Next Article in Journal
Protective Effects of a Jellyfish-Derived Thioredoxin Fused with Cell-Penetrating Peptide TAT-PTD on H2O2-Induced Oxidative Damage
Next Article in Special Issue
Resveratrol May Reduce the Degree of Periodontitis by Regulating ERK Pathway in Gingival-Derived MSCs
Previous Article in Journal
Identification of Signature Genes of Dilated Cardiomyopathy Using Integrated Bioinformatics Analysis
Previous Article in Special Issue
α,ω-Diacyl-Substituted Analogues of Natural and Unnatural Polyamines: Identification of Potent Bactericides That Selectively Target Bacterial Membranes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lichen-Derived Actinomycetota: Novel Taxa and Bioactive Metabolites

Yunnan Institute of Microbiology, School of Life Sciences, Yunnan University, Kunming 650091, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(8), 7341; https://doi.org/10.3390/ijms24087341
Submission received: 20 February 2023 / Revised: 13 March 2023 / Accepted: 10 April 2023 / Published: 16 April 2023
(This article belongs to the Special Issue Natural Bioactive Compounds: Design, Synthesis and Characterization)

Abstract

:
Actinomycetes are essential sources of numerous bioactive secondary metabolites with diverse chemical and bioactive properties. Lichen ecosystems have piqued the interest of the research community due to their distinct characteristics. Lichen is a symbiont of fungi and algae or cyanobacteria. This review focuses on the novel taxa and diverse bioactive secondary metabolites identified between 1995 and 2022 from cultivable actinomycetota associated with lichens. A total of 25 novel actinomycetota species were reported following studies of lichens. The chemical structures and biological activities of 114 compounds derived from the lichen-associated actinomycetota are also summarized. These secondary metabolites were classified into aromatic amides and amines, diketopiperazines, furanones, indole, isoflavonoids, linear esters and macrolides, peptides, phenolic derivatives, pyridine derivatives, pyrrole derivatives, quinones, and sterols. Their biological activities included anti-inflammatory, antimicrobial, anticancer, cytotoxic, and enzyme-inhibitory actions. In addition, the biosynthetic pathways of several potent bioactive compounds are summarized. Thus, lichen actinomycetes demonstrate exceptional abilities in the discovery of new drug candidates.

1. Introduction

Lichens form important symbiotic communities in the ecosystem and are characterized by a symbiotic association between fungi and algae. They occupy 8% of the earth’s surface [1]. Some bioactive compounds, such as usnic acid, gyrophoric acid, diffractaic acid, polysaccharides, anthraquinones, and terpenes, have been isolated from lichens, and some of these compounds have been employed in clinical treatments [2]. Organisms with a slower growth rate reportedly exhibit strong resistance to external secondary metabolism [3]. Further, organisms that move slowly and thrive in low-resource environments produce large amounts of defensive metabolites for protection against their many predators. Lichens and their symbiotic organisms, especially actinomycetes, grow slowly. They are natural habitats for the production of beneficial bioactive compounds or metabolites.
Antibiotics produced by microorganisms contribute significantly to human health. Actinomycetes are an essential resource for the discovery of drug-lead compounds. Actinomycete drug resources have been utilized and developed for many years. Hence, identifying new active structural substances has become increasingly difficult [4]. Lichens are a unique group of organisms formed by symbioses between fungi and algae or cyanobacteria [5]. The wide variety of lichens can provide new sources of actinomycetes for use in the discovery of novel drugs [6]. Only a few national and international research groups have investigated actinomycete lichen resources. Moreover, few groups have reported active metabolites derived from actinomycetes.
Lichen-derived actinomycetota are potent producers of bioactive metabolites. This review summarizes the compounds isolated from lichen-derived actinomycetes. The compounds are classified into 11 types based on their different structures. Some secondary metabolites exhibit various biological activities, such as anti-inflammatory, antimicrobial, anticancer, cytotoxic, and enzyme-inhibitory activities. The chemical structures of 114 secondary metabolites isolated from lichen-associated actinomycetes and 25 novel actinomycetota species are listed in this review. Furthermore, the biological activities of some lichen actinomycetota have also been investigated, although the effective chemical components of these strains are still unknown. The biosynthetic pathways of some unique secondary metabolites from lichen-derived actinomycetes are also reported here.

2. Novel Actinomycetota Taxa

The phylum “Actinobacteria” was modified to “Actinomycetota” by Goodfellow in 2021 [7]. Therefore, we use “actinomycetota” in the present article. Actinomycetes are extensively dispersed. The 25 novel species isolated from lichens between 2007 and 2022 are described in Table 1. In Figure 1, the red dots indicate the collection sites of the lichen samples from which new species of actinomycetes were identified. Most published literature suggests the distribution of new lichen-associated actinomycetota in Asia, especially in Yunnan Province, China. Among the 25 isolated species, 6 actinomycete species (24%) belonged to the family Microbacteriaceae. Three (12%) species belonged to each of the families of Micromonosporaceae, Pseudonocardiaceae, and Streptomycetaceae. The remaining actinomycetes came from the families Nakamurellaceae, Rhodobacteraceae, Streptomycetaceae, and others (Figure 2).
The humic acid–vitamin agar selective medium was mainly used to isolate most of the novel actinomycetota species (Table 1). A few novel species were isolated using standard growth media such as ISP2 medium and potato dextrose agar. Most species were incubated for 1–4 weeks at 25–30 °C.

3. Natural Products from Lichen-Associated Actinomycetota

A total of 114 compounds from lichen-associated actinomycetes were reported. Based on their structural characteristics, they were classified into diketopiperazines, peptides, indoles, furanones, quinones, isoflavonoids, etc. Some exhibit anti-inflammatory, antimicrobial, anticancer, cytotoxic, and enzyme-inhibitory activities. These compounds are promising small molecules for advancement as new drug and pesticide candidates.

3.1. N-Containing Compounds

3.1.1. Diketopiperazines

Diketopiperazines are the smallest cyclic dipeptides formed by the double condensation of two α-amino acids. They have a stable six-membered ring that serves as an important pharmacophore [32]. Diketopiperazines are structures with numerous biological functions of interest to natural product researchers.
Cyclo (Gly-L-Ala) (1) and 5-methyl-uracil (2) were isolated from the actinomycete Amycolatopsis sp. YIM 130687, collected from the Jinsha River region in Yunnan Province [33]. Compound 1 exhibited weak activity against Escherichia coli and Salmonella typhimurium, and 2 showed weak inhibitory activity against S. typhimurium. Cyclo (L-Pro-L-Val) (3) was isolated from Nocardia ignorata [34]. Cyclo-(Ala-Leu) (4), cyclo-(Gly-Phe) (5), cyclo-(Leu-Tyr) (6), and cyclo-(Phe-Tyr) (7) were obtained from Amycolatopsis sp. YIM 130932, associated with the lichen Punctelia rudecta found in Yunnan Province [35]. Cyclo (L-Pro-L-OMet) (8) was isolated from Nocardia ignorata [34]. Cyclo-(Phe-Pro) (9) and cyclo-(L-Leu-L-Pro) (10) were isolated from Streptomyces cyaneofuscatus MOLA 1488, associated with the marine lichen Lichina confinis [36]. The crude extract of this strain showed anticancer activity against murine melanoma cells (B16 cell line) and the HaCaT cell line (normal immortalized keratinocyte cell line). The half maximal inhibitory concentration (IC50) values for B16 and HaCaT cell lines were 0.33 ± 0.2 μM and 0.25 ± 0.1 μM, respectively. Cyclo-(Pro-Tyr) (11) was isolated from the actinobacterium QHHL-09 from the Tibetan Plateau [37]. Two new brominated diketopiperazines, cyclo (D-Pro-L-Br-Tyr) (12) and cyclo (L-Pro-L-Br-Tyr) (13), were isolated from Nocardia ignorata [34]. All 13 diketopiperazines from the lichen derived actinomycetota described above are presented in Figure 3.

3.1.2. Peptides

Peptide natural products (PNPs) represent a unique class of compounds with fascinating structural motifs that impart important biological activities [38]. Cis-3-isobutyl-tetrahydroimidazo [1,2-a] pyridine-2,5-dione (14) was isolated from the lichen-associated actinobacterium QHHL-09, found on the Tibetan Plateau. It inhibited HIV-1 reverse transcriptase, glutamate receptor 2, and protein kinase ck2 [37]. Turnagainolide B (15) was obtained from the strain Streptomyces sp. YIM 130597 from the lichen Punctelia rudecta, found in Yunnan Province, China [35]. It displayed weak antibacterial activity toward S. typhimurium with a MIC value of 64 μg/mL.
Antipain (16), V2 (antipain dehydration product) (17), and lichostatinal (18) were isolated from a British Columbian lichen associated with Streptomycetes sp. L91-3 [39]. Compound 16 was a known potent cathepsin K (CatK) inhibitor, whereas 17 was a new dehydrated analog of antipain and a much weaker CatK inhibitor. Compound 18 was identified as a new potent CatK inhibitor using affinity crystallography. N-methyldactinomycin (19) was collected from Streptomyces cyaneofuscatus MOLA1488 from Erquy (France) [36]. Geninthiocin B (20) was isolated from Streptomyces sp. YIM 130001 obtained from the tropical rainforest in Xishuangbanna (Yunnan, China). It showed antibacterial activity against Bacillus subtilis [28]. Skyllamycins A–E (2125) were isolated from a Streptomyces anulatus strain from the New Zealand lichen Pseudocyphellaria dissimilis. Antibacterial assays revealed that compound 24 possessed superior activity against B. subtilis E168 compared to previously reported congeners [40]. All 12 PNPs from the lichen derived actinomycetota described above are presented in Figure 4.

3.1.3. Indoles

Due to its diverse biological activities, the indole scaffold is a vital heterocyclic organic compound of medical and pharmaceutical interest [41]. Indole-carboxaldehyde (26) was isolated from the lichen Collema auriforme, found in the Austrian town of Kesselfallklamm [34]. Compound 26 showed weak cytotoxic activity against the HaCaT (IC50 = 79 ± 6 μM) and B16 (IC50 = 72 ± 6 μM) cell lines.
(3-Hydroxyacetyl) indole (27) and N-acetyl-β-oxotryptamine (28) were isolated from Lichina confinis, a marine lichen collected on the Brittany coast (Erquy, France) [36]. Acetotryptamide (29) was isolated from the fresh lichen Punctelia borreri from the Jinsha River region, Yunnan Province. It possessed antibacterial activity against S. typhimurium and E. coli [33]. 7-Prenylisatin (30), 5-isoprenylindole-3-carboxylate (31), JBIR-126 (32), and JBIR-149 (33) were isolated from culturable lichen actinomycetes found around Qinghai Lake on the Qinghai–Tibet Plateau [37]. Seven new compounds, cladoniamides A–G (3440) were isolated from cultures of Streptomyces uncialis found on the surface of the lichen Cladonia uncialis collected near Pitt River, British Columbia. Cladoniamide G displayed significant cytotoxicity against human breast cancer MCF-7 cells in vitro at 10 μg/mL [42]. All 15 indoles from the lichen derived actinomycetota described above are presented in Figure 5.

3.1.4. Pyrrole Derivatives

Pyrrole derivatives are a distinct class of heterocycle compounds that contribute significantly to natural products [43]. Mminaline (41) and 1H-pyrrole-2-carboxamide (42) were isolated from Amycolatopsis sp. YIM 130687, isolated from the fresh lichen Punctelia borreri found in the Jinsha River region, Yunnan Province. Compound 41 displayed weak antibacterial activity against MRSA, and 42 showed weak antibacterial activity against Staphylococcus aureus and F. solani [33]. Metacycloprodigiosin (43) and undecylprodigiosin (44) were extracted from the QHHL-18 isolate associated with lichens found around the Qinghai Lake on the Qinghai–Tibet Plateau [37]. All 4 pyrrole derivatives from the lichen derived actinomycetota described above are presented in Figure 6.

3.1.5. Pyridine Derivatives

Pyridine is a crucial heterocyclic framework found in natural products. Methods have been developed for pyridine synthesis because of their importance and appeal in organic chemistry and natural product research [44]. Four new echinosporins, amycolasporins A–D (4548), were derived from the lichen-associated actinomycete Amycolatopsis hippodrome. Compounds 46 and 47 demonstrated antibacterial activity against B. subtilis, S. aureus, and E. coli [45]. A novel compound, JBIR-120 (49), was isolated from Streptomyces sp. RI104-LiC104 from a lichen found on Rishiri Island, Hokkaido Prefecture, Japan [46]. It was weakly cytotoxic against 22Rv1 cells (human prostate carcinoma epithelial cell line) and effectively inhibited the growth of cells activated by dihydrotestosterone. Two known compounds, 4-methoxy-5-(methylthio)-[2,2′-bipyridine]-6-carbonitrile (50) and 7-methoxy-5-(pyridin-2-yl) isothiazolo [4,5-b] pyridine (51), were isolated from the Streptomyces strain YIM 130597 collected from the lichen Punctelia rudecta in Yunnan, China. Compound 51 exhibited strong antibacterial activity against S. aureus (MIC value of 64 μg/mL), E. coli (MIC value of 32 μg/mL), S. typhimurium (MIC value of 64 μg/mL), and MRSA (MIC value of 32 μg/mL) [35]. All 7 pyridine derivatives from the lichen derived actinomycetota described above are presented in Figure 7.

3.1.6. Aromatic Amides and Amines

One or both primary amino groups and the imino group in an aliphatic polyamine can interact with different acids, resulting in mono-, di-, or tri-substituted amide derivatives [47]. A new compound, (E)-3- hydroxy-2,4-dimethylhept-4-enamide (52), was derived from the marine actinomycete Streptomyces cavourensis YY01-17 [48]. 2-Acetamidophenol (53), phenacetamide (54), anthranilic acid (55), 4-(3-methylbut-2-enyloxy) benzamide (56), and 2-pyruvoylaminobenzamide (57) were isolated from the fresh lichen Punctelia borreri. Compound 53 effectively inhibited the growth of MCF-7 breast cancer cells. Compound 54 showed inhibitory activity against F. graminearum with a MIC value of 2 μg/mL and against S. aureus with a MIC value of 8 μg/mL [33]. Compounds 58 3-(4-hydroxyphenyl)-N-methylpropanamide and 59 N-(4-hydroxyphenethyl)-acetamide were isolated from actinomycetes from the Qinghai–Tibet Plateau near Qinghai Lake [37]. Amycophthalazinone A (60) was a new phthalazinone derivative isolated from Amycolatopsis sp. YIM 130642. It exhibited inhibitory activity against S. aureus, S. typhi, and Candida albicans with MIC values of 32, 32, and 64 μg/mL, respectively [49].
Compounds 61 2-carbamoyl-3-hydroxy-1,4-naphthoquinone and 62 (–)-chry-sogine were isolated from the fresh lichen Punctelia borreri found in the Jinsha River region of Yunnan. Compound 61 displayed antimicrobial activity against Botrytis cinerea, F. graminearum, S. aureus, and MRSA with MIC values of 1, 1, 2, and 2 μg/mL, respectively [33]. A new echinosporin derivative, amycolasporin E (63), and a known echinosporin (64) were obtained from the lichen-associated actinomycete Amycolatopsis sp. YIM 130415 [45]. All 13 aromatic amides and amines from the lichen derived actinomycetota described above are presented in Figure 8.

3.2. Furanones

Furanones are commonly utilized in synthesis. The products display important pharmacological properties such as antiviral, anticancer, and antimicrobial properties [50]. JBIR-89 (65) was a new butenolide from the lichen-derived Streptomyces sp. RI104-LiB101 collected from Rishiri Island, Hokkaido Prefecture, Japan [51]. (5S)-5-(6-Hydroxy-6-methyloctyl)-furan-2(5H)-one (66) and (5S)-5-(6-hydroxy-7- methyloctyl)-furan-2(5H)-one (67) were produced by lichen-associated actinomycetes collected on the Tibetan Plateau, China [37]. Six new compounds, actinofuranones D-I (6873), and three known compounds, JBIR-108 (74), E-975 (75), and E-492 (76), were obtained from S. gramineus derived from the lichen Leptogium trichophorum collected from an evergreen broad-leaf forest in Benzilan, Diqing, Yunnan, China [52]. Compounds 71, 72, 75, and 76 inhibited nitric oxide synthase expression (in OS) in RAW 264.7 cells after LPS induction. In addition, 71, 72, 75, and 76 inhibited the LPS-induced proinflammatory cytokines interleukin-6 (IL-6) and tumor necrosis factor α (TNF-α). All 12 furanones from the lichen derived actinomycetota described above are presented in Figure 9.

3.3. Aromatic Compounds

3.3.1. Quinones

Quinones exhibit various biological activities, including antibacterial, antiplasmodial, antioxidant, trypanocidal, anticancer, and anti-HIV activities. All these activities are linked to the redox properties of their carbonyl groups [53]. Compound 77 (+)-4-hydroxy-1-teralone was obtained from the lichen-derived actinomycete Amycolatopsis sp. YIM 130687 collected from Yunnan Province, China [33]. Four novel nanomycin compounds, 4aβ,10aα-dihydroxynanaomycin βA (78), 4aβ,10aβ-dihydroxynanaomycin βA (79), 4aα,10aβ-dihy-droxynanaomycin βA (80), and 10β-hydroxynanaomycin αE (81), and two known compounds, nanaomycin αA (82) and nanaomycin βA (83), were produced by Streptomyces hebeiensis [54]. Compounds 82 and 83 displayed antibacterial activity against S. aureus and B. subtilis with MIC values ranging from 3.13 to 100 μg/mL, and modest antifungal activity against C. albicans.
Compound 84 4-deoxy-ε-pyrromycinone was isolated from lichen actinomycetes collected around Qinghai Lake [37]. JBIR-88 (85) was a new angucycline produced by the lichen-derived Streptomyces spp. RI104-LiC106, and it exhibited antibacterial activity against Micrococcus luteus. Furthermore, 85 showed cytotoxicity against HeLa cells with a MIC of 36 μM and ACC-MESO-1 cells with a MIC of 52 μM [51]. BE-24566B (86) was a new antibiotic produced by Streptomyces violaceusniger A24566, which was isolated from a lichen collected in Jyogasaki, Shizuoka Prefecture, Japan. This compound inhibited Gram-positive bacteria, including methicillin-resistant S. aureus [55].
The lichen-derived actinomycete Steptomyces sp. 0630c, collected from Zhaosu County of the Xinjiang Uygur Autonomous Region, China, yielded three compounds, steffimycin D (87), steffimycin E (88), and steffimycin F (89) [56]. Compound 89 was a new steffimycin-type antibiotic with weak cytotoxicity towards MCF-7 (human breast adenocarcinoma), HepG-2 (human liver hepatocellular carcinoma), and A2780 (human ovarian carcinoma) cell lines. Two known compounds, 88 and 89, exhibited potent antibacterial action against S. aureus with MIC values of 2 μg/mL. (7S*, 9R*, 10R*)-Pyrromycin (90) was isolated from a lichen actinomycete collected from Qinghai Lake [37]. Uncialamycin (91) was a novel enediyne antibiotic isolated from the British Columbian lichen Cladonia uncialis collected near the Pitt River [57]. It exhibited antibacterial activity against S. aureus, E coli, and Burkholderia cepacia with MIC values of 0.0000064, 0.002, and 0.001 μg/mL, respectively. 7-O-methylkoninginin D (92) and koninginin E (93) were obtained from Streptomyces sp. from the lichen Punctelia rudecta collected in Yunnan Province, China [35]. All 17 quinones from the lichen derived actinomycetota described above are presented in Figure 10.

3.3.2. Isoflavonoids

Isoflavonoids have a B-ring connected to their C-ring at the C-3 position (3-phenylchroman skeleton). They display a wide range of biological activities including antioxidant, anticarcinogenic, and antiproliferative properties. They also possess the ability to reduce osteoporosis and cardiovascular disease [58]. Seven isoflavonoid glycosides, namely genistein (94), formononetin (95), prunetin (96), kakkatin (97), isoformononetin (98), 7-O-methyl-5-O-α-L-rhamnopyranosylgenestein (99), and 7-O-α-D-arabinofur-anosyldaidzein (100), were isolated from Amycolatopsis sp. YIM 130642. Compounds 94 and 99 inhibited S. aureus and E. coli. Compound 96 inhibited E. coli with a MIC of 32 μg/mL, while 100 demonstrated bacteriostatic activity against S. typhi with a MIC value of 64 μg/mL [49]. All 7 isoflavonoids from the lichen derived actinomycetota described above are presented in Figure 11.

3.3.3. Phenolic Derivatives

Phenolic compounds occur mainly in soluble conjugate and insoluble forms and are covalently bound to sugar moieties or structural components of the cell wall. Phenolic compounds have been extensively studied due to their varied health benefits as antioxidants and their roles in preventing chronic inflammation, cardiovascular disease, cancer, and diabetes [59,60]. The novel compound (R)-3-methyl-1,3-dihydroisobenzofuran-4,6-diol (101) was obtained from Amycolatopsis hippodrome [45]. P-hydroxyphenethyl alcohol (102) and sorbicillin (103) were obtained from Amycolatopsis sp. YIM 130687. Compound 103 showed cytotoxicity against the hepatocellular carcinoma cell line QGY-7703 and inhibited the growth of C. albicans [33]. 2-(4-Hydroxybenzylacetone)-5-methoxyphenol (104), amycolabenzoyl (105), and amycolabenzosides A–B (106107) were obtained from Amycolatopsis hippodrome. Compound 104 attenuated nitric oxide production by suppressing the expression of nitric oxide synthase (iNOS) in LPS-induced RAW 264.7 cells in a dose-dependent manner [44]. Usnic acid (108), a prevalent cytotoxic secondary metabolite in lichens, was isolated from Streptomyces cyaneofuscatus MOLA1488. It imparts a green to greenish yellow color to many lichens [36]. All 8 phenolic derivatives from the lichen derived actinomycetota described above are presented in Figure 12.

3.4. Linear Esters and Macrolides

Esters of linear long-chain unsaturated fatty acids with multiple alcohols, both linear and branched, are widely used in the lubricant industry [61]. Macrolides are a large and structurally diverse class of macrocyclic natural products. They are valuable targets in synthetic chemistry due to their biological and medicinal importance [62]. One new linear compound, 2(S)-3′-hydroxybutan-2′-yl 2-hydroxypropanoate (109), and a known compound, 2-hydroxy-3-methylbutanoic acid (110), were procured from the marine-derived actinomycete Streptomyces cavourensis YY01-17 [48]. Cyaneodimycin (111) and cyaneomycin (112) were isolated from marine-lichen-associated Streptomyces cyaneofuscatus. Compound 111 exhibited antiproliferative action against B16, HaCaT, and Jurkat cell lines with MIC values of 27 ± 4 μM, 47 ± 11 μM, and 18.5 ± 0.5 μM, respectively [35]. Macrolactin A (113) was isolated from a lichen actinomycete found on the Qinghai–Tibet Plateau [37]. All 5 linear esters and macrolides from the lichen derived actinomycetota described above are presented in Figure 13.

3.5. Sterols

Sterols are isoprenoid derivatives and structural components of biological membranes. They are currently being investigated for their structural, functional, and regulatory roles [63]. Campesterol (114) from the lichen-derived strain Amycolatopsis sp. YIM 130687 inhibited the growth of MRSA with a MIC of 128 μg/mL [33]. 1 sterol from the lichen derived actinomycetota described above are presented in Figure 14.

4. Bioactivity of Uncharacterized Compounds

The biological activities of many lichen actinomycetota have been investigated. However, these studies did not report any pure compounds or their structures. Several lichen-associated actinomycetota have been screened for their biological activities, such as antibacterial and antifungal activities, inhibition of β-glucosidase activity, etc. Such screenings without the structural elucidation of bioactive metabolites may not be useful for the discovery of new compounds. Nonetheless, these data highlight the possible utility of lichen-associated novel actinomycetota for the discovery of novel bioactive chemicals in the future [64].
Twelve actinomycete strains were isolated from lichens collected from the Maha Sarakham Province, Thailand. Among these, four Streptomyces isolates, LDG1-03, LDG1-15, LDG1-16, and LLG1-03, showed antimicrobial activity against B. subtilis ATCC 6633. LDG1-03 and LDG1-15 exhibited antimicrobial activity against S. aureus ATCC 25923, Kocuria rhizophila ATCC 9341, and C. albicans ATCC 10231. The Actinoplanes isolate LDG1-06 inhabited C. albicans ATCC 10231 [65]. Actinomycetes LC-23 was isolated from a lichen found growing on the bark of the Averrhoa carambola plant. Actinomycete pure strains were screened using agar diffusion on ISP2 agar medium to determine antimicrobial potency. The ethyl acetate extract of this strain displayed a positive inhibitory effect against S. aureus BTCC B-611 and M. luteus BTCC B-552 [66]. Lichen-associated Streptomyces olivaceus LEP7 was recovered from tree bark collected in the botanical garden of Nilgiris, Tamil Nadu, India. The extract of Streptomyces sp. LEP7 inhibited E. coli, S. aureus, and P. aeruginosa efficiently. The extract was found to contain cyclopentene upon GC-MS analysis. According to the report, the remarkable antimicrobial activity of Streptomyces olivaceus when tested against wound infections caused by microbial pathogens, and the derivation of cyclopentene from LEP7is a first step in this direction [67].
Extracts from the lichen Umbilicaria esculenta strongly inhibited mold and mammalian disaccharide hydrolytic enzymes (β-glucosidase). The inhibitory component of the extract was very stable, retaining more than 95% of its activity when treated with heat, acid, alkali, and some hydrolytic enzymes [68]. Streptomyces sp. DPUA 1542 and Nocardia sp. DPUA 1571, two actinomycetota strains isolated from Amazon River basin lichens, produced β-lactamase inhibitors which cured bovine mastitis [69]. Streptomyces sp. DPUA 1576, isolated from an Amazon basin lichen, yielded a fibrinolytic protease. This protease could potentially provide new and unexploited fibrinolytic enzymes for different therapeutic purposes [70,71].

5. Biosynthetic Pathways of Lichen Secondary Metabolites

Advances in synthetic biology and associated technologies such as DNA synthesis, sequencing, and analysis techniques have accelerated the DBT cycle for metabolic and protein engineering to the point where both can be deployed to engineer the biosynthesis of a particular molecule [72]. The genes encoding these natural products in actinomycetes tend to be clustered, which allows the transfer of entire biosynthetic pathways to an exogenous host for heterologous expression. This strategy also enables the genetic modifications of such pathways, allowing the generation of various natural product analogs as well as the optimization of production yield [73].
The production of the novel thiopetide antibiotic geninthiocin B (20) is due to the Gen B gene encoding a putative lantibiotic dehydratase in the biosynthetic gene cluster of the lichen-derived Streptomyces sp. YIM130001. As described in the literature, the production of associated genes includes precursor proteins, Yeao cyclodchydratase, lanthipeptide dehydratases, etc. (Figure 15A). The biosynthetic pathway of geninthiocin was proposed by Schneider et al. [28], and is exhibited in Figure 15B. The precursor peptide (GenA) unit, which possesses a 31 aa leader peptide (LP), is connected with a C-terminal 15 aa core peptide unit. The Yeao cyclodchydratase biosynthetic gene clusters GenG1 and GenG2 could catalyze the processing of azole rings formation. The proteins GenB and GenC show a high degree of similarity to lanthipeptide-like dehydratases and most likely catalyze the formation of the dehydroalanine (Dha) and dehydrobutyrine (Dhb) functional groups. The two Dha group residues from the serines Ser1 and Ser13 are then utilized by GenD for assembly of the central six-membered nitrogenous heterocycle. Finally, the cleavage of Ser15 to afford the C-terminus amide and the hydroxylation of Val7 is catalyzed by GenI and GenH, completing the biosynthesis of geninthiocin B.
The genome of Streptomyces uncialis includes halogenases and flavin reductase, indolocarbazole aglycone construction, new flavin-dependent oxygenases, and so on (Figure 16A). The biosynthetic pathway of cladoniamides were proposed by Ryan et al. [74] and is exhibited in Figure 16B. It shows that ClaH and ClaF are highly related to the characterized L-tryptophan chlorinases, and that chlorine is installed on the L-tryptophan in the first step of the related rebeccamycin pathway. Due to the action of ClaH and a partner flavin reductase ClaF, L-tryptophan is chlorinated at the C-5′ position. Then, 5-chloro-L-tryptophan reacts with ClaO, generating an indole-3-pyruvate imine. ClaD dimerizes two of these molecules to generate a chromo pyrrolic acid molecule. ClaY catalyzes the hydrolysis of an amide bond in the N-methylsuccinimide ring, which is followed by oxidative decarboxylation. Three enzymes unique to the indenotryptoline biosynthetic pathway include two putative flavin-dependent oxygenases (ClaX1 and ClaX2) and a putative α/β hydrolase (ClaY) shown in Figure 15A. The cladoniamide biosynthetic gene cluster is highly homologous to that of BE-54017. 1 (R1 = Cl, R2 = R3 = H) and a methylated derivative of 2 (R1 = R2 = H) separately accumulate in the BE-54017 heterologous expression system when the genes abeX1 and abeX2 are mutated, respectively. The route to generate downstream metabolites, indenotryptoline-containing molecules such as 3, from substrate 2 via cleavage of the epoxide could be driven via ketone formation from one tertiary alcohol, causing sigma-bond rupture and epoxide hydrolysis, opening the indolocarbazole scaffold. This cleaved molecule could then close through attack on the ketone by the indolic nitrogen, restoring the tertiary alcohol and arriving at the indenotryptoline scaffold 3. Each of these enzymes is thought to catalyze the transfer of a methyl group to a phenolic oxygen, consistent with the likely role of ClaM3 in cladoniamide biosynthesis of installing a methyl group on the appended hydroxyl group to produce cladoniamides A–C.

6. Conclusions

The current review focused on lichen actinomycetota from four different perspectives. (1) Lichen-associated actinomycetes represent a promising but underutilized resource. A wide variety of novel actinomycetes have been isolated from lichens. (2) The potential of bioactive metabolites from lichen actinomycetes has been explored, and a total of 114 secondary metabolites from lichen-associated actinomycetes are summarized here. (3) Although the biological activities of many lichen actinomycetota have been investigated, their definite chemical components are still undetermined. Thus, the discovery of more novel bioactive compounds reveals new research prospects. (4) The biosynthetic pathways of some unique secondary metabolites isolated from lichen-derived actinomycetes are discussed.
It has become increasingly difficult to isolate new sources of actinomycetes from common environments such as the soil, sea, and plants. These resources no longer meet the increasingly urgent demand for new drug-leading compounds [6]. Therefore, researchers must explore potent microbial resources from unique environments [64]. Current research on lichen-associated actinomycetes has focused mainly on Asia, whereas lichens are globally distributed. This wide distribution range enables researchers to search for novel species. Lichen environments are understudied in terms of microbiology, but they should not be disregarded in the hunt for novel actinomycetota and their diversity of beneficial chemical compounds [71]. The novelty and variety of lichen actinomycetota are evident in this review. Furthermore, the study of biosynthetic pathways is a crucial process in the excavation of bioactive natural products [75]. Biosynthesis of other biologically active compounds is relatively less studied and requires more attention from researchers. Future work could attempt to reveal more silent biosynthetic gene clusters, so that to uncover more and more novel and interesting biologically active natural products from lichen-associated actinomycetota.

Author Contributions

Q.Y. conceptualized and wrote the manuscript; Z.S. drew the figures; X.L., Y.H. and T.X. checked the paper; S.W. verified the content. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 81860634 and No. 32260110) and Yunnan Science and Technology Project (No. 202001BB050029 and No. 202201BF070001-003).

Conflicts of Interest

The authors declare that they have no known competing financial interests in this paper.

References

  1. Liu, C.B.; Jiang, Y.; Wang, X.Y.; Chen, D.B.; Chen, X.; Wang, L.S.; Han, L.; Huang, X.S.; Jiang, C.L. Diversity, antimicrobial activity, and biosynthetic potential of cultivable actinomycetes associated with lichen symbiosis. Microb. Ecol. 2017, 74, 570–584. [Google Scholar] [CrossRef] [PubMed]
  2. Jiang, Y.; Wang, X.Y.; Li, G.D.; Li, Q.Y.; Liu, C.B.; Chen, X.; Wang, L.S.; Yong, L.; Jiang, C.L. Diversity and anti-microbial activities of actinomycetes associated with three species of lichens. Am. J. Biosci. 2015, 3, 171. [Google Scholar] [CrossRef]
  3. Coley, P.D. Effect of plant-growth rate and leaf lifetime on the amount and type of anti-herbivore defense. Oecologia 1988, 74, 531–536. [Google Scholar] [CrossRef] [PubMed]
  4. Genilloud, O. Actinomycetes: Still a source of novel antibiotics. Nat. Prod. Rep. 2017, 34, 1203–1232. [Google Scholar] [CrossRef] [PubMed]
  5. Sigurbjornsdottir, M.A.; Vilhelmsson, O. Selective isolation of potentially phosphate-mobilizing, biosurfactant-producing and biodegradative bacteria associated with a sub-Arctic, terricolous lichen, Peltigera membranacea. Fems Microbiol. Ecol. 2016, 6, 92. [Google Scholar] [CrossRef] [PubMed]
  6. Ding, T.; Yang, L.J.; Zhang, W.D.; Shen, Y.H. The secondary metabolites of rare actinomycetes: Chemistry and bioactivity. RSC Adv. 2019, 9, 21964–21988. [Google Scholar] [CrossRef]
  7. Oren, A.; Garrity, G.M. Valid publication of the names of forty- two phyla of prokaryotes. Int. J. Syst. Evol. Microbiol. 2021, 71, 005056. [Google Scholar] [CrossRef]
  8. Zhang, K.; Jiang, L.Q.; Wang, L.S.; An, D.-F.; Lang, L.; Li, G.D.; Wang, X.Y.; Shi, S.B.; Li, Q.Y.; Jiang, C.L.; et al. Aureimonas leprariae sp. nov. isolated from a Lepraria sp. lichen. Curr. Microbiol. 2020, 77, 313–319. [Google Scholar] [CrossRef]
  9. Jiang, L.Q.; An, D.F.; Wang, X.Y.; Zhang, K.; Li, G.D.; Lang, L.; Wang, L.S.; Jiang, C.L.; Jiang, Y. Methylobacterium planium sp. nov. isolated from a lichen sample. Arch. Microbiol. 2020, 202, 1709–1715. [Google Scholar] [CrossRef]
  10. An, D.F.; Jiang, L.Q.; Zhang, K.; Li, G.D.; Wang, X.Y.; Jiang, M.G.; Lang, L.; Wang, L.S.; Imhoff, J.F.; Jiang, C.L.; et al. Glaciibacter flavus sp. nov. isolated from a lichen sample. Arch. Microbiol. 2021, 203, 2439–2444. [Google Scholar] [CrossRef]
  11. Cardinale, M.; Grube, M.; Berg, G. Frondihabitans cladoniiphilus sp nov. an actinobacterium of the family Microbacteriaceae isolated from lichen, and emended description of the genus Frondihabitans. Int. J. Syst. Evol. Microbiol. 2011, 61, 3033–3038. [Google Scholar] [CrossRef] [PubMed]
  12. An, S.Y.; Xiao, T.; Yokota, A. Leifsonia lichenia sp nov. isolated from lichen in Japan. J. Gen. Appl. Microbiol. 2009, 55, 339–343. [Google Scholar] [CrossRef] [PubMed]
  13. Jiang, L.Q.; Zhang, K.; Li, G.D.; Wang, X.Y.; Shi, S.B.; Li, Q.Y.; An, D.F.; Lang, L.; Wang, L.S.; Jiang, C.L.; et al. Naasia lichenicola sp. nov. an actinobacterium isolated from lichen. Int. J. Syst. Evol. Microbiol. 2020, 70, 1026–1030. [Google Scholar] [CrossRef]
  14. An, S.Y.; Xiao, T.; Yokota, A. Schumannella luteola gen. nov. sp nov. a novel genus of the family Microbacteriaceae. J. Gen. Appl. Microbiol. 2008, 54, 253–258. [Google Scholar] [CrossRef] [PubMed]
  15. Si, H.L.; Shi, F.X.; Zhang, L.L.; Yue, H.S.; Wang, H.Y.; Zhao, Z.T. Subtercola lobariae sp nov. an actinobacterium of the family Microbacteriaceae isolated from the lichen Lobaria retigera. Int. J. Syst. Evol. Microbiol. 2017, 67, 1516–1521. [Google Scholar] [CrossRef] [PubMed]
  16. Phongsopitanun, W.; Matsumoto, A.; Inahashi, Y.; Kudo, T.; Mori, M.; Shiomi, K.; Takahashi, Y.; Tanasupawat, S. Actinoplanes lichenis sp nov. isolated from lichen. Int. J. Syst. Evol. Microbiol. 2016, 66, 468–473. [Google Scholar] [CrossRef]
  17. Saeng-in, P.; Kanchanasin, P.; Yuki, M.; Kudo, T.; Ohkuma, M.; Phongsopitanun, W.; Tanasupawat, S. Actinoplanes lichenicola sp. nov. and Actinoplanes ovalisporus sp. nov. isolated from lichen in Thailand. Int. J. Syst. Evol. Microbiol. 2021, 7, 71. [Google Scholar] [CrossRef]
  18. Jiang, L.Q.; An, D.F.; Zhang, K.; Li, G.D.; Wang, X.; Lang, L.; Jiang, M.G.; Wang, L.; Jiang, C.L.; Jiang, Y. Nakamurella albus sp. nov.: A novel actinobacterium isolated from a lichen sample. Curr. Microbiol. 2020, 77, 1896–1901. [Google Scholar] [CrossRef]
  19. An, D.F.; Yang, S.J.; Jiang, L.Q.; Wang, X.Y.; Huang, X.Y.; Lang, L.; Chen, X.M.; Fan, M.Q.; Li, G.D.; Jiang, M.G.; et al. Nakamurella leprariae sp. nov. isolated from a lichen sample. Arch. Microbiol. 2022, 19, 204. [Google Scholar] [CrossRef]
  20. Li, B.; Xie, C.H.; Yokota, A. Nocardioides exalbidus sp. nov. a novel actinomycete isolated from lichen in Izu-Oshima Island, Japan. Actinomycetologica 2007, 21, 22–26. [Google Scholar] [CrossRef]
  21. Hamada, M.; Yamamura, H.; Komukai, C.; Tamura, T.; Suzuki, K.-i.; Hayakawa, M. Luteimicrobium album sp nov. a novel actinobacterium isolated from a lichen collected in Japan, and emended description of the genus Luteimicrobium. J. Antibiot. 2012, 65, 427–431. [Google Scholar] [CrossRef] [PubMed]
  22. Yamamura, H.; Ashizawa, H.; Nakagawa, Y.; Hamada, M.; Ishida, Y.; Otoguro, M.; Tamura, T.; Hayakawa, M. Actinomycetospora iriomotensis sp nov. a novel actinomycete isolated from a lichen sample. J. Antibiot. 2011, 64, 289–292. [Google Scholar] [CrossRef] [PubMed]
  23. Yamamura, H.; Ashizawa, H.; Nakagawa, Y.; Hamada, M.; Ishida, Y.; Otoguro, M.; Tamura, T.; Hayakawa, M. Actinomycetospora rishiriensis sp nov. isolated from a lichen. Int. J. Syst. Evol. Microbiol. 2011, 61, 2621–2625. [Google Scholar] [CrossRef] [PubMed]
  24. Vobis, G.; Solans, M.; Scervino, J.M.; Schumann, P.; Sproeer, C.; Messuti, M.I. Isolation and characterization of an endolichenic actinobacterium from the lichen thallus of Pseudocyphellaria berberina. Symbiosis 2020, 80, 43–51. [Google Scholar] [CrossRef]
  25. Lang, L.; An, D.F.; Jiang, L.Q.; Li, G.D.; Wang, L.S.; Wang, X.Y.; Li, Q.Y.; Jiang, C.L.; Jiang, Y. Paracoccus lichenicola sp. nov. Isolated from Lichen. Curr. Microbiol. 2021, 78, 816–821. [Google Scholar] [CrossRef]
  26. Jiang, L.Q.; Zhang, K.; Li, G.D.; Wang, X.Y.; Shi, S.B.; Li, Q.Y.; An, D.F.; Lang, L.; Wang, L.S.; Jiang, C.L.; et al. Rubellimicrobium rubrum sp. nov. a novel bright reddish bacterium isolated from a lichen sample. Antonie Van Leeuwenhoek Int. J. Gen. Mol. Microbiol. 2019, 112, 1739–1745. [Google Scholar] [CrossRef]
  27. Saeng-In, P.; Phongsopitanun, W.; Savarajara, A.; Tanasupawat, S. Streptomyces lichenis sp nov. isolated from lichen. Int. J. Syst. Evol. Microbiol. 2018, 68, 3641–3646. [Google Scholar] [CrossRef]
  28. Schneider, O.; Simic, N.; Aachmann, F.L.; Rueckert, C.; Kristiansen, K.A.; Kalinowski, J.; Jiang, Y.; Wang, L.; Jiang, C.-L.; Lale, R.; et al. Genome mining of Streptomyces sp. YIM 130001 isolated from lichen affords new thiopeptide antibiotic. Front. Microbiol. 2018, 9, 3139. [Google Scholar] [CrossRef]
  29. Somphong, A.; Poengsungnoen, V.; Buaruang, K.; Suriyachadkun, C.; Sripreechasak, P.; Tanasupawat, S.; Phongsopitanun, W. Diversity of the culturable lichen-derived actinobacteria and the taxonomy of Streptomyces parmotrematis sp. nov. Antonie Van Leeuwenhoek. Int. J. Gen. Mol. Microbiol. 2022, 115, 911–920. [Google Scholar] [CrossRef]
  30. Kanchanasin, P.; Phongsopitanun, W.; Yuki, M.; Kudo, T.; Ohkuma, M.; Nakashima, T.; Tanasupawat, S. Actinomadura violacea sp. nov. a madurastatin A1-producing strain isolated from lichen in Thailand. Int. J. Syst. Evol. Microbiol. 2021, 12, 71. [Google Scholar] [CrossRef]
  31. Somphong, A.; Poengsungnoen, V.; Buaruang, K.; Suriyachadkun, C.; Sripreechasak, P.; Tanasupawat, S.; Phongsopitanun, W. Actinomadura parmotrematis sp. nov. isolated from the foliose lichen, Parmotrema praesorediosum (Nyl.) Hale. Int. J. Syst. Evol. Microbiol. 2022, 7, 72. [Google Scholar] [CrossRef] [PubMed]
  32. Song, Z.Q.; Hou, Y.G.; Yang, Q.R.; Li, X.P.; Wu, S.H. Structures and biological activities of diketopiperazines from marine organisms: A review. Mar. Drugs 2021, 19, 403. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, C.S.; Jiang, Y.; Huang, R.; Jiang, B.G.; Zheng, K.X.; Wu, S.H. Diverse secondary metabolites from a lichen-derived Amycolatopsis Strain. Curr. Microbiol. 2020, 77, 2104–2110. [Google Scholar] [CrossRef]
  34. Noel, A.; Ferron, S.; Rouaud, I.; Gouault, N.; Hurvois, J.P.; Tomasi, S. Isolation and structure identification of novel brominated diketopiperazines from Nocardia ignorata a lichen-associated actinobacterium. Molecules 2017, 22, 371. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, B.G.; Wei, H.X.; Wang, Y.T.; Zheng, K.X.; Liu, S.S.; Zhang, S.P.; Jiang, Y.; Wu, S.H. Secondary metabolites of two lichen-derived Streptomyces. Chem. Nat. Compd. 2019, 55, 783–786. [Google Scholar] [CrossRef]
  36. Parrot, D.; Legrave, N.; Intertaglia, L.; Rouaud, I.; Legembre, P.; Grube, M.; Suzuki, M.T.; Tomasi, S. Cyaneodimycin, a bioactive compound isolated from the culture of Streptomyces cyaneofuscatus associated with Lichina confinis. Eur. J. Org. Chem. 2016, 2016, 3977–3982. [Google Scholar] [CrossRef]
  37. Hei, Y.Y.; Zhang, H.L.; Tan, N.N.; Zhou, Y.H.; Wei, X.; Hu, C.H.; Liu, Y.D.; Wang, L.; Qi, J.Z.; Gao, J.M. Antimicrobial activity and biosynthetic potential of cultivable actinomycetes associated with lichen symbiosis from Qinghai-Tibet Plateau. Microbiol. Res. 2021, 244, 126652. [Google Scholar] [CrossRef]
  38. Maini, R.; Umemoto, S.; Suga, H. Ribosome-mediated synthesis of natural product-like peptides via cell-free translation. Curr. Opin. Chem. Biol. 2016, 34, 44–52. [Google Scholar] [CrossRef]
  39. Aguda, A.H.; Lavallee, V.; Cheng, P.; Bott, T.M.; Meimetis, L.G.; Law, S.; Nguyen, N.T.; Williams, D.E.; Kaleta, J.; Villanueva, I.; et al. Affinity crystallography: A new approach to extracting high-affinity enzyme inhibitors from natural extracts. J. Nat. Prod. 2016, 79, 1962–1970. [Google Scholar] [CrossRef]
  40. Bracegirdle, J.; Hou, P.; Nowak, V.V.; Ackerley, D.F.; Keyzers, R.A.; Owen, J.G. Skyllamycins D and E, non-ribosomal cyclic depsipeptides from lichen-sourced Streptomyces anulatus. J. Nat. Prod. 2021, 84, 2536–2543. [Google Scholar] [CrossRef]
  41. Mahmoud, E.; Hayallah, A.M.; Kovacic, S.; Abdelhamid, D.; Abdel-Aziz, M. Recent progress in biologically active indole hybrids: A mini review. Pharmacol. Rep. 2022, 74, 570–582. [Google Scholar] [CrossRef] [PubMed]
  42. Williams, D.E.; Davies, J.; Patrick, B.O.; Bottriell, H.; Tarling, T.; Roberge, M.; Andersen, R.J. Cladoniamides A-G, tryptophan-derived alkaloids produced in culture by Streptomyces uncialis. Org. Lett. 2008, 10, 3501–3504. [Google Scholar] [CrossRef] [PubMed]
  43. Rostami, H.; Shiri, L. Review on synthesis of pyrrole derivatives promoted by nanoparticles. Appl. Organomet. Chem. 2021, 35, 6209. [Google Scholar] [CrossRef]
  44. Miyakoshi, T.; Konno, H. Improved synthesis of 2,4,6-trialkylpyridines from 1,5-diketoalkanes: The total synthesis of anibamine. Org. Biomol. Chem. 2019, 17, 2896–2905. [Google Scholar] [CrossRef]
  45. Jin, Y.; Aobulikasimu, N.; Zhang, Z.; Liu, C.; Cao, B.; Lin, B.; Guan, P.; Mu, Y.; Jiang, Y.; Han, L.; et al. Amycolasporins and dibenzoyls from lichen-associated Amycolatopsis hippodromi and their antibacterial and anti-inflammatory Activities. J. Nat. Prod. 2020, 83, 3545–3553. [Google Scholar] [CrossRef] [PubMed]
  46. Kawahara, T.; Hosoya, T.; Tsukamoto, M.; Okabe, S.; Yamamura, H.; Hayakawa, M.; Seimiya, H.; Takagi, M.; Shin-ya, K. JBIR-120: A new growth inhibitor of hormone-refractory prostate cancer cells. J. Antibiot. 2012, 65, 373–375. [Google Scholar] [CrossRef]
  47. Neu, D.; Lehmann, T.; Elleuche, S.; Pollmann, S. Arabidopsis amidase 1, a member of the amidase signature family. FEBS J. 2007, 274, 3440–3451. [Google Scholar] [CrossRef]
  48. Su, S.S.; Tian, L.; Chen, G.; Li, Z.Q.; Xu, W.F.; Pei, Y.H. Two new compounds from the metabolites of a marine-derived actinomycete Streptomyces cavourensis YY01-17. J. Asian Nat. Prod. Res. 2013, 15, 265–269. [Google Scholar] [CrossRef]
  49. Zheng, K.X.; Jiang, Y.; Jiang, J.X.; Huang, R.; He, J.; Wu, S.H. A new phthalazinone derivative and a new isoflavonoid glycoside from lichen-associated Amycolatopsis sp. Fitoterapia 2019, 135, 85–89. [Google Scholar] [CrossRef]
  50. Xu, T.C.; Song, Z.Q.; Hou, Y.G.; Liu, S.S.; Li, X.P.; Yang, Q.R.; Wu, S.H. Secondary metabolites of the genus Nigrospora from terrestrial and marine habitats: Chemical diversity and biological activity. Fitoterapia 2022, 161, 105254. [Google Scholar] [CrossRef]
  51. Motohashi, K.; Takagi, M.; Yamamura, H.; Hayakawa, M.; Shin-Ya, K. A new angucycline and a new butenolide isolated from lichen-derived Streptomyces spp. J. Antibiot. 2010, 63, 545–548. [Google Scholar] [CrossRef] [PubMed]
  52. Ma, J.; Cao, B.X.; Liu, C.B.; Guan, P.P.; Mu, Y.; Jiang, Y.; Han, L.; Huang, X.S. Actinofuranones D-I from a lichen-associated actinomycetes, Streptomyces gramineus, and their anti-inflammatory effects. Molecules 2018, 23, 2393. [Google Scholar] [CrossRef] [PubMed]
  53. Eyong, K.O.; Kuete, V.; Efferth, T. Quinones and benzophenones from the medicinal plants of Africa. Med. Plant Res. Afr. 2013, 10, 351–391. [Google Scholar]
  54. Liu, C.B.; Jiang, Y.; Lei, H.; Chen, X.; Ma, Q.J.; Han, L.; Huang, X.S. Four new nanaomycins produced by Streptomyces hebeiensis derived from Lichen. Chem. Biodivers. 2017, 14, 170057. [Google Scholar] [CrossRef] [PubMed]
  55. Kojiri, K.; Nakajima, S.; Fuse, A.; Suzuki, H.; Suda, H. BE-24566B, a new antibiotic Produced by Streptomyces violaceusniger. J. Antibiot. 1995, 48, 1506–1508. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, C.Y.; Li, Y.L.; Lu, J.H.; Qian, L.L.; Xu, K.; Wang, N.N.; Chang, W.Q.; Lou, H.X. Steffimycin F, a new steffimycin-type derivative from the lichen-derived actinomycetes streptomyces sp. J. Mol. Struct. 2021, 5, 1227. [Google Scholar]
  57. Davies, J.; Wang, H.; Taylor, T.; Warabi, K.; Huang, X.H.; Andersen, R.J. Uncialamycin, a new enediyne antibiotic. Org. Lett. 2005, 7, 5233–5236. [Google Scholar] [CrossRef]
  58. Al-Maharik, N. Isolation of naturally occurring novel isoflavonoids: An update. Nat. Prod. Rep. 2019, 36, 1156–1195. [Google Scholar] [CrossRef]
  59. Acosta-Estrada, B.A.; Gutierrez-Uribe, J.A.; Serna-Saldivar, S.O. Bound phenolics in foods, a review. Food Chem. 2014, 152, 46–55. [Google Scholar] [CrossRef]
  60. Singh, B.; Singh, J.P.; Kaur, A.; Singh, N. Phenolic composition and antioxidant potential of grain legume seeds: A review. Food Res. Int. 2017, 101, 1–16. [Google Scholar] [CrossRef]
  61. Sharma, N.; Thakre, G.D.; Ray, A. Hitherto unexplored three-membered heterocyclic rings favorably alter tribological properties of fatty acid linear esters. Tribol. Trans. 2021, 64, 996–1021. [Google Scholar] [CrossRef]
  62. Lambu, M.R.; Dash, A.K.; Hussain, N.; Bhat, S.; Raina, S.; Mukherjee, D. Synthesis of macrolide natural products from carbohydrates. Trends Carbohydr. Res. 2017, 9, 1–13. [Google Scholar]
  63. Valitova, J.N.; Sulkarnayeva, A.G.; Minibayeva, F.V. Plant sterols: Diversity, biosynthesis, and physiological functions. Biochem. -Mosc. 2016, 81, 819–834. [Google Scholar] [CrossRef]
  64. Rangseekaew, P.; Pathom-aree, W. Cave actinobacteria as producers of bioactive metabolites. Front. Microbiol. 2019, 10, 387. [Google Scholar] [CrossRef]
  65. Saeng-in, P.; Kuncharoen, N.; Chamroensaksri, N.; Tanasupawat, S. Identification and antimicrobial activity of Streptomyces and Actinoplanes strains from lichens. J. Appl. Pharm. Sci. 2015, 5, 23–29. [Google Scholar]
  66. Susanti, A.E.; Ratnakomala, S.; Mangunwardoyo, W.; Lisdiyanti, P.; Assoc Comp, M. In antimicrobialactivity of lichens-associated Actinomycetes Strain LC-23. Assoc. Comput. Mach. 2019, 19, 9–96. [Google Scholar]
  67. Rajaram, S.K.; Ahmad, P.; Keerthana, S.S.S.; Cressida, P.J.; Moorthy, I.G.; Suresh, R.S.S. Extraction and purification of an antimicrobial bioactive element from lichen associated Streptomyces olivaceus LEP7 against wound inhabiting microbial pathogens. J. King Saud Univ. Sci. 2020, 32, 2009–2015. [Google Scholar] [CrossRef]
  68. Lee, K.A.; Kim, M.S. Glucosidase inhibitor from Umbilicaria esculenta. Can. J. Microbiol. 2000, 46, 1077–1081. [Google Scholar] [CrossRef]
  69. Cunha, M.N.C.; Silva, N.M.V.; Teixeira, M.F.S.; Mota, R.A.; Lima, J.L.; Porto, T.S.; Porto, A.L.F. Actinomycetes producers of beta-lactamases inhibitors with antimicrobial activity against microorganisms isolated from bovine mastitis. Arq. Bras. Med. Vet. E Zootec. 2010, 62, 1312–1319. [Google Scholar] [CrossRef]
  70. Silva, G.M.M.; Bezerra, R.P.; Teixeira, J.A.; Porto, T.S.; Lima, J.L.; Porto, A.L.F. Fibrinolytic protease production by new Streptomyces sp. DPUA 1576 from Amazon lichens. Electron. J. Biotechnol. 2015, 18, 16–19. [Google Scholar] [CrossRef]
  71. Silva, G.M.M.; Bezerra, R.P.; Teixeira, J.A.; Silva, F.O.; Correia, J.M.; Porto, T.S.; Lima-Filho, J.L.; Porto, A.L.F. Screening, production and biochemical characterization of a new fibrinolytic enzyme produced by Streptomyces sp. (Streptomycetaceae) isolated from Amazonian lichens. Acta Amaz. 2016, 46, 323–331. [Google Scholar] [CrossRef]
  72. Huo, L.; Hug, J.J.; Fu, C.; Bian, X.; Zhang, Y.; Mueller, R. Heterologous expression of bacterial natural product biosynthetic pathways. Nat. Prod. Rep. 2019, 36, 1412–1436. [Google Scholar] [CrossRef]
  73. Cravens, A.; Payne, J.; Smolke, C.D. Synthetic biology strategies for microbial biosynthesis of plant natural products. Nat. Commun. 2019, 10, 2142. [Google Scholar] [CrossRef]
  74. Ryan, K.S. Biosynthetic gene cluster for the cladoniamides, bis-indoles with a rearranged scaffold. PLoS ONE 2011, 6, 23694. [Google Scholar] [CrossRef]
  75. Song, Z.Q.; Xu, T.C.; Wang, J.F.; Hou, Y.G.; Liu, C.S.; Liu, S.S.; Wu, S.H. Secondary metabolites of the genus Amycolatopsis: Structures, bioactivities and biosynthesis. Molecules 2021, 26, 1884. [Google Scholar] [CrossRef]
Figure 1. Collection points for lichen samples (the red dots represent the sampling sites).
Figure 1. Collection points for lichen samples (the red dots represent the sampling sites).
Ijms 24 07341 g001
Figure 2. The proportions of novel actinomycete species belonging to different families.
Figure 2. The proportions of novel actinomycete species belonging to different families.
Ijms 24 07341 g002
Figure 3. Chemical structures of compounds 113 from lichen-associated actinomycetota.
Figure 3. Chemical structures of compounds 113 from lichen-associated actinomycetota.
Ijms 24 07341 g003
Figure 4. Chemical structures of compounds 1425 from lichen-associated actinomycetota.
Figure 4. Chemical structures of compounds 1425 from lichen-associated actinomycetota.
Ijms 24 07341 g004
Figure 5. Chemical structures of compounds 2640 from lichen-associated actinomycetota.
Figure 5. Chemical structures of compounds 2640 from lichen-associated actinomycetota.
Ijms 24 07341 g005
Figure 6. Chemical structures of compounds 4144 from lichen-associated actinomycetota.
Figure 6. Chemical structures of compounds 4144 from lichen-associated actinomycetota.
Ijms 24 07341 g006
Figure 7. Chemical structures of compounds 4551 from lichen-associated actinomycetota.
Figure 7. Chemical structures of compounds 4551 from lichen-associated actinomycetota.
Ijms 24 07341 g007
Figure 8. Chemical structures of compounds 5264 from lichen-associated actinomycetota.
Figure 8. Chemical structures of compounds 5264 from lichen-associated actinomycetota.
Ijms 24 07341 g008
Figure 9. Chemical structures of compounds 6576 from lichen-associated actinomycetota.
Figure 9. Chemical structures of compounds 6576 from lichen-associated actinomycetota.
Ijms 24 07341 g009
Figure 10. Chemical structures of compounds 7793 from lichen-associated actinomycetota.
Figure 10. Chemical structures of compounds 7793 from lichen-associated actinomycetota.
Ijms 24 07341 g010
Figure 11. Chemical structures of compounds 94100 from lichen-associated actinomycetota.
Figure 11. Chemical structures of compounds 94100 from lichen-associated actinomycetota.
Ijms 24 07341 g011
Figure 12. Chemical structures of compounds 101108 from lichen-associated actinomycetota.
Figure 12. Chemical structures of compounds 101108 from lichen-associated actinomycetota.
Ijms 24 07341 g012
Figure 13. Chemical structures of compounds 109113 from lichen-associated actinomycetota.
Figure 13. Chemical structures of compounds 109113 from lichen-associated actinomycetota.
Ijms 24 07341 g013
Figure 14. Chemical structure of compound 114 from lichen-associated actinomycetota.
Figure 14. Chemical structure of compound 114 from lichen-associated actinomycetota.
Ijms 24 07341 g014
Figure 15. (A) The gene clusters of Streptomyces sp. YIM130001. (B) The biosynthetic pathway of geninthiocin B [28].
Figure 15. (A) The gene clusters of Streptomyces sp. YIM130001. (B) The biosynthetic pathway of geninthiocin B [28].
Ijms 24 07341 g015
Figure 16. (A) The gene clusters of cladoniamides. (B) The biosynthetic pathway of cladoniamides A–C [74].
Figure 16. (A) The gene clusters of cladoniamides. (B) The biosynthetic pathway of cladoniamides A–C [74].
Ijms 24 07341 g016
Table 1. Novel actinomycetota taxa isolated from lichens between 2007 and 2022.
Table 1. Novel actinomycetota taxa isolated from lichens between 2007 and 2022.
FamilyGenusSpeciesLichen HabitatsMediaRefs.
AurantimonadaceaeAureimonasAureimona leprariaeYunnan Province, south-west PR ChinaHumic acid–vitamin agar and ISP2[8]
MethylobacteriaceaeMethylobacteriumMethylobacterium planiumYunnan Province, south-west PR ChinaHumic acid–vitamin agar[9]
MicrobacteriaceaeGlaciibacterGlaciibacter flavusThe south bank forest of the Baltic Sea, GermanyHumic acid–vitamin agar[10]
FrondihabitansFrondihabitans cladoniiphilusThe natural spruce forest at Koralpe, in the Austrian AlpsTryptone–yeast extract medium and ISP2[11]
LeifsoniaLeifsonia licheniaThe Botanical Garden of the University of TokyoModified Detmer medium and nutrient agar[12]
NaasiaNaasia lichenicolaThe south bank of the Baltic Sea, GermanyHumic acid–vitamin agar and YIM 38 medium[13]
SchumannellaSchumannella luteolaTokyo, JapanModified Detmer medium and nutrient agar [14]
SubtercolaSubtercola lobariaeJiaozi Snow Mountain, Yunnan Province, ChinaPotato dextrose agar and ISP2[15]
MicromonosporaceaeActinoplanesActinoplanes lichenisMaha Sarakham Province, ThailandHumic acid–vitamin agar with nalidixic acid and cycloheximide and ISP2[16]
Actinoplanes lichenicolaMaha Sarakham Province, ThailandHumic acid–vitamin agar with nalidixic acid and cycloheximide and ISP2[17]
Actinoplanes ovalisporusMaha Sarakham Province, ThailandHumic acid–vitamin agar with nalidixic acid and cycloheximide and ISP2[17]
NakamurellaceaeNakamurellaNakamurella albusYunnan Province, south-west PR ChinaHumic acid–vitamin agar and ISP2[18]
NakamurellaNakamurella leprariaeYunnan Province, south-west PR ChinaHumic acid–vitamin agar and YIM 38 medium[19]
NocardioidaceaeNocardioidesNocardioides exalbidusIzu-Oshima Island, JapanIAM-A1 agar medium and trypticase soy agar[20]
PromicromonosporaceaeLuteimicrobiumLuteimicrobium albumRishiri Island, JapanHumic acid–vitamin agar with nalidixic acid and cycloheximide and nutrient agar[21]
PseudonocardiaceaeActinomycetosporaActinomycetospora iriomotensisIriomote Island, Japan.Humic acid–vitamin agar with nalidixic acid and cycloheximide and nutrient agar[22]
Actinomycetospora rishiriensisRishiri Island, Hokkaido, Japan.Humic acid–vitamin agar with nalidixic acid and cycloheximide[23]
PseudonocardiaPseudonocardiaNahuel Huapi National Park, PatagoniaArtificial soil agar and Emerson’s yeast extract–starch agar and KEHE agar[24]
RhodobacteraceaeParacoccusParacoccus lichenicolaYunnan Province, south-west PR ChinaHumic acid–vitamin agar and YIM 38 medium[25]
RubellimicrobiumRubellimicrobium rubrumThe south bank forest of the Baltic Sea, GermanyHumic acid–vitamin agar and YIM 38 medium [26]
StreptomycetaceaeStreptomycesStreptomyces lichenisChiang Rai Province, ThailandArginine–vitamin agar and ISP2[27]
StreptomycesThe tropical rainforest in Xishuangbanna, Yunnan, ChinaYIM 212 medium[28]
Streptomyces parmotrematisDoi Suthep-Pui National Park, Chiang Mai Province, Thailand.Humic acid–vitamin agar and starch casein nitrate agar with cycloheximide and nalidixic acid[29]
ThermomonosporaceaeActinomaduraActinomadura violaceaPong Phra Bat Waterfall, Chiang Rai Province, ThailandHumic acid–vitamin agar with nalidixic acid and cycloheximide[30]
Actinomadura parmotrematisChiang Rai Province, ThailandStarch–casein nitrate agar with nalidixic acid and cycloheximide[31]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, Q.; Song, Z.; Li, X.; Hou, Y.; Xu, T.; Wu, S. Lichen-Derived Actinomycetota: Novel Taxa and Bioactive Metabolites. Int. J. Mol. Sci. 2023, 24, 7341. https://doi.org/10.3390/ijms24087341

AMA Style

Yang Q, Song Z, Li X, Hou Y, Xu T, Wu S. Lichen-Derived Actinomycetota: Novel Taxa and Bioactive Metabolites. International Journal of Molecular Sciences. 2023; 24(8):7341. https://doi.org/10.3390/ijms24087341

Chicago/Turabian Style

Yang, Qingrong, Zhiqiang Song, Xinpeng Li, Yage Hou, Tangchang Xu, and Shaohua Wu. 2023. "Lichen-Derived Actinomycetota: Novel Taxa and Bioactive Metabolites" International Journal of Molecular Sciences 24, no. 8: 7341. https://doi.org/10.3390/ijms24087341

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

Article Metrics

Back to TopTop