspalinine (410), and penitrem A (411) were later found to be produced by the endophytic fungus Alternaria tenuissima EN-192. This fungus was isolated from the stem bark of Rhizophora stylosa.60 Carboline derivatives constitute an important group of natural products that possess various biological activities such as antiviral (e.g., euditomins, from tunicates),220,221 and cytotoxic and antimicrobial (e.g., manzamines, from sponges) activities.222 Although most tetrahydrocarbolines were previously believed to solely originate from marine animals, recent findings indicate that microorganisms are also capable of producing carbolines such as oxopropaline,223,224 bauerine,225 and β-carboline-1-propionic acid. The known synthetic compound 2-acetyl-1,2,3,4-tetrahydro-β-carboline (412) was isolated from a natural source for the first time from Fusarium incarnatum (HKI0504) isolated from Aegiceras corniculatum. However, it showed weak antiproliferative and cytotoxic activity against HUVEC, K-562, and HeLa human cell lines.197 Other carboline members such as harman (1-methyl-β-carboline) (413) and N9-methyl-1-methyl-β-carboline (414), were isolated from a culture of Penicillium sp. ZH58 of Avicennia, suggesting yet another possible method to produce these compounds.45 This was the first time that such metabolites had been reported from mangrove microbes.
image file: c4ra11756e-u20.tif

image file: c4ra11756e-s18.tif
Scheme 18 Plausible biosynthetic pathway to 393 and 394.
4.3.5 Quinazoline derivatives. A new moderate antibacterial agent, designated as aspergicin (415), was discovered from the mixed cultures of epiphytic Aspergillus sp. strain FSY-01 and FSW-02 occurring in Avicennia.192 Five new glyantrypine derivatives, including 3-hydroxyglyantrypine (416), oxoglyantrypine (417a, 417b), cladoquinazoline (418), epi-cladoquinazoline (419), and a new pyrazinoquinazoline derivative norquinadoline A (420) and eight known alkaloids (421–428) were isolated from the culture of the mangrove soil fungus Cladosporium sp. PJX-41 (Guangzhou, China). The absolute configurations of compounds 416–424 were established on the basis of CD, NOESY data, and single crystal X-ray diffraction analysis. Compounds 417b, 420, 422–424, and 426 showed significant activities against the influenza virus A (H1N1) with IC50 values of 82–89 μM.225 A series of quinazolinones, namely, aniquinazolines A–D (429–432) were isolated and identified from the culture of Aspergillus nidulans MA-143, an endophytic fungus obtained from the leaves of marine mangrove plant Rhizophora stylosa. As with some other classes of metabolites, these compounds showed potent lethality against brine shrimp with LD50 values of 1.27, 2.11, 4.95 and 3.42 μM, respectively, which were stronger than that of the positive control colchicine (with LD50 value of 88.4 μM). However, none of these displayed any antitumor (BEL-7402, MDA-MB-231, HL-60, and K562) or antibacterial (Escherichia coli and Staphylococcus aureus) activity.226
image file: c4ra11756e-u21.tif
4.3.6 Peptides. Culture of Bionectria ochroleuca, endogenous to the inner leaf tissues of Sonneratia caseolaris (Sonneratiaceae), yielded two new peptides designated pullularins E (433) and F (434) and two known congeners pullularins A (435) and C (436). In addition to the above mentioned compounds, the fungal epipolythiodioxopiperazine metabolite verticillin D (437) was also obtained from the Bionectria culture. All compounds were evaluated for cytotoxicity against the mouse lymphoma cells (L5178Y). Compounds 433, 435 and 436 showed moderate activity with EC50 value ranging between 2.6 and 6.7 μg mL−1. Interestingly, compound 437 showed pronounced activity against the mouse lymphoma cells with EC50 value less than 0.1 μg mL−1.227 Known pharmacological peptides were also isolated. A known cyclic tetrapeptide apicidin (438), previously characterized from Fusarium pallidoroseum, was reisolated from the culture of an unidentified endophytic fungus no. ZSU-H16 obtained from Avicennia. It was reported as a potent inhibitor of apicomplexan histone deacetylase (IC50 1–2 nM), a broad spectrum antiparasitic agent in vitro inhibits the activity of apicomplexan parasites and has shown in vivo efficacy against Plasmodium berghei malaria.65,228 Exumolide A (439), which can accelerate the growth of subintestinal vessel plexus (SIV) branches markedly, was identified from Phomopsis sp. (no. ZH-111) from sediment.35 Furthermore, two cyclic depsipeptides, guangomides A (440) and B (441) in addition to two diketopiperazines, Sch 54794 (442) and Sch 54796 (443) were purified from Acremonium sp. PSU-MA70 that were endophytes of Rhizophora apiculata.40 Compounds 440 and B 441 showed weak antibacterial activity against Staphylococcus epidermidis and Enterococcus durans229 while compound 442 exhibited inhibitory activity in the platelet activating factor (PAF) assay.230 A new diketopiperazine dimer, WIN 64821 (444) was isolated from the culture of Penicillium expansum 091006 associated with Excoecaria agallocha.80 Dimer 444 was described as a competitive antagonist to substance P (SP) at the human NK1 receptor with an inhibitor affinity constant of 230 ± 30 nM against [125I]SPin human astrocytoma cells.231 Two known cyclopentapeptides, malformins A1 and C (445 and 446) were identified from Aspergillus niger MA-132 isolated from Avicennia marina. Both compounds showed weak antibacterial activities against Staphylococcus aureus.232 More recently, a known insecticidal agent beauvericin (447) isolated from an Aspergillus terreus (no. GX7-3B) colonizing in Bruguiera gymnoihiza, displayed remarkable inhibition against α-acetylcholinesterase (AChE) with IC50 value 3.09 μM and strong cytotoxic activity against MCF-7 (IC50 2.02 μM), A549 (IC50 0.82 μM), HeLa (IC50 1.14 μM) and KB (IC50 1.10 μM) cell lines.61 Six other common compounds of this type, 3-(hydroxymethyl)-6-[4-(3-methylbut-2-enyloxy)benzyl]piperazine-2,5-dione (448), thymidine (449), and lumichrome (450), were first reported from different mangrove endophytes, Avicennia endophytic Penicillium sp. ZH58,49 Penicillium chrysogenum PXP-55 of Rhizophora stylosa,189 and Pestalotiopsis clavispora of Bruguiera sexangula.87
image file: c4ra11756e-u22.tif
4.3.7 Miscellaneous nitrogenated derivatives. Phomopsis-H76 C (451) from a Phomopsis sp. (#zsu-H76) originated from Excoecaria agallocha, featuring an unprecedented pyrano[4,3-b]pyran-5(2H)-one ring system. Primary bioassays showed that this metabolite inhibited subintestinal vessel plexus (SIV) branch growth.44 An unusual pyrrole fusarine (452) was characterized from endogenous Fusarium incarnatum (HKI0504) in Aegiceras corniculatum.197 Phomopsis sp. PSU-MA214 of Rhizophora apiculata yielded a known phenylethyl alcohol phomonitroester (453) previously described from another species Phomopsis PSU-D15 isolated from Garcinia dulcis.59,71 A previously known 8-O-methylbostrycoidin (454) has been isolated from Aspergillus terreus (no. GX7-3B) inhabitated in Bruguiera gymnoihiza as a strong inhibitor of α-acetylcholinesterase (AChE) with IC50 values 6.71 μM.61

The mycelium extract of a cultured Fusarium sp. from the bark of Kandelia candel (L.) Druce (Hainan, China) yielded fusaric acid (455) as the predominant constituent responsible for the antimycobacterial activity. A variety of metal complexes of fusaric acid were prepared. Antimycobacterial assays showed that Cadmium(II) and Copper(II) complexes exhibited potent inhibitory activity against the Mycobacterium bovis BCG (MIC = 4 μg mL−1) and the M. tuberculosis H37Rv (MIC = 10 μg mL−1). This is the first report of the antimycobacterial activity of a mangrove Fusarium metabolite and its coordinating metal complexes.233

image file: c4ra11756e-u23.tif

4.4. Steroids

Sterols with a 5,8-epidioxy moiety are well-known metabolites from marine organisms, such as corals, sponges, and mangrove-derived fungi, as well as terrestrial macrofungi.234–238 In contrast, sterols with the 5,9-epidioxy motif have been rarely reported. So far, only five 5,9-epidioxy-sterols have been isolated from several edible mushrooms; however, none of them were extracted from mangrove organisms. Nigerasterols A and B (456, 457) representing the first 5,9-epidioxy-sterols of mangrove origin were characterized from Aspergillus niger MA-132 isolated from Avicennia marina. Both 456 and 457 displayed potent activity against tumor cell lines HL60 (IC50 0.30/1.50 μM) and A549 (IC50 1.82/5.41 μM) in a preliminary bioassay. The absolute configuration of compound 456 was determined by the application of a modified version of Mosher's method.232 Apart from a known neuritogenic compound NGA0187 (458), compounds 3β,5α-dihydroxy-(22E,24R)-ergosta-7,22-dien-6-one (459), 3β,5α,14α-trihydroxy-(22E,24R)-ergosta-7,22-dien-6-one (460) were also purified from the culture of Aspergillus terreus (no. GX7-3B) obtained from Brugnieria gymnoihiza. Metabolite 458 induced significant neurite outgrowth in PC12 cells and was found to show potent inhibitory activity against MCF-7, A549, HL-60 and KB cell lines. MCF-7, A549, HL-60 and KB cell lines had IC50 values of 4.98, 1.95, 0.68, 1.50 μM, respectively.61,239,240 The 3β-hydroxy sterols and their oxygenated analogues include a large group of metabolites found in lower terrestrial organisms, marine organisms, lichens and fungi.241 Several ubiquitous microbial secondary metabolites such as ergosta-4,6,8(14),22-tetraen-3-one (461) were isolated from Pestalotiopsis clavispora from the stem of Bruguiera sexangula.87 The sterol 7,22-(E)-diene-3β,5α,6β-triol-ergosta (462) was discovered using a mixed fermentation technique on unidentified mangrove endophytic fungal strains Kandelia candel endophytic K38 and Eucheuma muricatum endophytic E33.37
image file: c4ra11756e-u24.tif

5. Concluding remarks

Mangrove-associated microbes continue to be the focus for much of natural products research that has been well documented. In this review, we systematically summarized the new findings regarding the chemistry and bioactivities of most natural products found in mangrove microbial ecosystems during the last three years. These include 464 naturally occurring small-biomolecules. The rate at which these metabolites have been discovered within the past three years has increased sharply (33%). Furthermore, they have contributed to increasing chemical diversity in the compounds discovered over the past twenty years. This increase is largely due to improvements in isolation procedures and structural analysis. These trends are shown graphically in Fig. 1, which compares outputs for 1989 through to Nov. 2010 with those for the recent three years. The compounds are shown according to their structural classification.
image file: c4ra11756e-f1.tif
Fig. 1 Comparison of mangrove-associated microbial metabolites distribution by backbone.

At the domain level, <5% of the species examined originated from the actinomyces or bacteria while the majority (>95%) originated from fungi. Actinomycetes are of special interests (4%) as they are known to produce chemically diverse compounds with an extraordinarily high proportion (60%) of their metabolites showing a wide range of biological activities. The endophytic fungi have been found to produce a significant number of interesting metabolites (>83%) with biologically active substances. The endophytic fungi have been found to produce a significant number of interesting metabolites >83% with 32% biologically active substances, occur in considerably higher numbers than those produced by rhizosphere soil or sediments microorganisms <12% but of 38% of bioactive substances. The high rate of inactive metabolite discovery in previous studies was probably because of bias in the screening programmes and limitations in analytical technology. It is noteworthy that Aspergillus, Penicillium, and Pestalotiopsis are predominant as producers of promising chemical diversity. Furthermore, they constitute 46% of the compounds reported, and if five other prolific genera such as Streptomyces, Acremonium, Phomopsis, Sporothrix, and Xylaria are considered, they account for nearly 68% of all of the metabolites. The remaining 26% are scattered across another 17 genera, and less than 6% of the metabolites are from unidentified microorganisms. Rhizophora apiculata (14%), rhizosphere soil and sediment (11%), Kandelia candel (10%), Bruguiera gymnoihiza (8%) and Rhizophora stylosa (6%) were considered to be the main sources of metabolic microbes. The emergence of plants native to China and Thailand as significant sources of new compounds may be of particular significance. Our last review discussed the problem of ambiguously described microorganisms; however, since then the ratio of ambiguous microbes and sources reported in the ensuing three years have declined dramatically, and most microbes are properly documented and characterized (Tables 1 and 2).

Some compounds intrigue the natural product researchers because of their unusual structures including the unusual 5/8/6/6 tetracyclic ring or 5/7/(3)6/5 pentacyclic sesterterpenoids asperterpenoid A (317), asperterpenols A (318) and B (319) from Aspergillus sp. Some metabolites have provided valuable access to novel hybrid chemotypes derived from different biosynthetic routes. These include NRPS, terpene, shikimate and polyketide mixed biosynthetic tetrasubstituted benzothiazole erythrazoles A (21) and B (22) from Erythrobacter sp., hybrid sesquiterpene-polyketide metabolites pestalotiopens A and B (336, 337) from Pestalotiopsis sp. JCM2A4. Among the 130 or so bioactive components presented in this review, several have fascinating bioactivities comparable to those of modern pharmacological products. These include the antitumor agents paeciloxocin A (179), sumalarins A–C (240–242), and merulin A and C (287, 289). Paeciloxocin A from Paecilomyces sp. exhibited strong cytotoxicity effects against the growth of hepG2 cell line at 1 μg mL−1. Sumalarins A–C (240–242) from Penicillium sumatrense MA-92 showed potent cytotoxicity against some of the tested tumor cell lines, whereas merulin A and C (282, 284) were identified from an unidentified fungus XG8D isolated from the leaves of Xylocarpus granatum, which displayed significant cytotoxicity against human breast cancer line (BT474) with IC50 values of 4.98 and 1.57 μg mL−1, respectively. Merulin C 284 displayed promising activity in a rat aortic ring sprouting (ex vivo) and a mouse Matrigel (in vivo) assay. Anti-H1N1 agents such as isoaspulvinone E (232) from Aspergillus terreus Gwq-48 showed significant anti-influenza A H1N1 virus activities with IC50 value of 32.3 μg mL−1 and exhibited effective inhibitory activity against H1N1 viral neuraminidase (NA). Indole-diterpenoid alkaloids 398–400 and 402–408 from Penicillium camemberti OUCMDZ-1492 exhibited significant in vitro anti-H1N1 with IC50 values of 6.6–89 μM, respectively. The compounds 417b, 420, 422–424, and 426 from Cladosporium sp. PJX-41 showed significant activities against the influenza virus A (H1N1) with IC50 values of 82–89 μM, which is much more effective than ribavirin (IC50 113.1 μM). Anti-enzyme agents such as p-terphenyls 125–127 showed strong inhibitory effects against R-glucosidase with IC50 values of 0.9, 4.9, and 2.5 μM, respectively. The compound 317 displayed strong inhibitory activity against Mycobacterium tuberculosis protein tyrosine phosphatase B (mPTPB) with an IC50 value of 2.2 μM while compounds 318 and 319 strongly inhibited acetylcholinesterase (AChE) with IC50 values of 2.3 and 3.0 μM, respectively. Arigsugacin I (338) from Penicillium sp. sk5GW1L showed potential anti-AChE activity with IC50 values of 0.64 ± 0.08 μM. Compounds, 27, 46 and 439 from Phomopsis sp. were demonstrated to accelerate the growth of subintestinal vessel plexus (SIV) markedly, whereas phomopsis-H76 C (451) was shown to be a subintestinal vessel plexus (SIV) branch growth inhibitor. Strong brine shrimp lethality was detected by aniquinazolines A–D (429–432) from Aspergillus nidulans. MA-143 showed potent lethality with LD50 values of 1.27, 2.11, 4.95 and 3.42 μM, respectively, which were over 40 times stronger than that of the positive control colchicines. This makes many of these compounds suitable candidates for drug discovery and will trigger groundbreaking synthesis studies of these compounds in the coming years.

Biogenetic hypotheses have become increasingly valuable in the screening of structurally related missing precursors/intermediates/products. Metabolites are biologically compatible with many living systems and often possess unique and useful biological activities, largely because their scaffolds are the result of evolutionarily significant selective pressures. Activation of these attenuated or silenced genes of biosynthetic pathways to obtain either improved titers of known compounds or new ones altogether has been the subject of particular interest. “One Strain Many Compounds” (OSMAC) approach has been popular towards the discovery of new metabolites. Consequently, the manipulation of biosynthetic pathways, epigenetic modifications (cytosporone R 150 from concomitant supplementation of Leucostoma persoonii), varying culture conditions (compounds 47–58 from Aspergillus tubingensis) and co-culturing two organisms together (compounds 221–227 from Pestalotiopsis sp. PSU-MA92/PSU-MA119) have been applied to stimulate the production of new compounds.

The anticancer drug taxol (316) was purified from Fusarium oxysporum of Rhizophora annamalayana. This drug implicated mangrove endophytes as a sustainable, economically feasible and alternative source of therapeutic compounds. The production of bioactive substances by mangrove microbes is directly related to the independent evolution of these microorganisms, which may have incorporated genetic information from their biotope and carry out certain functions. For example, austin-like derivatives 325–329 represent a class of compounds (identified from Emericella sp.) that are toxic to insects and play crucial ecological roles such as the capacity to protect its host plant against insect invaders. Thus, they contribute to a range of defense substances secreted by plants that can be further investigated. Interesting biotopes such as the mangrove microbial communities are especially productive since they accumulate a diverse array of bioactive compounds with novel scaffolds. As yet, the potential of this area remains virtually untapped. Substantial evidence now exists showing that in the long-term maintenance of the biodiversity profile, mangrove ecosystems can provide a valuable function in wave and storm surge attenuation, and erosion reduction. However, nowadays approximately one-third of the global cover of these ecosystems has been lost; this is particularly true in Southeast Asia. The most challenging goals concerning secondary metabolites are the elucidation of their true function in their native mangrove habitats and, closely associated with this, the identification of the physiological and ecological conditions that have led to the activation of secondary metabolism gene clusters. Many of the extensive areas of mangrove restoration throughout the world have cited expected ecosystem protection benefits, which will also support the continued provision of functional, diverse, inhabitated microbes. Multidisciplinary research and collaborative endeavors amongst biotechnologists, microbiologists, chemists and pharmacologists would result in the isolation of unusual or rare microorganisms that produce structurally interesting and biologically active molecules with potential use in medicinal and agricultural applications.

Acknowledgements

Co-Financed by grants of National Natural Science Foundation of China for Young Scholar (no. 81202456), Hainan Provincial Natural Science Foundation (no. 814285), Special Foundation for Modernization of Traditional Chinese Medicine of Hainan (ZY201429) and the programs for New Century Excellent Talents in University (NCET-13-0760) are gratefully acknowledged.

References

  1. I. C. Feller, C. E. Lovelock, U. Berger, K. L. McKee, S. B. Joye and M. C. Ball, Biocomplexity in mangrove ecosystems, Annu. Rev. Mar. Sci., 2010, 2, 395–417 CrossRef CAS PubMed.
  2. H. Thatoi, B. C. Behera, R. R. Mishra and S. K. Dutta, Biodiversity and biotechnological potential of microorganisms from mangrove ecosystems: a review, Ann. Microbiol., 2013, 63, 1–19 CrossRef CAS PubMed.
  3. J. W. Blunt, B. R. Copp, R. A. Keyzers, M. H. Munro and M. R. Prinsep, Marine natural products, Nat. Prod. Rep., 2013, 30, 237–323 RSC.
  4. J. W. Blunt, B. R. Copp, R. A. Keyzers, M. H. Munro and M. R. Prinsep, Marine natural products, Nat. Prod. Rep., 2014, 31, 160–258 RSC.
  5. X. Wang, Z. G. Mao, B. B. Song, C. H. Chen, W. W. Xiao, B. Hu, J. W. Wang, X. B. Jiang, Y. H. Zhu and H. J. Wang, 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 CAS.
  6. A. Debbab, A. H. Aly and P. Proksch, Bioactive secondary metabolites from endophytes and associated marine derived fungi, Fungal Diversity, 2011, 49, 1–12 CrossRef PubMed.
  7. A. Debbab, A. H. Aly and P. Proksch, Mangrove derived fungal endophytes-a chemical and biological perception, Fungal Diversity, 2013, 61, 1–27 CrossRef PubMed.
  8. J. Xu, Biomolecules produced by mangrove-associated microbes, Curr. Med. Chem., 2011, 18, 5224–5266 CrossRef CAS.
  9. R. Subramani and W. Aalbersberg, Culturable rare Actinomycetes: diversity, isolation and marine natural product discovery, Appl. Microbiol. Biotechnol., 2013, 97, 9291–9321 CrossRef CAS PubMed.
  10. G. D. Brown, The biosynthesis of artemisinin (Qinghaosu) and the phytochemistry of Artemisia annua L. (Qinghao), Molecules, 2010, 15, 7603–7698 CrossRef CAS PubMed.
  11. H. Q. Gong, Q. X. Wu, L. L. Liu, J. L. Yang, R. Wang and Y. P. Shi, Sesquiterpenoids from the aerial parts of Inula japonica, Helv. Chim. Acta, 2011, 94, 1269–1276 CrossRef CAS.
  12. O. Shirota, J. M. Oribello, S. Sekita and M. Satake, Sesquiterpenes from Blumea balsamifera, J. Nat. Prod., 2011, 74, 470–476 CrossRef CAS PubMed.
  13. F. R. Garcez, W. S. Garcez, L. Hamerski and A. C. d. M. Miranda, Eudesmane and rearranged eudesmane sesquiterpenes from Nectandra cissiflora, Quim. Nova, 2010, 33, 1739–1742 CrossRef CAS PubMed.
  14. S. J. Wu, S. Fotso, F. Li, S. Qin, G. Kelter, H. H. Fiebig and H. Laatsch, N-Carboxamido staurosporine and selina-4(14),7(11)-diene-8,9-diol, new metabolites from a marine Streptomyces sp., J. Antibiot., 2006, 59, 331–337 CrossRef CAS PubMed.
  15. Z. Yang, Y. Yang, X. Yang, Y. Zhang, L. Zhao, L. Xu and Z. Ding, Sesquiterpenes from the secondary metabolites of Streptomyces sp. (YIM 56130), Chem. Pharm. Bull., 2011, 59, 1430–1433 CrossRef CAS.
  16. L. Ding, A. Maier, H. H. Fiebig, W. H. Lin, G. Peschel and C. Hertweck, Kandenols A–E, eudesmenes from an endophytic Streptomyces sp. of the mangrove tree Kandelia candel, J. Nat. Prod., 2012, 75, 2223–2227 CrossRef CAS PubMed.
  17. L. Ding, J. Münch, H. Goerls, A. Maier, H. H. Fiebig, W. H. Lin and C. Hertweck, Xiamycin, a pentacyclic indolosesquiterpene with selective anti-HIV activity from a bacterial mangrove endophyte, Bioorg. Med. Chem. Lett., 2010, 20, 6685–6687 CrossRef CAS PubMed.
  18. L. Ding, A. Maier, H. H. Fiebig, W. H. Lin and C. Hertweck, A family of multicyclic indolosesquiterpenes from a bacterial endophyte, Org. Biomol. Chem., 2011, 9, 4029–4031 CAS.
  19. Z. L. Xu, M. Baunach, L. Ding and C. Hertweck, Bacterial synthesis of diverse indole terpene alkaloids by an unparalleled cyclization sequence, Angew. Chem., Int. Ed., 2012, 51, 10293–10297 CrossRef CAS PubMed.
  20. X. L. Li, M. J. Xu, Y. L. Zhao and J. Xu, A novel benzo[f][1,7]naphthyridine produced by Streptomyces Albogriseolus from mangrove sediments, Molecules, 2010, 15, 9298–9307 CrossRef CAS PubMed.
  21. M. J. Xu, X. J. Liu, Y. L. Zhao, D. Liu, Z. H. Xu, X. M. Lang, P. Ao, W. H. Lin, S. L. Yang, Z. G. Zhang and J. Xu, Identification and characterization of an anti-fibrotic benzopyran compound isolated from mangrove-derived Streptomyces xiamenensis, Mar. Drugs, 2012, 10, 639–654 CrossRef CAS PubMed.
  22. C. Tian, X. Jiao, X. Liu, R. Li, L. Dong, X. Liu, Z. G. Zhang, J. Xu, M. J. Xu and P. Xie, First total synthesis and determination of the absolute configuration of 1-N-methyl-3-methylamino-[N-butanoicacid-3-(9-methyl-8-propen-7-one)-amide]-benzo[f][1,7]naphthyridine-2-one, a novel benzonaphthyridine alkaloid, Tetrahedron Lett., 2012, 53, 4892–4895 CrossRef CAS PubMed.
  23. G. D. Chen, H. Gao, J. S. Tang, Y. F. Huang, Y. Chen, Y. Wang, H. N. Zhao, H. P. Lin, Q. Y. Xie, K. Hong, J. Li and X. S. Yao, Benzamides and quinazolines from a mangrove actinomycetes Streptomyces sp. (no. 061316) and their inhibiting caspase-3 catalytic activity in vitro, Chem. Pharm. Bull., 2011, 59, 447–451 CrossRef CAS.
  24. M. Ueki, K. Ueno, S. Miyadoh, K. Abe, K. Shibata, M. Taniguchi and S. Oi, UK-1, a novel cytotoxic metabolite from Streptomyces sp. 517-02. I. Taxonomy, fermentation, isolation, physico-chemical and biological properties, J. Antibiot., 1993, 46, 1089–1094 CrossRef CAS.
  25. P. S. Sommer, C. Almeida, K. Schneider, W. Beil, R. D. Süssmuth and H. P. Fiedler, Nataxazole, a new benzoxazole derivative with antitumor activity produced by Streptomyces sp. Tü 6176, J. Antibiot., 2008, 61, 683–686 CrossRef CAS PubMed.
  26. K. H. Michel, L. D. Boeck, M. M. Hoehn, N. D. Jones and M. O. Chaney, The discovery, fermentation, isolation, and structure of antibiotic A33853 and its tetraacetyl derivative, J. Antibiot., 1984, 37, 441–445 CrossRef CAS.
  27. C. Hohmann, K. Schneider, C. Bruntner, E. Irran, G. Nicholson, A. T. Bull and H. P. Fiedler, Caboxamycin, a new antibiotic of the benzoxazole family produced by the deep-sea strain Streptomyces sp. NTK 937*, J. Antibiot., 2009, 62, 99–104 CrossRef CAS PubMed.
  28. Y. W. Chu, J. J. Li, L. Wang and R. Yu, Antifungal antibiotic CAl189 produced by a mangrove endophyte Streptomyces sp. A1626, Chin. J. Antibiot., 2011, 36, 14–17 CAS.
  29. C. S. Neumann, D. G. Fujimori and C. T. Walsh, Halogenation strategies in natural product biosynthesis, Chem. Biol., 2008, 15, 99–109 CrossRef CAS PubMed.
  30. X. G. Li, X. M. Tang, J. Xiao, G. H. Ma, L. Xu, S. J. Xie, M. J. Xu, X. Xiao and J. Xu, Harnessing the potential of halogenated natural product biosynthesis by mangrove-derived actinomycetes, Mar. Drugs, 2013, 11, 3875–3890 CrossRef PubMed.
  31. E. Higashide, K. Hatano, M. Shibata and K. Nakazawa, Enduracidin, a new antibiotic. I. Streptomyces fungicidicus no. B5477, an enduracidin producing organism, J. Antibiot., 1968, 21, 126–137 CrossRef CAS.
  32. Y. Hu and J. B. MacMillan, Erythrazoles A-B, cytotoxic benzothiazoles from a marine-derived Erythrobacter sp., Org. Lett., 2011, 13, 6580–6583 CrossRef CAS PubMed.
  33. Z. Huang, J. Yang, X. Cai, Z. She and Y. Lin, A new furanocoumarin from the mangrove endophytic fungus Penicillium sp. (ZH16), Nat. Prod. Res., 2012, 26, 1291–1295 CrossRef CAS PubMed.
  34. B. Thongbai, F. Surup, K. Mohr, E. Kuhnert, K. D. Hyde and M. Stadler, Gymnopalynes A and B, chloropropynyl-isocoumarin antibiotics from cultures of the basidiomycete Gymnopus sp., J. Nat. Prod., 2013, 76, 2141–2144 CrossRef CAS PubMed.
  35. J. X. Yang, Y. Chen, C. Huang, Z. She and Y. Lin, A new isochroman derivative from the marine fungus Phomopsis sp. (no. ZH-111), Chem. Nat. Compd., 2011, 47, 13–16 CrossRef PubMed.
  36. R. K. Pettit, Mixed fermentation for natural product drug discovery, Appl. Microbiol. Biotechnol., 2009, 83, 19–25 CrossRef CAS PubMed.
  37. C. Y. Li, J. Zhang, J. S. Zhong, B. J. He, R. T. Li, Y. C. Liang, Y. C. Lin and S. N. Zhou, Isolation and identification of the metabolites from the mixed fermentation broth of two mangrove endophytic fungi, J. South China Agric. Univ., 2011, 32, 117–119 CAS.
  38. S. Klaiklay, V. Rukachaisirikul, Y. Sukpondma, S. Phongpaichit, J. Buatong and B. Bussaban, Metabolites from the mangrove-derived fungus Xylaria cubensis PSU-MA34, Arch. Pharmacal Res., 2012, 35, 1127–1131 CrossRef CAS PubMed.
  39. S. Li, M. Wei, G. Chen and Y. Lin, Two new dihydroisocoumarins from the endophytic fungus Aspergillus sp. collected from the south china sea, Chem. Nat. Compd., 2012, 48, 371–373 CrossRef CAS PubMed.
  40. V. Rukachaisirikul, A. Rodglin, Y. Sukpondma, S. Phongpaichit, J. Buatong and J. Sakayaroj, Phthalide and isocoumarin derivatives produced by an Acremonium sp. isolated from a mangrove Rhizophora apiculata, J. Nat. Prod., 2012, 75, 853–858 CrossRef CAS PubMed.
  41. L. Wen, Q. Wei, G. Chen, J. Cai and Z. She, Chemical constituents from the mangrove endophytic fungus Sporothrix sp., Chem. Nat. Compd., 2013, 49, 137–149 CrossRef CAS.
  42. R. A. Barrow and M. W. B. McCulloch, Linear naphtha-γ-pyrones: a naturally occurring scaffold of biological importance, Mini-Rev. Med. Chem., 2009, 9, 273–292 CrossRef CAS.
  43. Z. Huang, R. Yang, Z. Guo, Z. G. She and Y. C. Lin, A new naphtho-pyrone from mangrove endophytic fungus ZSU-H26, Chem. Nat. Compd., 2010, 46, 15–18 CrossRef CAS.
  44. J. Yang, F. Xu, C. Huang, J. Li, Z. She, Z. Pei and Y. Lin, Metabolites from the mangrove endophytic fungus Phomopsis sp. (# zsu-H76), Eur. J. Org. Chem., 2010, 3692–3695 CrossRef CAS.
  45. Z. Lin, T. Zhu, H. Wei, G. Zhang, H. Wang and Q. Gu, Spicochalasin A and new aspochalasins from the marine-derived fungus Spicaria elegans, Eur. J. Org. Chem., 2009, 3045–3051 CrossRef CAS.
  46. H. B. Huang, X. J. Feng, L. Liu, B. Chen, Y. J. Lu, L. Ma, Z. G. She and Y. C. Lin, Three dimeric naphtho-γ-pyrones from the mangrove endophytic fungus Aspergillus tubingensis isolated from Pongamia pinnata, Planta Med., 2010, 76, 1888–1891 CrossRef CAS PubMed.
  47. H. B. Huang, Z. E. Xiao, X. J. Feng, C. H. Huang, X. Zhu, J. H. Ju, M. F. Li, Y. C. Lin, L. Liu and Z. G. She, Cytotoxic naphtho-γ-pyrones from the mangrove endophytic fungus Aspergillus tubingensis (GX1-5E), Helv. Chim. Acta, 2011, 94, 1732–1740 CrossRef CAS.
  48. S. Klaiklay, V. Rukachaisirikul, K. Tadpetch, Y. Sukpondma, S. Phongpaichit, J. Buatong and J. Sakayaroj, Chlorinated chromone and diphenyl ether derivatives from the mangrove-derived fungus Pestalotiopsis sp. PSU-MA69, Tetrahedron, 2012, 68, 2299–2305 CrossRef CAS PubMed.
  49. J. Yang, R. Huang, S. X. Qiu, Z. She and Y. Lin, A new isobenzofuranone from the mangrove endophytic fungus Penicillium sp. (ZH58), Nat. Prod. Res., 2013, 27, 1902–1905 CrossRef CAS PubMed.
  50. Z. Huang, J. Yang, Z. She and Y. Lin, A new isoflavone from the mangrove endophytic fungus Fusarium sp. (ZZF60), Nat. Prod. Res., 2012, 26, 11–15 CrossRef CAS PubMed.
  51. K. K. Li, Y. J. Lu, X. H. Song, Z. G. She, X. W. Wu, L. K. An, C. X. Ye and Y. C. Lin, The metabolites of mangrove endophytic fungus Zh6-B1 from the South China Sea, Bioorg. Med. Chem. Lett., 2010, 20, 3326–3328 CrossRef CAS PubMed.
  52. C. H. Huang, J. H. Pan, B. Chen, M. Yu, H. B. Huang, X. Zhu, Y. J. Lu, Z. G. She and Y. C. Lin, Three bianthraquinone derivatives from the mangrove endophytic fungus Alternaria sp. ZJ9-6B from the South China Sea. Mar, Drugs, 2011, 9, 832–843 CAS.
  53. C. Huang, H. Jin, B. Song, X. Zhu, H. Zhao, J. Cai, Y. J. Lu, B. Chen and Y. Lin, The cytotoxicity and anticancer mechanisms of alterporriol L, a marine bianthraquinone, against MCF-7 human breast cancer cells, Appl. Microbiol. Biotechnol., 2012, 93, 777–785 CrossRef CAS PubMed.
  54. X. K. Xia, H. R. Huang, Z. G. She, C. L. Shao, F. Liu, X. L. Cai, L. L. P. Vrijmoed and Y. C. Lin, 1H and 13C NMR assignments for five anthraquinones from the mangrove endophytic fungus Halorosellinia sp. (no. 1403), Magn. Reson. Chem., 2007, 45, 1006–1009 CrossRef CAS PubMed.
  55. H. Chen, X. Zhu, L. L. Zhong, B. Yang, J. Li, J. H. Wu, S. P. Chen, Y. C. Lin, Y. H. Long and Z. G. She, Synthesis and antitumor activities of derivatives of the marine mangrove fungal metabolite deoxybostrycin, Mar. Drugs, 2012, 10, 2715–2728 CrossRef CAS PubMed.
  56. X. K. Xia, Q. Li, J. Li, C. L. Shao, J. Y. Zhang, Y. G. Zhang, X. Liu, Y. C. Lin, C. H. Liu and Z. G. She, Two new derivatives of griseofulvin from the mangrove endophytic fungus Nigrospora sp. (strain no. 1403) from Kandelia candel (L.) Druce, Planta Med., 2011, 77, 1735–1738 CrossRef CAS PubMed.
  57. C. Wang, J. Wang, Y. H. Huang, H. Chen, Y. Li, L. L. Zhong, Y. Chen, S. P. Chen, J. Wang, J. L. Kang, Y. Peng, B. Yang, Y. C. Lin, Z. G. She and X. M. Lai, Anti-mycobacterial activity of marine fungus-derived 4-deoxybostrycin and nigrosporin, Molecules, 2013, 18, 1728–1740 CrossRef CAS PubMed.
  58. Z. Shang, X. M. Li, C. S. Li and B. G. Wang, Diverse secondary metabolites produced by marine-derived fungus Nigrospora sp. MA75 on various culture media, Chem. Biodiversity, 2012, 9, 1338–1348 CAS.
  59. S. Klaiklay, V. Rukachaisirikul, S. Phongpaichit, C. Pakawatchai, S. Saithong, J. Buatong, S. Preedanon and J. Sakayaroj, Anthraquinone derivatives from the mangrove-derived fungus Phomopsis sp. PSU-MA214, Phytochem. Lett., 2012, 5, 738–742 CrossRef CAS PubMed.
  60. H. Sun, S. Gao, X. Li, C. Li and B. Wang, Chemical constituents of marine mangrove-derived endophytic fungus Alternaria tenuissima EN-192, Chin. J. Oceanol. Limnol., 2013, 31, 464–470 CrossRef CAS.
  61. C. M. Deng, S. X. Liu, C. H. Huang, J. Y. Pang and Y. C. Lin, Secondary metabolites of a mangrove endophytic fungus Aspergillus terreus (no. GX7-3B) from the South China Sea. Mar, Drugs, 2013, 11, 2616–2624 CAS.
  62. W. Ebrahim, A. H. Aly, A. Mándi, F. Totzke, M. H. Kubbutat, V. Wray, W. H. Lin, H. F. Dai, P. Proksch, T. Kurtán and A. Debbab, Decalactone derivatives from Corynespora cassiicola, an endophytic fungus of the mangrove plant Laguncularia racemosa, Eur. J. Org. Chem., 2012, 3476–3484 CrossRef CAS.
  63. H. J. Yan, X. M. Li, C. S. Li and B. G. Wang, Alkaloid and anthraquinone derivatives produced by the marine-derived endophytic fungus Eurotium rubrum, Helv. Chim. Acta, 2012, 95, 163–168 CrossRef CAS.
  64. K. S. Masters and S. Bräse, Xanthones from fungi, lichens, and bacteria: the natural products and their synthesis, Chem. Rev., 2012, 112, 3717–3776 CrossRef CAS PubMed.
  65. Z. Huang, R. Yang, Z. Guo, Z. G. She and Y. C. Lin, A new xanthone derivative from mangrove endophytic fungus no. ZSU-H16, Chem. Nat. Compd., 2010, 46, 348–351 CrossRef CAS.
  66. J. H. Pan, J. J. Deng, Y. G. Chen, J. P. Gao, Y. C. Lin, Z. G. She and Y. C. Gu, New lactone and xanthone derivatives produced by a mangrove endophytic fungus Phoma sp. SK3RW1M from the South China Sea, Helv. Chim. Acta, 2010, 93, 1369–1374 CrossRef CAS.
  67. C. Y. Li, J. Zhang, C. L. Shao, W. J. Ding, Z. G. She and Y. C. Lin, A new xanthone derivative from the co-culture broth of two marine fungi (strain no. E33 and K38), Chem. Nat. Compd., 2011, 47, 382–384 CrossRef CAS.
  68. Z. Huang, J. Yang, F. Lei, Z. She and Y. Lin, A New Xanthone O-glycoside from the mangrove endophytic fungus Phomopsis sp., Chem. Nat. Compd., 2013, 49, 27–30 CrossRef CAS.
  69. J. X. Yang, S. X. Qiu, Z. G. She and Y. C. Lin, A new xanthone derivative from the marine fungus Phomopsis sp. (no. SK7RN3G1), Chem. Nat. Compd., 2013, 49, 246–248 CrossRef CAS.
  70. M. Isaka, A. Jaturapat, K. Rukseree, K. Danwisetkanjana, M. Tanticharoen and Y. Thebtaranonth, Phomoxanthones A and B, novel xanthone dimers from the endophytic fungus Phomopsis species, J. Nat. Prod., 2001, 64, 1015–1018 CrossRef CAS PubMed.
  71. V. Rukachaisirikul, U. Sommart, S. Phongpaichit, J. Sakayaroj and K. Kirtikara, Metabolites from the endophytic fungus Phomopsis sp. PSU-D15, Phytochemistry, 2008, 69, 783–787 CrossRef CAS PubMed.
  72. Y. Shionoa, T. Sasakia, F. Shibuyaa, Y. Yasudaa, T. Kosekia and U. Supratman, Isolation of a phomoxanthone a derivative, a new metabolite of tetrahydroxanthone, from a Phomopsis sp. isolated from the mangrove, Rhizhopora mucronata, Nat. Prod. Commun., 2013, 8, 1735–1737 Search PubMed.
  73. M. Nawwar, S. Hussein, N. A. Ayoub, A. Hashim, G. Mernitz, B. Cuypers, M. Linscheid and U. Lindequist, Deuteromycols A and B, two benzofuranoids from a Red Sea marine-derived Deuteromycete sp., Arch. Pharmacal Res., 2010, 33, 1729–1733 CrossRef CAS PubMed.
  74. V. Cali, C. Spatafora and C. Tringali, Polyhydroxy-p-terphenyls and related p-terphenylquinones from fungi: overview and biological properties, Stud. Nat. Prod. Chem., 2003, 29, 263–307 CrossRef CAS.
  75. H. Wei, H. Inada, A. Hayashi, K. Higashimoto, P. Pruksakorn, S. Kamada, M. Arai, S. Ishida and M. Kobayashi, Prenylterphenyllin and its dehydroxyl analogs, new cytotoxic substances from a marine-derived fungus Aspergillus candidus IF10, J. Antibiot., 2007, 60, 586–590 CrossRef CAS PubMed.
  76. S. Cai, S. Sun, H. Zhou, X. Kong, T. Zhu, D. Li and Q. Gu, Prenylated polyhydroxy-p-terphenyls from Aspergillus taichungensis ZHN-7-07, J. Nat. Prod., 2011, 74, 1106–1110 CrossRef CAS PubMed.
  77. C. Paul and G. Pohnert, Production and role of volatile halogenated compounds from marine algae, Nat. Prod. Rep., 2011, 28, 186–195 RSC.
  78. Y. Zhang, X. M. Li, Z. Shang, C. S. Li, N. Y. Ji and B. G. Wang, Meroterpenoid and diphenyl ether derivatives from Penicillium sp. MA-37, a fungus isolated from marine mangrove rhizospheric soil, J. Nat. Prod., 2012, 75, 1888–1895 CrossRef CAS PubMed.
  79. J. Beau, N. Mahid, W. N. Burda, L. Harrington, L. N. Shaw, T. Mutka, D. E. Kyle, B. Barisic, A. Olphen and B. J. Baker, Epigenetic tailoring for the production of anti-infective cytosporones from the marine fungus Leucostoma persoonii, Mar. Drugs, 2012, 10, 762–774 CrossRef CAS PubMed.
  80. J. Wang, Z. Lu, P. Liu, Y. Wang, J. Li, K. Hong and W. Zhu, Cytotoxic polyphenols from the fungus Penicillium expansum 091006 endogenous with the mangrove plant Excoecaria agallocha, Planta Med., 2012, 78, 1861–1866 CrossRef CAS PubMed.
  81. E. L. Teuten, L. Xu and C. M. Reddy, Two abundant bioaccumulated halogenated compounds are natural products, Science, 2005, 307, 917–920 CrossRef CAS PubMed.
  82. A. Shimada, I. Takahashi, T. Kawano and Y. Z. Kimura, Chloroisosulochrin, chloroisosulochrin dehydrate, and pestheic acid, plant growth regulators, produced by Pestalotiopsis theae, Z. Naturforsch., B: J. Chem. Sci., 2001, 56, 797–803 CAS.
  83. J. Y. Li and G. A. Strobel, Jesterone and hydroxy-jesterone antioomycete cyclohexenone epoxides from the endophytic fungus Pestalotiopsis jesteri, Phytochemistry, 2001, 57, 261–265 CrossRef CAS.
  84. J. Wang, W. Ding, C. Li, S. Huang, Z. She and Y. Lin, A new polysubstituted benzaldehyde from the co-culture broth of two marine fungi (strains nos E33 and K38), Chem. Nat. Compd., 2013, 49, 799–802 CrossRef CAS.
  85. B. S. Gould and H. Raistrick, Studies in the biochemistry of micro-organisms: the crystalline pigments of species in the Aspergillus glaucus series, Biochem. J., 1934, 28, 1640–1656 CAS.
  86. H. Gao, T. Zhu, D. Li, Q. Gu and W. Liu, Prenylated indole diketopiperazine alkaloids from a mangrove rhizosphere soil derived fungus Aspergillus effuses H1-1, Arch. Pharmacal Res., 2013, 36, 952–956 CrossRef CAS PubMed.
  87. H. Y. Deng, J. G. Xing and D. Q. Luo, Metabolites of endophytic fungus Pestalotiopsis clavispora isolated from the stem of Bruguiera sexangula, Mycosystema, 2011, 30, 263–267 CAS.
  88. S. A. Ahmed, E. Bardshiri, C. R. McIntyre and T. J. Simpson, Biosynthetic-studies on tajixanthone and shamixanthone, polyketide hemiterpenoid metabolites of Aspergillus variecolor, Aust. J. Chem., 1992, 45, 249–274 CrossRef CAS.
  89. L. Wen, G. Chen, Z. She, C. Yan, J. Cai and L. Mu, Two new paeciloxocins from a mangrove endophytic fungus Paecilomyces sp., Russ. Chem. Bull., 2010, 59, 1656–1659 CrossRef CAS.
  90. J. G. Xing, H. Y. Deng and D. Q. Luo, Two new compounds from an endophytic fungus Pestalotiopsis heterocornis, J. Asian Nat. Prod. Res., 2011, 13, 1069–1073 CrossRef CAS PubMed.
  91. E. Yoshida, H. Fujimoto and M. Yamazaki, Isolation of three new azaphilones, luteusins C, D, and E, from an ascomycete, talaromyces luteus, Chem. Pharm. Bull., 1996, 44, 284–287 CrossRef CAS.
  92. K. Matsuzaki, H. Tahara, J. Inokoshi, H. Tanaka, R. Masuma and S. Omura, New brominated and halogen-less derivatives and structure–activity relationship of azaphilones inhibiting gp120-CD4 binding, J. Antibiot., 1998, 51, 1004–1011 CrossRef CAS.
  93. J. Y. Nam, H. K. Kim, J. Y. Kwon, M. Y. Han, K. H. Son, U. C. Lee, J. D. Choi and B. M. Kwon, 8-O-Methylsclerotiorinamine, antagonist of the Grb2-SH2 domain, isolated from Penicillium multicolor, J. Nat. Prod., 2000, 63, 1303 CrossRef CAS PubMed.
  94. R. C. Clark, S. Y. Lee, M. Searcey and D. L. Boger, The isolation, total synthesis and structure elucidation of chlorofusin, a natural product inhibitor of the p53-MDM2 protein–protein interaction, Nat. Prod. Rep., 2009, 26, 465–477 RSC.
  95. L. Musso, S. Dallavalle, L. Merlini, A. Bava, G. Nasini, S. Penco, G. Giannini, C. Giommarelli, A. De Cesare, V. Zuco, L. Vesci, C. Pisano, F. Dal Piaz, N. De Tommasi and F. Zunino, Natural and semisynthetic azaphilones as a new scaffold for Hsp90 inhibitors, Bioorg. Med. Chem., 2010, 18, 6031–6043 CrossRef CAS PubMed.
  96. N. Osmanova, W. Schultze and N. Ayoub, Azaphilones: a class of fungal metabolites with diverse biological activities, Phytochem. Rev., 2010, 9, 315–342 CrossRef CAS.
  97. J. M. Gao, S. X. Yang and J. C. Qin, Azaphilones: chemistry and biology, Chem. Rev., 2013, 113, 4755–4811 CrossRef CAS PubMed.
  98. R. Chong, R. R. King and W. B. Whalley, The chemistry of fungi. Part LXI. The synthesis of (±)-sclerotiorin, of (±)-4,6-dimethylocta-trans-2,trans-4-dienoic acid, and of an analogue of rotiorin, J. Chem. Soc. C, 1971, 3566–3571 RSC.
  99. J. Zhu, A. R. Germain and J. A. Porco, Synthesis of azaphilones and related molecules by employing cycloisomerization of O-alkynylbenzaldehydes, Angew. Chem., 2004, 116, 1259–1263 CrossRef.
  100. M. A. Marsini, K. M. Gowin and T. R. Pettus, Total synthesis of (±)-mitorubrinic acid, Org. Lett., 2006, 8, 3481–3483 CrossRef CAS PubMed.
  101. A. R. Germain, D. M. Bruggemeyer, J. Zhu, C. Genet, P. O’Brien and J. A. Porco Jr, Synthesis of the azaphilones (+)-sclerotiorin and (+)-8-O-methylsclerotiorinamine utilizing (+)-sparteine surrogates in copper-mediated oxidative dearomatization, J. Org. Chem., 2011, 76, 2577–2584 CrossRef CAS PubMed.
  102. H. B. Huang, X. J. Feng, Z. E. Xiao, L. Liu, H. X. Li, L. Ma, Y. J. Liu, J. H. Ju, Z. G. She and Y. C. Lin, Azaphilones and p-terphenyls from the mangrove endophytic fungus Penicillium chermesinum (ZH4-E2) isolated from the South China Sea, J. Nat. Prod., 2011, 74, 997–1002 CrossRef CAS PubMed.
  103. D. Liu, X. M. Li, L. Meng, C. S. Li, S. S. Gao, Z. Shang, P. Proksch, C. G. Huang and B. G. Wang, Nigerapyrones A–H, α-pyrone derivatives from the marine mangrove-derived endophytic fungus Aspergillus niger MA-132, J. Nat. Prod., 2011, 74, 1787–1791 CrossRef CAS PubMed.
  104. M. Kitagawa, T. Okabe, H. Ogino, H. Matsumoto, I. Suzuki-Takahashi, T. Kokubo, H. Higashi, S. Saitoh, Y. Taya, H. Yasuda, Y. Ohba, S. Nishimura, N. Tanaka and A. Okuyama, Butyrolactone I, a selective inhibitor of cdk2 and cdc2 kinase, Oncogene, 1993, 8, 2425–2432 CAS.
  105. Y. Shen, J. Zou, D. Xie, H. Ge, X. Cao and J. Dai, Butyrolactone and cycloheptanetrione from mangrove-associated fungus Aspergillus terreus, Chem. Pharm. Bull., 2012, 60, 1437–1441 CrossRef CAS PubMed.
  106. W. Ebrahim, A. H. Aly, V. Wray, P. Proksch and A. Debbab, Unusual octalactones from Corynespora cassiicola, an endophyte of Laguncularia racemosa, Tetrahedron Lett., 2013, 54, 6611–6614 CrossRef CAS PubMed.
  107. J. B. Helms and J. E. Rothman, Inhibition by brefeldin A of a Golgi membrane enzyme that catalyses exchange of guanine nucleotide bound to ARF, Nature, 1992, 360, 352–354 CrossRef CAS PubMed.
  108. S. F. Brady, M. M. Wagenaar, M. P. Singh, J. E. Janso and J. Clardy, The cytosporones, new octaketide antibiotics isolated from an endophytic fungus, Org. Lett., 2000, 2, 4043–4046 CrossRef CAS PubMed.
  109. J. D. Hall, N. W. Duncan-Gould, N. A. Siddiqi, J. N. Kelly, L. A. Hoeferlin, S. J. Morrison and J. K. Wyatt, Cytosporone E: racemic synthesis and preliminary antibacterial testing, Bioorg. Med. Chem., 2005, 13, 1409–1413 CrossRef CAS PubMed.
  110. T. Ohzeki and K. Mori, Synthetic racemate and enantiomers of cytosporone E, a metabolite of an endophytic fungus, show indistinguishably weak antimicrobial activity, Biosci., Biotechnol., Biochem., 2003, 67, 2584–2590 CrossRef CAS.
  111. V. Rukachaisirikul, A. Rodglin, S. Phongpaichit, J. Buatong and J. Sakayaroj, α-Pyrone and seiricuprolide derivatives from the mangrove-derived fungi Pestalotiopsis spp. PSU-MA92 and PSU-MA119, Phytochem. Lett., 2012, 5, 13–17 CrossRef CAS PubMed.
  112. J. Xu, A. H. Aly, V. Wray and P. Proksch, Polyketide Derivatives of Endophytic Fungus Pestalotiopsis sp. Isolated from the Chinese Mangrove Plant Rhizophora mucronata, Tetrahedron Lett., 2011, 52, 21–25 CrossRef CAS PubMed.
  113. Y. Song, J. Wang, H. Huang, L. Ma, J. Wang, Y. Gu, L. Liu and Y. Lin, Four eremophilane sesquiterpenes from the mangrove endophytic fungus Xylaria sp. BL321, Mar. Drugs, 2012, 10, 340–348 CrossRef CAS PubMed.
  114. N. Kiriyama, K. Nitta, Y. Sakacuchi, Y. Taguchi and Y. Yamamoto, Studies on the metabolic products of Aspergillus terreus. III. Metabolites of the strain IFO 8835. (1), Chem. Pharm. Bull., 1977, 25, 2593–2601 CrossRef CAS.
  115. R. Haritakun, P. Rachtawee, R. Chanthaket, N. Boonyuen and M. Isaka, Butyrolactones from the fungus Aspergillus terreus BCC 4651, Chem. Pharm. Bull., 2010, 58, 1545–1548 CrossRef CAS.
  116. K. Nishio, T. Ishida, H. Arioka, H. Kurokawa, K. Fukuoka, T. Nomoto, H. Fukumoto, H. Yokote and N. Saijo, Antitumor effects of butyrolactone I, a selective cdc2 kinase inhibitor, on human lung cancer cell lines, Anticancer Res., 1995, 16, 3387–3395 Search PubMed.
  117. T. G. Schimmel, A. D. Coffman and J. Parsons, Effect of butyrolactone I on the producing fungus, Aspergillus terreus, Appl. Environ. Microbiol., 1998, 64, 3707–3712 CAS.
  118. N. Ojima, S. Takenaka and S. Seto, New butenolides from Aspergillus terreus, Phytochemistry, 1973, 12, 2527–2529 CrossRef CAS.
  119. C. E. Caufield, S. A. Antane, K. M. Morris, S. M. Naughton, D. A. Quagliato, P. M. Andrae, A. Enos and J. F. Chiarello, PCT Int. Appl., WO 2005019196, 118, 2005.
  120. P. G. Cruz, D. S. Auld, P. J. Schultz, S. Lovell, K. P. Battaile, R. MacArthur, M. Shen, G. Tamayo-Castillo, J. Inglese and D. H. Sherman, Titration-based screening for evaluation of natural product extracts: identification of an aspulvinone family of luciferase inhibitors, Chem. Biol., 2011, 18, 1442–1452 CrossRef CAS PubMed.
  121. L. Vertesy, H. J. Burger, J. Kenja, M. Knauf, H. Kogler, E. F. Paulus, N. V. Ramakrishna, K. H. Swamy, E. K. Vijayakumar and P. Hammann, Kodaistatins, novel inhibitors of glucose-6-phosphate translocase T1 from Aspergillus terreus thom DSM 11247: isolation and structural elucidation, J. Antibiot., 2000, 53, 677–686 CrossRef CAS.
  122. H. Gao, W. Guo, Q. Wang, L. Zhang, M. Zhu, T. Zhu, Q. Gu and D. Li, Aspulvinones from a mangrove rhizosphere soil-derived fungus Aspergillus terreus Gwq-48 with anti-influenza A viral (H1N1) activity, Bioorg. Med. Chem. Lett., 2013, 23, 1776–1778 CrossRef CAS PubMed.
  123. M. S. C. Pedras and P. B. Chumala, Phomapyrones from blackleg causing phytopathogenic fungi: isolation, structure determination, biosyntheses and biological activity, Phytochemistry, 2005, 66, 81–87 CrossRef CAS PubMed.
  124. X. L. Yang and Z. Z. Li, New spiral γ-lactone enantiomers from the plant endophytic fungus Pestalotiopsis foedan, Molecules, 2013, 18, 2236–2242 CrossRef CAS PubMed.
  125. J. Dai, K. Krohn, U. Flörke, G. Pescitelli, G. Kerti, T. Papp, K. E. Kövér, A. C. Bényei, S. Draeger, B. Schulz and T. Kurtán, Curvularin-type metabolites from the fungus Curvularia sp. isolated from a marine alga, Eur. J. Org. Chem., 2011, 6928–6937 Search PubMed.
  126. K. Rudolph, A. Serwe and G. Erkel, Inhibition of TGF-β signaling by the fungal lactones (S)-curvularin, dehydrocurvularin, oxacyclododecindione and galiellalactone, Cytokine, 2013, 61, 285–296 CrossRef CAS PubMed.
  127. A. H. Aly, R. Edrada-Ebel, V. Wray and P. Proksch, Antitrypanosomal metabolites from an endophytic Penicillium sp. isolated from Limonium tubiflorum, Planta Med., 2010, 76, 402 CrossRef PubMed.
  128. N. Schmidt, J. Art, I. Forsch, A. Werner, G. Erkel, M. Jung, S. Horke, H. Kleinert and A. Pautz, The anti-inflammatory fungal compound (S)-curvularin reduces proinflammatory gene expression in an in vivo model of rheumatoid arthritis, J. Pharmacol. Exp. Ther., 2012, 343, 106–114 CrossRef CAS PubMed.
  129. Y. Xu, P. Espinosa-Artiles, V. Schubert, Y. M. Xu, W. Zhang, M. Lin, A. A. L. Gunatilaka, R. Süssmuth and I. Molnár, Characterization of the biosynthetic genes for 10,11-dehydrocurvularin, a heat shock response-modulating anticancer fungal polyketide from Aspergillus terreus, Appl. Environ. Microbiol., 2013, 79, 2038–2047 CrossRef CAS PubMed.
  130. L. H. Meng, X. M. Li, C. T. Lv, C. S. Li, G. M. Xu, C. G. Huang and B. G. Wang, 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 CrossRef CAS PubMed.
  131. D. Rönsberg, A. Debbab, A. Mándi, V. Wray, H. F. Dai, T. Kurtán, P. Proksch and A. H. Aly, Secondary metabolites from the endophytic fungus Pestalotiopsis virgatula isolated from the mangrove plant Sonneratia caseolaris, Tetrahedron Lett., 2013, 54, 3256–3259 CrossRef PubMed.
  132. G. C. Jiang, Y. C. Lin, S. N. Zhou, L. L. P. Vrijmoed and E. D. G. Jones, Studies on the secondary metabolites of mangrove fungus no. 1403 from the South China Sea, Acta Sci. Nat. Univ. Sunyatseni, 2000, 39, 68–72 CAS.
  133. M. H. Rønnest, B. Rebacz, L. Markworth, A. H. Terp, T. O. Larsen, A. Krämer and M. H. Clausen, Synthesis and structure–activity relationship of griseofulvin analogues as Inhibitors of centrosomal clustering in cancer cells, J. Med. Chem., 2009, 52, 3342–3347 CrossRef PubMed.
  134. L. Wen, Z. Guo, Q. Li, D. Zhang, Z. She and L. L. Vrijmoed, A new griseofulvin derivative from the mangrove endophytic fungus Sporothrix sp., Chem. Nat. Compd., 2010, 46, 363–365 CrossRef CAS.
  135. F. Liu, X. L. Cai, H. Yang, X. K. Xia, Z. Y. Guo, J. Yuan, M. F. Li, Z. G. She and Y. C. Lin, The bioactive metabolites of the mangrove endophytic fungus Talaromyces sp. ZH-154 isolated from Kandelia candel (L.) Druce, Planta Med., 2010, 76, 185–189 CrossRef CAS PubMed.
  136. Y. B. Zeng, H. Wang, W. J. Zuo, B. Zheng, T. Yang, H. F. Dai and W. L. Mei, A fatty acid glycoside from a marine-derived fungus isolated from mangrove plant Scyphiphora hydrophyllacea, Mar. Drugs, 2012, 10, 598–603 CrossRef CAS PubMed.
  137. D. H. Li, Z. Y. Liang, M. F. Guo, J. Zhou, X. B. Yang and J. Xu, Study on the chemical composition and extraction technology optimization of essential oil from Wedelia trilobata (L.) Hitchc, Afr. J. Biotechnol., 2012, 11(20), 4513–4517 CAS.
  138. S. Klaiklay, V. Rukachaisirikul, S. Phongpaichit, J. Buatong, S. Preedanon and J. Sakayaroj, Flavodonfuran: a new difuranylmethane derivative from the mangrove endophytic fungus Flavodon flavus PSU-MA201, Nat. Prod. Res., 2012, 1–5 Search PubMed.
  139. F. L. Sebastianes, N. Cabedo, N. El Aouad, A. M. Valente, P. T. Lacava, J. L. Azevedo, A. A. Pizzirani-Kleiner and D. Cortes, 3-Hydroxypropionic acid as an antibacterial agent from endophytic fungi Diaporthe phaseolorum, Curr. Microbiol., 2012, 65, 622–632 CrossRef CAS PubMed.
  140. B. V. Bhimba, A. C. Pushpam, P. Arumugam and S. Prakash, Phthalate derivatives from the marine fungi Phoma herbarum VB7, Int. J. Biol. Pharma. Res., 2012, 3, 507–512 Search PubMed.
  141. S. Nakahara, M. Kusano, S. Fujioka, A. Shimada and Y. Kimura, Penipratynolene, a novel nematicide from Penicillium bilaiae Chalabuda, Biosci., Biotechnol., Biochem., 2004, 68, 257–259 CrossRef CAS.
  142. D. C. Baldoqui, M. J. Kato, A. J. Cavalheiro, V. D. S. Bolzani, M. C. M. Young and M. Furlan, A chromene and prenylated benzoic acid from Piper aduncum, Phytochemistry, 1999, 51, 899–902 CrossRef CAS.
  143. D. A. Casteel, Peroxy natural products, Nat. Prod. Rep., 1999, 16, 55–73 RSC.
  144. V. M. Dembitsky, T. A. Gloriozova and V. V. Poroikov, Natural peroxy anticancer agents, Mini-Rev. Med. Chem., 2007, 7, 571–589 CrossRef CAS.
  145. V. M. Dembitsky, Bioactive peroxides as potential therapeutic agents, Eur. J. Med. Chem., 2008, 43, 223–251 CrossRef CAS PubMed.
  146. S. R. Meshnick, T. E. Taylor and S. Kamchonwongpaisan, Artemisinin and the antimalarial endoperoxides: from herbal remedy to targeted chemotherapy, Microbiol. Rev., 1996, 60, 301–315 CAS.
  147. S. Chokpaiboon, D. Sommit, T. Teerawatananond, N. Muangsin, T. Bunyapaiboonsri and K. Pudhom, Cytotoxic nor-chamigrane and chamigrane endoperoxides from a basidiomycetous fungus, J. Nat. Prod., 2010, 73, 1005–1007 CrossRef CAS PubMed.
  148. S. Chokpaiboon, D. Sommit, T. Bunyapaiboonsri, K. Matsubara and K. Pudhom, Antiangiogenic effect of chamigrane endoperoxides from a Thai mangrove-derived fungus, J. Nat. Prod., 2011, 74, 2290–2294 CrossRef CAS PubMed.
  149. H. Li, H. Huang, C. Shao, H. Huang, J. Jiang, X. Zhu, Y. Liu, L. Liu, Y. J. Lu, M. F. Li, Y. C. Lin and Z. She, Cytotoxic norsesquiterpene peroxides from the endophytic fungus Talaromyces flavus isolated from the mangrove plant Sonneratia apetala, J. Nat. Prod., 2011, 74, 1230–1235 CrossRef CAS PubMed.
  150. L. Y. Zang, W. Wei, Y. Guo, T. Wang, R. H. Jiao, S. W. Ng, R. X. Tan and H. M. Ge, Sesquiterpenoids from the mangrove-derived endophytic fungus Diaporthe sp., J. Nat. Prod., 2012, 75, 1744–1749 CrossRef CAS PubMed.
  151. B. M. Fraga, Natural sesquiterpenoids, Nat. Prod. Rep., 2011, 28, 1580–1610 RSC.
  152. B. M. Fraga, Natural sesquiterpenoids, Nat. Prod. Rep., 2013, 30, 1226–1264 RSC.
  153. Y. X. Song, J. Wang, S. W. Li, S. W. Li, B. Cheng, L. Li, B. Chen, L. Liu, Y. C. Lin and Y. C. Gu, Metabolites of the mangrove fungus Xylaria sp. BL321 from the south China sea, Planta Med., 2012, 78, 172–176 CrossRef CAS PubMed.
  154. W. A. Ayer and E. R. Cruz, The tremulanes, a new group of sesquiterpenes from the aspen rotting fungus Phellinus tremulae, J. Org. Chem., 1993, 58, 7529–7534 CrossRef CAS.
  155. S. B. Singh, P. Felock and D. J. Hazuda, Chemical and enzymatic modifications of integric acid and HIV1 integraseinhibitory activity, Bioorg. Med. Chem. Lett., 2000, 10, 235–238 CrossRef CAS.
  156. A. Ichihara, S. Sawamura and S. Sakamura, Structures of altiloxins A and B, phytotoxins from Phoma asparagi sacc, Tetrahedron Lett., 1984, 25, 3209–3212 CrossRef CAS.
  157. Y. Hemberger, J. Xu, V. Wray, P. Proksch, J. Wu and G. Bringmann, Pestalotiopens A and B: stereochemically challenging flexible sesquiterpene-cyclopaldic acid hybrids from Pestalotiopsis sp., Chem.–Eur. J., 2013, 19, 15556–15564 CrossRef CAS PubMed.
  158. C. M. Deng, C. H. Huang, Q. L. Wu, J. Y. Pang and Y. C. Lin, A new sesquiterpene from the mangrove endophytic fungus Aspergillus terreus (no. GX7-3B), Nat. Prod. Res., 2013, 27, 1882–1887 CrossRef CAS PubMed.
  159. M. Chinworrungsee, S. Wiyakrutta, N. Sriubolmas, P. Chuailua and A. Suksamrarn, Cytotoxic activities of trichothecenes isolated from an endophytic fungus belonging to order Hypocreales, Arch. Pharmacal Res., 2008, 31, 611–616 CrossRef CAS PubMed.
  160. W. A. Ayer and S. Miao, Secondary metabolites of the aspen fungus Stachybotrys cylindrospora, Can. J. Chem., 1993, 71, 487–493 CrossRef CAS.
  161. J. W. Harrison, R. M. Scrowston and B. Lythgoe, Taxine. Part IV. The constitution of taxine-I, J. Chem. Soc. C, 1966, 1933–1945 RSC.
  162. M. C. Wani, H. L. Taylor, M. E. Wall, P. Coggon and A. T. McPhail, Plant antitumor agents. VI. Isolation and structure of taxol, a novel antileukemic and antitumor agent from Taxus brevifolia, J. Am. Chem. Soc., 1971, 93, 2325–2327 CrossRef CAS.
  163. Y. F. Wang, Q. W. Shi, M. Dong, H. Kiyota, Y. C. Gu and B. Cong, Natural taxanes: developments since 1828, Chem. Rev., 2011, 111, 7652–7709 CrossRef CAS PubMed.
  164. A. Stierle, G. Strobel and D. Stierle, Taxol and taxane production by Taxomyces andreanae, Science, 1993, 260, 214–216 CAS.
  165. A. Stierle, G. Strobel, D. Stierle, P. Grothaus and G. Bignami, The search for taxol producing microorganisms among the endophytic fungi of the Pacific yew, Taxus brevifolia, J. Nat. Prod., 1995, 58, 1315–1324 CrossRef CAS.
  166. G. Strobel, X. S. Yang, J. Sears, R. Kramer, R. S. Sidhu and W. M. Hess, Taxol from Pestalotiopsis microspora, an endophytic fungus of Taxus wallachiana, Microbiology, 1996, 142, 435–440 CrossRef CAS PubMed.
  167. J. Chen, X. Qiu, R. Wang, L. Duan, S. Chen, J. Luo and L. Kong, Inhibition of human gastric carcinoma cell growth in vitro and in vivo by cladosporol isolated from the paclitaxel-producing strain Alternaria alternata var. monosporus, Biol. Pharm. Bull., 2009, 32, 2072–2074 CAS.
  168. F. Xu, W. Tao, L. Cheng and L. Guo, Strain improvement and optimization of the media of taxol-producing fungus Fusarium maire, Biochem. Eng. J., 2006, 31, 67–73 CrossRef CAS PubMed.
  169. J. Y. Li, R. S. Sidhu, E. J. Ford, D. M. Long, W. M. Hess and G. A. Strobel, The induction of taxol production in the endophytic fungus—Periconia sp. from Torreya grandifolia, J. Ind. Microbiol. Biotechnol., 1998, 20, 259–264 CrossRef CAS.
  170. A. Elavarasi, G. S. Rathna and M. Kalaiselvam, Taxol producing mangrove endophytic fungi Fusarium oxysporum from Rhizophora annamalayana, Asian Pac. J. Trop. Biomed., 2012, 2, S1081–S1085 CrossRef.
  171. R. Ebel, Terpenes from marine-derived fungi, Mar. Drugs, 2010, 8, 2340–2368 CrossRef CAS PubMed.
  172. M. Cueto, P. R. Jensen and W. Fenical, Aspergilloxide, a novel sesterterpene epoxide from a marine-derived fungus of the genus Aspergillus, Org. Lett., 2002, 4, 1583–1585 CrossRef CAS PubMed.
  173. M. K. Renner, P. R. Jensen and W. Fenical, Mangicols: structures and biosynthesis of a new class of sesterterpene polyols from a marine fungus of the genus Fusarium, J. Org. Chem., 2000, 65, 4843–4852 CrossRef CAS PubMed.
  174. X. S. Huang, H. B. Huang, H. X. Li, X. F. Sun, H. R. Huang, Y. J. Lu, Y. C. Lin, Y. H. Long and Z. G. She, 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 CrossRef CAS PubMed.
  175. Z. E. Xiao, H. R. Huang, C. L. Shao, X. K. Xia, L. Ma, X. S. Huang, Y. J. Lu, Y. C. Lin and Z. G. She, Asperterpenols A and B, new sesterterpenoids isolated from a mangrove endophytic fungus Aspergillus sp. 085242, Org. Lett., 2013, 15, 2522–2525 CrossRef CAS PubMed.
  176. D. Q. Luo, H. Y. Deng, X. L. Yang, B. Z. Shi and J. Z. Zhang, Oleanane-type triterpenoids from the endophytic fungus Pestalotiopsis clavispora isolated from the Chinese mangrove plant Bruguiera sexangula, Helv. Chim. Acta, 2011, 94, 1041–1047 CrossRef CAS.
  177. M. Menna, C. Imperatore, F. D'Aniello and A. Aiello, Meroterpenes from Marine invertebrates: structures, occurrence, and ecological implications, Mar. Drugs, 2013, 11, 1602–1643 CrossRef CAS PubMed.
  178. B. T. M. Schurmann, W. S. T. Sallum and J. A. Takahashi, Austin, dehydroaustin and other metabolites from Penicillium brasilianum, Quim. Nova, 2010, 33, 1044–1046 CrossRef CAS PubMed.
  179. R. M. Geris dos Santos and E. Rodrigues-Filho, Meroterpenes from Penicillium sp. found in associating with Melia azedarach, Phytochemistry, 2002, 61, 907–912 CrossRef CAS.
  180. R. M. Geris dos Santos and E. Rodrigues-Filho, Structures of meroterpenes produced by Penicillium sp., an endophytic fungus found associated with Melia azedarach, J. Braz. Chem. Soc., 2003, 14, 722–727 CrossRef CAS PubMed.
  181. K. K. Chexal, J. P. Springer, J. Clardy, R. J. Cole, J. W. Kirksey, J. W. Domer, H. G. Cutler and B. J. Sterawter, Austin, a novel polyisoprenoid mycotoxin from Aspergillus ustus, J. Am. Chem. Soc., 1976, 98, 6748–6750 CrossRef CAS.
  182. R. Geris, E. Rodrigues-Fo, H. H. Garcia da Silva and I. Garcia da Silva, Larvicidal effects of fungal meroterpenoids in the control of Aedes aegypti L., the main vector of dengue and yellow fever, Chem. Biodiversity, 2008, 5, 341–345 CAS.
  183. H. Hayashi, M. Mukaihara, S. Murao, M. Arai, A. Y. Lee and J. Clardy, Acetoxydehydroaustin, a new bioactive compound, and related compound neoaustin from Penicillium sp. MG-11, Biosci., Biotechnol., Biochem., 1994, 58, 334–338 CrossRef CAS.
  184. Y. X. Song, L. T. Qiao, J. J. Wang, H. M. Zeng, Z. G. She, C. D. Miao, K. Hong, Y. C. Gu, L. Liu and Y. C. Lin, Two new meroterpenes from the mangrove endophytic fungus Aspergillus sp. 085241B, Helv. Chim. Acta, 2011, 94, 1875–1880 CrossRef CAS.
  185. G. J. Zhang, S. W. Sun, T. J. Zhu, Z. J. Lin, J. Y. Gu, D. H. Li and Q. Q. Gu, Antiviral isoindolone derivatives from an endophytic fungus Emericella sp. associated with Aegiceras corniculatum, Phytochemistry, 2011, 72, 1436–1442 CrossRef CAS PubMed.
  186. S. Kataoka, S. Furutani, K. Hirata, H. Hayashi and K. Matsuda, Three Austin family compounds from Penicillium brasilianum exhibit selective blocking action on cockroach nicotinic acetylcholine receptors, Neurobehav. Toxicol., 2011, 32, 123–129 CAS.
  187. J. Xu, S. S. Ebada and P. Proksch, Chemistry and bioactivity of natural products from the fungal genus Pestalotiopsis, Fungal Diversity, 2010, 44, 15–31 CrossRef.
  188. X. Huang, X. Sun, B. Ding, M. Lin, L. Liu, H. Huang and Z. She, A new anti-acetylcholinesterase α-pyrone meroterpene, arigsugacin I, from mangrove endophytic fungus Penicillium sp. sk5GW1L of Kandelia candel, Planta Med., 2013, 79, 1572–1575 CrossRef CAS PubMed.
  189. X. Peng, Y. Wang, K. Sun, P. Liu, X. Yin and W. Zhu, Cerebrosides and 2-pyridone alkaloids from the halotolerant fungus Penicillium chrysogenum grown in a hypersaline medium, J. Nat. Prod., 2011, 74, 1298–1302 CrossRef CAS PubMed.
  190. A. H. Aly, A. Debbab and P. Proksch, Fungal endophytes: unique plant inhabitants with great promises, Appl. Microbiol. Biotechnol., 2011, 90, 1829–1845 CrossRef CAS PubMed.
  191. R. G. Micetich and J. C. MacDonald, Metabolites of Aspergillus sclerotiorum Huber, J. Chem. Soc., 1964, 1507–1510 RSC.
  192. F. Zhu, G. Chen, X. Chen, M. Huang and X. Wan, Aspergicin, a new antibacterial alkaloid produced by mixed fermentation of two marine-derived mangrove epiphytic fungi, Chem. Nat. Compd., 2011, 47, 767–769 CrossRef CAS PubMed.
  193. J. Wang, L. Wang, M. Chang, R. F. He, Q. H. Zhang and B. P. Ye, The endophytic fungus AGR12 in the stem of Rhizophora stylosa Grif and its antibacterial metabolite, Chin. J. Antibiot., 2011, 36, 102–106 CAS.
  194. M. H. Wheeler, R. D. Stipanovic and L. S. Puckhaber, Phytotoxicity of equisetin and epi-equisetin isolated from Fusarium equiseti and F. pallidoroseum, Mycol. Res., 1999, 103, 967–973 CrossRef CAS.
  195. H. R. Burmeister, G. A. Bennett, R. F. Vesonder and C. W. Hesseltine, Antibiotic produced by Fusarium equiseti NRRL 5537, Antimicrob. Agents Chemother., 1974, 5, 634–639 CrossRef CAS.
  196. T. König, A. Kapus and B. Sarkadi, Effects of equisetin on rat liver mitochondria: evidence for inhibition of substrate anion carriers of the inner membrane, J. Bioenerg. Biomembr., 1993, 25, 537–545 CrossRef.
  197. L. Ding, H. M. Dahse and C. Hertweck, Cytotoxic alkaloids from Fusarium incarnatum associated with the mangrove tree Aegiceras corniculatum, J. Nat. Prod., 2012, 75, 617–621 CrossRef CAS PubMed.
  198. S. P. Putri, H. Kinoshita, F. Ihara, Y. Igarashi and T. Nihira, Farinomalein, a maleimide-bearing compound from the entomopathogenic fungus Paecilomyces farinosus, J. Nat. Prod., 2009, 72, 1544–1546 CrossRef CAS PubMed.
  199. J. Xu, J. Kjer, J. Sendker, V. Wray, H. S. Guan, R. A. Edrada, W. E. G. Müller, M. Bayer, W. H. Lin, J. Wu and P. Proksch, Cytosporones, coumarins, and an alkaloid from the endophytic fungus Pestalotiopsis sp. isolated from the Chinese mangrove plant Rhizophora mucronata, Bioorg. Med. Chem., 2009, 17, 7362–7367 CrossRef CAS PubMed.
  200. M. El Amrani, A. Debbab, A. H. Aly, V. Wray, S. Dobretsov, W. E. Müller, W. H. Lin, D. W. Lai and P. Proksch, Farinomalein derivatives from an unidentified endophytic fungus isolated from the mangrove plant Avicennia marina, Tetrahedron Lett., 2012, 53, 6721–6724 CrossRef CAS PubMed.
  201. K. Scherlach and C. Hertweck, Discovery of aspoquinolones A–D, prenylated quinoline-2-one alkaloids from Aspergillus nidulans, motivated by genome mining, Org. Biomol. Chem., 2006, 4, 3517–3520 CAS.
  202. T. Liu, Z. L. Li, Y. Wang, L. Tian, Y. H. Pei and H. M. Hua, A new alkaloid from the marine-derived fungus Hypocrea virens, Nat. Prod. Res., 2011, 25, 1596–1599 CrossRef CAS PubMed.
  203. C. Y. An, X. M. Li, H. Luo, C. S. Li, M. H. Wang, G. M. Xu and B. G. Wang, 4-Phenyl-3,4-dihydroquinolone derivatives from Aspergillus nidulans MA-143, an endophytic fungus isolated from the mangrove plant Rhizophora stylosa, J. Nat. Prod., 2013, 76, 1896–1901 CrossRef CAS PubMed.
  204. F. Zhu, G. Chen, J. Wu and J. Pan, Structure revision and cytotoxic activity of marinamide and its methyl ester, novel alkaloids produced by co-cultures of two marine-derived mangrove endophytic fungi, Nat. Prod. Res., 2013, 27, 1960–1964 CrossRef CAS PubMed.
  205. D. L. Li, X. M. Li, T. G. Li, H. Y. Dang and B. G. Wang, Dioxopiperazine alkaloids produced by the marine mangrove derived endophytic fungus Eurotium rubrum, Helv. Chim. Acta, 2008, 91, 1888–1893 CrossRef CAS.
  206. M. T. Liu, S. Lin, M. L. Gan, M. H. Chen, L. Li, S. J. Wang, J. C. Zi, X. N. Fan, Y. Liu, Y. K. Si, Y. C. Yang, X. G. Chen and J. G. Shi, Yaoshanenolides A and B: new spirolactones from the bark of Machilus yaoshansis, Org. Lett., 2012, 14, 1004–1007 CrossRef CAS PubMed.
  207. A. Dossena, R. Marchelli and A. Pochini, New metabolites of Aspergillus amstelodami related to the biogenesis of neoechinulin, J. Chem. Soc., Chem. Commun., 1974, 771–772 RSC.
  208. S. W. Sun, C. Z. Ji, Q. Q. Gu, D. H. Li and T. J. Zhu, Three new polyketides from marine-derived fungus Aspergillus glaucus HB1-19, J. Asian Nat. Prod. Res., 2013, 15, 956–961 CrossRef CAS PubMed.
  209. L. B. Sokolov, L. E. Alekseeva, V. O. Kul'bakh, N. A. Kuznetsova and V. S. Nyn, Study of the chemical composition of aspergin – a metabolite from Aspergillus sp., Antibiotiki, 1971, 16, 504–510 CAS.
  210. H. Gao, W. Liu, T. Zhu, X. Mo, A. Mándi, T. Kurtán, J. Li, J. Ai, Q. Q. Gu and D. Li, Diketopiperazine alkaloids from a mangrove rhizosphere soil derived fungus Aspergillus effuses H1-1, Org. Biomol. Chem., 2012, 10, 9501–9506 CAS.
  211. S. Saikia, M. J. Nicholson, C. Young, E. J. Parker and B. Scott, The genetic basis for indole-diterpene chemical diversity in filamentous fungi, Mycol. Res., 2008, 112, 184–199 CrossRef CAS PubMed.
  212. K. M. Byrne, S. K. Smith and J. G. Ondeyka, Biosynthesis of nodulisporic acid A: precursor studies, J. Am. Chem. Soc., 2002, 124, 7055–7060 CrossRef CAS PubMed.
  213. C. Li, J. B. Gloer, D. T. Wicklow and P. F. Dowd, Thiersinines A and B: novel antiinsectan indole diterpenoids from a new fungicolous Penicillium species (NRRL 28147), Org. Lett., 2002, 18, 3095–3098 CrossRef PubMed.
  214. J. G. Ondeyka, G. L. Helms, O. D. Hensens, M. A. Goetz, D. L. Zink, A. Tsipouras, W. L. Shoop, L. Slayton, A. W. Dombrowski, J. D. Polishook, D. A. Ostlind, N. N. Tsou, R. G. Ball and S. B. Singh, Nodulisporic acid A, a novel and potent insecticide from a Nodulisporium sp. Isolation, structure determination, and chemical transformations, J. Am. Chem. Soc., 1997, 119, 8809–8816 CrossRef CAS.
  215. O. D. Hensens, J. G. Ondeyka, A. W. Dombrowski, D. A. Ostlind and D. L. Zink, Isolation and structure of nodulisporic acid A1 and A2, novel insecticides from a Nodulisporium sp., Tetrahedron Lett., 1999, 40, 5455–5458 CrossRef CAS.
  216. H. G. Knaus, A. Eberhart, H. Glossmann, P. Munujos, G. J. Kaczorowski, O. B. McManus, S. H. Lee, W. A. Schmalhofer, M. Garcia-Calvo, L. M. H. Helms, M. Sanchez, K. Giangiacomo, J. P. Reuben, A. B. Smith, G. Kaczorowshi and M. L. Garcia, Tremorgenic indole alkaloids potently inhibit smooth muscle high-conductance calcium-activated potassium channels, Biochemistry, 1994, 33, 5819–5828 CrossRef CAS.
  217. M. F. Qiao, N. Y. Ji, X. H. Liu, K. Li, Q. M. Zhu and Q. Z. Xue, Indoloditerpenes from an algicolous isolate of Aspergillus oryzae, Bioorg. Med. Chem. Lett., 2010, 20, 5677–5680 CrossRef CAS PubMed.
  218. L. Ding, A. Maier, H. H. Fiebig, H. Görls, W. H. Lin, G. Peschel and C. Hertweck, Divergolides A–D from a mangrove endophyte reveal an unparalleled plasticity in ansa-macrolide biosynthesis, Angew. Chem., Int. Ed., 2011, 50, 1630–1634 CrossRef CAS PubMed.
  219. Y. Fan, Y. Wang, P. Liu, P. Fu, T. Zhu, W. Wang and W. Zhu, Indole-diterpenoids with anti-H1N1 activity from the aciduric fungus Penicillium camemberti OUCMDZ-1492, J. Nat. Prod., 2013, 76, 1328–1336 CrossRef CAS PubMed.
  220. J. Kobayashi, G. C. Harbour, J. Gilmore and K. L. Rinehart, Eudistomins A, D, G, H, I, J, M, N, O, P, and Q, bromo-, hydroxyl-, pyrrolyl-, and 1-pyrrolinyl-β-carbolines from the antiviral caribbean tunicate Eudistoma olivaceum, J. Am. Chem. Soc., 1984, 106, 1526–1528 CrossRef CAS.
  221. H. Nakamura, S. Deng, J. I. Kobayashi, Y. Ohizumi, Y. Tomotake, T. Matsuzaki and Y. Hirata, Keramamine-A and -B, novel antimicrobial alkaloids from the Okinawan marine sponge Pellina sp., Tetrahedron Lett., 1987, 28, 621–624 CrossRef CAS.
  222. N. Abe, Y. Nakakita, T. Nakamura, N. Enoki, H. Uchida, S. Takeo and M. Munekata, Novel cytocidal compounds, oxopropalines from Streptomyces sp. G324 producing lavendamycin. I. Taxonomy of the producing organism, fermentation, isolation and biological activities, J. Antibiot., 1993, 46, 1672–1677 CrossRef CAS.
  223. N. Abe, N. Enoki, Y. Nakakita, H. Uchida, T. Nakamura and M. Munekata, Novel cytocidal compounds, oxopropalines from Streptomyces sp. G324 producing lavendamycin. II. Physico-chemical properties and structure elucidations, J. Antibiot., 1993, 46, 1678–1686 CrossRef CAS.
  224. L. K. Larsen, R. E. Moore and G. M. Patterson, β-Carbolines from the blue-green alga Dichothrix baueriana, J. Nat. Prod., 1994, 57, 419–421 CrossRef CAS.
  225. J. Peng, T. Lin, W. Wang, Z. Xin, T. Zhu, Q. Gu and D. Li, Antiviral alkaloids produced by the mangrove-derived fungus Cladosporium sp. PJX-41, J. Nat. Prod., 2013, 76, 1133–1140 CrossRef CAS PubMed.
  226. C. Y. An, X. M. Li, C. S. Li, M. H. Wang, G. M. Xu and B. G. Wang, Aniquinazolines A–D, four new quinazolinone alkaloids from marine-derived endophytic fungus Aspergillus nidulans, Mar. Drugs, 2013, 11, 2682–2694 CrossRef CAS PubMed.
  227. W. Ebrahim, J. Kjer, M. El Amrani, V. Wray, W. Lin, R. Ebel, D. W. Lai and P. Proksch, Pullularins E and F, two new peptides from the endophytic fungus Bionectria ochroleuca isolated from the mangrove plant Sonneratia caseolaris, Mar. Drugs, 2012, 10, 1081–1091 CrossRef CAS PubMed.
  228. S. B. Singh, D. L. Zink, J. D. Polishook, A. W. Dombrowski, S. J. Darkin-Rattray, D. M. Schmatz and M. A. Goetz, Apicidins: novel cyclic tetrapeptides as coccidiostats and antimalarial agents from Fusarium pallidoroseum, Tetrahedron Lett., 1996, 37, 8077–8080 CrossRef CAS.
  229. T. Amagata, B. I. Morinaka, A. Amagata, K. Tenney, F. A. Valeriote, E. Lobkovsky, J. Clardy and P. Crews, A chemical study of cyclic depsipeptides produced by a sponge-derived fungus, J. Nat. Prod., 2006, 69, 1560–1565 CrossRef CAS PubMed.
  230. M. Chu, R. Mierzwa, I. Truumees, F. Gentile, M. Patel, V. Gullo, T.-M. Chan and M. S. Puar, Two novel diketopiperazines isolated from the fungus Tolypocladium sp., Tetrahedron Lett., 1993, 34, 7537–7540 CrossRef CAS.
  231. C. J. Barrow, P. Cai, J. K. Snyder, D. M. Sedlock, H. H. Sun and R. Cooper, WIN 64821, a new competitive antagonist to substance P, isolated from an Aspergillus species: structure determination and solution conformation, J. Org. Chem., 1993, 58, 6016–6021 CrossRef CAS.
  232. D. Liu, X. M. Li, C. S. Li and B. G. Wang, Nigerasterols A and B, Antiproliferative sterols from the mangrove-derived endophytic fungus Aspergillus niger MA-132, Helv. Chim. Acta, 2013, 96, 1055–1061 CrossRef CAS.
  233. J. H. Pan, Y. Chen, Y. H. Huang, Y. W. Tao, J. Wang, Y. Li, T. Dong, X. M. Lai and Y. C. Lin, Antimycobacterial activity of fusaric acid from a mangrove endophyte and its metal complexes, Arch. Pharmacal Res., 2011, 34, 1177–1181 CrossRef CAS PubMed.
  234. Y. Yaoita, K. Amemiya, H. Ohnuma, K. Furumura, A. Masaki, T. Matsuki and M. Kikuchi, Sterol constituents from five edible mushrooms, Chem. Pharm. Bull., 1998, 46, 944–950 CrossRef CAS.
  235. Y. Yaoita, K. Matsuki, T. Iijima, S. Nakano, R. Kakuda, K. Machida and M. Kikuchi, New sterols and triterpenoids from four edible mushrooms, Chem. Pharm. Bull., 2001, 49, 589–594 CrossRef CAS.
  236. Y. Sera, K. Adachi and Y. Shizuri, A new epidioxy sterol as an antifouling substance from a Palauan marine sponge, Lendenfeldia chondrodes, J. Nat. Prod., 1999, 62, 152–154 CrossRef CAS PubMed.
  237. S. Yu, Z. Deng, L. van Ofwegen, P. Proksch and W. Lin, 5,8-Epidioxysterols and related derivatives from a Chinese soft coral Sinularia flexibilis, Steroids, 2006, 71, 955–959 CrossRef CAS PubMed.
  238. M. E. Rateb and R. Ebel, Secondary metabolites of fungi from marine habitats, Nat. Prod. Rep., 2011, 28, 290–344 RSC.
  239. Y. Nozawa, N. Sakai, K. Matsumoto and K. Mizoue, A novel neuritogenic compound, NGA0187, J. Antibiot., 2002, 55, 629–634 CrossRef CAS.
  240. Z. Hua, D. A. Carcache, Y. Tian, Y. M. Li and S. J. Danishefsky, The synthesis and preliminary biological evaluation of a novel steroid with neurotrophic activity: NGA0187, J. Org. Chem., 2005, 70, 9849–9856 CrossRef CAS PubMed.
  241. A. A. L. Gunatilaka, Y. Gopichand, F. J. Schmitz and C. Djerassi, Minor and trace sterols in marine invertebrates. 26. Isolation and structure elucidation of nine new 5α,8α-epidoxy sterols from four marine organisms, J. Org. Chem., 1981, 46, 3860–3866 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2015