Next Article in Journal
A Comparative Study of the In Vitro Intestinal Permeability of Pinnatoxins and Portimine
Previous Article in Journal
Identification of Novel LCN2 Inhibitors Based on Construction of Pharmacophore Models and Screening of Marine Compound Libraries by Fragment Design
Previous Article in Special Issue
New Cyclopiane Diterpenes and Polyketide Derivatives from Marine Sediment-Derived Fungus Penicillium antarcticum KMM 4670 and Their Biological Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bioactive Angucyclines/Angucyclinones Discovered from 1965 to 2023

School of Biological Science and Technology, University of Jinan, 336 West Road of Nan Xinzhuang, Jinan 250022, China
*
Author to whom correspondence should be addressed.
Mar. Drugs 2025, 23(1), 25; https://doi.org/10.3390/md23010025
Submission received: 9 November 2024 / Revised: 25 December 2024 / Accepted: 30 December 2024 / Published: 5 January 2025
(This article belongs to the Special Issue Natural Products Isolated from Marine Sediment)

Abstract

:
Angucyclines/angucyclinones, a class of polyketides with diverse chemical structures, display various bioactivities including antibacterial or antifungal, anticancer, anti-neuroinflammatory, and anti-α-glucosidase activities. Marine and terrestrial microorganisms have made significant contributions to the discovery of bioactive angucyclines/angucyclinones. This review covers 283 bioactive angucyclines/angucyclinones discovered from 1965 to 2023, and the emphasis is on the biological origins, chemical structures, and biological activities of these interesting natural products.

1. Introduction

Nature has always been a significant source in the history of drug discovery, and the ocean has long been recognized as a reservoir of numerous lead compounds owing to its unique environment [1]. Over the past two decades, a substantial number of new marine natural products (MNPs) have been discovered [2]. Among them, over 45% of bioactive molecules sourced from microorganisms are produced by actinomycetes [3]. As the most well-known genus of actinomycetes, Streptomyces continues to yield a great diversity of novel bioactive compounds with varied chemical structures. Increasing evidence suggests that Streptomyces spp. are prolific producers of secondary metabolites with antibacterial/antifungal [4], anticancer [5], anti-neuroinflammatory [6], and anti-α-glucosidase [7] activities.
Angucyclines/angucyclinones are a class of polyaromatic polyketides that exhibit a huge diversity in chemical structures [8]. The first report of angucyclines/angucyclinones from Streptomyces species could date back to 1965 [9]. Angucyclines/angucyclinones could be isolated from both marine and terrestrial actinomycetes, especially Streptomyces. The decanone derived from acetyl-CoA is cyclized by polyketide cyclases to form the tetraene core of angucyclines/angucyclinones [10,11] with a characteristic angular benz[α]anthraquinone framework (the classical type) [12]. In certain instances, the typical angular tetracyclic angucyclines/angucyclinones undergo rearrangement into linear tetracyclic or tricyclic systems through enzymatic or non-enzymatic modifications, resulting in oxidized or rearranged benz[α]anthraquinone frameworks (the non-classical type) [13]. The oxidation state of the framework and the positions of substituents, combined with the presence of different types and numbers of sugars in O- and C-glycosides, contribute to the structural diversity of the angucyclines/angucyclinones [12]. Previous reviews published in 1992, 2012, and 2020 provided detailed and comprehensive summaries of their structures [10,12,14].
The structural diversity of angucyclines/angucyclinones confers upon them a variety of biological and pharmacological activities. The diverse chemical structures and biological activities of the compounds make them a focal point in drug discovery. Herein, the bioactive angucyclines/angucyclinones are categorized into compounds displaying both cytotoxic and antimicrobial activities, cytotoxic or antimicrobial activities only, and other activities, encompassing 283 angucyclines/angucyclinones discovered from 1965 to 2023. In this paper, the biological activities, chemical structures, and biological sources of these fascinating molecules are introduced.

2. Bioactive Angucyclines/Angucyclinones

The search for and screening of angucyclines/angucyclinones with bioactivity were conducted under the guidance of Preferred Reporting Items for Systematic reviews and Meta-Analyses (PRISMA) statement [15] (Figure 1). Cytotoxic activity with an IC50 value less than 10 μM, antibacterial or antifungal activity with an MIC value less than 128 μg/mL, and other activity that was comparable to or stronger than the positive control are considered as a bioactive compound.

2.1. Cytotoxic and Antibacterial or Antifungal Activities

2.1.1. Marine-Derived Angucyclines/Angucyclinones

SS-228 Y (1) was obtained from sediment-derived Chainia purpurogena SS-228 (Sagami Bay, Japan) and exhibited various bioactivities. Compound 1 could prolong the survival period of mice inoculated with Ehrlich ascites tumor when the dosage was above 1.56 μg/piece/day in 10 days. Meanwhile, it showed broad-spectrum inhibition against Gram-positive bacteria with the minimum inhibitory concentrations (MICs) falling in the range of 0.78–12.5 μg/mL, except Mycobacterium tuberculosis [16] (Figure 2, Table 1). Streptomyces sp. HB202, which was isolated from the marine sponge Halichondria panicea, could generate a benz[a]anthracene mayamycin (2). Compound 2 exerted significant cytotoxic activities against HepG2 (hepatocellular carcinoma cells), HT-29 (colon cancer cells), GXF251L (gastric cancer cells), LXF529L (non-small-cell lung cancer cells), MAXF401NL (mammary cells), MEXF462NL (melanoma cells), PAXF1657L (pancreatic cancer cells), and RXF486L (renal carcinoma cells) with semi-inhibitory concentration (IC50) values in the range of 0.13–0.3 μM. Meanwhile, it could also inhibit Bacillus subtilis DSM 347, Brevibacterium epidermidis DSM 20660, Dermabacter hominis DSM 7083, Klebsiella pneumoniae, Pseudomonas aeruginosa DSM 50071, Staphylococcus aureus ATCC 12600, S. epidermidis DSM 20044, and S. lentus DSM 6672, with the IC50 values within 0.31–8.4 μM, comparable to the positive control chloramphenicol [17]. A 3 m deep-soil-derived Streptomyces sp. QD01-2 collected in Qingdao, China, produced gilvocarcin HE (3) together with gilvocarcins H (4), V (5), and M (6), and all of them showed antibacterial activities against S. aureus, B. subtilis, Escherichia coli, and Candida albicans with MIC values of 0.1–25 μM. Gilvocarcin V (5) was cytotoxic to MCF-7 (breast cancer cells), K562 (leukemic cells), and P388 (mouse leukemia cells) with IC50 values ranging within 0.8–1.8 μM [18]. The marine S. fradiae PTZ0025 was the producer of the fradimycins A (7), B (8), and MK844-mF10 (9), which exhibited inhibition of human colon cancer HCT-15 (colon cancer cells), SW620 (colon cancer cells), and rat glioma C6 cells (IC50 values = 0.13–6.46 μM) as well as S. aureus (MICs = 6.0, 2.0, and 4.0 μg/mL, respectively) [19].
Rabelomycin (10) and phenanthroviridone (11) were isolated from the culture of Micromonospora rosaria SCSIO N160, which was obtained from a sediment sample from the South China Sea. Both of them could inhibit SF-268 (neurocarcinoma cells), MCF-7, and NCI-H460 (large-cell lung cancer cells) with IC50 values of 0.09–9.91 μM, and they also displayed antibacterial activities against E. coli ATCC 25922, S. aureus ATCC 29213, B. thuringiensis SCSIO BT01, and B. subtilis SCSIO BS01 with MICs of 0.25–60 μg/mL [20]. The marine-derived Micromonospora echinospora SCSIO 04089 generated homophenanthroviridone (12), homophenanthridonamide (13), nenesophanol (14), rabelomycin E (15), and homorabelomycin (16), which exhibited activities against cancer cells and pathogenic bacteria or fungi. Compound 12 could inhibit SF-268, MCF-7, and HepG2 cells with IC50 values ranging from 1.4 to 5.4 µM, and the IC50 values of 14 and 16 ranged from 7.6 to 12.5 µM, while 13 could inhibit only HepG2 cells (IC50 = 4.0 µM). Compound 12 also displayed inhibition to S. aureus ATCC 29213, B. thuringensis SCSIO BT01, B. subtilis 1064, M. luteus SCSIO ML01, and methicillin-resistant S. aureus (MRSA) shhs-A1, and their MICs ranged from 2 to 4 μg/mL, while 15 could inhibit S. aureus ATCC 29213 and M. luteus SCSIO ML01 at the concentrations of 4 and 8 μg/mL, respectively [21].
(±)-Actinoxocine (17), actinaphthorans A (18) and B (19) [22], as well as (±)-pratenone A (20) [23], were isolated from S. pratensis KCB-132, which was associated with sediment collected in Jiaozhou Bay, China. Compound 17 is characterized by a unique epoxybenzo[f]naphtho[1,8-bc]oxocine carbon skeleton. Compounds 18 and 19 were two unusual C-ring cleavage analogues with cytotoxicities and antibacterial activities against human colon cancer cells LS180 (IC50 values = 1.9 μM), B. cereus (MIC = 2 μg/mL), and Colletotrichum lagenarium (MIC = 2 μg/mL). Enantiomers of 17 could inhibit various bacteria and fungi with MICs of 8–32 μg/mL [22], while (±)-pratenone A (20) revealed antibacterial activity against S. aureus CMCC 26003 with an MIC of 8 μg/mL [23]. Further study of Streptomyces sp. KCB-132 led to the isolation of the nitrogen-containing enantiomers (±)-pratensilin D (21) and compound 22, featuring an A-ring cleavage structural property. Compound (–)-21 exhibited moderate cytotoxicity to the NCI-H460 and HepG2 cell lines, with respective IC50 values of 4.6 and 9.3 µg/mL, while (+)-21 was active only against NCI-H460 cells (IC50 = 9.2 µg/mL). Compound 22 displayed cytotoxicities to colon 38 (colon cancer) and HeLa (cervical cancer) cells, with IC50 values of 7.3 and 10.3 µg/mL, respectively. Compound (–)-21 also exhibited selective inhibitory activity against B. cereus CMCC 32210 with an MIC value of 4 µg/mL, while (+)-21 showed no efficacy against all tested microbial strains, up to 64 µg/mL [24]. The culture extract of Streptomyces sp. KCB-132 also contained an antibiotic compound actetrophenol A (23), which displayed moderate activities against Gram-positive strains with MICs ranging within 1–16 μg/mL. Meanwhile, 23 also showed inhibition toward multiple resistant strains, especially S. aureus and Enterococcus faecium, with an MIC of 4 μg/mL, better than the positive control, penicillin (MIC > 32 μg/mL) [25].
The research group of 23 also discovered 24, from S. pratensis KCB-132, with moderate activities against multiple resistant “ESKAPE” pathogens (E. faecium, S. aureus, K. pneumoniae, Acinetobacter baumannii, P. aeruginosa, and Enterobacter species); the MICs of 24 ranged within 3.1–21.4 μg/mL, comparable to the positive controls ampicillin, amikacin, and ciprofloxacin [26]. S. ardesiacus 156VN-095 was isolated from a sample collected near Nha Trang Bay, Vietnam, and the fermentation extract contained urdamycins W (25) and X (26), grincamycin U (27), as well as an analogue, urdamycin E (28). Compounds 25, 26, and 28 were cytotoxic to ACHN (renal adenocarcinoma cells), HCT-15, MDA-MB-231 (breast cancer cells), NCI-H23 (non-small-cell lung cancer cells), NUGC-3 (gastric cancer cells), and PC-3 (prostate cancer cells) with GI50 (50% growth inhibition concentration) values of 0.019–0.150 μM, comparable to those (0.140–0.162 μM) of Adriamycin. In addition, 2527 displayed antibacterial activities against B. subtilis KCTC 1021, Micrococcus luteus KCTC 1915, and S. aureus KCTC 1927, and the MICs ranged from 8 to 64 µg/mL [27]. Marine-derived Streptomyces sp. BCC45596 collected from Sichang Island (5 m deep, Chonburi province, Thailand) generated C-glycosylated benz[α]anthraquinone urdamycinone E (29), urdamycinone G (30), dehydroxyaquayamycin (31), and urdamycin E (28). Compounds 2931 displayed inhibition against KB (oral epidermoid cancer cells), MCF-7, NCI-H187 (retinoblastoma cells) and Vero (African green monkey kidney cells) (IC50 values = 0.092–15.46 μg/mL); M. tuberculosis (IC50 values = 3.13–12.50 μg/mL) as well as Plasmodium falciparum (IC50 values = 0.0534–22.93 μg/mL) [28]. Streptomyces sp. SCSIO 11594 was isolated from a 2403 m deep sediment sample collected from the South China Sea, and the fermentation broth contained marangucyclines A (32) and B (33). The ketose-containing compound 33 exhibited inhibition against A594 (non-small-cell lung cancer cells), CNE2 (parotid cyst cancer cells), HepG2, and MCF-7, and the IC50 values ranged from 0.24 to 0.56 μM; it also displayed selectivity between cancer cells and normal cells. Marangucyclines A (32) and B (33) also showed weak antibacterial activities against E. faecalis ATCC29212 (both MICs at 64 μg/mL) [29]. S. lusitanus SCSIO LR32 was isolated from a deep-sea sediment sample collected in the South China Sea, China. A-7884 (34) and grincamycin J (35), produced by SCSIO LR3, exhibited cytotoxic activities against human cancer cells MDA-MB-435 (melanoma), MDA-MB-231, NCI-H460, HCT-116 (colon cancer), and HepG2 as well as MCF-10A (normal breast epithelial cells) with IC50 values ranging from 0.4 to 6.9 µM. In addition, A-7884 (34) demonstrated antimicrobial activity against M. luteus, with an MIC value of 1.95 µg/mL [30].

2.1.2. Terrestrial-Derived Angucyclines/Angucyclinones

The fermentation of S. gilvotanarens NRRL 11382, a new species isolated from a soil sample collected in Kochi, Japan, led to the discovery of gilvocarcins V (5) and M (6). Both were bioactivated toward tumor cells such as sarcoma 180 (mouse malignant sarcoma cells) and P388. Mice treated with compound 5 lived significantly longer than control animals. Compounds 5 and 6 also showed inhibition against S. aureus ATCC 6538P and B. subtilis No. 10707, with MICs of 0.05/1.6 and 0.78/25 μg/mL, respectively [31]. Saquayamycins A–D (3639) were produced by S. nodosus MH190-16F3, associated with a soil sample taken from tobacco growth areas (Kitakyushu, Japan), and could inhibit Adriamycin-sensitive (P388/S) and Adriamycin-resistant (P388/ADR) sublines of P388 in vitro, with IC50 values of 0.06–0.15 μg/mL. The LD50 (median lethal dose) by intraperitoneal injection of saquayamycins A (36) and B (37) in mice were 6.25–12.5 mg/kg. In addition, 3639 exhibited antibacterial activities against S. aureus FDA209P, M. lysodeikticus IFO 3333, M. luteus PCI1001, and B. subtilis PCI 219 with MICs of 1.56–6.25 μg/mL [32] (Figure 3). Further research revealed that saquayamycins A (36) and B (37) exhibited remarkable activities against L-1210 (mouse leukemia cells), A549, and HT-29 (IC50 values = 0.004, 0.2, and 0.06 μg/mL, respectively), and demonstrated distinct toxicity in vivo [33]. Aquayamycin (40) and Adriamycin (positive control) could also inhibit the aforementioned two cells with IC50 values at 2.0/2.2 and 0.01/0.55 μg/mL, respectively [32]. S. antibioticus Tü 6040 was isolated from a soil sample from Iguaguu, Argentina, and the mycelium extract contained the simocyclinones D4 (41) and D8 (42), both of which showed cytotoxicities and antibacterial activities, while the GI50s against HMO2 (human milk oligosaccharides) and MCF-7 cell lines ranged from 0.3 to 5.6 μM, better than 5-fluorouracil (positive control, GI50 = 1.2 and 50 μM). The MICs against B. brevis DSM30 were 30 and 10 μg/mL, respectively [34,35].
Kerriamycins A–C (4345) (produced by S. violaceolatus) [36] and capoamycin (46) (produced by soil-derived S. capoamus collected in Fujioka, Japan) [37], which were discovered by the same research group, could prolong the survival periods of mice bearing Ehrlich ascites carcinoma when they were subjected to intraperitoneal injections on days 1 and 5. The LD50 of 46 was 15 mg/kg (ip), and the antitumor activity was based on an induction effect on the differentiation process of mouse myeloid leukemia cells (M1). Meanwhile, 4346 displayed inhibitory activities against S. aureus FDA 209P, B. subtilis ATCC 6633, B. cereus IAM 1729, and M. luteus ATCC 9341, with the MICs at 1.65–25 μg/mL. In addition, 46 showed activity against Penicillium chrysogentrrn ATCC 10002 and Trichophyton mentagrophytes, and the MICs were 1.56 and 12.5 μg/mL, respectively. Grincamycin (47), produced by S. griseoincarnatus, was also isolated by the same group above and was revealed to exert significant cytotoxicity toward the P388 cell line (IC50 = 13 ng/mL) and moderate antibacterial activities against S. aureus FDA 209P (MIC = 50 μg/mL), M. luteus ATCC 9341 (MIC = 25 μg/mL), and B. cereus IAM 1729 (MIC = 50 μg/mL) [38].
The fermentation of S. venezuelae ISP5230 led to the discovery of jadomycins S (48), T (49), DT (50), B (51), L (52), F (53), DM (54), DS (55), and Y (56), with cytotoxicities and antibacterial activities. Compounds containing L-serine and D/L-threonine displayed higher cytotoxicities against MDA-MB-435; the IC50 values of 4850 ranged from 1.06 to 2.82 μM. Jadomycins containing aromatic side chains showed the lowest activities against T-47D (breast cancer cells), indicating that the activities of 4850 are related to the hydrogen donor of the hydroxyl side chain. All these jadomycins (4856) displayed inhibitory activities against S. aureus C622 (ATCC 25923), S. aureus 305, S. aureus BeckerCP8 (ATCC 49525), S. aureus BeckerLyc12CP336 (ATCC 55804), S. epidermidis C960 (ATCC 14990), S. epidermidis C621 (clinical isolate), and B. subtilis C971 (ATCC 6633), with MICs of <1–64 μg/mL. Especially, the MICs against S. aureus C623 (MRSA) were all less than 1–8 μg/mL, better than that of the positive control erythromycin (MIC > 128 μg/mL) [39]. Jadomycins retain their cytotoxic properties toward multi-drug-resistant (MDR) breast cancer cells because they cannot be expelled by ATP-binding cassette (ABC) transporters, which is an important reason for tumor resistance to doxorubicin [40]. Further study confirmed that the cytotoxicities of jadomycins 48 and 5153 were minimally affected by the efflux transporter functions of ABCB1, ABCC1, and ABCG2 [41]. Jadomycin DS (55) could bind to a variety of proteins, likely in a non-specific manner. The quality and quantity of direct binding between topoisomerase IIβ and jadomycin DS (55) were demonstrated by WaterLOGSY NMR spectroscopy [42]. The addition of Nε-trifluoroacetyl-L-lysine in the fermentation of S. venezuelae ISP5230 led to the isolation of 57 and 58, containing amide and furan rings. The oxazolone-ring-containing 57 was active against MRSA (MIC = 3 μg/mL), S. warneri (3 μg/mL), and vancomycin-resistant Enterococcus faecium (VRE, 13 μg/mL). In contrast, 58 was much less active (MIC ≥ 100 μg/mL) against the three Gram-positive strains. The enhanced antibiotic activity of 57 in comparison with 58 implies that the hemiaminal ether functionality plays an important role in the antimicrobial properties of the jadomycins [43]. Streptomyces sp. AC113 was isolated from the root of Taxus chinensis (Bakata) and could produce (–)-8-O-methyltetrangomycin (59), 8-O-methyltetrangulol (60), and 8-O-methyl-7-deoxo-7-hydroxytetrangomycin (61). These three compounds were cytotoxic to mouse melanoma B16 (IC50 = 0.054–7.13 μg/mL) and HT-29 (IC50 = 8.59–66.9 μg/mL) cells, and they showed antibacterial activities against P. aeruginosa CCM 3955, S. aureus CCM 3953, E. coli CCM 3988, L. monocytogenes NCTC 4886, B. subtilis CCM 2216, and B. cereus, with MIC values ranging from 0.6 to 78.6 μg/mL [44].
In addition to fermenting the producer of jadomycins in the presence of amino acid analogues, semi-synthesis, structural gene deletion, and deletion or heterologous expression of sugar biosynthetic genes led to the discovery and isolation of more than 70 jadomycins [45]. This enabled a comprehensive evaluation of the cytotoxic and antibacterial activity of jadomycins and facilitated the study of the mechanism of bioactivity [46]. The cytotoxicity of jadomycins involves the generation of cytosolic superoxide via a Cu(II)–jadomycin reaction, a mechanism common to all the jadomycins tested and observed in MCF7-CON and drug-resistant MCF7-TXL cells. The generation of intracellular ROS in the superoxide dismutase 1, glutathione, and peroxiredoxin/thioredoxin cellular antioxidant enzyme pathways was scavenged by jadomycin treatment. The blocking of these antioxidant pathways may enhance the cytotoxic potency of jadomycin in both drug-sensitive and drug-resistant breast cancers [47]. The breast cancer cell death induced by jadomycins is independent of ROS activity through the inhibition or poisoning of type II topoisomerases and the induction of DNA damage and apoptosis, and jadomycins B (51) and F (53) selectively poison topoisomerase IIb to induce DNA damage and apoptosis. [48]. The pharmacokinetics, toxicities, and antitumoral effects in zebrafish larvae and mice showed that jadomycin B (51) had a good safety profile and provided partial antitumoral effects [49] together with the generation of reactive oxygen species (ROS) induced by copper [47], the inhibition of topoisomerase IIα and IIβ [48], and the avoidance of ABC transporters [40]. All these observations suggest that jadomycins may be used as a breast cancer chemotherapy in clinical practice, while further studies on their ability to penetrate the blood–brain barrier are required [50].
Langkocyclines A1–A3 (6264) were obtained from the extract of Streptomyces sp. Acta 3034, which was associated with the rhizospheric soil of Clitorea sp. Compound 64 displayed inhibitory activities against HepG2 and NIH 3T3 (mouse embryonic fibroblast) cells, with IC50 values of 2.5–5.0 μM, while 6264 showed inhibition toward B. subtilis, with IC50 values at 40.7, 4.07, and 2.17 μM, respectively [51] (Figure 4). The fermentation of soil-derived Saccharopolyspora BCC 21906 (Chanthaburi, Thailand) led to the isolation of saccharosporones A (65) and B (66), as well as (+)-ochromycinone (67) and tetrangulol methyl ether (68). Compounds 65 and 66 exhibited cytotoxic activities against KB, MCF-7, and NCI-H187 cell lines with IC50 values ranging within 3.4–9.1µM. Compounds 6568 showed growth inhibition against M. tuberculosis, with IC50 values of 76.2, 72.7, 40.8, and 19.7 µM, respectively [52]. Heterologous expression of two eDNA-derived KSβ sequences associated with the biosynthesis of (C24)-pradimicin and (C26)-xantholipin-type metabolites in Streptomyces salbus led to the isolation of calixanthomycin A (69) and the arenimycins C (70) and D (71), and all of them were cytotoxic toward HCT-116 cells, with respective IC50 values at 0.43 nM, 0.17 μM, and 2.8 μM. Meanwhile, 6971 also could inhibit MRSA and B. subtilis RM125 with MICs of 0.0015–50 μg/mL [53]. The overexpression in S. chattanoogensis L10 (CGMCC 2644) of a pathway-specific activator gene under the constitutive ermE* promoter successfully triggered the expression of the angucycline biosynthetic genes and led to the discovery of chattamycins A (72) and B (73). Compound 72 was cytotoxic to MCF-7 (IC50 = 6.46 μM), while 73 showed inhibitory activities against MCF-7 (IC50 = 1.08 μM) and HepG2 (IC50 = 5.93 μM) cells. In addition, 73 exhibited activity against B. subtilis ATCC 67736 (IC50 = 102.59 μM) [54]. The method of site-directed mutagenesis led to the generation of ten mutants of S. chattanoogensis L10 (CGMCC 2644) with point mutations in the highly conserved region of rpsL (encoding the ribosomal protein S12) or rpoB (encoding the RNA polymerase β-subunit). L10/RpoB (H437Y) accumulated anthrachamycin (74), which was absent in the wild type. In the 2,2′-amino-di(2-ethyl-benzothiazoline sulfonic acid-6) ammonium salt (ABTS) free radical scavenging assay and the ferric ion reducing antioxidant power (FRAP) iron reduction assay, 74 showed antioxidant activity at 67.28 and 24.31 mg VCE/g LP, respectively [55].
C-glycosylated benz[α]anthraquinone, dehydroxyaquayamycin B (75) was isolated from the fermentation broth of S. blastomycetica F4-20 associated with the root of Tripterygium wilfordii Hook. f. Compound 75 showed cytotoxic activities against the BGC823 (gastric adenocarcinoma) and HeLa cell lines with IC50 values of 0.71 and 1.34 μg/mL, respectively, and it also displayed antifungal activities against Valsa mali, C. orbiculare, and Fusarium graminearum at 50 μg/mL with inhibition rates of 41.5%, 58.3%, and 51.0%, respectively [56]. S. bulli GJA1, associated with Gardenia jasminoides, was the producer of 76 and 77, both of which were cytotoxic toward OV90 and ES2 ovarian cancer cells with MICs of 0.36/0.55 µM and 2.42/1.69µM, respectively, better than paclitaxel and cisplatin. Compound 76 also showed antivirulence activity by inhibiting the phenol-soluble modulin (PSM) production and the biofilm formation of MRSA [57]. Strain NJES-13T, a newly established actinobacteria genera Aptenodytes in the family Dermatophilaceae, was isolated from the gut microbiota of the Antarctic emperor penguin. The fermentation broth of NJES-13T contained 2-hydroxy-frigocyclinone (78) and 2-hydroxy-tetrangomycin (79). Compound 78 showed inhibitory activities against HL-60 (leukemia), Bel-7402 (hepatocellular carcinoma), and A549 cells, with IC50 values ranging from 4.2 to 8.5 µM, while 78 and 79 could inhibit S. aureus, B. subtilis, and C. albicans with MICs of 5.7–27.2 µg/mL [58]. A soil-derived S. cellulosae YIM PH20352 (Yunnan province, China) produced rabelomycin (10), dehydrorabelomycin (80) [59], urdamycinone B (81) and dehydroxyaquayamycin (31) [60]. Compounds 10 and 80 could inhibit the root rot pathogens of Panax notoginseng, including Plectosphaerella cucumerina, Alternaria panax, F. oxysporum, and F. solani, with MICs of 32–128 μg/mL. Compound 81 exhibited antifungal activities against A. panax and P. cucumerina with MICs at 16 and 64 μg/mL, respectively, and 31 showed inhibitory activity toward A. panax with the MIC at 64 μg/mL. Streptomyces sp. XZHG99T was isolated from a soil sample collected from the Color desert (Tibet Autonomous Region, China), and produced grincamycins L–N (8284) as well as the known compounds rabelomycin (10), moromycin B (85), fridamycin D (86), and saquayamycin B1 (87), all of which showed inhibitions against A549, H157 (non-small-cell lung cancer), MCF-7, MDA-MB-231, and HepG2 cells with the IC50 values ranging from 1.52 to 17.3μM. Compound 10 also exhibited antibacterial activities toward Mycolicibacterium smegmatis and S. aureus, with IC50 values from 0.12 to 23.1 μM [61].
Streptomyces sp. IB201691-2A, which was obtained from the endemic mollusk Benedictia baicalensis of Lake Baikal, was the producer of baikalomycin C (88), rabelomycin (10), and 5-hydroxy-rabelomycin (89). Baikalomycin C (88) displayed inhibition of Huh7.5 (hepatocellular carcinoma) and SW620 cells (IC50 values = 7.62 and 3.87 µM, respectively) as well as S. carnosus DSMZ 20501 (MIC = 62 µM), while 10 and 89 showed inhibitory activities against A549, Huh7.5, and SW620 cells (IC50 values = 7.21–13.43 µM) as well as Erwinia persicina DSMZ 19328, S. carnosus DSMZ 20501, and M. smegmatis DSMZ 43286 (MICs = 31–125 µM) [62]. 12-Deoxo-12-hydroxy-8-O-methyltetrangomycin (90), the C-ring cleavage product of angucyclinone C (91), tetrangomycin (92), and 8-O-methyltetrangomycin (59) were isolated from the secondary metabolites of Streptomyces sp. CB01913 (soil sample of Weishan County, Yunnan Province, China). Compounds 90, 92, and 59 exhibited inhibitory activities against SF295 (malignant glioma cells) and H226 (lung squamous cells) with the IC50 values ranging from 3.1 to 10 μM. Compounds 92 and 59 also inhibited M14 (melanoma cells) with IC50 values of 2.4 and 9.7 μM, respectively. Meanwhile, 91, 92, and 59 displayed antibacterial activities toward S. aureus ATCC 25923, B. subtilis ATCC 23857, and M. smegmatis ATCC 607, with the MIC values ranging from 8.1 to 93 μg/mL [63]. 6,9-Dihydroxytetrangulol (93) was isolated from S. lividans TK23 transformed with a kinanthraquinone biosynthetic gene cluster in which the kiqO gene was disrupted. Compound 93 revealed both cytotoxicity and antibacterial activity; the IC50 toward HL-60 cells was 5.1 μM, and the IC50 values toward S. aureus and C. albicans were 1.9 and 1.1 μM, respectively, better than chloramphenicol [64].

2.1.3. Angucyclines/Angucyclinones from Other Sources

Nocardia lurida was the producer of benzanthrins A (94) and B (95), which exhibited antibacterial activities against various Gram-positive bacteria, with MIC values between 0.2 and 3.1 μg/mL [65], and cytotoxicities against 9KB (nasopharyngeal carcinoma cells) and 9PS (IC50 values = 0.3 and 0.01 μg/mL, respectively) (Figure 5). It was interesting that benzanthrin A (94) caused a reversal of adenosine cyclic 3′,5′-monophosphate-induced morphological changes in AC glioma tumor (9ASK) cells at 10 μg/mL, while no reversal was observed with benzanthrin B (95) [66]. The fermentation broth of S. matensis A-6621 contained PI-083 (96), which exhibited cytotoxicity against the KB cell line with the IC50 at 0.026 μM and inhibitory activities toward S. aureus 209P-JC, Sepidermidis IID 866, E. faecium ATCC 8043, B. cereus S 1101, and B. subtilis ATCC 6633 with the MICs at 0.39, 1.56, 3.13, 12.5, and 1.56 μg/mL, respectively [67]. Brasiliquinones A–C (9799), which were isolated from the culture broth of the pathogenic Nocardia sp. IFM 0089, displayed inhibitory activities toward the L-1210 and P388 cell lines (IC50 values = 2.9–7.0 μg/mL) and were also active against P388/ADR cells, with IC50 values ranging from 3.0 to 3.8 μg/mL. Compounds 9799 also showed antibacterial activities against S. aureus 209P, S. aureus MRSAIFM 62971, M. smegmatis ATCC 607, and M. luteus IFM 2066, with the MICs ranging from 0.39 to 50 μg/mL [68].
Kinamycins A–D (100103), isolated from S. murayamaensis, have a highly unusual and potentially reactive diazo group. Kinamycins A (100) and C (102) showed IC50 values of 10 μM and 0.3 μM, respectively, against Chinese hamster ovary (CHO) cancer cells. Kinamycins A (100) and C (102) also could inhibit the catalytic decatenation activity of DNA topoisomerase IIα, but showed no activity as a topoisomerase II poison. Meanwhile, their inhibition of catalytic activity was not correlated with a cell growth inhibitory effect [69]. Kinamycins A–D (100103) also showed antibacterial activities against Gram-positive bacteria [70]. 4′-acetylated-chrysomycins A (104) and B (105) were discovered during the screening for antitumor agents from the metabolites of actinomycetes, and both compounds showed high cytotoxicities toward most of the tested cancer cells, with IC50 values less than 10 ng/mL. Compound 104 showed strong anti-Gram-positive-bacterial activities toward MRSA and VRE, with MIC values of 0.5–2 μg/mL [71]. S. aureofaciens CCM 3239, received from the Czech Collection of Microorganisms (CCM, Brno, Czech Republic), produced auricin (106) with cytotoxicities against the human ovarian carcinoma cell line A2788 (IC50 = 1.05 μM), cisplatin-resistant cells A2780/CP (IC50 = 0.7 μM), MDA-MB-231 (IC50 = 4.19 μM), and MCF-7 (IC50 = 2.8μM). Compound 106 was active against B. subtilis and S. aureus Newman, with MICs at 4.6 and 9.2 μM, respectively [72].

2.2. Cytotoxicities

2.2.1. Marine-Derived Angucyclines/Angucyclinones

A sediment-derived actinomycete, Streptomyces CNH990, produced marmycins A (107) and B (108), which exhibited cytotoxicities against HCT-116 cells with IC50 values at 60.5 and 1.09 μM, respectively. For marmycin A (107), tumor cell cytotoxicity appeared to coincide with the induction of modest apoptosis and arrest in the G1 phase of the cell cycle [73] (Figure 6). A sponge-derived Saccharopolyspora taberi PEM-06-F23-019B (Tanzanian) produced PM070747 (109), which displayed cytotoxicities against MDA-MB-231, HT-29, and A549 cells with the IC50 values at 0.71, 1.42, and 3.28 μM, respectively [74]. The secondary metabolites of S. lusitanus SCSIO LR32 contained grincamycin (47); grincamycins B (110), C (111), and E (112) [75]; grincamycins H–J (113, 116, and 35), congers P-1894B (vineomycin A1, 114), saquayamycin B (37) [29], vineomycin B2 (115), and A-7884 (34) [30]. Grincamycins B (110), C (111), and E (112) and grincamycin (47) displayed cytotoxicities against the B16 and HepG2 cell lines with IC50 values of 1.1–11 μM. Compounds 110 and 47 could inhibit SW-1990 (pancreatic cancer) and HeLa cell lines with IC50 values of 5.4–11 μM [75], while 37 and 113115 showed inhibitory activities against Jurkat T (acute T-cell leukemia cells) with IC50 values of 0.011–3.0 μM (positive control, doxorubicin, 0.034 μM) [13]. Grincamycins I (116), J (35), and A-7884 (34) were cytotoxic to tumor cells MDA-MB-435, MDA-MB-231, NCI-H460, HCT-116, and HepG2, with the IC50 values at 0.4–6.9 µM, and they also showed toxicity to the normal cells MCF-10A with IC50 values of 22.43–2.90 µM [30]. Meanwhile, saquayamycin B (37), which was isolated from an intertidal sediment-derived Streptomyces sp., displayed significant cytotoxicities against HepG2, SMMC-7721 (hepatocellular carcinoma cells), and PLC-PRF-5 (hepatoma cells Alexander) with the respective IC50 at 0.135, 0.033, and 0.244 μM, better than the positive control doxorubicin (0.706–2.16 μM) [76].
The fermentation broth of the marine Streptomyces sp. M268 contained kiamycin (117), possessing a 1,12-epoxybenz[a]anthracene ring system. Compound 117 showed inhibitory activities against the human cell lines HL-60, A549, and BEL-7402, with respective inhibition rates of 68.2%, 55.9%, and 31.7% at 100 µM [77]. Micromonospora sp., which was isolated from sediment collected off the CátBà peninsula in the East Sea of Vietnam, produced dehydrorabelomycin (80), phenanthroviridone (11), and WS-5995 A (118). Compound 80 showed inhibition against Kuramochi (ovarian cancer cells) with the IC50 at 6.72 μM, and 11 could inhibit Kuramochi and high-grade ovarian cancer cells (OVCAR4) with the respective IC50s of 1.11 and 4.82 μM. Compounds 11, 80, and 118 could inhibit murine ovarian surface epithelial (MOSE) and murine oviductal epithelial (MOE) cells with the LC50 of 2.85–9.80 μM, while 118 displayed cytotoxicity against L-1210 cells with the IC50 value about 0.5 μM [78]. The secondary metabolites of Streptomyces sp. SS13I contained gephyromycin C (119), which exhibited cytotoxicities against PC3 (prostate cancer cells, IC50 = 1.3 μM) and H1975 (lung adenocarcinoma cells, inhibition rate = 48% at 5 μM) [79]. A sediment-derived Streptomyces sp. HN-A124 (Hainan province, China) produced cysrabelomycin (120), which showed inhibitory activity against A2780 cells with the IC50 at 10.23 μM [80]. Vineomycin E (121), together with moromycin B (85) and saquayamycins B1 (87) and B (37), were generated by the marine-derived Streptomyces sp. OC1610.4, and all these compounds displayed potent anti-proliferation against MCF-7, MDA-MB-231, and BT-474 (breast cancer cells), with the IC50 values ranging from 0.16 to 7.72 µM. Meanwhile, saquayamycin B (37) inhibited the migration and invasion of MDA-MB-231 cells in a dose-dependent manner [81]. Moromycin B (85) and saquayamycins B1 (87) and B (37) were also isolated from the secondary metabolites of another marine Streptomyces sp. and exhibited cytotoxicity against SW480 (colon cancer cells), SW620, LoVo (colon cancer cells), HT-29, and QSG-7701 (normal hepatocyte cells), with IC50 values of 0.18–1.57 µM, which were comparable to or better than the positive control doxorubicin. Saq B1(87) could not only induce apoptosis but also inhibit invasion and metastasis in CRC (colon cancer cells) through the PI3K/AKT signaling pathway [82].
Streptomyces sp. XS-16 was obtained from a marine sediment sample (Naozhou Island, China) and generated compound 122, which showed growth inhibitory activities against MDA-MB-231, K562, ASPC-1 (pancreatic cancer cells), H69AR (Adriamycin-resistant small-cell lung cancer cells), and H69 (small-cell lung cancer cells) with IC50 values at 0.32–5.33 µM [83]. Kumemicinones A (123), B (124), and E–G (125127), as well as SF2315B (128), were isolated from the Actinomadura sp. KD439, associated with marine suspended matter near the coast of Kumejima Island (612 m deep, Okinawa, Japan), and all of them could inhibit P388 cells, with the IC50 values ranging from 1.7 to 10.7 μM [84]. Rearranged angucyclinones donghaecyclinones B (129) and C (130) were isolated from the marine sediment-derived Streptomyces sp. SUD119 (volcanic island, Korea). Compound 129 could inhibit hepatocellular carcinoma (SK-HEP1) cells, and 130 could inhibit HCT-116, MDA-MB-231, SNU638 (gastric cancer cells), A549, and SK-HEP1 cells, with IC50 values ranging from 6.0 to 9.6 μM [85]. The Streptomyces sp. HDN15129 isolated from a sediment sample collected in the South China Sea produced monacycliones I (131) and J (132), both of which showed inhibition against multiple human cancer cell lines such as HL-60, K562, SH-SY5Y (neuroblastoma), BEL-7402, U87 (glioblastoma), ASPC-1, and HCT-116 cells, with the IC50 values ranging from 3.5 to 10 μM [86].
The screening of marine actinomycete extracts against the pseudomyxoma peritonei (PMP) cell line ABX023-1 led to the isolation of grincamycins R–T (133135) from Streptomyces sp. CNZ-748. Compounds 133135 showed inhibitory activities against PMP501-1 and PMP457-2 cells, with IC50 values of 1.9–7.9 μM, and could inhibit ABX023-1 and C09-1, with IC50 values of 1.4–10 μM (5-fluorouracil, 2.0–4.0 μM) [87]. Caribbean sponges-derived Streptomyces sp. M7_15 was the producer of frigocyclinone (136) and monacyclinone F (137), which exhibited cytotoxicity against SJCRH30 (rhabdomyosarcoma cells) with the respective EC50 (median effect concentration) values at 5.2 μM and 0.73 μM. The result suggested that additional amino deoxy sugar subunits may be important for the activity of this class of molecules [88]. A gut-derived (Oxya chinensis) Amycolatopsis sp. HCa1 generated (2R, 3R)-2-hydroxy-5-O-methyltetrangomycin (138) together with tetrangomycin (92), PD116779 (139), and sakyomicin A–C (140142), which displayed cytotoxicities against HeLa cells with IC50 values ranging from 0.11 to 0.59 µM. Compound 142 could inhibit SGC-7901 (gastric adenocarcinoma cells) with the IC50 of 4.41 µM, while 140 could inhibit SPC-A-1 (lung cancer cells) with IC50 at 8.34 µM [89].

2.2.2. Terrestrial-Derived Angucyclines/Angucyclinones

OM-4842 (143) was isolated from a soil-derived Streptomyces sp. Om-4842 (Chiba, Japan) and displayed inhibition toward doxorubicin-resistant cells of P388 at 1.5 μg/mL [90] (Figure 7). Rubiginones A1 (144), A2 (145), B1 (146), B2 (147), C1 (148), and C2 (149), secondary metabolites of the soil-derived S. griseorubiginosus No. Q144-2 (Andhra Pradesh, India), displayed significant potentiated cytotoxicities against vincristine-resistant P388 cells, with IC50 values ranging within 0.007–0.23 μg/mL [91]. KY002 and KY40-1 were both soil-derived Streptomyces sp. discovered in the Appalachian Mountains, USA. Streptomyces sp. KY002 produced moromycin B (85), which exhibited inhibitory activities against H-460 and MCF-7 cell lines with GI50s of 5.6 and 5.6 µM, respectively [92]. Streptomyces sp. KY40-1 generated saquayamycins G–K (150154) as well as the known compounds saquayamycins B1(87), A (36), and B (37), which displayed significant cytotoxicities against PC3 cells (IC50 values = 0.0075–1.759 μM) and moderate activities against H-460 cells (IC50 values = 3.30–7.28 μM) [93]. Polycarcin V (155) was produced by S. polyformus sp. nov. YIM 33176, which was associated with a soil sample of Vietnam, and it revealed inhibitory activities against 37 different human tumor cell lines representing 14 different solid tumor types, with the IC70 values ranging from 0.3 to 431.0 ng/mL, indicating a pronounced antitumor specificity [94].
The fermentation broth of Streptomyces sp. N05WA963 contained N05WA963 A (156), B (157), and D (158), which exhibited cytotoxicities against SW620, YES-4 (esophageal cancer cells), U251SP (glioma cells), K562, MDA-MB-231, and T-98 (glioma cells) with IC50 values of 1.0–10.3 µM [95]. Alkaline soil-derived Streptomyces Acta 2930 (Northumberland, UK) generated warkmycin A (159), with antiproliferative activities against NIH-3T3, HepG2, and HT-29 cells; the IC50 values were 2.74, 1.26, and 1.61 µM, respectively [96]. The Himalayan-based Streptomyces sp. PU-MM59 was the producer of himalaquinone G (160), which exhibited cytotoxicities against the PC3 and A549 cell lines with IC50 values of 0.32 and 1.88 μM, respectively [97]. Vineomycin A1 (P-1894B, 114), a noncompetitive prolyl hydroxylase inhibitor (2.2 × 10−6 M, 50%), was isolated from the secondary metabolites of a soil-derived S. albogriseolus subsp. No. 1894 and was necessary for collagen biosynthesis [98]. Compound 114 showed a significant inhibitory effect on Jurkat T-cell proliferation with an IC50 at 0.011 μM [13]. The first total synthesis of vineomycin A1 (114) was accomplished in 2019, and its cytotoxicities were evaluated by MTT assay against A549, HCT-116, and Capan-1 (pancreatic cancer cells), with the IC50 values ranging from 0.01 to 0.64 μM. The test indicated that vineomycin A1 (114) effectively induced cancer cell death via apoptosis, not by acting as a DNA intercalating agent [99].

2.2.3. Angucyclines/Angucyclinones from Other Sources

Landomycin E (161) was produced by S. globisporus 1912 [100], and it displayed inhibitory activity toward tumor cell lines via induction of apoptosis in the low micromolar range for MDA-MB-231 (IC50 = 0.76 mg/mL), HL-60 (1.87 mg/mL), and KB-3-1 (4.3 mg/mL) [101] (Figure 8). Landomycins I (162) and J (163), together with landomycins A (164), B (165), E (161), and D (166), and one landomycinone (167) [102,103,104,105,106,107], were generated by a mutant strain of S. cyanogenus whose glycosyltransferase encoded by lanGT3 was over-expressed [108]. Complementation of gilvocarcin for the mutant S. lividans TK24 (cosG9B3-U), in which the biosynthesis of the natural sugar donor substrate was compromised with various deoxy sugar plasmids, led to the production of the gilvocarcin analogues gilvocarcin V (5), 4′-OH-gilvocarcin V (168), D-olivosyl-gilvocarcin V (169), and polycarcin V (155) with altered saccharide moieties [109]. Compounds that differed in their sugar moieties showed inhibition against LL/2 (mouse lung cancer), MCF-7, and NCI-H460 cell lines, indicating that the anticancer activity of landomycins did not increase simultaneously with the elongation of their oligosaccharide chain lengths [108,109]. However, other studies showed different results in the structure–activity relationship of landomycins.
Landomycins R–W (170175) along with tetrangulol (176) [110]; 5,6-anhydrolandomycinone (177) [102]; landomycinone (167); landomycins A (164), B (165), D (166), F (178) [111], M (179) [112], and O (180) [113]; and tetrangomycin (92) were isolated from the culture broth of S. cyanogenus S-136 [114]. 11-Deoxylandomycinone (181) and landomycins X–Z (182184) were produced by the mutant strain of S. cyanogenus K62 [115]. These compounds showed varying degrees of cytotoxic activity toward MCF-7 (estrogen-sensitive) and MDA-MB-231 (estrogen-insensitive) cell lines. Compounds 164, 167, and 177 showed the best combined activities to both MCF-7 and MDA-MB-231 cells, with 177 for the former and 167 and 164 for the latter. Compounds 173175 and 181184 showed activities against MCF-7, with IC50 values of 1.0–6.7 µM, while compounds 172175 and 181184 could inhibit MDA-MB-231, with IC50 values of 1.2–2.5 µM. Compounds with shorter saccharidal moieties were less potent against MCF-7. The fact that most landomycins with bioactivities had either long penta- or hexasaccharide chains or no sugars at all suggests that the large molecules may act by a different mode of action compared with their small sugar-free congeners [114,115].
The fermentation broth of Streptomyces contained gilvocarcin V (5), which displayed cytotoxicities against sarcoma 180, Ehrlich carcinoma, Meth I fibrosarcoma, MH134 (mouse hepatoma), and P388 cells. After intraperitoneal administration of gilvocarcin V (5) to mice bearing Ehrlich ascites carcinoma, 40% of the treated mice survived for 60 days [31]. The inactivation of the gilU gene in the mutant S. lividans TK24 (cosG9B3-U) led to the production of three analogues of the gilvocarcin-type aryl-C-glycoside compounds, 4′-hydroxy gilvocarcins E (185), M (186), and V (168), which showed different degrees of activity in the anticancer assay. The activity of 185, which lacks an essential vinyl residue for DNA binding, was lower than those of 168 and 186. Nevertheless, the introduction of the 4′-OH group changed an inactive gilvocarcin E (161) into a moderately active one (185) [109,116]. Kinamycin F (187), as a secondary metabolite of S. murayamaensis, was found to induce apoptosis and downregulate cyclin D3 in K562 cells, and it induced single-stranded DNA breaks and inhibited the activity of topoisomerase IIα with an IC50 of 0.33 μM [69,117]. Salinispora pacifica DPJ-0019 (NRRL 50168), which was acquired from the USDA Agricultural Research Service, generated (−)-lomaiviticins C–E (188190) as well as lomaiviticin A (191) and kinamycin C (102), all of which displayed significant cytotoxicities against K562, LNCaP (prostate cancer), HCT-116, and HeLa cells, with the IC50 values ranging from 2 to 589 nM [118]. Inactivation of the flavoenzyme-encoding gene of lsO1 in fluostatin biosynthesis led to the isolation of fluostarenes A (192), B (193), and PK1 (194). Fluostarene B (193) was cytotoxic toward SF-268, MCF-7, HepG2, and A549 cell lines, with IC50 values at 7–10μM, not as good as Adriamycin (1.13–1.42 μM) [119]. BE-7585A (195), which is characterized by a 2-thiosugarand, was isolated from a culture broth of Amycolatopsis orientalis subsp. vinearia and exerted cell inhibitory effects against mouse Ehrlich ascites carcinoma, with an IC50 of 8.0 μg/mL. The antitumor mechanism of 195 might be based on the inhibition toward thymic acid synthase, one of the key enzymes of nucleic acids [120].

2.3. Antibacterial or Antifungal Activities

2.3.1. Marine-Derived Angucyclines/Angucyclinones

Fujianmycin C (196) was isolated from the fermentation broth of the marine actinomycetes Streptomyces sp. B6219, which was isolated from the sediment of the Galapagos mangrove (Figure 9). Fujianmycin C (196) showed weak antibacterial activity against S. viridochromogenes Tṻ57, with an inhibition zone of 14 mm at 40 μg/tablet [121]. Saccharothrix espanaensis AN113 was isolated from the marine mollusk Anadara broughtoni and yielded three antibiotics, saccharothrixins A–C (197199), which displayed moderate activities against B. subtilis, E. faecium, and Xanthomonas sp. pv. Badrii at 100 μg/mL [122]. In the process of S. pratensis NA-ZhouS1’s culture, the addition of 100 µM nickel ion led to the production of antibacterial gypenocyclins stremycins A (200) and B (201), both of which could inhibit P. aeruginosa CMCC (B) 10104, MRSA, K. pneumonia CMCC (B) 46117, and E. coli CMCC (B) 44102 with equal MIC values of 16 µg/mL, and they showed inhibition against B. subtilis CMCC (B) 63501, with MIC values of 8–16 µg/mL. This is the first report that a new angucycline compound has been discovered through a “metal stress technique” [123]. The Nocardiopsis sp. HB-J378 is associated with a marine sponge Theonella sp. and produced nocardiopsistins A–C (202204), which displayed activities against MRSA with MICs of 3.12–12.5 μg/mL. Among them, the MIC of nocardiopsistin B (203) was comparable to chloramphenicol (positive control, 3.12 μg/mL) [124]. The brominated nocardiopsistin D (205) and sulfur-containing nocardiopsistins E (206) and F (207) were also identified from Nocardiopsis sp. HB-J378; all of them showed anti-MRSA activities, with MICs at 0.098, 3.125, and 0.195 μg/mL, respectively. The single bromination in 205 drastically enhanced the anti-MRSA activity by 128-fold, and it acquired activities against vancomycin-resistant S. aureus (VRSA), E. faecium, and B. cereus [125].
S. lusitanus OUCT16-27, isolated from deep-sea sediment (4495 m deep, the Indian Ocean), produced the antibiotics grincamycins L (82) and I (116), and both of them displayed bioactivities against E. faecium, E. faecalis, and S. aureus, with MICs of 3.12–6.25 μg/mL [126]. A type II PKS gene cluster harboring genes to encode several distinct oxidoreductases were identified from a rare marine actinomycete Saccharothrix sp. D09 by genome mining. The study of the gene cluster led to the isolation of the angucycline derivatives 208210, all of which showed bioactivities toward Helicobacter pylori with MIC values ranging from 16 to 32 μg/mL [127]. The same research group’s study of sediment-derived Streptomyces sp. BHB-032 (Bohai Gulf, China) led to the discovery of atramycin C (211), bearing an O-6 rhamnose side chain, and a highly hydroxylated angucyclinone emycin G (212). Compounds 211 and 212 exhibited moderate activities toward S. aureus CMCC 26003, Nocardia, B. cereus CMCC 32210, and B. subtilis CMCC 63501, with MICs of 16–64 μg/mL, but not as active as the positive control (ampicillin, MIC < 1 μg/mL) [128]. Study of M. rosaria SCSIO N160 led to the discovery of pyrazolofluostatins A–C (213215), which possess a benzo[cd]indeno[2,1-f]indazol skeleton with a pyrazole-fused 6/5/6/6/5 pentacyclic ring system. Compounds 213215 showed weak bioactivities against pathogens including E. coli ATCC 25922, S. aureus ATCC 29213, B. thuringensis SCSIO BT01, B. subtilis SCSIO BS01, and C. albicans ATCC 10231. Pyrazolofluostatin A (213) also exhibited moderate antioxidant activity (EC50 = 48.6 μM) [129]. Further, the expression of the fluostatin structural genes of M. rosaria SCSIO N160 in a heterologous host, S. coelicolor YF11, led to the isolation of an unusual heterodimer difluostatin A (216). Compound 216 exhibited antibacterial activities against K. pneumoniae ATCC 13883, Aeromonas hydrophila ATCC 7966, and S. aureus ATCC 29213, with respective MICs of 4, 4, and 8 µg/mL, while the MIC values of the positive control trimethoprim (TMP) were 0.25, 0.25, and 4 µg/mL, respectively [130].

2.3.2. Terrestrial-Derived Angucyclines/Angucyclinones

Sakyomicins A–C (140142), which were isolated from the fermentation extract of soil-derived actinomycete Nocardia. sp. M-53, displayed selective inhibitory activities against several Gram-positive bacteria (Bacillus, Staphylococcus, Micrococcus, Cotrnebacterium, Mycobacteriu), with MIC values ranging from 0.78 to 12.5 μg/mL [131]. The research on soil-derived Streptomyces sp. DSM 4769 (Adamata, India) led to the discovery of the antibiotics SM 196 A (217) and B (218), both of which could inhibit S. aureus H 503 (MIC = 100 and 25 μg/mL, respectively), S. pyogenes (MIC = 12.5–25 and 6.25 μg/mL, respectively) [132] (Figure 10). Streptomyces sp. WK-6326, which was isolated from a soil sample collected in Utah, USA, produced deacetylravidomycin M (219) and deacetylravidomycin (220). Compound 219 could inhibit the growth of B. subtillis and M. luteus (MIC = 25 μM/mL), and 220 displayed antimicrobial activities against the Gram-positive bacteria B. subtillis, S. aureus, M. luteus, and M. smegmatis, with MICs ranging from 3.0 to 5.0 μM/mL. Meanwhile, 219 inhibited IL-4-induced CD23 expression in U937 cells but had no cytotoxic effect, while 220 was identified as an interleukin-4 (IL-4) signal transduction inhibitor [133]. Seitomycin (221) and tetrangulol methyl ether (68) were isolated from the fermentation extracts of two terrestrial Streptomyces spp., GW19/1251 and GW10/1118. Compounds 221 and 68 exhibited moderate antibacterial activities in the agar diffusion assay toward B. subtilis, S. viridochromogenes Tü57, S. aureus, and E. coli, with the inhibition zones of 8–29 and 17–20 mm at 5 μg/disk, respectively. Compound 221 also showed weak phytotoxicity against Chlorella vulgaris and C. sorokiniana [134].
Streptosporangium sp. Sg3, a soil-derived actinomycete from Algeria, produced angucyclinone R2 (222), with antimicrobial activity [135]. Compound 222 significantly inhibited M. luteus ATCC 9314 and B. subtilis ATCC 6633, with MICs of 0.5 and 1.0 μg/mL, and it could also moderately inhibit S. aureus CIP 7625, Listeria monocytogenes CIP 82110, and M. smegmatis ATCC 607, with MICs of 10, 40, and 50 μg/mL, respectively [136]. Waldiomycin (223) was isolated from the strain MK844-mF10 which was associated with soil collected at Shiogama, Miyagi, Japan. Waldiomycin (223) exhibited activities against S. aureus and B. subtilis, with IC50 values at 8.8 and 10.2 μM, respectively, and could also inhibit the methicillin-resistant ones, with MICs ranging from 4 to 8 μg/mL [137]. Studies on the antibacterial mechanism of waldiomycin (223) showed that it targeted WalK histidine kinases and inhibited the WalR regulon genes expression, thereby affecting both cell wall metabolism and cell division [138]. An angucycline containing O-glycosylated 6-deoxy-a-L-talose, amycomycin D (224), produced by Kitasatospora sp., displayed inhibition toward S. aureus Newman, Pichia anomala, Mucor hiemalis, and E. coli ToIC, with MICs of 9.21–14.6 μM [139]. The expression of the landomycin A structural genes LanI and LanK in a heterologous host, S. albus J1074, led to the isolation of 6,11-dihydroxytetrangulol (93), 11-hydroxyrabelomycin (225), and fridamycin G (226). Compounds 93 and 225 exhibited activities against B. subtilis DSM 1092 and M. luteus DSM 20030, while 226 could inhibit S. aureus Newman; all the MICs were 1 µg/mL [140]. Streptomyces sp. KMC004, which was associated with acid wastewater collected from coal mines (Yeongdong, Gangneung, Republic of Korea), produced angumycinones A (227) and B (228), and both compounds showed comparable inhibitory activities against M. luteus, E. hirae, and MRSA with ampicillin (MICs = 0.78–12.5 μg/mL); the MIC values ranged from 0.78 to 12.5 μg/mL [141]. Actinoallomurus sp. ID145698 was the producer of angucyclinone allocyclinones A–D (229232), which contained chlorine atom substitutions. Compounds 229232 exhibited antibacterial activities against S. aureus ATCC 6538P, S. pyogenes L49, E. faecalis L560, and E. faecium L569; the MIC values ranged from 0.25 to 4 μg/mL, and the antibacterial activity increased as the number of chlorine atoms increased [142].
A Saharan soil-derived actinobacterium PAL114 (Mzab region, southern Algeria) produced mzabimycins A (233) and B (234), which contained L-tryptophan and glucoside derivatized chromophore on account of the addition of L-tryptophan in the fermentation; both of them exhibited antibacterial activities against M. flavus ATCC 9314 and L. monocytogenes ATCC 13932, with MIC values ranging from 15 to 40 μg/mL [143]. Based on a bioassay-guided isolation, 235 was discovered from the stem bark extracts of Stereospermum fimbriatum and exhibited bioactivities against S. epidermidis ATCC 12228, MRSA, and S. aureus ATCC 25923, with MIC values of 3.13–6.25 μg/mL [144]. Actinomycetes strain RI104-LiC106, associated with lichen, generated a 1,1-dichlorocyclopropane-containing angucycline, JBIR-88 (236), which exhibited antibacterial activity against M. luteus when the paper contained 25 μg of the compound (inhibition zone, 11mm) [145]. A soil-derived Streptomyces sp. TK08046 (Shizuoka, Japan) was the producer of saprolmycins A–E (237241), which displayed inhibition against Saprolegnia parasitica, with respective MICs of 0.0039, 8, 1, 1, and 0.0078 µg/mL. Compound 241 could also inhibit S. aureus, B. subtilis, and Daphnia pulex, with MICs at 15.6, 7.8, and 4.5 µg/mL, respectively [146]. S. griseus NTK 97 was isolated from a terrestrial sample of Terra Nova Bay at Edmondson Point, Antarctica and was the producer of frigocyclinone (136), which could inhibit B. subtilis DSM 10 and S. aureus DSM 20231, with MICs of 10 and 33 μM, respectively (positive control vancomycin and erythromycin, MICs = 1 μM) [147].

2.3.3. Angucyclines/Angucyclinones from Other Sources

6-Deoxy-8-O-methylrabelomycin I (242), produced by S. tsusimaensis MI310-38F7, showed inhibitory activities against various Gram-positive bacteria (S. aureus Smith, multi-resistant S. aureus MS9610, M. luteus PCI 1001, and B. subtilis NRRLB-558), with MIC values between 12.5 and 25.0 μg/mL, but it displayed no activity against Gram-negative bacteria [148] (Figure 11). Ecological cultivation of Actinomadura sp. RB29 and mass spectral-mediated molecular network analysis led to the expression of a silent gene cluster and the discovery of maduralactomycin A (243), which exhibited moderate activities against VRE (few colonies in the inhibition zone, 13 mm) and M. vaccae (12 mm) using the broth dilution method [149]. Acidonemycins A (244) and B (245) were discovered from the acidic culture (pH 5.4) of S. indonesiensis DSM 41759 and exhibited in vitro antivirulence activities against MRSA. Both compounds could inhibit the production of PSM and the formation of biofilm but not a significant growth inhibition. Further study indicated that the PSM and biofilm inhibitory activities of 244 and 245 were due to the (+)-ochromycinone aglycone moiety [150].

2.4. Other Bioactivities

2.4.1. Marine-Derived Angucyclines/Angucyclinones

The inhibition of dopamine S-hydroxylase caused by 1 was examined according to the method of NAGATSU, and the inhibition percentage at 0.1 μg/mL was 65.2% [16]. Actinokineospora spheciospongiae EG49 was isolated from the Red Sea sponge Spheciospongia vagabunda, and the fermentation broth contained actinosporins A (246) [151], C (247), and D (248) [152]. Actinosporin A (246) exhibited selective inhibitory activity against Trypanosoma brucei brucei with an IC50 value of 15 µM. The antioxidant potential of actinosporins C (247) and D (248) was demonstrated using the FRAP assay. Meanwhile, at 1.25 µM, actinosporins C (247) and D (248) showed significant antioxidant and protective capacity against the genomic damage induced by hydrogen peroxide in the HL-60 cell line. Furthermore, co-cultivation of Actinokineospora sp. EG49 with Rhodococcus sp. UR59 and antimalarial-guide isolation led to the discovery of actinosporins E (249), H (250), and G (251), and tetragulol (252), which exhibited antimalarial activities and good binding affinity to lysyl-tRNA synthetase (PfKRS1), with IC50 values of 9–13.5 μg/mL [153]. Solid cultivation of Actinokineospora sp. led to the generation of fridamycin H (253), which exhibited growth inhibition toward T. brucei TC221; the IC50 values after 48 h and 72 h were 7.18 and 3.35 μM, respectively [154].

2.4.2. Terrestrial-Derived Angucyclines/Angucyclinones

Highly oxygenated grecocycline D (254) was obtained from the extract of soil-derived Streptomyces sp. KCB15JA014 (Jeju Island, Republic of Korea), and showed a 46.2% inhibition rate at 50 μM against the IDO (indoleamine 2,3-dioxygenase) enzyme [155]. Streptomycete Acta 1362, which was isolated from pine rhizosphere soil on Crete, was the producer of grecocycline B (255), inhibiting protein tyrosine phosphatase 1B (PTP1B), with an IC50 at 0.52±0.17 μM [156]. Highly oxygenated gephyromycin (256) was isolated from the fermentation broth of Antarctic soil-derived S. griseus and demonstrated glutaminergic agonistic properties. When 256 was incubated with neurons for 5 min at 3 mg/mL, the concentration of intracellular Ca2+ increased twofold [157].
Streptomyces sp. KCB15JA151 isolated from soil samples collected in Jeju Island, Republic of Korea, produced pseudonocardone D (257). Compound 257 could inhibit cell proliferation induced by 17β-estradiol, which suggested that 257 might be an ER-α (estrogen receptor) antagonist [158]. The research on a soil-derived Streptomyces sp. KIB-M10 led to the isolation of cangumycins B (258) and E (259), which exhibited potent immunosuppressive activities (IC50 values = 8.1 and 2.7 μM, respectively) against human T-cell proliferation at a non-cytotoxic concentration. [159]. Fermentation of soil-derived Streptomyces sp. #AM1699 (Queensland, Australia) led to the isolation of saquayamycin A1 (260) and A-7884 (34) and vineomycin C (261), which showed inhibitory activities in the inducible nitric oxide synthase (iNOS) assay with IC50 values of 101.2, 43.5, and 25.3 μM, respectively, comparable to the known standard inhibitors NG-monomethyl-L-arginine (L-NMMA, 17.0μM) and NG-nitro-L-arginine (L-NNA, 71.0 μM) [160]. Compounds 262, 263, and 147 exhibited antimalarial activities against P. falciparum K1, with IC50 values of 3.9–6.0 µM [161]. The solid-state fermentation of soil-derived Streptomyces sp. P294 led to the isolation of X-14881 E (264) [162], which could inhibit P. burgneri hepatis with an IC50 at 3 μM [36]. OM-4842 (143) showed an inhibitory effect on platelet aggregation induced by ADP (adenosine diphosphate), arachidonic acid, PAF (platelet-activating factor), or collage; the MICs were 5.0, 12.5, and 25 μg/mL, respectively, better than kerriamycins B (44) and C (45), produced by S. violaceolatus [36,90]. Compound 265 showed a significant inhibitory activity against the DNA viruses Herpes simplex I and II, with MIC values of 0.55 and 4.54 μg/mL, respectively [132].

2.4.3. Angucyclines/Angucyclinones from Other Sources

Aquayamycin 266 is a noncompetitive inhibitor of tyrosine hydroxylase and dopamine β-hydroxylase, and it can inhibit the activity of enzymes by 50% with Ki values of 0.36 μM, 0.21 μM at concentrations of 0.37 μM, 0.40 μM, respectively. The inhibition of 266 was not affected by cofactors such as ascorbic acid and fumarate, and the inhibitory mechanism was possibly due to the chelating action of 266 on protein-bound copper. However, the inhibition could be reversed by the addition of Fe2+ but not Cu2+ [163,164]. Meanwhile, 266 also showed noncompetitive inhibition against tryptophan 5-mono-oxygenase (1.0 × 10−7 μM, 40%) [165]. Saquayamycins E (267) and F (268), produced by actinomyces MK290-AF1, were reported to inhibit the FPTase (farnesyl protein transferase) from bovine brains, with IC50 values of 1.8 and 2.0 μM, respectively, and they had a noncompetitive inhibitory effect on this enzyme [166]. The heterogeneous expression of the biosynthetic gene cluster from simocyclinones in S. coelicolor YF11 M1152, and the deletion of the keto-reducing gene simC7, related to angucyclinone, led to the production of 7-oxo-simocyclinone D8 (269), while simocyclinone D8 (42) was produced by S. antibioticus Tü 6040. Both compounds displayed inhibitory activities against DNA gyrase with the IC50 values of 50 μM and 0.1–0.6 μM, respectively. The production of 269 was related to the absence of simC7, indicating that SimC7 catalyzed the conversion of 269 to 42 as an NAD(P) (nicotinamide adenine dinucleotide phosphate) H-dependent ketoreductase, and the reduction of the carbonyl group by SimC7 was essential for the compound to bind to the enzyme with high affinity [167]. The mycelium extract of Streptomyces sp. DSM 17045 contained the PPAR-γ (proliferator-activated receptor gamma) antagonists chlorocyclinones A–D (270273). When using an AlphaScreen assay, which was able to displace rosiglitazone from the PPAR-γ ligand-binding domain (LBD) in a scintillation proximity assay (SPA), 270273 antagonized rosiglitazone-induced peroxisome PPAR-γ activation with an IC50 < 0.4 µM in vitro. The compounds were also proved to be active in a cell-based reporter gene assay, antagonizing rosiglitazone-induced PPAR-γ activities with IC50 values between 0.60 and 7.0 µM [168].
High-throughput screening of microbial metabolites led to the discovery of the IDO1 inhibitors 274276, which showed inhibition of the production of kynurenine, with respective IC50 values at 0.230, 0.067, and 5.88 µM. Enzyme kinetics experiments showed that compound 274 was a reversible noncompetitive inhibitor of IDO1 [169]. Fluostarenes A (192), B (193), and PK1 (194) showed α-glucosidase inhibitory activities, with IC50 values of 0.89, 1.58, and 0.13 μM, respectively (positive control acarbose, 0.015μM) [119]. The secondary metabolites of S. matensis A-6621 contained PI-080 (277), PI-083 (96), PI-085 (278), and PI-087 (279), with inhibitory effects on platelet aggregation in rabbits. Using ADP, collagen, and arachidonic acid as aggregating agents, the IC50 values ranged from 1.56 to 30.4 μg/mL [170]. P371A1 (280) and P371A2 (281), produced by Streptomyces sp. P371, exhibited inhibitory activities against pentagastrin-stimulated acid secretion and also showed protective activities against HCl/ethanol- and indomethacin-induced gastric lesions [171]. P371A1 (280) showed an inhibition rate of 61% on pentagastrin-stimulated acid secretion, suggesting the compound served as a CCK B/gastrin receptor antagonist. When administered interperitoneally at a dose of 10 mg/kg, 280 provided 83.6% inhibition against HCl/ethanol (60% ethanol in 150 mL HCl)-induced lesions, and at a dose of 25 mg/kg, 280 provided 72.8% inhibition against indomethacin-induced lesions [172]. Glycosylated streptocyclinones A (282) and B (283) were isolated from Streptomyces sp. and displayed antioxidant properties and modulation of the inflammatory response. Streptocyclinones A (282) and B (283) were able to protect SH-SY5Y neuroblastoma cells from H2O2-induced oxidative injury by activating the nuclear factor E2-related factor (Nrf2), and they were also able to inhibit the activity of β-secretase 1 and decrease the release of reactive oxygen species in BV2 (mouse glioma cells) stimulated with Aβ (amyloid β protein) [173]. Compound 208 displayed anti-inflammatory activity by inhibiting the production of NO with an IC50 at 28 μM [128]. Further research showed that 242 could also inhibit liver-stage P. burgneri with the IC50 of 18.5 μM [162].
Table 1. Angucyclines/angucyclinones with bioactivities.
Table 1. Angucyclines/angucyclinones with bioactivities.
Cytotoxic Activities
Compound No.ProducerModel of BioactivitiesReference
1Chainia purpurogenaEHRLICH ascites[16]
2Streptomyces sp. HB202HepG2, HT-29, GXF251L, LXF529L, MAXF401NL,
MEXF462NL, PAXF1657L, RXF486L
[17]
5Streptomyces sp. QD01-2
S. gilvotanarens NRRL 11382
Mutant strain of S. cyanogenus
MCF-7, K562, P388
Sarcoma 180, P388, Ehrlich carcinoma, Meth I fibrosarcoma, MH134
LL/2, MCF-7, NCI-H460
[18]
[31]
[108,109]
6S. gilvotanarens NRRL 11382Sarcoma 180, P388[31]
79S. fradiae PTZ0025HCT-15, SW620, C6[19]
10Micromonospora rosaria SCSIO N160
Streptomyces sp. XZHG99T
Streptomyces sp. IB201691-2A
SF-268, MCF-7, NCI-H460
A549, H157, MCF-7, MDA-MB-231, HepG2
Huh7.5, SW620, A549
[20]
[61]
[62]
11M. rosaria SCSIO N160
Micromonospora sp.
SF-268, MCF-7, NCI-H460
Kuramochi, OVCAR4, MOSE, MOE
[20]
[78]
12, 1416M. echinospora SCSIO 04089SF-268, MCF-7, HepG2[21]
13M. echinospora SCSIO 04089HepG2[21]
18, 19S. pratensis KCB-132LS180[22]
(+)-21S. pratensis KCB-132NCI-H460[24]
(–)-21S. pratensis KCB-132NCI-H460, HepG2[24]
22S. pratensis KCB-132Colon 38, HeLa cells[24]
25, 26S. ardesiacus 156VN-095ACHN, HCT-15, MDA-MB-231, NCI-H23, NUGC-3, PC-3[27]
28S. ardesiacus 156VN-095
Streptomyces sp. BCC45596
ACHN, HCT-15, MDA-MB-231, NCI-H23, NUGC-3, PC-3
KB, MCF-7, NCIH187, Vero
[27]
[28]
2931Streptomyces sp. BCC45596KB, MCF-7, NCIH187, Vero[28]
33Streptomyces sp. SCSIO 11594A594, CNE2, HepG2, MCF-7[29]
34, 35S. lusitanus SCSIO LR32MDA-MB-435, MDA-MB-231, NCI-H460,
HCT-116, HepG2, MCF10A
[30]
36S. nodosus MH190-16F3
Streptomyces sp. KY40-1
P388/S, P388/ADR, L-1210, A-549, HT-29
PC3, H-460
[32,33]
[93]
37S. nodosus MH190-16F3
S. lusitanus SCSIO LR32
Streptomyces sp.
Streptomyces sp. OC1610.4
Streptomyces sp.
Streptomyces sp. KY40-1
P388/S, P388/ADR, L-1210, A-549, HT-29
Jurkat T cells
HepG-2, SMMC-7721, PLC-PRF-5
MCF-7, MDA-MB-231, BT-474, MDA-MB-231
SW480, SW620, LoVo, HT-29, QSG-7701, CRC
PC3, H-460
[32,33]
[13]
[76]
[81]
[82]
[93]
38Amycolatopsis sp. HCa1HeLa[89]
40S. nodosus MH190-16F3P388/S, P388/ADR[32]
41, 42S. antibioticus Tü 6040HMO2, MCF-7[34,35]
4346S. capoamusM1[36]
47S. griseoincarnatus
S. lusitanus SCSIO LR32
P388
B16, HepG2, SW-1990, HeLa
[38]
[75]
4850S. venezuelae ISP5230MDA-MB-435, T-47D[39]
59Streptomyces sp. AC113
Streptomyces sp. CB01913
B16, HT29
SF295, H226, M14
[44]
[63]
60, 61Streptomyces sp. AC113B16, HT29[44]
64Streptomyces sp. Acta 3034HepG2, NIH 3T3[51]
65, 66Saccharopolyspora BCC 21906KB, MCF-7, NCI-H187[52]
6971S. salbusHCT-116[53]
72S. chattanoogensis L10 (CGMCC 2644)MCF-7[54]
73S. chattanoogensis L10 (CGMCC 2644)MCF-7, HepG2[54]
75S. blastomycetica F4-20BGC823, HeLa[56]
76, 77S. bulli GJA1, Gardenia jasminoidesOV90, ES2[57]
78Dermatophilaceae Aptenodytes NJES-13THL-60, Bel-7402, A549[58]
80Micromonospora sp.Kuramochi, MOSE, MOE[78]
8284, 86Streptomyces sp. XZHG99TA549, H157, MCF7, MDA-MB-231, HepG2[61]
85Streptomyces sp. XZHG99T
Streptomyces sp. OC1610.4
Streptomyces sp.
Streptomyces sp. KY002
A549, H157, MCF7, MDA-MB-231, HepG2
MCF-7, MDA-MB-231, BT-474, MDA-MB-231
SW480, SW620, LoVo, HT-29, QSG-7701, CRC
H-460 and MCF-7
[61]
[81]
[82]
[92]
87Streptomyces sp. XZHG99T
Streptomyces sp. OC1610.4
Streptomyces sp.
Streptomyces sp. KY40-1
A549, H157, MCF7, MDA-MB-231, HepG2
MCF-7, MDA-MB-231, BT-474, MDA-MB-231
SW480, SW620, LoVo, HT-29, QSG-7701, CRC
PC3, H-460
[61]
[81]
[82]
[93]
88Streptomyces sp. IB201691-2AHuh7.5, SW620[62]
89Streptomyces sp. IB201691-2AHuh7.5, SW620, A549[62]
90Streptomyces sp. CB01913SF295, H226[63]
92Streptomyces sp. CB01913SF295, H226, M14[63]
93S. lividans TK23HL-60[64]
94Nocardia lurida9KB, 9PS[65]
95N. lurida9KB, 9PS, 9ASK[65,66]
96Streptomyces matensis A-6621KB[67]
97–99Nocardia sp. IFM 0089L1210, P388, P388/ADR[68]
100S. murayamaensisCHO[69]
102S. murayamaensis
Salinispora pacifica DPJ-0019
CHO
K562, LNCaP, HCT-116, HeLa
[69]
[118]
104, 105Unknown actinomycetesMost of the cancer cells[71]
106S. aureofaciens CCM 3239A2788, A2780/CP, MDA-MB-231, MCF-7[72]
107, 108Streptomyces sp. CNH990HCT-116[73]
109Saccharopolyspora taberi PEM-06-F23-019BMDA-MB-231, HT-29, A-549[74]
110112S. lusitanus SCSIO LR32B16, HepG2, SW-1990, HeLa[75]
113, 115S. lusitanus SCSIO LR32Jurkat T cells[13]
114S. lusitanus SCSIO LR32
S. albogriseolus subsp. No. 1894
Jurkat T cells
Jurkat T-cells, A549, HCT-116, Capan-1
[13]
[13,99]
116S. lusitanus SCSIO LR32MDA-MB-435, MDA-MB-231, NCI-H460,
HCT-116, HepG2, MCF10A
[30]
117Streptomyces sp. M268HL-60, A549, BEL-7402[77]
118Micromonospora sp.L1210, MOSE, MOE[78]
119Streptomyces sp. SS13IPC3, H1975[79]
120Streptomyces sp. HN-A124A2780[80]
121Streptomyces sp. OC1610.4MCF-7, MDA-MB-231, BT-474, MDA-MB-231[81]
122Streptomyces sp. XS-16MDA-MB-231, K562, ASPC-1, H69AR, H69[83]
123128Actinomadura sp. KD439P388[84]
129Streptomyces sp. SUD119SK-HEP1[85]
130Streptomyces sp. SUD119HCT-116, MDA-MB-231, SNU638, A549, SK-HEP1[85]
131, 132Streptomyces sp. HDN15129HL-60, K562, SH-SY5Y, BEL-7402, U87, ASPC-1, HCT-116[86]
133Streptomyces sp. CNZ-748PMP501-1, PMP457-2[87]
134, 135Streptomyces sp. CNZ-748PMP501-1, PMP457-2, ABX023-1, C09-1[87]
136, 137Streptomyces sp. M7_15SJCRH30[88]
138, 141Amycolatopsis sp. HCa1HeLa[89]
140Amycolatopsis sp. HCa1SPC-A-1, HeLa[89]
142Amycolatopsis sp. HCa1SGC-7901, HeLa[89]
143Streptomyces sp. Om-4842P388[90]
144149S. griseorubiginosus No. Q144-2VCR-resistant P388[91]
150154Streptomyces sp. KY40-1PC3, H-460[93]
155S. polyformus sp. nov. YIM 33176
Mutant strain of S. cyanogenus
37 different human tumor cells
LL/2, MCF-7, NCI-H460
[94]
[108,109]
156, 157Streptomyces sp. N05WA963SW620, YES-4, U251SP, K562, MDA-MB-231, T-98[95]
158Streptomyces sp. N05WA963SW620, YES-4, U251SP, K562, MDA-MB-231, T-98[95]
159Streptomyces sp.Acta 2930NIH-3T3, HepG2, HT-29[96]
160Streptomyces sp. PU-MM59PC3, A549[97]
161169Mutant strain of S. cyanogenusLL/2, MCF-7, NCI-H460[108,109]
164, 167, 173175, 177S. cyanogenus S-136MCF-7, MDA-MB-231[114]
172S. cyanogenus S-136MDA-MB-231[114]
181184S. cyanogenus K62MCF-7, MDA-MB-231[115]
185, 186Mutant strain of S. lividans TK24Several cancer cells[109,116]
187S. murayamaensisK562[69,117]
188191S. pacifica DPJ-0019 (NRRL 50168)K562, LNCaP, HCT-116, HeLa[118]
193Unknown actinomycetesSF-268, MCF-7, HepG2, A549[119]
195Amycolatopsis orientalis subsp. vineariaEhrlich ascites carcinoma[120]
Antibacterial or antifungal activities
1Chainia purpurogenaGram-positive bacteria except M. tuberculosis[16]
2Streptomyces sp. HB202B. subtilis DSM 347, B. epidermidis DSM 20660,
D. hominis DSM 7083, K. pneumoniae, P. aeruginosa DSM 50071,
S. aureus ATCC 12600, S. aureus, S. epidermidis DSM 20044,
S. lentus DSM 6672
[17]
3, 4Streptomyces sp. QD01-2S. aureus, B. subtilis, Escherichia coli, Candida albicans[18]
5, 6Streptomyces sp. QD01-2
S. gilvotanarens NRRL 11382
S. aureus, B. subtilis, Escherichia coli, Candida albicans
S. aureus ATCC 6538P, B. subtilis No. 10707
[18]
[31]
79S. fradiae PTZ0025S. aureus[19]
10M. rosaria SCSIO N160
 
S. cellulosae YIM PH20352
 
Streptomyces sp. XZHG99T
Streptomyces sp. IB201691-2A
E. coli ATCC 25922, S. aureus ATCC 29213,
B. thuringiensis SCSIO BT01, B. subtilis SCSIO BS01
P. cucumerina, Alternaria panax, F. oxysporum,
F. solani, M. smegmatis, S. aureus
S. carnosus DSMZ 20501, Erwinia persicina DSMZ 19328
S. carnosus DSMZ 20501, M. smegmatis DSMZ 43286
[20]
 
[59,60]
[61]
[62]
11M. rosaria SCSIO N160E. coli ATCC 25922, S. aureus ATCC 29213,
B. thuringiensis SCSIO BT01, B. subtilis SCSIO BS01
[20]
12M. echinospora SCSIO 04089S. aureus ATCC 29213, B. thuringensis SCSIO BT01,
B. subtilis 1064, M. luteus SCSIO ML01, MRSA shhs-A1
[21]
15M. echinospora SCSIO 04089M. luteus SCSIO ML01,[21]
17S. pratensis KCB-132A variety of bacteria and fungi[22]
18, 19S. pratensis KCB-132B. cereus, C. lagenarium[22]
20S. pratensis KCB-132S. aureus CMCC 26003[23]
(–)-21Streptomyces sp. KCB-132B. cereus CMCC 32210[24]
23Streptomyces sp. KCB-132S. aureus, Enterococcus faecium[25]
24S. pratensis KCB-132E. faecium, S. aureus, K. pneumoniae,
A. baumannii, P. aeruginosa, E. species
[26]
2527S. ardesiacus 156VN-095B. subtilis KCTC 1021, M.s luteus KCTC 1915,
S. aureus KCTC 1927
[27]
2830Streptomyces sp. BCC45596M. tuberculosis, P. falciparum[28]
31Streptomyces sp. BCC45596
S. cellulosae YIM PH20352
M. tuberculosis, P. falciparum
A. panax
[28]
[59,60]
32, 33Streptomyces sp. SCSIO 11594E. faecalis ATCC29212[29]
34S. lusitanus SCSIO LR32M. luteus[30]
3639Streptomyces nodosus MH190-16F3S. aureus FDA209P, S. aureus, M. 1ysodeikticus IFO 3333,
M. luteus PCI1001, B. subtilis PCI 219
[32]
41, 42S. antibioticus Tü 6040B. brevis DSM30[34,35]
4346S. violaceolatusS. aureus FDA 209P, B. subtilis ATCC 6633,
B. cereus IAM 1729, M. luteus ATCC 9341
[36]
46S. capoamusP. chrysogentrrn ATCC 10002, T. mentagrophytes[37]
47S. griseoincarnatusS. aureus FDA 209P, M. luteus ATCC 9341, B. cereus IAM 1729[38]
4856S. venezuelae ISP5230S. aureus C622 (ATCC25923), S. aureus 305,
S. aureus BeckerCP8 (ATCC49525),
S. aureus BeckerLyc12CP336 (ATCC55804),
S. epidermidis C960 (ATCC14990),
S. epidermidis C621 (clinical isolate), B. subtilis C971 (ATCC6633), S. aureus C623(MRSA)
[39]
57, 58S. venezuelae ISP5230MRSA, S. warneri, VRE[43]
59Streptomyces sp. AC113
 
Streptomyces sp. CB01913
P. aeruginosa CCM 3955, S. aureus CCM 3953, E. coli CCM 3988,
L. monocytogenes NCTC 4886, B. subtilis CCM 2216, B. cereus
S. aureus ATCC 25923, B. subtilis ATCC 23857,
M. smegmatis ATCC 607
[44]
 
[63]
6061Streptomyces sp. AC113P. aeruginosa CCM 3955, S. aureus CCM 3953, E. coli CCM 3988,
L. monocytogenes NCTC 4886, B. subtilis CCM 2216, B. cereus
[44]
6264Streptomyces sp. Acta 3034B. subtilis[51]
6567Saccharopolyspora BCC 21906M. tuberculosis[52]
68Saccharopolyspora BCC 21906
Streptomyces spp. GW19/1251 and GW10/1118
M. tuberculosis
B. subtilis, S. viridochromogenes Tü57, S. aureus, E. coli
[52]
[134]
6971S. salbusMRSA, B. subillis RM125[53]
73S. chattanoogensis L10 (CGMCC 2644)B. subtilis ATCC 67736[54]
75S. blastomycetica F4-20Valsa mali, C. orbiculare, F. graminearumat[56]
76S. bulliGJA1, Gardenia jasminoidesMRSA[57]
78, 79D. Aptenodytes NJES-13TS. aureus, B. subtilis, C. albicans[58]
80S. cellulosae YIM PH20352P. cucumerina, A. panax, F. oxysporum, F. solani with[59,60]
81S. cellulosae YIM PH20352P. cucumerina, A. panax[59,60]
82S. lusitanus OUCT16-27E. faecium, E. faecalis, S. aureus[126]
88, 89Streptomyces sp. IB201691-2AS. carnosus DSMZ 20501, E. persicina DSMZ 19328,
M. smegmatis DSMZ 43286
[62]
90, 92Streptomyces sp. CB01913S. aureus ATCC 25923, B. subtilis ATCC 23857,
M. smegmatis ATCC 607
[63]
93S. lividans TK23
S. albus J1074
S. aureus, C. albicans
B. subtilis DSM 1092, M. luteus DSM 20030
[64]
[140]
94, 95N. luridaGram-positive bacteria[65]
96S. matensis A-6621S. aureus 209P-JC, Sepidermidis IID 866, E. faecium ATCC8043,
B. cereus S 1101, B. subtilis ATCC6633
[67]
9799Nocardia sp. IFM 0089S. aureus 209P, S. aureus MRSAIFM 62971,
M. smegmatis ATCC607, M. luteus IFM 2066
[68]
100103S. murayamaensisGram-positive bacteria[70]
104Unknown actinomycetesMRSA, VRE[71]
106S. aureofaciens CCM 3239B. subtilis, S. aureus[72]
116S. lusitanus OUCT16-27E. faecium, E. faecalis, S. aureus[126]
136S. griseus NTK 97B. subtilis DSM 10, S. aureus DSM 20231[147]
140142Nocardia. sp. M-53Bacillus, Staphylococcus, Micrococcus, Cotrnebacterium, Mycobacteriu[131]
196Streptomyces sp. B6219S. viridochromogenes Tü57[121]
197199S. espanaensis AN113B. subtilis, E. faecium, Xanthomonas sp. pv. Badrii[122]
200, 201S. pratensis NA-ZhouS1’sP. aeruginosa CMCC (B) 10104, MRSA,
K. pneumonia CMCC (B) 46117, E. coli CMCC (B) 44102,
B. subtilis CMCC (B) 63501
[123]
202204Nocardiopsis sp. HB-J378MRSA[124]
205Nocardiopsis sp. HB-J378MRSA, VRE, B. cereus[125]
206, 207Nocardiopsis sp. HB-J378MRSA[124]
208210Saccharothrix sp. D09H. pylori[127]
211, 212Streptomyces sp. BHB-032S. aureus CMCC 26003, Nocardia,
B. cereus CMCC 32210, B. subtilis CMCC 63501
[128]
213215M. rosaria SCSIO N160E. coli ATCC 25922, S. aureus ATCC 29213,
B. thuringensis SCSIOBT01, B. subtilis SCSIO BS01,
C. albicans ATCC 10231.
[129]
216M. rosaria SCSIO N160 in a heterologous host S. coelicolor YF11K. pneumoniae ATCC 13883, A. hydrophila ATCC 7966,
S. aureus ATCC 29213
[130]
217, 218Streptomyces sp. DSM 4769S. aureus H 503, S. pyogenes[132]
219, 220Streptomyces sp. WK-6326B. subtillis, M. luteus, B. subtillis, S. aureus, M. luteus, M. smegmatis[133]
221Streptomyces spp. GW19/1251 and GW10/1118B. subtilis, S. viridochromogenes Tü57, S. aureus,
E. coli, Chlorella vulgaris, C. sorokiniana
[134]
222Streptosporangium sp. Sg3M. luteus ATCC 9314, B. subtilis ATCC 6633, S. aureus CIP 7625,
L. monocytogenes CIP 82110, M. smegmatis ATCC 607
[135,136]
223Streptomyces sp. MK844-mF10S. aureus, B. subtilis[137]
224Kitasatospora sp.S. aureus Newman, P. anomala, M. hiemalis, E. coli ToIC[139]
225S. albus J1074B. subtilis DSM 1092, M. luteus DSM 20030[140]
226S. albus J1074S. aureus Newman[140]
227, 228Streptomyces sp. KMC004M. luteus, E. hirae, MRSA[141]
229232Actinoallomurus sp. ID145698S. aureus ATCC 6538P, S. pyogenes L49,
E. faecalis L560, E. faecium L569
[142]
233, 234Actinobacterium PAL114M. flavus ATCC 9314, L. monocytogenes ATCC 13932[143]
235S. fimbriatumS. epidermidis ATCC 12228, MRSA, S. aureus ATCC 25923[144]
236Actinomycetes RI104-LiC106M. luteus[145]
237240Streptomyces sp. TK08046S. parasitica[146]
241Streptomyces sp. TK08046S. parasitica, S. aureus, B. subtilis, D. pulexwith[146]
242S. tsusimaensis MI310-38F7S. aureus Smith, S. aureus MS9610 (multi-resistant),
M. luteus PCI 1001, B. subtilis NRRLB-558
[148]
243Actinomadura sp. RB29VRE, M. vaccae[149]
244, 245S. indonesiensis DSM41759MRSA[150]
Enzyme inhibitory activities
1C. purpurogenaDopamine S-hydroxylase inhibition[16]
42S. coelicolor YF11 M1152
S. antibioticus Tü 6040
DNA gyrase inhibition[167]
192194Unknown actinomycetesα-glucosidase inhibition[119]
254Streptomyces sp. KCB15JA014IDO1 inhibition[155]
255Streptomycete Acta 1362PTP1B inhibition[156]
266Unknown actinomycetesTyrosine hydroxylase inhibition,
dopamine β-hydroxylase inhibition,
tryptophan 5-mono-oxygenase inhibition
[163,164,165]
267, 268Actinomyces MK290-AF1,FPTase inhibition[166]
269S. coelicolor YF11 M1152
S. antibioticus Tü 6040
DNA gyrase inhibition[167]
270, 273Streptomyces sp. DSM 17045Antagonize rosiglitazone-induced
peroxisome PPAR-γ activation
[168]
274276High-throughput screening of microbialIDO1 inhibition[169]
Other activities
34Streptomyces sp. #AM1699Nitric oxide inhibition[160]
74Mutant strain of S. chattanoogensis L10 (CGMCC 2644)Antioxidant activity[55]
96S. matensis A-6621Platelet aggregation inhibition[170]
143Streptomyces sp. P294Platelet aggregation inhibition[36]
208Unknown actinomycetesNitric oxide inhibition[128]
213M. rosaria SCSIO N160Antioxidant activity[129]
219, 220Streptomyces sp. WK-6326IL-4 inhibition[133]
242Unknown actinomycetesP. burgneri inhibition[162]
246A. heciospongiae EG49T. brucei brucei inhibition[151]
247, 248A. heciospongiae EG49Antioxidant activity[152]
249252Actinokineospora sp. EG49
with Rhodococcus sp. UR59
Antimalarial activity[153]
253Actinokineospora sp.T. brucei TC221 inhibition[154]
256S. griseusGlutaminergic agonist[157]
257Streptomyces sp. KCB15JA151Cell proliferation inhibition[158]
258, 259Streptomyces sp. KIB-M10Human T-cell proliferation inhibition[159]
260, 261Streptomyces sp. #AM1699Nitric oxide inhibition[160]
262, 263Unknown actinomycetesP. falciparum K1 inhibition[161]
264Streptomyces sp. P294P. burgneri inhibition[36]
265Streptomyces sp. DSM 4769DNA viruses Herpes simplex I and II[132]
277279S. matensis A-6621Platelet aggregation inhibition[170]
280, 281Streptomyces sp. P371Inhibitory activity against pentagastrin-stimulated acid secretion, protective activity against HCl/ethanol- and indomethacin-induced gastric lesions[171]
280Streptomyces sp. P371CCK B/gastrin receptor antagonist[172]
282, 283Streptomyces sp.SH-SY5Y neuroblastoma cells protection[173]

3. Biosynthesis of Landomycins

Landomycins are one of the best-known subgroups of the angucyclines/angucyclinones, with glycosylation at C8 and hydroxylation at C11 positions [10]. Microbial fermentation, heterologous expression of biosynthetic gene clusters (BGCs), and modification of culture conditions have resulted in the identification of over 30 landomycins. A total of 20 active landomycins (161167, 170175, and 176184) are involved in this review. Landomycins were first isolated from the fermentation broth of S. cyanogenus S136 in 1990 [103], including the hexasaccharidal landomycin A (164), the pentasaccharidal landomycin B (165), and the disaccharidal landomycin D (166). Landomycin A BGC (lan) from S. cyanogenus S136 [174], landomycin E BGC (lnd) from S. globisporus 1912 [175], and lnd-like BGC [176], identified in the metagenomic DNA, are related to the landomycin biosynthetic pathway. The homologous genes that relate to the biosynthesis of the basic skeleton of the landomycins are almost the same in the three pathways, while there are some differences in details. For example, the putative hexose synthase genes lanZ2, glycosyltransferase lanGT3, and TetR-like repressor gene lanK in the landomycin A BGCs are absent in the landomycin E BGCs, while prx, lndW-W2, and lndY-lndYR, related to the regulation and export of compounds, are missing in the landomycin A BGCs.
Isotope labeling experiments confirmed the origins of the carbon and oxygen atoms in the landomycin angucyclic scaffold [104]. The carbons, together with two oxygen atoms at the C-1 and C-8 positions, were proved to originate from acetate, while the oxygen atoms at C-7 and C-12 were found to originate from molecular oxygen. The functions of genes in the BGCs were elucidated by analyzing the products of mutants in which specific genes were inactivated [177]. Products encoded by lanA to lanD are typically found in the biosynthesis of aromatic polyketide, indicating that the genes are used for the synthesis of ketoacyl, the extension of the carbon chain, the transport of acyl groups, as well as the reduction of the ketone group, respectively. The biosynthesis of landomycin involved a decaketide intermediate (Figure 12), which needs to undergo further cyclization, aromatization, oxidation, and reduction to establish the backbone of landomycinone. LanL (homolog of lndL) was suggested to relate to the first cyclization–aromatization during the biosynthesis of landomycins, while lanF (homolog of lndF) was proposed to catalyze the formation of the third and the fourth ring. The products of lnd/lanF and lnd/lanL are also homologous to other cyclases found in type II polyketide gene clusters [177].
Genetic analysis and in vitro assays showed that the tetracyclic framework needs to undergo a cascade of hydrolysis and decarboxylation to form UWM6 after the cyclization. Oxygenases and reductases encoded by lan/lndZ4, Z5, E, M2, and V were proposed for the conversion of UWM6 [178]. Dehydration catalyzed by lan/lndM2 leads to the production of prejadomycin (2,3-dehydroUWM6). C-12 oxygenase encoded by lan/lndE, the first oxygenase in the pathway, could transform prejadomycin to tetrangomycin (92) [179]. The 6-keto group was reduced by lan/lndV to form 11-deoxylandomycinone, and lanV also contributed to the aromatization of the landomycin angucyclic scaffold [179]. The inactivation of lndM2 in S. globisporus 1912 and the feeding of the intermediate showed that there were at least two paths for tetrangomycin (92) to convert to 11-deoxylandomycinone (181) [180]. LndM2 catalyzes the hydroxylation at the C-6 position of 92 to yield rabelomycin (10), then catalyzes the reduction of the C5–C6 double bond and 2,3-dehydration of 10 to yield 181. Alternatively, lndM2 reduces the C5–C6 double bond, which converts 92 to 5,6-dihydrotetrangomycin, and then the oxygenation at C-6 and aromatization also generates 181. LanZ4 and LanZ5 are related to the hydroxylation of the C-11 position. Because of broad substrate specificity, hydroxylation can occur at different stages of glycosylation and independent of the length of the sugar chain [178]. In the presence of multiple substrates, what is the preferred substrate still needs further investigation.
Hexasaccharidal landomycin A is composed of a repetitive sequence of NDP-D-olivose and NDP-L-rhodinose, and four glycosyltransferase genes (lanGT1, lanGT2, lanGT3, and lanGT4) are responsible for six glycosyl transfer steps. The functions were elucidated by expressing individual genes, isolating the products and intermediates, and feeding the intermediates back to the knockout mutant strains [174]. LanGT2 catalyzes the glycosylation step at the C8 position with NDP-D-olivose to produce landomycin H [181]; when flavin reductase LanZ4 and bifunctional oxygenase-dehydratase LanZ5 are functioning, C-11 oxidation can convert landomycin H into landomycin I (162). The remarkable feature of LanGT1 and LanGT4 is that they are used twice during the hexasaccharidal landomycin biosynthesis [102]. LanGT1 encodes a D-olivosyltransferase, which is responsible for the attachment of the D-olivose moiety as the second and fifth sugars, converting landomycin I (162) to landomycins D (166) and B (165). LanGT4 is responsible for the attachment of the L-rhodinose moiety (the third and sixth sugars) to produce landomycins E (161) and A (164). LanGT1 and lanGT4 display a relaxed substrate specificity toward the sugar acceptor substrates. LanGT3 is related to the attachment of the fourth sugar (D-olivose) moiety to yield landomycin J (163) [182]. Based on the functions of LanGTs, their unbalancing may lead to the accumulation of specific products. LanGTs can also convert landomycin H directly to landomycins F (178), G, V (174), and S (171) without the oxidation of the C11 position. Dehydration of 11-deoxylandomycinone (181) yields tetrangulol (176), and oxidation catalyzed by lanZ4/lanZ5 converts tetrangulol (176) to 5, 6-anhydrolandomycinone (177). Compound 176 could be catalyzed by LanGTs to form landomycins O (180), P, M (179), and T (172), while the glycation products of 5, 6-anhydrolandomycinone were landomycins R (170), Q, U (173), and W (175), respectively (Figure 12).
There are also landomycins [103] wherein the order of L-rhodinose and D-olivose in the sugar chain is different from the landomycins in Figure 12, or the D-olivose is replaced with D-amicetose [115]. Some biosynthetic genes are involved in the export of landomycins. As substances with antibacterial and cytotoxic activity, landomycins are also toxic to their producer. The detoxification mechanism may be related to the effluence of the compounds. LanJ encodes a proton-dependent antiporter protein in S. cyanogenus S136. Overexpression of lanJ confers resistance to landomycin A (164) and increased accumulation of landomycin with shorter glycoside chains. The TetR-type repressor gene lanK shares the common promoter lanKJp with lanJ and negatively regulates the expression of lanJ [183]. Binding of landomycins to lanK relieves its inhibition and thus triggers the biosynthesis and export of landomycins. lndW has been identified at the end of the landomycin BGCs, and it encodes an ATP-binding subunit of the ABC transporters protein, which is related to the resistance against landomycin [184].

4. Discussion

This review covers 283 angucyclines/angucyclinones discovered from 1965 to 2023 with various bioactivities. Marine and terrestrial microorganisms have made nearly equal contributions to the production of these bioactive compounds, affording 100 (35%) and 113 (40%) of them, respectively (Figure 13a). The bioactivities of the compounds are related to their sources. Sixty-five percent of the marine-derived angucyclines/angucyclinones have been reported to show only one type of bioactivity (cell toxicity, antibacterial activity, enzyme activity, anti-inflammation, etc.). Only 35% of the marine-derived compounds show both cytotoxicity and antimicrobial activity, while this figure is 51% for terrestrial-derived angucyclines/angucyclinones (Figure 13b). Streptomyces undoubtedly is the most important producer of bioactive angucyclines/angucyclinones (73%), and the gene cluster related to angucyclines/angucyclinones’ biosynthesis in Streptomyces has been described in detail in previous reviews [10]. In the future, more angucyclines/angucyclinones might be obtained through gene knockout or the activation of silent genes. The genera Nocardia and Saccharoployspora follow closely behind, producing 4% of the total included compounds (Figure 13c). Taking 10 years as a statistical unit, the largest number of these molecules was found in the 2010s (Figure 13d), which might relate to the development of separation and structural identification technologies in the 21st century as well as to the increasing number of bioassays. Therefore, it is reasonable to predict that more bioactive novel angucyclines/angucyclinones will be discovered, or new bioactivities might be found for known compounds, in the next ten years.
In addition to landomycins, which are characterized by glycosylation at the C8 position, angucyclines/angucyclinones like urdamycins, saquayamycins, and sangkocyclines also can be glycosylated. The number, type, and order of sugars as well as the position of glycoside chains are variable. Among the 283 active angucyclines/angucyclinones collected in this review, 127 compounds contain glycoside chains in their chemical structures, accounting for 45% of the total number. Furthermore, 42 compounds contain 2 glycoside chains, accounting for 15% of the total angucyclines/angucyclinones and 33% of the glycosylated ones (Figure 14a). Disaccharidal members account for the biggest number among all the glycosylated angucyclines/angucyclinones, followed by monosaccharidal and trisaccharidal ones. Disaccharidal angucyclines/angucyclinones tend to display cytotoxic activities, while monosaccharidal ones tend to show antibacterial activities, suggesting that the number of sugars in the glycoside chains may be related to bioactivities (Figure 14b). For example, landomycins with glycosylation at the C8 position display cytotoxicity only. Some researchers believe that most landomycins with bioactivities have long oligosaccharide chains [114,115], while others believe that the cytotoxicity does not increase simultaneously with the elongation of the chain [108,109]. Further research is thus needed to investigate the effects between the oligosaccharide chains and bioactivities of landomycins and the relationship between oligosaccharide chains and glycosylated angucyclines/angucyclinones.
International Whole Genome Sequencing efforts and comparative bioinformatics studies have revealed that the biosynthetic potential of existing microorganisms has not been fully exploited [185,186], and translating genetic blueprints into single compounds remains a significant scientific challenge. Angucyclines/angucyclinones were first discovered in the 1960s and have been a hot topic in the field of drug discovery due to their diverse chemical structures and biological activities. With the rapid development and the application of microbial isolation techniques such as in situ culture [187], analytical techniques like molecular network analysis [188], culturing techniques such as simulated ecological cultivation [149], heterologous gene expression [64], metal stress induction [123], as well as isolation techniques like activity-guided isolation [189] and high-throughput screening [190], the current strategies for discovering microbial natural products have become more specific and efficient compared with traditional methods. This has further enriched the structural diversity of the angucyclines/angucyclinones. Additionally, Global Natural Products Social Molecular Networking (GNPS) has been used widely to analyze the culture extracts of microorganisms and elucidate the important intermediates in the biosynthesis of metabolites [149]. At the same time, the halogenation or substitution with other heteroatoms could modify and enhance the activities of the compounds, and the synthetic derivatization could expand the diversity of molecular structures. All the studies are important to the discovery of more angucyclines/angucyclinones with novel structures and activities.
The biosynthesis toward skeleton and glycosylation of landomycins has been well established, while the functions of genes related to regulation and export of landomycins (such as prx, lndI, lanK, lanJ, lanW, etc.) still need further investigation. At the same time, the differences in biosynthetic pathways between landomycin and other angucyclines/angucyclinones (such as urdamycins, saquayamycins, langkocyclines) are also valuable for further study.

5. Conclusions

Endophytic microorganisms, particularly endophytic actinomycetes, have long been regarded as an important source of leading compounds of drugs with broad biological activities [191]. One class of secondary metabolites from microorganisms, angucyclines/angucyclinones, have always been a focus of drug development research due to their unique molecular structures and favorable biological activities. The difference in living environments between marine and terrestrial actinomycetes has led to the isolation of many new microorganisms and the discovery of novel compounds, while research on the isolated microorganisms is still ongoing.
The methods mentioned in this paper for discovering new compounds are also applicable to the discovery of other types of microbial natural products, providing valuable references for finding new natural molecules with diverse structural features. This review summary on the structure and source of active angucyclines/angucyclinones can also provide researchers with new research directions and inspirations for transformation.

Author Contributions

Writing—original draft preparation, H.-S.L., H.-R.C.; writing—structure preparation of compounds, H.-R.C., S.-S.H., Z.-H.L., C.-Y.W.; writing—review and editing, H.-S.L., H.Z.; supervision, H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Shandong Province, China (ZR2021QD109) and the Project of Shandong Province Higher Educational Youth Innovation Science and Technology Program (2022KJ096).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs over the Nearly Four Decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef]
  2. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2017, 34, 235–294. [Google Scholar] [CrossRef] [PubMed]
  3. Jakubiec-Krzesniak, K.; Rajnisz-Mateusiak, A.; Guspiel, A.; Ziemska, J.; Solecka, J. Secondary Metabolites of Actinomycetes and their Antibacterial, Antifungal and Antiviral Properties. Pol. J. Microbiol. 2018, 67, 259–272. [Google Scholar] [CrossRef] [PubMed]
  4. Procópio, R.E.; Silva, I.R.; Martins, M.K.; Azevedo, J.L.; Araújo, J.M. Antibiotics produced by Streptomyces. Braz. J. Infect. Dis. 2012, 16, 466–471. [Google Scholar] [CrossRef] [PubMed]
  5. Sivalingam, P.; Hong, K.; Pote, J.; Prabakar, K. Extreme Environment Streptomyces: Potential Sources for New Antibacterial and Anticancer Drug Leads. Int. J. Microbiol. 2019, 2019, 5283948. [Google Scholar] [CrossRef]
  6. Hu, J.; Wang, Z.X.; Li, P.M.; Qian, P.Y.; Liu, L.L. Structural identification of pyridinopyrone compounds with anti-neuroinflammatory activity from Streptomyces sulphureus DSM 40104. Front. Microbiol. 2023, 14, 1205118. [Google Scholar] [CrossRef] [PubMed]
  7. Castro, J.F.; Razmilic, V.; Gomez-Escribano, J.P.; Andrews, B.; Asenjo, J.; Bibb, M. The “gifted” actinomycete Streptomyces leeuwenhoekii. Antonie. Van. Leeuwenhoek 2018, 111, 1433–1448. [Google Scholar] [CrossRef] [PubMed]
  8. Vysloužilová, D.; Kováč, O. The Chemistry of Angucyclines. ChemPlusChem 2024, 89, e202400307. [Google Scholar] [CrossRef]
  9. Dann, M.; Lefemine, D.V.; Barbatschi, F.; Shu, P.; Kunstmann, M.P.; Mitscher, L.A.; Bohonos, N. Tetrangomycin, a new quinone antibiotic. Antimicrob. Agents Chemother 1965, 5, 832–835. [Google Scholar] [PubMed]
  10. Kharel, M.K.; Pahari, P.; Shepherd, M.D.; Tibrewal, N.; Nybo, S.E.; Shaaban, K.A.; Rohr, J. Angucyclines: Biosynthesis, mode-of-action, new natural products, and synthesis. Nat. Prod. Rep. 2012, 29, 264–325. [Google Scholar] [CrossRef]
  11. Xu, X.; Chang, Y.; Chen, Y.; Zhou, L.; Zhang, F.; Ma, C.; Che, Q.; Zhu, T.; Pfeifer, B.A.; Zhang, G.; et al. Biosynthesis of Atypical Angucyclines Unveils New Ring Rearrangement Reactions Catalyzed by Flavoprotein Monooxygenases. Org. Lett. 2024, 26, 7489–7494. [Google Scholar] [CrossRef] [PubMed]
  12. Rohr, J.; Thiericke, R. Angucycline group antibiotics. Nat. Prod. Rep. 1992, 9, 103–137. [Google Scholar] [CrossRef]
  13. Zhu, X.; Duan, Y.; Cui, Z.; Wang, Z.; Li, Z.; Zhang, Y.; Ju, J.; Huang, H. Cytotoxic rearranged angucycline glycosides from deep sea-derived Streptomyces lusitanus SCSIO LR32. J. Antibiot. 2017, 70, 819–822. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, J.; Duan, Y.; Zhu, X.; Yan, X. Angucycline/angucyclinone natural products research (2010–2020). Chin. J. Biotech. 2021, 37, 2147–2165. [Google Scholar] [CrossRef]
  15. Page, M.J.; Moher, D.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. PRISMA 2020 explanation and elaboration: Updated guidance and exemplars for reporting systematic reviews. The BMJ. 2021, 372, n160. [Google Scholar] [CrossRef] [PubMed]
  16. Okazaki, T.; Kitahara, T.; Okami, Y. Studies on marine microorganisms. IV. A new antibiotic SS-228 Y produced by Chainia isolated from shallow sea mud. J. Antibiot. 1975, 28, 176–184. [Google Scholar] [CrossRef] [PubMed]
  17. Schneemann, I.; Kajahn, I.; Ohlendorf, B.; Zinecker, H.; Erhard, A.; Nagel, K.; Wiese, J.; Imhoff, J.F. Mayamycin, a Cytotoxic Polyketide from a Streptomyces Strain Isolated from the Marine Sponge Halichondria panicea. J. Nat. Prod. 2010, 73, 1309–1312. [Google Scholar] [CrossRef]
  18. Hou, J.; Liu, P.; Qu, H.; Fu, P.; Wang, Y.; Wang, Z.; Li, Y.; Teng, X.; Zhu, W. Gilvocarcin HE: A new polyketide glycoside from Streptomyces sp. J. Antibiot. 2012, 65, 523–526. [Google Scholar] [CrossRef]
  19. Xin, W.; Ye, X.; Yu, S.; Lian, X.; Zhang, Z. New capoamycin-type antibiotics and polyene acids from marine Streptomyces fradiae PTZ0025. Mar. Drugs 2012, 10, 2388–2402. [Google Scholar] [CrossRef] [PubMed]
  20. Zhang, W.; Liu, Z.; Li, S.; Lu, Y.; Chen, Y.; Zhang, H.; Zhang, G.; Zhu, Y.; Zhang, G.; Zhang, W. Fluostatins I–K from the South China Sea-derived Micromonospora rosaria SCSIO N160. J. Nat. Prod. 2012, 75, 1937–1943. [Google Scholar] [CrossRef]
  21. Fang, Z.; Jiang, X.; Zhang, Q.; Zhang, L.; Zhang, W.; Yang, C.; Zhang, H.; Zhu, Y.; Zhang, C. S-Bridged Thioether and Structure-Diversified Angucyclinone Derivatives from the South China Sea-Derived Micromonospora echinospora SCSIO 04089. J. Nat. Prod. 2020, 83, 3122–3130. [Google Scholar] [CrossRef]
  22. Zhang, S.; Zhang, L.; Fu, X.; Li, Z.; Guo, L.; Kou, L.; Liu, M.; Xie, Z. (+)- and (–)-actinoxocine, and actinaphthorans A–B, C-ring expansion and cleavage angucyclinones from a marine-derived Streptomyces sp. Org. Chem. Front. 2019, 6, 3925–3928. [Google Scholar] [CrossRef]
  23. Zhang, S.; LuKou, L.; Qiao-LiQu, B.; GennaroXie, Z. Isolation, stereochemical study, and racemization of (+/–)-pratenone A, the first naturally occurring 3-(1-naphthyl)-2-benzofuran-1(3H)-one polyketide from a marine-derived actinobacterium. Chirality 2020, 32, 299–307. [Google Scholar] [CrossRef]
  24. Guo, L.; Zhang, L.; Yang, Q.; Xu, B.; Fu, X.; Liu, M.; Li, Z.; Zhang, S.; Xie, Z. Antibacterial and cytotoxic bridged and ring cleavage angucyclinones from a marine Streptomyces sp. Front. Chem. 2020, 8, 586. [Google Scholar] [CrossRef] [PubMed]
  25. Guo, L.; Yang, Q.; Wang, G.; Zhang, S.; Liu, M.; Pan, X.; Pescitelli, G.; Xie, Z. Ring D-modified and highly reduced angucyclinones from marine sediment-derived Streptomyces sp. Front. Chem. 2021, 9, 756962. [Google Scholar] [CrossRef] [PubMed]
  26. Fu, X.; Zhang, S.; Wang, G.; Yang, Q.; Guo, L.; Pescitelli, G.; Xie, Z. Atypical Angucyclinones with Ring Expansion and Cleavage from a Marine Streptomyces sp. J. Org. Chem. 2022, 87, 15998–16010. [Google Scholar] [CrossRef]
  27. Anh, C.V.; Kwon, J.; Kang, J.S.; Lee, H.; Heo, C.; Shin, H.J. New Angucycline Glycosides from a Marine-Derived Bacterium Streptomyces ardesiacus. Int. J. Mol. Sci. 2022, 23, 13779. [Google Scholar] [CrossRef] [PubMed]
  28. Supong, K.; Thawai, C.; Suwanborirux, K.; Choowong, W.; Supothina, S.; Pittayakhajonwut, P. Antimalarial and antitubercular C-glycosylated benz[α]anthraquinones from the marine-derived Streptomyces sp. BCC45596. Phytochem. Lett. 2012, 5, 651–656. [Google Scholar] [CrossRef]
  29. Song, Y.; Liu, G.; Li, J.; Huang, H.; Zhang, X.; Zhang, H.; Ju, J. Cytotoxic and antibacterial angucycline- and prodigiosin- analogues from the deep-sea derived Streptomyces sp. SCSIO 11594. Mar. Drugs 2015, 13, 1304–1316. [Google Scholar] [CrossRef]
  30. Lai, Z.; Yu, J.; Ling, H.; Song, Y.; Yuan, J.; Ju, J.; Tao, Y.; Huang, H. Grincamycins I–K, Cytotoxic Angucycline Glycosides Derived from Marine-Derived Actinomycete Streptomyces lusitanus SCSIO LR32. Planta Med. 2018, 84, 201–207. [Google Scholar] [CrossRef]
  31. Morimoto, M.; Okubo, S.; Tomita, F.; Marumo, H. Gilvocarcins, new antitumor antibiotics. 3. Antitumor activity. J. Antibiot. 1981, 34, 701–707. [Google Scholar] [CrossRef]
  32. Uchida, T.; Imoto, M.; Watanabe, Y.; Miura, K.; Dobashi, T.; Matsuda, N.; Sawa, T.; Naganawa, H.; Hamada, M.; Takeuchi, T.; et al. Saquayamycins, new aquayamycin-group antibiotics. J. Antibiot. 1985, 38, 1171–1181. [Google Scholar] [CrossRef]
  33. Henkel, T.; Zeeck, A. Derivatives of saquayamycins A and B. Regio- and diastereoselective addition of alcohols to the L-aculose moiety. J. Antibiot. 1990, 43, 830–837. [Google Scholar] [CrossRef]
  34. Schimana, J.; Walker, M.; Zeeck, A.; Fiedler, H. Simocyclinones: Diversity of metabolites is dependent on fermentation conditions. J. Ind. Microbiol. Biotechnol. 2001, 27, 144–148. [Google Scholar] [CrossRef]
  35. Schimana, J.; Fiedler, H.P.; Groth, I.; Suessmuth, R.; Beil, W.; Walker, M.; Zeeck, A. Simocyclinones, novel cytostatic angucyclinone antibiotics produced by Streptomyces antibioticus Tü 6040. I. Taxonomy, fermentation, isolation and biological activities. J. Antibiot. 2010, 32, 301–307. [Google Scholar] [CrossRef]
  36. Hayakawa, Y.; Iwakiri, T.; Imamura, K.; Seto, H.; Otake, N. Studies on the isotetracenone antibiotics. II. Kerriamycins A, B and C, new antitumor antibiotics. J. Antibiot. 1985, 38, 960–963. [Google Scholar] [CrossRef]
  37. Hayakawa, Y.; Iwakiri, T.; Imamura, K.; Seto, H.; Otake, N. Studies on the isotetracenone antibiotics. I. Capoamycin, a new antitumor antibiotic. J. Antibiot. 1985, 38, 957–959. [Google Scholar] [CrossRef]
  38. Hayakawa, Y.; Iwakiri, T.; Imamura, K.; Seto, H.; Otake, N. Studies on the isotetracenone antibiotics. III. A new isotetracenone antibiotic, grincamycin. J. Antibiot. 1987, 40, 1785–1787. [Google Scholar] [CrossRef]
  39. Jakeman, D.L.; Bandi, S.; Graham, C.L.; Reid, T.R.; Wentzell, J.R.; Douglas, S.E. Antimicrobial activities of jadomycin B and structurally related analogues. Antimicrob. Agents Ch. 2009, 3, 1245–1247. [Google Scholar] [CrossRef]
  40. Bonitto, E.P.; McKeown, B.T.; Goralski, K.B. Jadomycins: A potential chemotherapy for multi-drug-resistant metastatic breast cancer. Pharmacol. Res. Perspect. 2021, 9, e00886. [Google Scholar] [CrossRef]
  41. Issa, M.E.; Hall, S.R.; Dupuis, S.N.; Graham, C.L.; Jakeman, D.L.; Goralski, K.B. Jadomycins are cytotoxic to ABCB1-, ABCC1-, and ABCG2-overexpressing MCF7 breast cancer cells. Anticancer Drugs. 2014, 25, 255–269. [Google Scholar] [CrossRef] [PubMed]
  42. Martinez-Farina, C.F.; McCormick, N.; Robertson, A.W.; Clement, H.; Jee, A.; Ampaw, A.; Chan, N.L.; Syvitski, R.T.; Jakeman, D.L. Investigations into the binding of jadomycin DS to human topoisomerase IIβ by WaterLOGSY NMR spectroscopy. Org. Biomol. Chem. 2015, 13, 10324–10327. [Google Scholar] [CrossRef]
  43. Forget, S.M.; Robertson, A.W.; Overy, D.P.; Kerr, R.G.; Jakeman, D.L. Furan and Lactam Jadomycin Biosynthetic Congeners Isolated from Streptomyces venezuelae ISP5230 Cultured with Nε-Trifluoroacetyl-L-lysine. J. Nat. Prod. 2017, 80, 1860–1866. [Google Scholar] [CrossRef] [PubMed]
  44. Maruna, M.; Sturdikova, M.; Liptaj, T.; Godany, A.; Muckova, M.; Certik, M.; Pronayova, N.; Proksa, B. Isolation, structure, elucidation and biological activity of angucycline antibiotics from an epiphytic yew Streptomycete. J. Basic Microbiol. 2010, 50, 135–142. [Google Scholar] [CrossRef]
  45. Macleod, J.M.; Forget, S.M.; Jakeman, D.L. The Expansive Library of Jadomycins. Can. J. Chem. 2018, 96, 495–501. [Google Scholar] [CrossRef]
  46. de Koning, C.B.; Ngwira, K.J.; Rousseau, A.L. Biosynthesis, synthetic studies, and biological activities of the jadomycin alkaloids and related analogues. Alkaloids Chem. Biol. 2020, 84, 125–199. [Google Scholar] [CrossRef] [PubMed]
  47. Hall, S.R.; Blundon, H.L.; Ladda, M.A.; Robertson, A.W.; Martinez-Farina, C.F.; Jakeman, D.L.; Goralski, K.B. Jadomycin breast cancer cytotoxicity is mediated by a copper-dependent, reactive oxygen species-inducing mechanism. Pharmacol. Res. Perspect. 2015, 3, e00110. [Google Scholar] [CrossRef]
  48. Hall, S.R.; Toulany, J.; Bennett, L.G.; Martinez-Farina, C.F.; Robertson, A.W.; Jakeman, D.L.; Goralski, K.B. Jadomycins Inhibit Type II Topoisomerases and Promote DNA Damage and Apoptosis in Multidrug-Resistant Triple-Negative Breast Cancer Cells. J. Pharmacol. Exp. Ther. 2017, 363, 196–210. [Google Scholar] [CrossRef] [PubMed]
  49. McKeown, B.T.; Relja, N.J.; Hall, S.R.; Gebremeskel, S.; MacLeod, J.M.; Veinotte, C.J.; Bennett, L.G.; Ohlund, L.B.; Sleno, L.; Jakeman, D.L.; et al. Pilot study of jadomycin B pharmacokinetics and anti-tumoral effects in zebrafish larvae and mouse breast cancer xenograft models. Can. J. Physiol. Pharmacol. 2022, 100, 1065–1076. [Google Scholar] [CrossRef] [PubMed]
  50. Iwasaki, E.; Shimizu, Y.; Akagi, Y.; Komatsu, T. Synthesis and in Vitro Cytotoxicity Evaluation of Jadomycins. Chem. Pharm. Bull. 2023, 71, 730–733. [Google Scholar] [CrossRef] [PubMed]
  51. Kalyon, B.; Tan, G.A.; Pinto, J.M.; Foo, C.; Wiese, J.; Imhoff, J.F.; Suessmuth, R.D.; Sabaratnam, V.; Fiedler, H. Langkocyclines: Novel angucycline antibiotics from Streptomyces sp. Acta 3034. J. Antibiot. 2013, 66, 609–616. [Google Scholar] [CrossRef] [PubMed]
  52. Boonlarppradab, C.; Suriyachadkun, C.; Rachtawee, P.; Choowong, W. Saccharosporones A, B and C, cytotoxic antimalarial angucyclinones from Saccharopolyspora sp. BCC 21906. J. Antibiot. 2013, 66, 305–309. [Google Scholar] [CrossRef]
  53. Kang, H.; Brady, S.F. Mining soil metagenomes to better understand the evolution of natural product structural diversity: Pentangular polyphenols as a case study. J. Am. Chem. Soc. 2014, 136, 18111–18119. [Google Scholar] [CrossRef]
  54. Zhou, Z.; Xu, Q.; Bu, Q.; Guo, Y.; Liu, S.; Liu, Y.; Du, Y.; Li, Y. Genome Mining-Directed Activation of a Silent Angucycline Biosynthetic Gene Cluster in Streptomyces chattanoogensis. ChemBioChem 2015, 16, 496–502. [Google Scholar] [CrossRef]
  55. Li, Z.; Bu, Q.; Wang, J.; Liu, Y.; Chen, X.; Mao, X.; Li, Y. Activation of anthrachamycin biosynthesis in Streptomyces chattanoogensis L10 by site-directed mutagenesis of rpoB. J. Zhejiang Univ. Sci. B 2019, 20, 983–994. [Google Scholar] [CrossRef]
  56. Yan, H.; Li, Y.; Zhang, X.Y.; Zhou, W.Y.; Feng, T.J. A new cytotoxic and anti-fungal C-glycosylated benz[α]anthraquinone from the broth of endophytic Streptomyces blastomycetica strain F4-20. J. Antibiot. 2017, 70, 301–303. [Google Scholar] [CrossRef]
  57. Kim, J.W.; Kwon, Y.; Bang, S.; Kwon, H.E.; Park, S.; Lee, Y.H.; Deyrup, S.; Song, G.; Lee, D.; Joo, H.S. Unusual bridged angucyclinones and potent anticancer compounds from Streptomyces bulli GJA1. Org. Biomol. Chem. 2020, 18, 8443–8449. [Google Scholar] [CrossRef] [PubMed]
  58. Zhu, W.Z.; Wang, S.H.; Gao, H.M.; Ge, Y.M.; Dai, J.; Zhang, X.L.; Yang, Q. Characterization of Bioactivities and Biosynthesis of Angucycline/Angucyclinone Derivatives Derived from Gephyromycinifex aptenodytis gen. nov., sp. nov. Mar. Drugs 2021, 20, 34. [Google Scholar] [CrossRef]
  59. Xu, X.; Zhao, Y.; Bao, K.; Miao, C.; Zhao, L.; Chen, Y.; Wu, S.; Li, Y. Purification and characterization of anti-phytopathogenic fungi angucyclinone from soil-derived Streptomyces cellulosae. Folia Microbiol. 2022, 67, 517–522. [Google Scholar] [CrossRef]
  60. Xu, X.D.; Zhao, Y.; Bao, K.; Miao, C.P.; Tang, S.K.; Wu, S.H.; Li, Y.Q. Isolation, Structure Elucidation and Antifungal Activity of Angucycline Antibiotics from Streptomycete cellulosae. Appl. Biochem. Microbiol. 2023, 59, 456–461. [Google Scholar] [CrossRef]
  61. Bao, J.; He, F.; Li, Y.; Fang, L.; Wang, K.; Song, J.; Zhou, J.; Li, Q.; Zhang, H. Cytotoxic antibiotic angucyclines and actinomycins from the Streptomyces sp. XZHG99T. J. Antibiot. 2018, 71, 1018–1024. [Google Scholar] [CrossRef]
  62. Voitsekhovskaia, I.; Paulus, C.; Dahlem, C.; Rebets, Y.; Nadmid, S.; Zapp, J.; Axenov-Gribanov, D.; Ruckert, C.; Timofeyev, M.; Kalinowski, J.; et al. Baikalomycins A–C, new aquayamycin-type angucyclines isolated from lake baikal derived Streptomyces sp. IB201691-2A. Microorganisms 2020, 8, 680. [Google Scholar] [CrossRef]
  63. Ma, M.; Rateb, M.E.; Teng, Q.; Yang, D.; Shen, B. Angucyclines and Angucyclinones from Streptomyces sp. CB01913 Featuring C-Ring Cleavage and Expansion. J. Nat. Prod. 2015, 78, 2471–2480. [Google Scholar] [CrossRef]
  64. Sakai, K.; Takao, R.; Koshino, H.; Futamura, Y.; Osada, H.; Takahashi, S. 6,9-Dihydroxytetrangulol, a novel angucyclinone antibiotic accumulated in kiqO gene disruptant in the biosynthesis of kinanthraquinone. J. Antibiot. 2021, 74, 593–595. [Google Scholar] [CrossRef] [PubMed]
  65. Theriault, R.J.; Rasmussen, R.R.; Kohl, W.L.; Prokop, J.F.; Hutch, T.B.; Barlow, G.J. Benzanthrins A and B, a new class of quinone antibiotics. I. Discovery, fermentation and antibacterial activity. J. Antibiot. 1986, 39, 1509–1514. [Google Scholar] [CrossRef]
  66. Rasmussen, R.R.; Nuss, M.E.; Scherr, M.H.; Mueller, S.L.; Mcalpine, J.B.; Mitscher, L.A. benzanthrins a and b, a new class of quinone antibiotics ii. isolation, elucidation of structure and potential antitumor activity. J. Antibiot. 1986, 39, 1515–1526. [Google Scholar] [CrossRef]
  67. Kawashima, A.; Yoshimura, Y.; Goto, J.; Nakaike, S.; Mizutani, T.; Hanada, K.; Omura, S. PI-083, a new platelet aggregation inhibitor. J. Antibiot. 1988, 41, 1913–1914. [Google Scholar] [CrossRef]
  68. Nemoto, A.; Tanaka, Y.; Karasaki, Y.; Komaki, H.; Yazawa, K.; Mikami, Y.; Tojo, T.; Kadowaki, K.; Tsuda, M.; Kadowaki, K. Brasiliquinones A, B and C, New benz[α]anthraquinone Antibiotics from Nocardia brasiliensis. J. Antibiot. 1997, 50, 18–21. [Google Scholar] [CrossRef]
  69. Hasinoff, B.B.; Wu, X.; Yalowich, J.C.; Goodfellow, V.; Laufer, R.S.; Adedayo, O.; Dmitrienko, G.I. Kinamycins A and C, bacterial metabolites that contain an unusual diazo group, as potential new anticancer agents: Antiproliferative and cell cycle effects. Anti-Cancer Drugs 2006, 17, 825–837. [Google Scholar] [CrossRef]
  70. Omura, S.; Nakagawa, A.; Yamada, H.; Hata, T.; Furusaki, A.; Watanabe, T. Structure of Kinamycin C, and the Structural Relationship among Kinamycin A, B, C, and D. Chem. Pharm. Bull. 1971, 19, 2428–2430. [Google Scholar] [CrossRef]
  71. Wada, S.-I.; Sawa, R.; Iwanami, F.; Nagayoshi, M.; Kubota, Y.; Iijima, K.; Hayashi, C.; Shibuya, Y.; Hatano, M.; Igarashi, M.; et al. Structures and biological activities of novel 4′-acetylated analogs of chrysomycins A and B. J. Antibiot. 2017, 70, 1078–1082. [Google Scholar] [CrossRef]
  72. Matulova, M.; Feckova, L.; Novakova, R.; Mingyar, E.; Kormanec, J. A Structural Analysis of the Angucycline-Like Antibiotic Auricin from Streptomyces lavendulae Subsp. Lavendulae CCM 3239 Revealed Its High Similarity to Griseusins. J. Antibiot. 2019, 8, 102. [Google Scholar] [CrossRef]
  73. Martin, G.D.A.; Tan, L.T.; Jensen, P.R.; Dimayuga, R.E.; Fairchild, C.R.; Raventos-Suarez, C.; Fenical, W. Marmycins A and B, cytotoxic pentacyclic C-glycosides from a marine sediment-derived actinomycete related to the genus Streptomyces. J. Nat. Prod. 2007, 70, 1406–1409. [Google Scholar] [CrossRef] [PubMed]
  74. Pérez, M.; Schleissner, C.; Rodríguez, P.; Zúñiga, P.; Benedit, G.; Sánchez-Sancho, F.; de la Calle, F. PM070747, a new cytotoxic angucyclinone from the marine-derived Saccharopolyspora taberi PEM-06-F23-019B. J. Antibiot. 2009, 62, 167–169. [Google Scholar] [CrossRef] [PubMed]
  75. Huang, H.; Yang, T.; Ren, X.; Liu, J.; Song, Y.; Sun, A.; Ma, J.; Wang, B.; Zhang, Y.; Huang, C.; et al. Cytotoxic Angucycline Class Glycosides from the Deep Sea Actinomycete Streptomyces lusitanus SCSIO LR32. J. Nat. Prod. 2012, 75, 202–208. [Google Scholar] [CrossRef]
  76. Peng, A.; Qu, X.; Liu, F.; Li, X.; Li, E.; Xie, W. Angucycline glycosides from an intertidal sediments strain Streptomyces sp. and their cytotoxic activity against hepatoma carcinoma cells. Mar. Drugs 2018, 16, 470. [Google Scholar] [CrossRef] [PubMed]
  77. Xie, Z.; Liu, B.; Wang, H.; Yang, S.; Zhang, H.; Wang, Y.; Ji, N.; Qin, S.; Laatsch, H. Kiamycin, a unique cytotoxic angucyclinone derivative from a marine Streptomyces sp. Mar. Drugs 2012, 10, 551–558. [Google Scholar] [CrossRef] [PubMed]
  78. Mullowney, M.W.; hAinmhire, E.O.; Tanouye, U.; Burdette, J.E.; Pham, V.C.; Murphy, B.T. A pimarane diterpene and cytotoxic angucyclines from a marine-derived Micromonospora sp. in Vietnam’s East Sea. Mar. Drugs 2015, 13, 5815–5827. [Google Scholar] [CrossRef]
  79. Jiang, Y.; Gan, L.; Ding, W.; Chen, Z.; Ma, Z. Cytotoxic gephyromycins from the Streptomyces sp. SS13I. Tetrahedron Lett. 2017, 58, 3747–3750. [Google Scholar] [CrossRef]
  80. Zhou, B.; Ji, Y.; Zhang, H.; Shen, L. Gephyyamycin and cysrabelomycin, two new angucyclinone derivatives from the Streptomyces sp. HN-A124. Nat. Prod. Res. 2021, 35, 2117–2122. [Google Scholar] [CrossRef] [PubMed]
  81. Qu, X.; Ren, J.; Peng, A.; Lin, S.; Lu, D.; Du, Q.; Liu, L.; Li, X.; Li, E.; Xie, W. Cytotoxic, anti-migration, and anti-invasion activities on breast cancer cells of angucycline glycosides isolated from a marine-derived Streptomyces sp. Mar. Drugs 2019, 17, 277. [Google Scholar] [CrossRef]
  82. Li, J.; Han, N.; Zhang, H.; Xie, X.; Zhu, Y.; Zhang, E.; Ma, J.; Shang, C.; Yin, M.; Xie, W.; et al. Saquayamycin B1 Suppresses Proliferation, Invasion, and Migration by Inhibiting PI3K/AKT Signaling Pathway in Human Colorectal Cancer Cells. Mar. Drugs 2022, 20, 570. [Google Scholar] [CrossRef] [PubMed]
  83. Xu, X.; Zhang, F.; Zhou, L.; Chang, Y.; Che, Q.; Zhu, T.; Li, D.; Zhang, G. Overexpression of Global Regulator SCrp Leads to the Discovery of New Angucyclines in Streptomyces sp. XS-16. Mar. Drugs 2023, 21, 240. [Google Scholar] [CrossRef]
  84. Zhang, Z.; In, Y.; Fukaya, K.; Yang, T.; Harunari, E.; Urabe, D.; Imada, C.; Oku, N.; Igarashi, Y. Kumemicinones A–G, Cytotoxic Angucyclinones from a Deep Sea-Derived Actinomycete of the Genus. J. Nat. Prod. 2022, 85, 1098–1108. [Google Scholar] [CrossRef] [PubMed]
  85. Bae, M.; An, J.S.; Hong, S.H.; Bae, E.S.; Chung, B.; Kwon, Y.; Hong, S.; Oh, K.B.; Shin, J.; Lee, S.K. Donghaecyclinones A–C: New Cytotoxic Rearranged Angucyclinones from a Volcanic Island-Derived Marine Streptomyces sp. Mar. Drugs 2020, 18, 121. [Google Scholar] [CrossRef] [PubMed]
  86. Chang, Y.; Xing, L.; Sun, C.; Liang, S.; Liu, T.; Zhang, X.; Zhu, T.; Pfeifer, B.A.; Che, Q.; Zhang, G.; et al. Monacycliones G–K and ent-Gephyromycin A, Angucycline Derivatives from the Marine-Derived Streptomyces sp. HDN15129. J. Nat. Prod. 2020, 83, 2749–2755. [Google Scholar] [CrossRef]
  87. Shang, Z.; Ferris, Z.E.; Sweeney, D.; Chase, A.B.; Li, J. Grincamycins P–T: Rearranged Angucyclines from the Marine Sediment-Derived Streptomyces sp. CNZ-748 Inhibit Cell Lines of the Rare Cancer Pseudomyxoma Peritonei. J. Nat. Prod. 2021, 84, 1638–1648. [Google Scholar] [CrossRef] [PubMed]
  88. Vicente, J.; Stewart, A.K.; van Wagoner, R.M.; Elliott, E.; Bourdelais, A.J.; Wrigh, J.L.C. Monacyclinones, new angucyclinone metabolites isolated from Streptomyces sp. M7_15 associated with the Puerto Rican Sponge Scopalina ruetzleri. Mar. Drugs 2015, 13, 4682–4700. [Google Scholar] [CrossRef]
  89. Guo, Z.K.; Wang, T.; Guo, Y.; Song, Y.C.; Tan, R.X.; Ge, H.M. Cytotoxic angucyclines from Amycolatopsis sp. HCa1, a rare actinobacteria derived from Oxya chinensis. Planta Med. 2011, 77, 2057–2060. [Google Scholar] [CrossRef]
  90. Omura, S.; Nakagawa, A.; Fukamachi, N.; Miura, S.; Takahashi, Y.; Komiyama, K.; Kobayashi, B. OM-4842, a new platelet aggregation inhibitor from Streptomyces. J. Antibiot. 1988, 41, 812–813. [Google Scholar] [CrossRef]
  91. Oka, M.; Kamei, H.; Hamagishi, Y.; Tomita, K.; Miyaki, T.; Konishi, M.; Oki, T. Chemical and biological properties of rubiginone, a complex of new antibiotics with vincristine-cytotoxicity potentiating activity. J. Antibiot. 1990, 43, 967–976. [Google Scholar] [CrossRef]
  92. Abdelfattah, M.S.; Kharel, M.K.; Hitron, J.A.; Baig, I.; Rohr, J. Moromycins A and B, Isolation and Structure Elucidation of C-Glycosylangucycline-Type Antibiotics from Streptomyces sp. KY002. J. Nat. Prod. 2008, 71, 1569–1573. [Google Scholar] [CrossRef]
  93. Shaaban, K.A.; Ahmed, T.A.; Leggas, M.; Rohr, J. Saquayamycins G–K, Cytotoxic Angucyclines from Streptomyces sp. Including Two Analogues Bearing the Aminosugar Rednose. J. Nat. Prod. 2012, 75, 1383–1392. [Google Scholar] [CrossRef] [PubMed]
  94. Li, Y.; Huang, X.; Ishida, K.; Maier, A.; Kelter, G.; Jiang, Y.; Peschel, G.; Menzel, K.; Li, M.; Wen, M.; et al. Plasticity in gilvocarcin-type C-glycoside pathways: Discovery and antitumoral evaluation of polycarcin V from Streptomyces polyformus. Org. Biomol. Chem. 2008, 6, 3601–3605. [Google Scholar] [CrossRef] [PubMed]
  95. Ren, X.; Lu, X.; Ke, A.; Zheng, Z.; Lin, J.; Hao, W.; Zhu, J.; Fan, Y.; Ding, Y.; Jiang, Q.; et al. Three novel members of angucycline group from Streptomyces sp. N05WA963. J. Antibiot. 2011, 64, 339–343. [Google Scholar] [CrossRef]
  96. Helaly, S.E.; Goodfellow, M.; Zinecker, H.; Imhoff, J.F.; Suessmuth, R.D.; Fiedler, H. Warkmycin, a novel angucycline antibiotic produced by Streptomyces sp. Acta 2930. J. Antibiot. 2013, 66, 669–674. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, Y.; Cheema, M.T.; Ponomareva, L.V.; Ye, Q.; Shaaban, K.A. Himalaquinones A–G, Angucyclinone-Derived Metabolites Produced by the Himalayan Isolate Streptomyces sp. PU-MM59. J. Nat. Prod. 2021, 84, 1930–1940. [Google Scholar] [CrossRef] [PubMed]
  98. Okazaki, H.; Ohta, K.; Kanamaru, T.; Ishimaru, T.; Kishi, T. A potent prolyl hydroxylase inhibitor, P-1894B, produced by a strain of Streptomyces. J. Antibiot. 1981, 34, 1355–1356. [Google Scholar] [CrossRef] [PubMed]
  99. Matsumoto, Y.; Kuriki, H.; Kitamura, T.; Takahashi, D.; Toshima, K. Total Synthesis and Structure-Activity Relationship Study of Vineomycin A1. J. Org. Chem. 2019, 84, 14724–14732. [Google Scholar] [CrossRef]
  100. Matseliukh, B.P.; Lavrinchuk, V.I. The isolation and characteristics of mutant Streptomyces globisporus 1912 defective for landomycin E biosynthesis. Mikrobiolohichnyi Zhurnal 1999, 61, 22. [Google Scholar] [PubMed]
  101. Korynevska, A.; Heffeter, P.; Matselyukh, B.; Elbling, L.; Micksche, M.; Stoika, R.; Berger, W. Mechanisms underlying the anticancer activities of the angucycline landomycin E. Biochem. Pharmacol. 2007, 74, 1713–1726. [Google Scholar] [CrossRef] [PubMed]
  102. Ostash, B.; Rix, U.; Rix, L.L.R.; Liu, T.; Lombo, F.; Luzhetskyy, A.; Gromyko, O.; Wang, C.; Braña, A.F.; Méndez, C. Generation of new landomycins by combinatorial biosynthetic manipulation of the LndGT4 gene of the landomycin E cluster in S. globisporus. Chem. Biol. 2004, 11, 547–555. [Google Scholar] [CrossRef]
  103. Henkel, T.; Jürgen, R.; Beale, J.M.; Schwenen, L. Landomycins, new angucycline antibiotics from Streptomyces sp. I. Structural studies on landomycins A–D. J. Antibiot. 1990, 43, 492. [Google Scholar] [CrossRef] [PubMed]
  104. Weber, S.; Zolke, C.; Rohr, J.; Beale, J.M. Investigations of the biosynthesis and structural revision of landomycin A. J. Org. Chem. 1994, 59, 4211–4214. [Google Scholar] [CrossRef]
  105. Ostash, B.; Korynevska, A.; Stoika, R.; Fedorenko, V. Chemistry and Biology of Landomycins, an Expanding Family of Polyketide Natural Products. Mini-Rev. Med. Chem. 1040. [Google Scholar] [CrossRef]
  106. Mulert, V.U.; Luzhetskyy, A.; Hofmann, C.; Mayer, A.; Bechthold, A. Expression of the landomycin biosynthetic gene cluster in a PKS mutant of Streptomyces fradiae is dependent on the coexpression of a putative transcriptional activator gene. FEMS Microbiol. Lett. 2004, 230, 91–97. [Google Scholar] [CrossRef]
  107. Zhu, L.; Luzhetskyy, A.; Luzhetska, M.; Mattingly, C.; Adams, V.; Bechthold, A.; Rohr, J. Generation of New Landomycins with Altered Saccharide Patterns through Over-expression of the Glycosyltransferase Gene lanGT3 in the Biosynthetic Gene Cluster of Landomycin A in Streptomyces cyanogenus S-136. ChemBioChem 2007, 8, 83–88. [Google Scholar] [CrossRef]
  108. Luzhetskyy, A.; Liu, T.; Fedoryshyn, M.; Ostash, B.; Bechthold, A. Function of lanGT3, a glycosyltransferase gene involved in landomycin a biosynthesis. ChemBioChem 2004, 5, 1567–1570. [Google Scholar] [CrossRef] [PubMed]
  109. Shepherd, M.D.; Liu, T.; Mendez, C.; Salas, J.A.; Rohr, J. Engineered biosynthesis of gilvocarcin analogues with altered deoxyhexopyranose moieties. Appl. Environ. Microbiol. 2011, 77, 435–441. [Google Scholar] [CrossRef]
  110. Luzhetskyy, A.; Zhu, L.; Gibson, M.; Fedoryshyn, M.; Bechthold, A. Generation of Novel Landomycins M and O through Targeted Gene Disruption. ChemBioChem 2010, 6, 675–678. [Google Scholar] [CrossRef] [PubMed]
  111. Krohn, K.; Böker, N.; Flörke, U.; Freund, C. Synthesis of Angucyclines. 8. Biomimetic-Type Synthesis of Rabelomycin, Tetrangomycin, and Related Ring B Aromatic Angucyclinones. J. Org. Chem. 1997, 62, 2350–2356. [Google Scholar] [CrossRef]
  112. Kuntsmann, M.P.; Mitscher, L.A. The Structural Characterization of Tetrangomycin and Tetrangulol. J. Org. Chem. 1966, 31, 2920–2925. [Google Scholar] [CrossRef]
  113. Krohn, K.; Khanbabaee, D.C.K. First Total Synthesis of (±)-Rabelomycin. Angew. Chem. Int. Ed. 1994, 33, 99–100. [Google Scholar] [CrossRef]
  114. Shaaban, K.A.; Srinivasan, S.; Kumar, R.; Damodaran, C.; Rohr, J. Landomycins P–W, cytotoxic angucyclines from Streptomyces cyanogenus S-136. J. Nat. Prod. 2011, 74, 2–11. [Google Scholar] [CrossRef]
  115. Shaaban, K.A.; Stamatkin, C.; Damodaran, C.; Rohr, J. 11-Deoxylandomycinone and landomycins X–Z, new cytotoxic angucyclin(on)es from a Streptomyces cyanogenus K62 mutant strain. J. Antibiot. 2011, 64, 141–150. [Google Scholar] [CrossRef]
  116. Liu, T.; Kharel, M.K.; Zhu, L.; Bright, S.A.; Mattingly, C.; Adams, V.R.; Rohr, J. Inactivation of the Ketoreductase gilU Gene of the Gilvocarcin Biosynthetic Gene Cluster Yields New Analogues with Partly Improved Biological Activity. ChemBioChem 2009, 10, 278–286. [Google Scholar] [CrossRef]
  117. O’Hara, K.A.; Wu, X.; Patel, D.; Liang, H.; Yalowich, J.C.; Chen, N.; Goodfellow, V.; Adedayo, O.; Dmitrienko, G.I.; Hasinoff, B.B. Mechanism of the cytotoxicity of the diazoparaquinone antitumor antibiotic kinamycin F. Free Radical Biol. Med. 2007, 43, 1132–1144. [Google Scholar] [CrossRef]
  118. Woo, C.M.; Beizer, N.E.; Janso, J.E.; Herzon, S.B. Isolation of Lomaiviticins C–E, Transformation of Lomaiviticin C to Lomaiviticin A, Complete Structure Elucidation of Lomaiviticin A, and Structure-Activity Analyses. J. Am. Chem. Soc. 2012, 134, 15285–15288. [Google Scholar] [CrossRef]
  119. Yang, C.; Huang, C.; Fang, C.; Zhang, L.; Chen, S.; Zhang, Q.; Zhang, C.; Zhang, W. Inactivation of Flavoenzyme-Encoding Gene flsO1 in Fluostatin Biosynthesis Leads to Diversified Angucyclinone Derivatives. J. Org. Chem. 2021, 86, 11019–11028. [Google Scholar] [CrossRef] [PubMed]
  120. Sasaki, E.; Liu, H.W. Mechanistic studies of the biosynthesis of 2-thiosugar: Evidence for the formation of an enzyme-bound 2-ketohexose intermediate in BexX-catalyzed reaction. J. Am. Chem. Soc. 2010, 132, 15544–15546. [Google Scholar] [CrossRef]
  121. Abdalla, M.A.; Helmke, E.; Laatsch, H. Fujianmycin C, A bioactive angucyclinone from a marine derived Streptomyces sp. B6219. Nat. Prod. Commun. 2010, 5, 1917–1920. [Google Scholar] [CrossRef] [PubMed]
  122. Kalinovskaya, N.I.; Kalinovsky, A.I.; Romanenko, L.A.; Pushilin, M.A.; Dmitrenok, P.S.; Kuznetsova, T.A. New angucyclinones from the marine mollusk-associated actinomycete Saccharothrix espanaensis An 113. Nat. Prod. Commun. 2008, 3, 1611–1616. [Google Scholar] [CrossRef]
  123. Akhter, N.; Liu, Y.; Auckloo, B.N.; Shi, Y.; Wang, K.; Chen, J.; Wu, X.; Wu, B. Stress-driven discovery of new angucycline-type antibiotics from a marine Streptomyces pratensis NA-ZhouS1. Mar. Drugs 2018, 16, 331/1–331/16. [Google Scholar] [CrossRef]
  124. Xu, D.; Nepal, K.K.; Harmody, D.; McCarthy, P.J.; Wright, A.E.; Wang, G.; Chen, J.; Zhu, H. Nocardiopsistins A–C: New angucyclines with anti-MRSA activity isolated from a marine sponge-derived Nocardiopsis sp. HB-J378. Synth. Syst. Biotechnol. 2018, 3, 246–251. [Google Scholar] [CrossRef]
  125. Xu, D.; Metz, J.; Harmody, D.; Peterson, T.; Winder, P.; Guzman, E.A.; Russo, R.; McCarthy, P.J.; Wright, A.E.; Wang, G. Brominated and Sulfur-Containing Angucyclines Derived from a Single Pathway: Identification of Nocardiopsistins D–F. Org. Lett. 2022, 24, 7900–7904. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, L.; Hou, L.; Li, H.; Li, W. Antibiotic angucycline derivatives from the deepsea-derived Streptomyces lusitanus. Nat. Prod. Res. 2020, 34, 3444–3450. [Google Scholar] [CrossRef] [PubMed]
  127. Shen, Q.; Dai, G.; Li, A.; Liu, Y.; Zhong, G.; Li, X.; Ren, X.; Sui, H.; Fu, J.; Jiao, N.; et al. Genome-Guided Discovery of Highly Oxygenated Aromatic Polyketides, Saccharothrixins D–M, from the Rare Marine Actinomycete Saccharothrix sp. D09. J. Nat. Prod. 2021, 84, 2875–2884. [Google Scholar] [CrossRef] [PubMed]
  128. Liu, M.; Yang, Y.; Gong, G.; Li, Z.; Zhang, L.; Guo, L.; Xu, B.; Zhang, S.; Xie, Z. Angucycline and angucyclinone derivatives from the marine-derived Streptomyces sp. Chirality 2022, 34, 421–427. [Google Scholar] [CrossRef] [PubMed]
  129. Zhang, W.; Yang, C.; Huang, C.; Zhang, L.; Zhang, H.; Zhang, Q.; Yuan, C.; Zhu, Y.; Zhang, C.; Zhang, W.; et al. Pyrazolofluostatins A–C, Pyrazole-Fused Benzoafluorenes from South China Sea-Derived Micromonospora rosaria SCSIO N160. Org. Lett. 2017, 19, 592–595. [Google Scholar] [CrossRef]
  130. Yang, C.; Huang, C.; Zhang, W.; Zhu, Y.; Zhang, C. Heterologous expression of fluostatin gene cluster leads to a bioactive heterodimer. Org. Lett. 2015, 17, 5324–5327. [Google Scholar] [CrossRef]
  131. Nagasawa, T.; Fukao, H.; Irie, H.; Yamada, H. Sakyomicins A, B, C and D: New quinone-type antibiotics produced by a strain of Nocardia. Taxonomy, production, isolation and biological properties. J. Antibiot. 1984, 37, 693–699. [Google Scholar] [CrossRef]
  132. Grabley, S.; Hammann, P.; Hütter, K.; Kluge, H.; Thiericke, R.; Wink, J.; Zeeck, A. Secondary metabolites by chemical screening. Part 19. SM 196 A and B, novel biologically active angucyclinones from Streptomyces sp. J. Antibiot. 1991, 44, 670. [Google Scholar] [CrossRef] [PubMed]
  133. Arai, M.; Tomoda, H.; Tabata, N.; Ishiouro, N.; Kobayashi, S.; Omura, S. Deacetylravidomycin M, a new inhibitor of IL-4 signal transduction, produced by Streptomyces sp. WK-6326. II. Structure elucidation. J. Antibiot. 2001, 54, 554–561. [Google Scholar] [CrossRef] [PubMed]
  134. Abdelfattah, M.; Maskey, R.P.; Asolkar, R.N.; Gruen-Wollny, I.; Laatsch, H. Seitomycin: Isolation, structure elucidation and biological activity of a new angucycline antibiotic from a terrestrial streptomycete. J. Antibiot. 2003, 56, 539–542. [Google Scholar] [CrossRef]
  135. Boudjella, H.; Bouti, K.; Zitouni, A.; Mathieu, F.; Lebrihi, A.; Sabaou, N. Isolation and partial characterization of pigment-like antibiotics produced by a new strain of Streptosporangium isolated from an Algerian soil. J. Appl. Microbiol. 2007, 103, 228–236. [Google Scholar] [CrossRef] [PubMed]
  136. Boudjella, H.; Zitouni, A.; Coppel, Y.; Mathieu, F.; Monje, M.; Sabaou, N.; Lebrihi, A. Antibiotic R2, a new angucyclinone compound from Streptosporangium sp. Sg3. J. Antibiot. 2010, 63, 709–711. [Google Scholar] [CrossRef] [PubMed]
  137. Igarashi, M.; Watanabe, T.; Hashida, T.; Umekita, M.; Hatano, M.; Yanagida, Y.; Kino, H.; Kimura, T.; Kinoshita, N.; Inoue, K.; et al. Waldiomycin, a novel WalK-histidine kinase inhibitor from Streptomyces sp. MK844-mF10. J. Antibiot. 2013, 66, 459–464. [Google Scholar] [CrossRef] [PubMed]
  138. Fakhruzzaman, M.; Inukai, Y.; Yanagida, Y.; Kino, H.; Igarashi, M.; Eguchi, Y.; Utsumi, R. Study on in vivo effects of bacterial histidine kinase inhibitor, Waldiomycin, in Bacillus subtilis and Staphylococcus aureus. J. Gen. Appl. Microbiol. 2015, 61, 177–184. [Google Scholar] [CrossRef] [PubMed]
  139. Broetz, E.; Bilyk, O.; Kroeger, S.; Paululat, T.; Bechthold, A.; Luzhetskyy, A. Amycomycins C and D, new angucyclines from Kitasatospora sp. Tetrahedron Lett. 2014, 55, 5771–5773. [Google Scholar] [CrossRef]
  140. Myronovskyi, M.; Broetz, E.; Rosenkraenzer, B.; Manderscheid, N.; Tokovenko, B.; Rebets, Y.; Luzhetskyy, A. Generation of new compounds through unbalanced transcription of landomycin A cluster. Appl. Microbiol. Biotechnol. 2016, 100, 9175–9186. [Google Scholar] [CrossRef] [PubMed]
  141. Park, H.B.; Lee, J.K.; Lee, K.R.; Kwon, H.C. Angumycinones A and B, two new angucyclic quinones from Streptomyces sp. KMC004 isolated from acidic mine drainage. Tetrahedron Lett. 2014, 55, 63–66. [Google Scholar] [CrossRef]
  142. Cruz, J.C.S.; Sosio, M.; Donadio, S.; Cruz, J.C.S.; Maffioli, S.I.; Wellington, E.; Maffioli, S.I.; Bernasconi, A.; Brunati, C.; Gaspari, E.; et al. Allocyclinones, hyperchlorinated angucyclinones from Actinoallomurus. J. Antibiot. 2017, 70, 73–78. [Google Scholar] [CrossRef] [PubMed]
  143. Tata, S.; Aouiche, A.; Bijani, C.; Bouras, N.; Pont, F.; Mathieu, F.; Sabaou, N. Mzabimycins A and B, novel intracellular angucycline antibiotics produced by Streptomyces sp. PAL114 in synthetic medium containing L-tryptophan. Saudi Pharm. J. 2019, 27, 907–913. [Google Scholar] [CrossRef]
  144. Izyani Awang, A.F.; Ahmed, Q.U.; Shah, S.A.A.; Jaffri, J.M.; Ghafoor, K.; Uddin, A.B.M.H.; Ferdosh, S.; Islam Sarker, M.Z. Isolation and characterization of novel antibacterial compound from an untapped plant, Stereospermum fimbriatum. Nat. Prod. Res. 2020, 34, 629–637. [Google Scholar] [CrossRef]
  145. Keiichiro; Motohashi; Motoki; Takagi; Hideki; Yamamura; Masayuki; Hayakawa; Kazuo; Shin-ya. A new angucycline and a new butenolide isolated from lichen-derived Streptomyces spp. J. Antibiot. 2010, 63, 545–548. [Google Scholar] [CrossRef]
  146. Nakagawa, K.; Hara, C.; Tokuyama, S.; Takada, K.; Imamura, N. Saprolmycins A–E, new angucycline antibiotics active against Saprolegnia parasitica. J. Antibiot. 2012, 65, 599–607. [Google Scholar] [CrossRef] [PubMed]
  147. Bruntner, C.; Binder, T.; Pathom-Aree, W.; Goodfellow, M.; Bull, A.T.; Potterat, O.; Puder, C.; Hoerer, S.; Schmid, A.; Bolek, W.; et al. Frigocyclinone, a novel angucyclinone antibiotic produced by a Streptomyces griseus strain from Antarctica. J. Antibiot. 2005, 58, 346–349. [Google Scholar] [CrossRef] [PubMed]
  148. Shigihara, Y.; Koizumi, Y.; Tamamura, T.; Homma, Y.; Isshiki, K.; Dobashi, K.; Naganawa, H.; Takeuchi, T. 6-Deoxy-8-O-methylrabelomycin and 8-O-methylrabelomycin from a Streptomyces species. J. Antibiot. 1988, 41, 1260–1264. [Google Scholar] [CrossRef] [PubMed]
  149. Guo, H.; Schwitalla, J.W.; Benndorf, R.; Baunach, M.; Steinbeck, C.; Görls, H.; de Beer, Z.W.; Regestein, L.; Beemelmanns, C. Gene Cluster Activation in a Bacterial Symbiont Leads to Halogenated Angucyclic Maduralactomycins and Spirocyclic Actinospirols. Org. Lett. 2020, 22, 2634–2638. [Google Scholar] [CrossRef]
  150. Kim, H.; Kim, J.; Ji, C.; Lee, D.; Shim, S.H.; Joo, H.; Kang, H. Acidonemycins A–C, Glycosylated Angucyclines with Antivirulence Activity Produced by the Acidic Culture of Streptomyces indonesiensis. J. Nat. Prod. 2023, 86, 2039–2045. [Google Scholar] [CrossRef]
  151. Abdelmohsen, U.R.; Cheng, C.; Viegelmann, C.; Zhang, T.; Grkovic, T.; Ahmed, S.; Quinn, R.J.; Hentschel, U.; Edrada-Ebel, R. Dereplication strategies for targeted isolation of new antitrypanosomal actinosporins A and B from a marine sponge associated-Actinokineospora sp. EG49. Mar. Drugs 2014, 12, 1220–1244. [Google Scholar] [CrossRef]
  152. Grkovic, T.; Abdelmohsen, U.R.; Othman, E.M.; Stopper, H.; Edrada-Ebel, R.A.; Hentschel, U.; Quinn, R.J. Two new antioxidant actinosporin analogues from the calcium alginate beads culture of sponge-associated Actinokineospora sp. strain EG49. Bioorg. Med. Chem. Lett. 2014, 24, 5089–5092. [Google Scholar] [CrossRef] [PubMed]
  153. Alhadrami, H.A.; Thissera, B.; Hassan, M.H.A.; Behery, F.A.; Ngwa, C.J.; Hassan, H.M.; Pradel, G.; Abdelmohsen, U.R.; Rateb, M.E. Bio-guided isolation of antimalarial metabolites from the coculture of two red sea sponge-derived Actinokineospora and Rhodococcus spp. Mar. Drugs 2021, 19, 109. [Google Scholar] [CrossRef]
  154. Tawfike, A.; Attia, E.Z.; Desoukey, S.Y.; Hajjar, D.; Makki, A.A.; Schupp, P.J.; Edrada-Ebel, R.; Abdelmohsen, U.R. New bioactive metabolites from the elicited marine sponge-derived bacterium Actinokineospora spheciospongiae sp. nov. AMB Express 2019, 9, 1–9. [Google Scholar] [CrossRef] [PubMed]
  155. Kim, G.S.; Kim, G.J.; Lee, B.; Oh, T.H.; Kwon, M.; Lee, J.; Jang, J.; Choi, H.; Ko, S.; Hong, Y.; et al. Highly oxygenated angucycline from Streptomyces sp. KCB15JA014. J. Antibiot. 2020, 73, 859–862. [Google Scholar] [CrossRef] [PubMed]
  156. Paululat, T.; Kulik, A.; Hausmann, H.; Karagouni, A.D.; Zinecker, H.; Imhoff, J.F.; Fiedler, H.P. Grecocyclines: New Angucyclines from Streptomyces sp. Acta 1362. Eur. J. Org. Chem. 2010, 12, 2344–2350. [Google Scholar] [CrossRef]
  157. Bringmann, G.; Lang, G.; Maksimenka, K.; Hamm, A.; Gulder, T.A.M.; Dieter, A.; Bull, A.T.; Stach, J.E.M.; Kocher, N.; Müller, W.E.G. Gephyromycin, the first bridged angucyclinone, from Streptomyces griseus strain NTK 14. Phytochem. 2005, 66, 1366–1373. [Google Scholar] [CrossRef]
  158. Kim, G.S.; Jang, J.P.; Oh, T.H.; Kwon, M.; Jang, J.H. Angucyclines Containing β-Glucuronic Acid from Streptomyces sp. KCB15JA151. Bioorg. Med. Chem. Lett. 2021, 48, 128237. [Google Scholar] [CrossRef] [PubMed]
  159. Wang, L.; Wang, L.; Zhou, Z.; Wang, Y.; Huang, J.; Ma, Y.; Liu, Y.; Huang, S. Cangumycins A–F, six new angucyclinone analogues with immunosuppressive activity from Streptomyces. Chin. J. Nat. Med. 2019, 17, 982–987. [Google Scholar] [CrossRef]
  160. Alvi, K.A.; Baker, D.D.; Stienecker, V.; Hosken, M.; Nair, B.G. Identification of inhibitors of inducible nitric oxide synthase from microbial extracts. J. Antibiot. 2000, 53, 496–501. [Google Scholar] [CrossRef]
  161. Su, H.; Shao, H.; Zhang, K.; Li, G. Antibacterial metabolites from the Actinomycete Streptomyces sp. P294. J. Microbiol. 2016, 54, 131–135. [Google Scholar] [CrossRef] [PubMed]
  162. Carr, G.; Derbyshire, E.R.; Caldera, E.; Currie, C.R.; Clardy, J. Antibiotic and antimalarial quinones from fungus-growing ant-associated Pseudonocardia sp. J. Nat. Prod. 2012, 75, 1806–1809. [Google Scholar] [CrossRef] [PubMed]
  163. Ayukawa, S.; Takeuchi, T.; Sezaki, M.; Hara, T.; Umezawa, H.; Nagatsu, T. Inhibition of tyrosine hydroxylase by aquayamycin. J. Antibiot. 1968, 21, 350–353. [Google Scholar] [CrossRef] [PubMed]
  164. Nagatsu, T.; Ayukawa, S.; Umezawa, H. Inhibition of dopamine β-hydroxylase by aquayamycin. J. Antibiot. 1968, 21, 354–357. [Google Scholar] [CrossRef]
  165. Hayaishi, O.; Okuno, S.; Fujisawa, H.; Umezawa, H. Inhibition of brain tryptophan 5-monooxygenase by aquayamycin. Biochem. Biophys. Res. Commun. 1970, 39, 643–650. [Google Scholar] [CrossRef] [PubMed]
  166. Sekizawa, R.; Iinuma, H.; Naganawa, H.; Hamada, M.; Takeuchi, T.; Yamaizumi, J.; Umezawa, K. Isolation of novel saquayamycins as inhibitors of farnesyl-protein transferase. J. Antibiot. 1996, 49, 487–490. [Google Scholar] [CrossRef]
  167. Schafer, M.; Le, T.B.K.; Hearnshaw, S.J.; Maxwell, A.; Challis, G.L.; Wilkinson, B.; Buttner, M.J. SimC7 is a novel NAD(P)H-dependent ketoreductase essential for the antibiotic activity of the DNA gyrase inhibitor simocyclinone. J. Mol. Biol. 2015, 427, 2192–2204. [Google Scholar] [CrossRef]
  168. Potterat, O.; Puder, C.; Wagner, K.; Bolek, W.; Vettermann, R.; Kauschke, S.G. Chlorocyclinones A–D, chlorinated angucyclinones from Streptomyces sp. strongly antagonizing rosiglitazone-induced PPAR-gamma activation. J. Nat. Prod. 2007, 70, 1934–1938. [Google Scholar] [CrossRef]
  169. Zhu, J.; Yu, M.; Shen, W.; Ren, X.; Zheng, H.; Mu, Y.; Lu, X.; Zhai, L. A new series of IDO1 inhibitors derived from microbial metabolites. Phytochem. Lett. 2023, 54, 76–80. [Google Scholar] [CrossRef]
  170. Kawashima, A.; Kishimura, Y.; Tamai, M.; Hanada, K. New platelet aggregation inhibitors. Chem. Pharm. Bull. 1989, 37, 3429–3431. [Google Scholar] [CrossRef]
  171. Uesato, S.; Tokunaga, T.; Takeuchi, K. Novel angucycline compound with both antigastrin- and gastric mucosal protective-activities. Bioorg. Med. Chem. Lett. 1998, 8, 1969–1972. [Google Scholar] [CrossRef]
  172. Uesato, S.; Tokunaga, T.; Mizuno, Y.; Fujioka, H.; Kuwajima, H. Absolute Stereochemistry of Gastric Antisecretory Compound P371A1 and Its Congener P371A2 from Streptomyces Species P371. J. Nat. Prod. 2000, 63, 787–792. [Google Scholar] [CrossRef] [PubMed]
  173. Alvarino, R.; Alonso, E.; Lacret, R.; Oves-Costales, D.; Genilloud, O.; Reyes, F.; Alfonso, A.; Botana, L.M. Streptocyclinones A and B ameliorate Alzheimer’s disease pathological processes in vitro. Neuropharmacology 2018, 141, 283–295. [Google Scholar] [CrossRef] [PubMed]
  174. Westrich, L.; Domann, S.; Faust, B.; Bedford, D.; Hopwood, D.A.; Bechthold, A. Cloning and characterization of a gene cluster from Streptomyces cyanogenus S136 probably involved in landomycin biosynthesis. FEMS Microbiol. Lett. 1999, 170, 381–387. [Google Scholar] [CrossRef]
  175. Matselyukh, B.P.; Polishchuk, L.V.; Lukyanchuk, V.V.; Golembiovska, S.L.; Lavrenchuk, V.Y. Sequences of Landomycin E and Carotenoid Biosynthetic Gene Clusters, and Molecular Structure of Transcriptional Regulator of Streptomyces globisporus 1912. Mikrobiol. Z. 2016, 78, 60–70. [Google Scholar] [CrossRef]
  176. Feng, Z.; Kallifidas, D.; Brady, S.F. Functional analysis of environmental DNA-derived type II polyketide synthases reveals structurally diverse secondary metabolites. Proc. Natl. Acad. Sci. USA 2011, 108, 12629–12634. [Google Scholar] [CrossRef] [PubMed]
  177. Ostash, B.; Rebets, Y.; Yuskevich, V.; Luzhetskyy, A.; Tkachenko, V.; Fedorenko, V. Targeted disruption of Streptomyces globisporus lndF and lndL cyclase genes involved in landomycin E biosynthesis. Folia. Microbiol. 2003, 48, 484–488. [Google Scholar] [CrossRef] [PubMed]
  178. Zhu, L.; Ostash, B.; Rix, U.; Nur-E-Alam, M.; Mayers, A.; Luzhetskyy, A.; Mendez, C.; Salas, J.A.; Bechthold, A.; Fedorenko, V.; et al. Identification of the function of gene lndM2 encoding a bifunctional oxygenase-reductase involved in the biosynthesis of the antitumor antibiotic landomycin E by Streptomyces globisporus 1912 supports the originally assigned structure for landomycinone. J. Org. Chem. 2005, 70, 631–638. [Google Scholar] [CrossRef]
  179. Luzhetskyy, A.; Taguchi, T.; Fedoryshyn, M.; Dürr, C.; Wohlert, S.E.; Novikov, V.; Bechthold, A. LanGT2 Catalyzes the First Glycosylation Step during landomycin A biosynthesis. Chembiochem. 2005, 6, 1406–1410. [Google Scholar] [CrossRef]
  180. Yushchuk, O.; Kharel, M.; Ostash, I.; Ostash, B. Landomycin biosynthesis and its regulation in Streptomyces. Appl. Microbiol. Biotechnol. 2019, 103, 1659–1665. [Google Scholar] [CrossRef] [PubMed]
  181. Luzhetskyy, A.; Fedoryshyn, M.; Dürr, C.; Taguchi, T.; Novikov, V.; Bechthold, A. Iteratively acting glycosyltransferases involved in the hexasaccharide biosynthesis of landomycin A. Chem. Biol. 2005, 12, 725–729. [Google Scholar] [CrossRef]
  182. Rebets, Y.; Ostash, B.; Luzhetskyy, A.; Hoffmeister, D.; Brana, A.; Mendez, C.; Salas, J.A.; Bechthold, A.; Fedorenko, V. Production of landomycins in Streptomyces globisporus 1912 and S. cyanogenus S136 is regulated by genes encoding putative transcriptional activators. FEMS Microbiol. Lett. 2003, 222, 149–153. [Google Scholar] [CrossRef]
  183. Deneka, M.; Ostash, I.; Yalamanchili, S.; Bennett, C.S.; Ostash, B. Insights into the Biological Properties of Ligands and Identity of Operator Site for LanK Protein Involved in Landomycin Production. Curr. Microbiol. 2023, 81, 5. [Google Scholar] [CrossRef] [PubMed]
  184. Ostash, I.; Rebets, Y.; Ostash, B.; Kobylyanskyy, A.; Myronovskyy, M.; Nakamura, T.; Walker, S.; Fedorenko, V. An ABC transporter encoding gene lndW confers resistance to landomycin E. Arch. Microbiol. 2008, 190, 105–109. [Google Scholar] [CrossRef]
  185. Deane, C.D.; Mitchell, D.A. Lessons learned from the transformation of natural product discovery to a genome-driven endeavor. J. Ind. Microbiol. Biotechnol. 2014, 41, 315–331. [Google Scholar] [CrossRef] [PubMed]
  186. Van Lanen, S.G.; Shen, B. Microbial genomics for the improvement of natural product discovery. Curr. Opin. Microbiol. 2006, 9, 252–260. [Google Scholar] [CrossRef]
  187. Modolon, F.; Schultz, J.; Duarte, G.; Vilela, C.; Thomas, T.; Peixoto, R.S. In situ devices can culture the microbial dark matter of corals. iScience 2023, 26, 108374. [Google Scholar] [CrossRef]
  188. Perez De Souza, L.; Alseekh, S.; Brotman, Y.; Fernie, A.R. Network-based strategies in metabolomics data analysis and interpretation: From molecular networking to biological interpretation. Expert. Rev. Proteomics. 2020, 17, 243–255. [Google Scholar] [CrossRef]
  189. He, S.; Cui, X.; Khan, A.; Liu, Y.; Wang, Y.; Cui, Q.; Zhao, T.; Cao, J.; Cheng, G. Activity Guided Isolation of Phenolic Compositions from Anneslea fragrans Wall. and Their Cytoprotective Effect against Hydrogen Peroxide Induced Oxidative Stress in HepG2 Cells. Molecules 2021, 26, 3690. [Google Scholar] [CrossRef]
  190. Meena, S.N.; Wajs-Bonikowska, A.; Girawale, S.; Imran, M.; Poduwal, P.; Kodam, K.M. High-Throughput Mining of Novel Compounds from Known Microbes: A Boost to Natural Product Screening. Molecules 2024, 29, 3237. [Google Scholar] [CrossRef]
  191. Matsumoto, A.; Takahashi, Y. Endophytic actinomycetes: Promising source of novel bioactive compounds. J. Antibiot. 2017, 70, 514–519. [Google Scholar] [CrossRef]
Figure 1. PRISMA 2020 flow diagram for systematic reviews.
Figure 1. PRISMA 2020 flow diagram for systematic reviews.
Marinedrugs 23 00025 g001
Figure 2. Structures of compounds 135.
Figure 2. Structures of compounds 135.
Marinedrugs 23 00025 g002
Figure 3. Structures of compounds 3661.
Figure 3. Structures of compounds 3661.
Marinedrugs 23 00025 g003
Figure 4. Structures of compounds 6293.
Figure 4. Structures of compounds 6293.
Marinedrugs 23 00025 g004
Figure 5. Structures of compounds 94106.
Figure 5. Structures of compounds 94106.
Marinedrugs 23 00025 g005
Figure 6. Structures of compounds 107142.
Figure 6. Structures of compounds 107142.
Marinedrugs 23 00025 g006
Figure 7. Structures of compounds 143160.
Figure 7. Structures of compounds 143160.
Marinedrugs 23 00025 g007
Figure 8. Structures of compounds 161195.
Figure 8. Structures of compounds 161195.
Marinedrugs 23 00025 g008
Figure 9. Structures of compounds 196216.
Figure 9. Structures of compounds 196216.
Marinedrugs 23 00025 g009
Figure 10. Structures of compounds 217241.
Figure 10. Structures of compounds 217241.
Marinedrugs 23 00025 g010
Figure 11. Structures of compounds 242283.
Figure 11. Structures of compounds 242283.
Marinedrugs 23 00025 g011
Figure 12. Biosynthesis of landomycins.
Figure 12. Biosynthesis of landomycins.
Marinedrugs 23 00025 g012
Figure 13. (a) Proportion of angucyclines/angucyclinones from different sources. (b) Active types of angucyclines/angucyclinones from different sources. (c) Producer of bioactive angucyclines/angucyclinones. (d) The discovery time of active angucyclines/angucyclinones.
Figure 13. (a) Proportion of angucyclines/angucyclinones from different sources. (b) Active types of angucyclines/angucyclinones from different sources. (c) Producer of bioactive angucyclines/angucyclinones. (d) The discovery time of active angucyclines/angucyclinones.
Marinedrugs 23 00025 g013
Figure 14. (a) Proportion of glycosylated angucyclines/angucyclinones. (b) Active types of angucyclines/angucyclinones with different lengths of oligosaccharide chains.
Figure 14. (a) Proportion of glycosylated angucyclines/angucyclinones. (b) Active types of angucyclines/angucyclinones with different lengths of oligosaccharide chains.
Marinedrugs 23 00025 g014
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

Liu, H.-S.; Chen, H.-R.; Huang, S.-S.; Li, Z.-H.; Wang, C.-Y.; Zhang, H. Bioactive Angucyclines/Angucyclinones Discovered from 1965 to 2023. Mar. Drugs 2025, 23, 25. https://doi.org/10.3390/md23010025

AMA Style

Liu H-S, Chen H-R, Huang S-S, Li Z-H, Wang C-Y, Zhang H. Bioactive Angucyclines/Angucyclinones Discovered from 1965 to 2023. Marine Drugs. 2025; 23(1):25. https://doi.org/10.3390/md23010025

Chicago/Turabian Style

Liu, Hai-Shan, Hui-Ru Chen, Shan-Shan Huang, Zi-Hao Li, Chun-Ying Wang, and Hua Zhang. 2025. "Bioactive Angucyclines/Angucyclinones Discovered from 1965 to 2023" Marine Drugs 23, no. 1: 25. https://doi.org/10.3390/md23010025

APA Style

Liu, H.-S., Chen, H.-R., Huang, S.-S., Li, Z.-H., Wang, C.-Y., & Zhang, H. (2025). Bioactive Angucyclines/Angucyclinones Discovered from 1965 to 2023. Marine Drugs, 23(1), 25. https://doi.org/10.3390/md23010025

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