Next Article in Journal
Absolute Structure Determination and Kv1.5 Ion Channel Inhibition Activities of New Debromoaplysiatoxin Analogues
Next Article in Special Issue
Identification and Heterologous Expression of the Kendomycin B Biosynthetic Gene Cluster from Verrucosispora sp. SCSIO 07399
Previous Article in Journal
Structure Characteristics, Biochemical Properties, and Pharmaceutical Applications of Alginate Lyases
Previous Article in Special Issue
Zhaoshumycins A and B, Two Unprecedented Antimycin-Type Depsipeptides Produced by the Marine-Derived Streptomyces sp. ITBB-ZKa6
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Products from Actinomycetes Associated with Marine Organisms

Department of Development Technology of Marine Resources, College of Life Sciences and Medicine, Zhejiang Sci-Tech University, Hangzhou 310018, China
*
Author to whom correspondence should be addressed.
Mar. Drugs 2021, 19(11), 629; https://doi.org/10.3390/md19110629
Submission received: 11 October 2021 / Revised: 5 November 2021 / Accepted: 6 November 2021 / Published: 10 November 2021
(This article belongs to the Special Issue Bioactive Compounds from Marine Streptomyces)

Abstract

:
The actinomycetes have proven to be a rich source of bioactive secondary metabolites and play a critical role in the development of pharmaceutical researches. With interactions of host organisms and having special ecological status, the actinomycetes associated with marine animals, marine plants, macroalgae, cyanobacteria, and lichens have more potential to produce active metabolites acting as chemical defenses to protect the host from predators as well as microbial infection. This review focuses on 536 secondary metabolites (SMs) from actinomycetes associated with these marine organisms covering the literature to mid-2021, which will highlight the taxonomic diversity of actinomycetes and the structural classes, biological activities of SMs. Among all the actinomycetes listed, members of Streptomyces (68%), Micromonospora (6%), and Nocardiopsis (3%) are dominant producers of secondary metabolites. Additionally, alkaloids (37%), polyketides (33%), and peptides (15%) comprise the largest proportion of natural products with mostly antimicrobial activity and cytotoxicity. Furthermore, the data analysis and clinical information of SMs have been summarized in this article, suggesting that some of these actinomycetes with multiple host organisms deserve more attention to their special ecological status and genetic factors.

1. Introduction

Actinomycetes are Gram-positive bacteria with a GC-rich linear genome and have proven to be a rich source of secondary metabolites (SMs) of broad structural diversity and biological properties [1]. The ocean has been demonstrated as an ecosystem with many unique forms of actinomycetes [2]. The diversity of marine actinomycetes is determined by the uniqueness of the marine environment: some live freely in seawater, some in the seafloor sediments or sea mud; and some are symbiotic, parasitic, endophytic, or epiphytic with marine organisms [2,3]. Compared with actinomycetes isolated from seawater and sediment samples, recent reports of secondary metabolites from marine actinomycetes associated with a variety of aquatic organisms, including invertebrates such as sponges, corals, ascidians, echinoderms, and vertebrates such as pufferfish, as well as algae and seaweed, have increased significantly [4]. Studies have indicated that multiple active compounds previously isolated from marine invertebrates were possibly produced by their symbiotic microorganisms, especially actinomycetes [5,6,7]. With interactions of the host and having special ecological status, the actinomycetes associated with marine organisms have more potential to produce active metabolites acting as chemical defenses to protect the host from predators and microbial infection.
The objective of this article is to provide an overview of the natural products from actinomycetes associated with marine animals, marine plants, macroalgae, cyanobacteria, and lichens. The present review was not only summarizing the structural classes and biological activities of SMs but also highlighted the taxonomic diversity of actinomycetes, as well as the data analysis of integrated above information. Some of these metabolites with excellent activity are expected to become new drugs such as antibiotics, antineoplastic drugs, or anticancer drugs. Therefore, actinomycetes with multiple host organisms deserve more attention to their special ecological status and genetic factors.

2. Biology of Actinomycetes Associated with Marine Animals, Marine Plants, Macroalgae, Cyanobacteria and Lichens

Marine actinomycetes are abundant in species and widely inhabit sediments, seawater, and aquatic organisms. At present, these actinomycetes are mainly separated from marine invertebrates especially sponges, ascidians, corals as well as brown algae. It is presently estimated that only 1% of microbes can be separated using traditional culturing techniques, making the potential for this field more compelling [2,6]. However, emerging technologies provide us with the tools to determine overall microbial diversity. The Ribosomal Database Project (RDP) classification of 16S rRNA sequences of marine organism-associated actinomycetes in 2014, revealed 136 genera within the subclass of Acidimicrobidae, Actinobacteridae, and Rubrobacteridea, of which Streptomyces, Micromonospora, Microbacterium, and Nocardiopsis were abundant [8]. In addition, rare actinomycetes including Actinoalloteichus, Agromyces, Agrococcus, Amycolatopsis, Nonomuraea, Phycicoccus, Saccharothrix, and Serinicoccus were discovered from these marine organisms [8]. And uncultured actinomycetes such as Iamia, Aciditerrimonas, and Illumatobacter were also detected [8]. As the diversity of metabolites is closely related to biodiversity, the potential for obtaining abundant and novel SMs from actinomyces associated with various marine hosts is relatively high.
The biosynthesis ability of actinomycetes in the production of complex natural products has been known for a long time. Not all actinomycetes, however, are prolific secondary metabolite producers. To assess the diversity and distribution of natural product-producing actinomycetes associated with various marine hosts, we constructed a neighbor-joining phylogenetic tree of 16S rRNA gene sequences. We obtained these associated actinomycetes producing natural products from literature and the antiSMASH database and selected strains with a length of 16S rRNA sequences > 900 bp that were available in the NCBI database to construct the phylogenetic tree. The phylogenetic tree involved 84 strains within nine families, and the family Streptomycetaceae (Streptomyces) represents 67% of the actinomycetes, followed by Micromonosporaceae (Salinispora and Micromonospora), Pseudonocardiaceae (Saccharopolyspora) with 18%, 5% of actinomycetes, respectively, indicating the better potential of those three families to produce SMs (Figure 1). Currently, actinomycetes that produce natural products are mainly isolated from sponges. Therefore, taking sponge-derived actinomycetes as an example, almost half of the genera belong to Micrococcineae, while their potential for secondary metabolism appears to be limited to few reports. On the contrary, Streptomycetaceae, Micromonosporaceae, and Pseudonocardiaceae with low abundance in sponges have more potential for secondary metabolism, which reveals the clinical and pharmaceutical importance of these three families for providing novel lead compounds [9].

3. Chemical Structures and Biological Properties of the Actinomycetes Associated with Marine Animals, Marine Plants, Macroalgae, Cyanobacteria, and Lichens

3.1. Natural Products of the Actinomycetes Derived from Marine Animals

Among the 34 animal phyla known on the earth, marine animals account for 33 species, 15 of which are unique to the ocean. Studies have shown that some symbiotic microorganisms are species-specific [8], indicating that marine animals may be rich in microbial resources. And the active substances in marine animals are mostly produced by their associated microorganisms, of which actinomycetes are an important group [5,6,7]. Therefore, actinomycetes associated with marine animals are a flourishing source for novel natural products. (The structure of compounds with ‘S’ added before the number is shown in supporting information.)

3.1.1. Alkaloids

Alkaloids are a kind of nitrogen-containing organic compounds existing in nature, which have alkali-like properties. They are secondary metabolites with significant biological activities, most of which have complex ring structures and the nitrogen is mostly contained in the ring.

Alkaloids Derived from the Sponge-Associated Actinomycetes

Two novel indolocarbazole alkaloids, 4′-N-methyl-5′-hydroxystaurosporine (1) and 5′-hydroxystaurosporine (2), as well as the known staurosporine (3) (Figure 2) were purified from the culture broth of Micromonospora sp. L-31-CLCO-002, which was associated with marine sponge Clathrina coriacea collected offshore Fuerteventura (Canary Islands). These compounds displayed cytotoxicity against various tumor cell lines [9,10]. The analysis of structure-activity relationships of staurosporine and its derivatives demonstrated that hydroxylation at C-3 of the indolocarbazole moiety led to the increase in anti-proliferative activity, while hydroxylation at C-11 caused a decrease in activity. The results suggested that not only the presence/absence of hydrophilic substitutions but also the position of the alteration within the molecule is significant in the anti-proliferative activities of the various staurosporine analogs [11,12].
In 2005, an ongoing screening program for in vitro anticancer microbial extracts led to the discovery of two prodigiosin analogs metacycloprodigiosin (4) and undecylprodigiosin (5) (Figure 2) from a CHCl3 extract of a strain of Saccharopolyspora sp. nov., which was associated with the marine sponge Mycale plumose collected on the coast of Qingdao, China. Both compounds 4 and 5 showed potent in vitro cytotoxicity against five cancer cell lines (P388, HL60, A-549, BEL-7402, and SPCA4), which was the first report on metacycloprodigiosin with anti-cancer activity [13]. The Micromonospora sp. strain M42 obtained from Indonesian sponge Acanthostrongylophora sp. produced manzamine A (6) and 8-hydroxy manzamine (7) (Figure 2), which had demonstrated activities against malaria, tuberculosis, and HIV [9].
Isolation of two rifamycins B and SV (8, 9) (Figure 2) was reported in 2006 from the Salinispora sp. strain M403 associated with sponge Pseudoceratina clavata. This is the first recorded source of rifamycins from marine bacteria and Salinispora sp. represents a potential new source of rifamycins outside the genus Amycolatopsis. Rifamycins are a group of antibiotics that belong to the ansamycin family with pronounced activities against Gram-positive bacteria [9,14]. The structure and activity relationship of rifamycins with many different targets have been extensively studied [15,16]. The rifamycin SV has been successfully widely used in the clinical treatment of tuberculosis, leprosy, and mycobacterial infections.
Urauchimycins A and B (10, 11) (Figure 2) were the first antimycin type antibiotics that possess a branched side chain and the side chain contains an odd number of carbons. The two novel antimycins were obtained from the Streptomyces sp. strain Ni-80 cultivated from an unidentified sponge collected at Urauchicove, Iriomote, Japan. They showed antifungal activity against Candida albicans by inhibiting the morphological differentiation of Candida albicans [9,17].
New phenazines streptophenazines A-H (1219) (Figure 2) were isolated from Streptomyces sp. strain HB202 cultivated from the marine sponge Halichondria panicea collected from the Baltic Sea (Germany). Streptophenazines 1214, 17, and 18 showed a wide range of antibacterial activity against both Gram-positive and Gram-negative bacteria, while compounds 15, 16, and 19 were only against B. subtilis [18]. In addition, streptophenazine B exhibited weak cytotoxicity against both cancer cells and normal cells [19].
In 2008, two novel macrolactams cebulactams A1 and A2 (S1, S2) featured with a six-membered cyclic ether as part of the macrocycle were obtained from the Saccharopolyspora cebuensis SPE 10-1 associated with the sponge Haliclona sp. which was collected from Cebu, Philippines [9,20]. The indolocarbazole alkaloid staurosporine (3) was purified from Streptomyces sp. strain 11, which had been cultivated from the sponge Tedania sp. It displayed significant anti-parasitic activity against Leishmania major and Trypanosoma brucei brucei with IC50 values of 5.30 and 0.022 μM, respectively. In addition, staurosporine also exhibited general cytotoxicity against 293T kidney epithelial cells (IC50 1.30 µM) and J774.1 macrophages (IC50 < 0.13 µM) [9,21].
A new anthracycline 5-iminoaranciamycin (S3) was separated from Streptomyces sp. strain Sp080513GE-26 associated with the sponge Haliclona sp. collected from Tateyama, Japan [9,22]. Three novel isoprenoids JBIR 46–48 (2022) (Figure 2), phenazine derivatives harboring dimethylallyl moieties, were produced by Streptomyces sp. SpC080624SC-11 cultivated from the sponge Cinachyra sp. collected from the seashore at Nagura Bay, Ishigaki, Japan [9,23]. JBIR 46–48 displayed cytotoxic activity against HL-60 cells (IC50 = 189, 226 and 96 μM, respectively) [24].
Two new antibacterial phenazines, 6-hydroxymethyl-1-phenazine- carboxamide (23) and 1,6-phenazinedimethanol (24) (Figure 2) were discovered from the Brevibacterium sp. KMD 003 is derived from Callyspongia sp. (Kyeongpo, Gangwondo, Korea). The two compounds demonstrated antibacterial activities against Enterococcus hirae and Micrococcus luteus with an IC50 value of 5 μM [9]. JBIR-58 (25) (Figure 2), a new salicylamide derivative, was purified from a sponge-associated actinomycetes Streptomyces sp. SpD081030ME-02 was collected offshore from Ishigaki Island, Japan. JBIR-58 exhibited cytotoxic activity against human cervical carcinoma HeLa cells with an IC50 value of 28 μM [9,25].
A novel indole alkaloid streptomycindole (S4) and a known related compound N-phenyl acetyl-L-tryptophan (S5) were isolated from Streptomyces sp. DA22 is associated with Craniella australiensis collected at the South China Sea [26]. In 2011, Rong-Bian Wei et al. reported isolation and identification of two new kijanimicin derivatives lobophorins C (26) and D (27) (Figure 2) from Streptomyces carnosus strain AZS17 associated with the marine sponge Hymeniacidon sp. collected from coastal waters of the East China Sea. Lobophorin C demonstrated potent cytotoxic activity against the human liver cancer cell line 7402 with IC50 values of 0.6 μg/mL and lobophorin D showed a significant inhibitory effect on the growth of human breast cancer cells MDA-MB 435 with an IC50 value of 7.5 μM [27].
In 2012, Abdelmohsen et al. reported unprecedented antioxidant and anti-protease activities of a dibenzodiazepine alkaloid diazepinomicin (28) (Figure 2) separated from the Micromonospora sp. RV115 recovered from the Mediterranean sponge Aplysina aerophoba. Additionally, diazepinomicin displayed broad-spectrum antitumor activity and antiparasitic activity against trypomastigote forms of Trypanosoma brucei (IC50 = 13.5 μM) [9,28].
Compound WS-9659 A (29) (Figure 2) separated from Streptomyces sp. CMS JV M18_3 associated with marine sponge Agelas sceptrum collected in Mona Island exhibited inhibitory activity on testosterone 5α-reductase [3,29]. Three new trichostatin analogs JBIR-109 (30), JBIR-110 (31), and JBIR-111 (32) (Figure 2) were purified together with trichostatin A (S6) and trichostatic acid (S7) from Streptomyces sp. RM72 is associated with an unidentified marine sponge collected near Takara Island, Japan. The compounds 3032 have relatively weak inhibitory effects on histone deacetylase with IC50 of 48 μg/mL, 74 μg/mL, and 57 μg/mL, respectively [9,30]. In addition, a new metabolite JBIR-107 (S8) was produced by sponge-derived Streptomyces tateyamensis NBRC 105047 in 2013 [31]. Four glutarimide-derived compounds (33 and S9S11) were purified from Streptomyces anulatus S71 cultivated from a marine sponge Aplysina aerophoba collected at the Yongxing Island in the South China Sea. Compound 33 (Figure 2) was identified as a new member of the glutarimide antibiotics family [32].
Salinispora sp FS-0034 collected from a sponge Theonella sp. produced rifamycin W (34) (Figure 2) with antibacterial activity. Rifamycin W showed potent antibacterial activity against multi-drug resistant human pathogens such as methicillin-resistant Staphylococcus aureus (MRSA), wild type Staphylococcus aureus (WTSA), and vancomycin-resistant Enterococcus faecium (VREF) with MICs of 15.62, 7.80 and 250.00 µg/mL, respectively [33].
Streptomyces sp. M7_15 associated with the Puerto Rican Sponge Scopalina ruetzleri produced an angucyclinone derivative frigocyclinone (35) and six new angucyclinone derivatives monacyclinones A–F (3641) (Figure 2). Frigocyclinone (35) with an unusual C-glycosidic linked aminodeoxy sugar moiety and monacyclinones A–F exhibited biological activity against rhabdomyosarcoma cancer cells (SJCRH30). Monacyclinones A–F also inhibits Gram-positive bacteria. Among them, monacyclinone F (41) showed the strongest bioactivity against SJCRH30 (EC50 = 0.73 µM) and Gram-positive bacteria. The analysis of their structures and bioactivities suggested that the aminodeoxysugar subunit, the epoxide groups, and the ketone moiety could all be important for biological activity. [34].
Naphthacene glycoside SF2446 A2 (42) (Figure 2) was isolated from Streptomyces sp. strain RV15 obtained from the sponge Dysidea tupha collected offshore Rovinj, Croatia. The SF2446 A2, previously reported against Gram-positive bacteria and several mycoplasma strains, has proven to possess two novel activities. One is the potential of inhibiting Chlamydia trachomatis and inhibiting the primary infection and progeny formation in a dose-dependent manner. In addition, it can destroy the surface area of Schistosoma mansoni and affect the gonad by impairing the formation of oogenesis and spermatogenesis [35].
In 2015, Xin Zhen et al. explicitly disclosed for the first time that tirandamycin B (44) is a 1-keto-4′-enol form, which is different from the 1-enol-4′-keto form of tirandamycin A (43) (Figure 2). The two antibiotic tirandamycins A (43) and B (44), together with staurosporine (3), were discovered from Streptomyces sp. LS298 is associated with marine sponge Gelliodes carnosa [36]. Tirandamycin A has inhibitory activity on bacterial RNA polymerase, and tirandamycin B inhibited the asparaginyl-tRNA synthetase (AsnRS) of Brugia malayi with an IC50 value of 30 μM [37,38]. In addition, staurosporine showed antifungal activity, protein kinase C inhibitory activity (IC50 = 2.7 nM) and great cytotoxic activity against HeLa S3 cells (IC50 = 4 × 10−12 M) [39].
Two novel metabolites, strepoxazine A (45) and ageloline A (46), together with two known antibiotic phenazines phencomycin (47) and tubermycin B (48) (Figure 2), were isolated from Streptomyces sp. SBT345 cultivated from Mediterranean sponge Agelas oroides. Strepoxazine A (45) exhibited a significant cytotoxic effect against human promyelocytic leukemia cells HL-60 with IC50 at 8 μg/mL. Ageloline A showed antioxidant potential and further reduce oxidative stress and genomic damage induced by 4-nitroquinoline-1-oxide (NQO). In addition, ageloline A also showed antichlamydial activity by inhibiting the formation and growth of Chlamydia trachomatis inclusion in a dose-dependent manner with an IC50 value of 9.54 ± 0.36 μM [40,41].
In 2016, Sekurova et al. reported isolation and identification of four new deferoxamine analogs (49, S12S14) with additional acyl and sugar moieties from upon overexpression of Pathway-Specific Regulatory Gene in the Streptomyces albus PVA94-07 associated with sponge collected in the Trondheim fjord (Norway). Compound 49 (Figure 2) showed 52–56% inhibition of E. coli at 16 µg/mL [42]. Comparing the structure and activity of deferoxamine analogs 49 and S14, it was found that the position of the hydroxyl group in the enamine group affected the antibacterial activity.
A new azepino-diindole alkaloid rhodozepinone (50) (Figure 2), along with two known compounds 2-amino-3-[2(1H)-quinolinon-4-yl]propionic acid (S15) and indole-3-acetic acid (S16), was isolated from Rhodococcus sp. UA13 obtained from the Red Sea sponge Callyspongia aff. implexa. Rhodozepinone (50) exhibited significant antibacterial and antitrypanosomal activities against Staphylococcus aureus NCTC 8325 (IC50 = 8.9 µg/mL) and Trypanosoma brucei brucei TC221 [IC50 = 16.3 (48 h) and 11.8 (72 h) µg/mL], respectively [43].
Jian Lin Li et al. separated three dimeric indole derivatives (5153) (Figure 2) from an actinomycete strain Rubrobacter radiotolerans cultivated from sponge Petrosia sp. obtained off the coast of Xisha Islands, China. All three metabolites suppressed chlamydial growth in a concentration-dependent manner. Among them, the novel one (51) exhibited the most effective antichlamydial activity with IC50 values of 46.6–96.4 µM in the production of infectious progeny [44].
2,3-dihydroxybenzamide (54) (Figure 2) was isolated from Streptomyces sp. SBT348 associated with Petrosia ficiformis from Milos, Greece. It exhibited significant cytotoxicity against human promyelocytic HL-60 and human colon adenocarcinoma HT-29 cell lines [45].
Isolation of three new anthranilic acid derivatives anthranosides A−C (S17, S18, 55) were reported in 2018 from the culture of sponge-derived Streptomyces sp. CMN-62 was collected at NaoZhou Island of the Guangdong Province, China. Anthranoside C (55, Figure 2) with a unique indole-containing scaffold showed anti-influenza H1N1 activity with an IC50 =171 μM (ribavirin as positive control, IC50 133 μM) [46].
Saccharomonosporine A (56) (Figure 2) and a novel brominated oxo-indole alkaloidconvolutamydine F (S19), along with other two known induced metabolites (57, S20), were identified from the co-culture of Saccharomonospora sp. UR22 and Dietzia sp. UR66, cultivated from the Red Sea sponge Callyspongia siphonella. Additionally, the axenic culture of Saccharomonospora sp. UR22 led to the isolation of four known microbial metabolites S21S23 and 58 (Figure 2). Compounds 56 and 57 (Figure 2) were potent Pim-1 kinase inhibitors and displayed significant antiproliferative activities against HT-29 and HL-60 cell lines [47].
Tetrocarcin Q (59) is a novel spirotetronate glycoside with a unique glycosyl group 2-deoxy-allose at the C-9 position, which was separated together with six known analogs tetrocarcin A (60), AC6H (61), tetrocarcin N (62), tetrocarcin H (63), arisostatin A (64) (Figure 2), and tetrocarcin F1 (S24) from Micromonospora carbonacea LS276 (FJ937935) associated with Gelliodes carnosa from Ling Shui Bay. Compound 59 displayed moderate antibacterial activity against Bacillus subitlis ATCC 63501 with a MIC value of 12.5 µM. Compounds 6064 also showed potent antibacterial activity against Bacillus subitlis ATCC 63,501 with MICs of <0.048, 0.5, 1.562, 50, 0.048 µM, respectively. In addition, compounds 60 and 64 showed moderate activity against four cell lines (A549, BGC823, HCT116, HepG2) with the IC50 values ranging from 5.33 µM to 19.7 µM, and also exhibited the most potent antitumor activity against U87MG cell line with IC50 values of 0.50 µM and 2.42 µM, respectively. On comparing the structure and activity of analogs (6064), it was found that the modification of the tetronolide skeleton affected the in vitro antitumor activity to some extent. Tetrocarcin F1 (S24) was inactive, indicating that the sugar moiety at C-9 position could play an important role in the antitumor activity. In addition, compound 59 showed no or weak in vitro antitumor activity, suggesting that the deoxy sugar analog may also influence the antitumor activity [48].
Tirandamycins A and B (43, 44) were isolated from strain HNM0039T, a novel Streptomyces sp. named Streptomyces tirandamycinicus sp. nov., which was obtained from a marine sponge collected from the coast of Wenchang, Hainan Province of China. Tirandamycins A and B displayed potent inhibitory activity against Streptococcus agalactiae HNe0 and showed antibacterial activity against Bacillus subtilis GIM1.222 [49].
2-ethylhexyl-1H-imidazole-4-carboxylate (65) (Figure 2) and a known alkaloid butyl 1H-imidazole-4-carboxylate (S25), were isolated from Verrucosispora sp. FIM06-0036, associated with marine sponge sample from the East China Sea. Compound 65 was active against H. pylori, K. pneumonia, S. aureus, and E. faecium with MIC values of 8 µg/mL, 64µg/mL, 16 µg/mL, and 256 µg/mL, respectively [50]. The Verrucosispora sp. FIM06025 obtained from a sponge sample collected from the East China Sea, led to the isolation of two new alkaloids (66 and S26). Among them, compound 66 (Figure 2) exhibited a broad spectrum of antimicrobial activity with MIC values ranging from 3.4 to 200 μg/mL [51]. Fridamycin I (S27) was first isolated from sponge-derived Actinokineospora spheciospongiae sp. nov., obtained from the Red Sea sponge Spheciospongia vagabunda collected from offshore Ras Mohamed, Egypt [52].
In 2018, a sponge-derived strain MCCB267 was obtained from Mycale sp., collected in the Indian Ocean off the southeast coast of India. The strain MCCB267 designated as S. zhaozhouensis subsp. mycale. subsp. nov. led to the discovery of four the polycyclic tetramate macrolactam (PTM) family: ikarugamycin (IK) (67), clifednamide A (CF) (68), 30-oxo-28-N-methylikarugamycin (OI) (69), and 28-N-methylikarugamycin (MI) (70) (Figure 2). The four compounds 6770 were active against NCI-H460 lung carcinoma cells in vitro, by inducing apoptosis. Compounds 67, 69, and 70 induced cell cycle arrest during the G1 phase in the NCI-H460 cell line, whereas 68 induced cell arrest in the S phase [53].
Borrelindine J and K (71 and 72) (Figure 2) with rare nitrile group were obtained only in the co-culture of sponge-associated Streptomyces rochei MB037 with fungus Rhinocladiella similis 35 derived from gorgonian, along with two known 18-membered macrolides, borrelidin (S28), and borrelidin F (S29). Compounds 71 and 72 showed potent antibacterial activity against methicillin-resistant Staphylococcus aureus with MIC values of 0.195 and 1.563 µg/mL, respectively [54].
Isolation of 9H-pyrido[3,4-b]indole (S30) and 9H-purin-6-amine (S31) was reported in 2019 from sponge-derived Streptomyces sp. G248 collected in the East Vietnam Sea [55]. In addition, the 9H-pyrido[3,4-b]indole (S30) was identified from sponge-associated Streptomyces sp. G246 collected at Son Tra island (Da Nang)-Vietnam. Along with 9H-pyrido[3,4-b]indole, another known compound indole-3-acetic acid (S16) was discovered from this strain [56].
Two quinocycline antibiotics strains, quinocycline B (kosinostatin) (73) and isoquinocycline B (74) (Figure 2), were separated from the strain 28ISP2-46T assigned to a novel species Micromonospora ferruginea sp. which was recovered from a mid-Atlantic deep-sea sponge. Compounds 73 and 74 showed antibiotic activity and also inhibited DNA topoisomerase IIα. Isoquinocycline B (74) showed good antibiotic activity not only against the Gram-positive strain S. aureus SH1000, but also the Gram-negative strains efflux knockout (KO) mutant strains of K. pneumoniae ATCC10031 and A. baumannii ATCC 17978. It was also active against the human liver cancer cell line HepG2 with an IC50 of 13.3 µM [57].
Streptomyces sp. RM66 derived from the Red Sea sponge Amphimedon sp. in Hurghada (Egypt), led to the discovery of eight phenazine analogs (S32S37, 47, 48) through the addition of GlcNAc. Tubermycin B (48) showed environmentally compatible antimicrobial activity against various phytopathogens. Phencomycin (47) exhibited antibacterial activities against Escherichia coli and Bacillus subtilis [58].
Marine sponge Callyspongia sp. collected from Hurghada (Red Sea, Egypt) led to the isolation of two actinobacteria Rhodococcus sp. UR59 and Actinokineospora spheciospongiae strain EG49. In addition, the strain EG49 was previously discovered the Red Sea sponge Spheciospongia vagabunda. UK-2B (75) with antifungal activity was isolated from the axenic cultures of strain EG49. And the separate cultivation of strain UR59 leads to the identification of Mitomycin K (antitumor activity) (76), Piericidin F (anticancer activity) (77), Migrastatin (anticancer activity) (78) (Figure 2) [59].

Alkaloids Derived from the Coral-Associated Actinomycetes

Three thiazole derivatives known as watasemycin A (79) (Figure 3), pulicatin G (S38), and aerugine (80) (Figure 3) were identified together with pyrrole-2-carboxamide (S39) and furan-2-carboxamide (S40) from Streptomyces sp. OUCMDZ-1703 associated with a soft coral sample collected from the South China Sea. Thiazole derivatives 79 and 80 displayed moderate antibacterial activity against S. aureus and three methicillin-resistant strains MRSA082, MRSA111, and MRSA234 [60].
Lobophorin K (81) (Figure 3), a novel metabolites separated from Streptomyces sp. M-207 associated with Lophelia pertusa collected at submarine Avilés Canyon exhibited cytotoxicity against two human tumor cell lines pancreatic carcinoma (MiaPaca-2) and breast adenocarcinoma (MCF-7). It also showed moderate and selective antibiotic activity against pathogenic Gram-positive bacteria such as Staphylococcus aureus [61].
Ziwen Cong et al. reported the isolation of isotirandamycin B (82) (Figure 3) from Streptomyces sp. SCSIO 41399 cultivated from the Porites sp. coral collected in Wenchang, Hainan province of China. Along with isotirandamycin B, four other known compounds anthracycline derivatives (83, 84) (Figure 3) and tirandamycin derivatives (43 and 44) were also discovered from this strain. Compounds 82, 43, and 44 were active against Streptococcus agalactiae with MIC values of 11.5, 5.9, and 5.7 μM, respectively. Besides, compounds 83 and 84 displayed moderate in vitro cytotoxicity against the K562 cell lines with IC50 values of 1.80 ± 0.01 and 12.1 ± 0.07 μM, respectively [62].
Two novel metabolites uridine derivative 11457A (S42) and indole derivative 11457B (S43), together with 1H-indole-2-carbal-dehyde (S44) were isolated from Pseudonocardia sp. SCSIO 11457, which had been cultivated from the scleractinian coral Galaxea fascicularis [63].

Alkaloids Derived from the Ascidian-Associated Actinomycetes

Two novel antitumor antibiotics, lomaiviticins A (85) and B (86) (Figure 4) with unique dimeric diazobenzofluorene glycosides were discovered from Salinispora pacifica LL-37I366 derived from marine ascidian Polysyncraton lithostrotum. Lomaiviticins A and B were demonstrated to be significant DNA-damaging agents with a minimum induction concentration ≤0.1 ng/spot. They also exhibited potent antimicrobial activities against Gram-positive bacteria Staphylococcus aureus and Enterococcus faecium (MICs, 6–25 ng/spot). In addition, lomaiviticin A also displayed cytotoxicity against a number of cancer cell lines with IC50 values ranging from 0.01 to 98 ng/mL [64,65,66].
A novel dibenzodiazepine alkaloid diazepinomicin (28) was obtained from Micromonospora sp. DPJ12 separated from marine ascidian Didemnum proliferum Kott which was collected at Shishijima Island, Japan. Diazepinomicin showed modest antimicrobial activity against selected Gram-positive bacteria with MIC of approximately 32 µg/mL [67].
Two new piericidin compounds C7 (87) and C8 (88), together with previously identified piericidins A1 (89) and A2 (90) (Figure 4) were isolated from the Streptomyces sp. YM14-060, which is associated with an unidentified greenish ascidian from Iwayama Bay, Palau. The compounds 8790 exhibited cytotoxicity against RG-E1A-7 rat glial cells and inhibited the growth of Neuro-2a mouse neuroblastoma cells [64,68]. Streptomyces sp. JP90 separated from marine ascidian Aplidium lenticulum (Great Barrier Reef, Australia) produced a new organophosphate (S)-cinnamoyl-phosphoramide (91) (Figure 4), which displayed inhibitory activity towards Butyrylcholinesterase (BChE) [64].
Isolation of a new antibiotic arenimycin (92) (Figure 4) was reported in 2009 from Salinispora arenicola strain CNR-647 associated with sea squirt Ecteinascidia turbinate collected at Sweetings Cay, Grand Bahama Island. Arenimycin exhibited potent activities against drug-resistant Staphylococci, some other Gram-positive microanimals and plants, and one Mycobacterium strain. In addition, arenimycin was also active against eukaryotic cell division, which may lead to non-selective cytotoxicity [64,69].
The marine bacterium Aeromicrobium halocynthiae KME 001T, which was separated from sea squirt Halocynthia roretzi (Gangneung, Korea), led to the discovery of taurocholic acid (S45) [64,70]. Bohemamine (S46) was isolated from Streptomyces sp., a marine actinomycete associated with an unidentified ascidian collected from Lyttelton Harbor, New Zealand [64]. Forazolines A (93) (Figure 4) and B (S47) were isolated from Actinomadura sp. WMMB-499 that was obtained from the marine ascidian Ecteinascidia turbinate. Forazoline A (93) displayed inhibitory activity against the fungal pathogen Candida albicans with a MIC value of 16 μg/mL [71].
Three new 2(1H)-pyrazinone derivatives including (S)-6-(sec-butyl)-3- isopropylpyrazin-2(1H)-one (94) (Figure 4), (S)-3-(sec-butyl)-6-isopropylpyrazin-2(1H)-one (S48) and (S)-6-(sec-butyl)-3-isobutylpyrazin-2(1H)-one (95), together with the known (1H)-pyrazinones analogues deoxymutaaspergillic acid (96), 3,6-diisobutyl-2(1H)- pyrazinone (97) and 3,6-disec-butyl-2(1H)-pyrazinone (98) (Figure 4), were isolated from Streptomyces sp. Did-27, which is associated with the marine ascidian Didemnum sp. Expect for compound S48, all other compounds presented cytotoxic activities against cancer cell lines HCT-116, HepG2 and MCF-7 [64,72].
Compound 1, 6-dihydroxyphenazine (99) (Figure 4) was produced by Nocardiopsis dassonvillei HQA404 derived from star ascidian Botryllus schlosseri. This phenazine (99) has antimicrobial activity against Vibrio anguillarum and Vibrio parahaemolyticus, lethal activity against Artemia salina, and enzyme inhibiting activity against Alpha-glucosidase. 2-(acetylamino)-phenol (100) (Figure 4) isolated from Nocardiopsis dassonvillei HQA404 showed lethality against brine shrimp Artemia salina [64].

Alkaloids Derived from the Actinomycetes Associated with Other Marine Animals

Halichomycin (101) (Figure 5) was a novel structurally unique macrolide with potent cytotoxicity, separated from Streptomyces hygroscopicus OUPS-N92 obtained from the gastrointestinal tract of marine fish Halichoeres bleekeri. Halichomycin exhibited significant cytotoxicity in the P-388 lymphocytic leukemia test system with ED50 of 0.13 μg/mL [73].
Isolation of three novel cytotoxic indolic metabolites 3,6-disubstituted indoles (102104) (Figure 5) were reported in 2003 from Streptomyces sp. BL-49-58-005, which was isolated from an unidentified Mexican marine invertebrate. Indole A (6-prenyltryptophol) (102) exhibited the best activity against K-562 (leukemia) with a GI50 value of 8.46 µM. Aldoxime indole B (103) showed activity with GI50 values within the micromolar range against different cancer cell lines [74].
Nocardiopsis dassonvillei RG-33B isolated from the ovary of pufferfish Fugu rubripes from the Bohai Sea of China produced tetrodotoxin (105) (Figure 5). Tetrodotoxin (TTX) is one of the most potent nonprotein neurotoxins [75]. Two known inactive metabolites bohemamine (S46) and bohemamine B (S49) were identified from the small-scale extract of Streptomyces sp. (LA3L2) derived from New Zealand and Malaysian marine invertebrates. Additionally, thiazostatin B (S50) was produced by Streptomyces sp. (LA5L4) [76].
A new metabolite JBIR-66 (106) (Figure 5) isolated from Saccharopolyspora sp. strain SS081219 JE-28 associated with an unidentified tunicate collected at the seashore of Tateyama City, Chiba Prefecture, Japan, displayed relatively weak activity against human lymphoblastoid Namalwa cells [64,77].
Streptomyces sp. 1053U.I.1a.3b, cultivated from Lienardia totopotens, a new species of conoidean mollusk collected in the Philippines, led to the identification of two novel lobophorins H-I (S51, 107) and three known lobophorins F (108), B (109), C (26). Compounds 107-109 (Figure 5) and 26 showed strong inhibitory activity to M. tuberculosis and B. subtilis with MIC values ranging from 1.3 to 24 μM. In addition, all these active compounds also showed strong cytotoxicity against the human CEM-TART cancer cell line [78].
A novel hydroxamate metabolite MBJ-0003 (110) (Figure 5) was produced by Micromonospora sp. 29867 derived from a shellfish collected in Suruga Bay, Shizuoka Prefecture, Japan. MBJ-0003 displayed moderate cytotoxicity against human ovarian adenocarcinoma SKOV-3 cells with the IC50 values of 11 μM [79].
Caerulomycin A (111) (Figure 5), which was extracted from the Actinoalloteichus sp. strain PM0525875 associated with a marine invertebrate collected from the deep-sea (Anjuna Beach, Goa, India), showed potent in vitro antifungal activity against drug-resistant fungal strains and its minimum inhibitory concentration (MIC) was found in the range of 0.39–1.56 µg/mL against pathogenic fungal test strains [80].
Keyicin (112) (Figure 5), a novel and otherwise unattainable bisnitroglycosylated anthracycline antibiotic, was discovered from the producer Micromonospora sp. co-culturing with Rhodococcus sp. associated with marine invertebrates. The Keyicin was selectively active against Gram-positive bacteria including Rhodococcus sp. and Mycobacterium sp. E. coli-based chemical genomics studies revealed that keyicin’s MOA, in contrast to many other anthracyclines, does not invoke nucleic acid damage [81].
Compounds 58, 113116 (Figure 5), S21, and S52 were isolated from Streptomyces sp. G278 that was obtained from echinoderm Holothuria edulis collected in Cu Lao Cham–Quang Nam. Compounds 113 and 114 were isolated from a natural source for the first time. Compounds 113 selectively inhibited Enterococcus faecalis. Compound 113 was proven to have antibacterial and antifungal activity and the known metabolites (115, 116, 58) exhibited antimicrobial activity. In addition, compound 115 possessed antifouling activities [82].
The isolation of a new pyrazolidine derivative, 1-acetyl-2-isobutyrylpyrazolidine-4- carboxylic acid (S53) was reported in 2018 from a sea anemone-associated actinomycete Streptomyces sp. ZZ406 separated from Haliplanella lineate [83]. An unusual macrodilactone streptoseomycin (117) (Figure 5) with potent bioactivity against Helicobacter pylori with a MIC value of 2 μg/mL was discovered from Streptomyces seoulensis A01 associated with a marine prawn collected in the Yellow Sea in China. It also exhibited antibacterial activities against a panel of microaerophilic bacteria with MICs in the range of 4–8 μg/mL [84].
Five novel 5-hydroxyanthranilic acid derivatives anthocidins A-D (S54S57) and crassilin (S58) were purified from a sea urchin-derived actinomycete Streptomyces sp. HDa1, which was obtained from the gut of Anthocidaris crassispina collected from Hainan Island, China. Additionally, two known analogs n-lauryl 5-hydroxyanthranilate (118) and isolauryl 5-hydroxyanthranilate (119) (Figure 5), along with benzamide (S59), 3-hydroxy-4-methoxycinnamamide (120) (Figure 5), and oxachelin (S60) were also discovered from this strain. Compounds 118 and 119 were demonstrated to possess potent in vitro 5-lipoxygenase inhibitory activity, and herein, compound 118 was isolated and reported as a natural product for the first time. Compound 120 showed weak activity against the Gram-positive bacterium Bacillus subtilis with an inhibition zone of 3 mm [85,86].
In 2020, Zhenbin Zhou et al. reported isolation and identification of three novel borrelidins M-O (121, S61, S62), together with four previously known borrelidins CR1, A, E, and K (122, 123, S63, S64) from Streptomyces olivaceus SCSIO LO13 associated with pulmonated mollusks Onchidium sp. collected at Daya Bay, South China Sea. Borrelidin A (123) (Figure 5) has a variety of biological activities such as antibacterial, anti-parasite, and cytotoxic activities. It showed significant activity against Micrococcus luteus with an MIC value of <0.5 μM. Borrelidins M (121) and CR1 (122) (Figure 5) displayed weak cytotoxicity against normal human hepatic cell line L02, but no inhibitory effect on cancer cell lines was detected. In addition, borrelidins M, CR1, and K exhibited moderate activity against Micrococcus luteus with MIC values of 33 μM [87]. The analysis of structure-activity relationships revealed that the carboxyl moiety at C-22 and the position of hydroxylation are significant for both cytotoxic and antibacterial activities. In addition, the cytotoxicities of borrelidins were correlated to the cyano moiety, stereo configurations of the hydroxyl moiety at C-11, and the C-C double bonds. And the steric structural arrangement within the C-17 side chain is important for differentiating cytotoxic and antiangiogenic activities [87,88,89].

3.1.2. Polyketides

Polyketides are a large class of secondary metabolites produced by bacteria, fungi, actinomycetes, or plants. They are synthesized by simple fatty acids under the catalysis of polyketide biosynthase (PKS) through a synthesis pathway similar to long-chain fatty acids. Its structural types include macrolides, aromatic polyketides, polyether, pyranones, and other polyketides.

Polyketides Derived from the Sponge-Associated Actinomycetes

IB-96212 (124) (Figure 6) is a 26-membered spiroketal macrolide produced by Micromonospora sp. L-25-ES25-008 was obtained from an unidentified sponge from the Indian Ocean, Mozambique. It exhibited cytotoxic activity against mouse leukemia P-388, human lung non-small carcinoma A-549, colon adenocarcinoma HT-29, and melanoma MEL-28 cell lines [9,90].
The identification of a new angucyclinone PM070747 (125) (Figure 6) was reported from sponge-derived Saccharopolyspora taberi PEM-06-F23-019B obtained from near the coast of Tanzania, together with the known angucyclinone PD116740 (126) (Figure 6). The two angucyclinones showed antitumor activity and compound 126 was active against leukemia and adenocarcinoma cell lines [91].
Tetracenoquinocin (127) (Figure 6) was separated from sponge-associated Streptomyces sp. Sp080513GE-26 and exhibited cytotoxicity against human cervical carcinoma HeLa cells and human acute myelogenous leukemia LH-60 cells with IC50 values of 2.7 μM. Additionally, aranciamycin (128) and antibiotic SM 173B (129) (Figure 6) were also discovered from strain Sp080513GE-26. Aranciamycin was active against human cervical carcinoma HeLa cells and human acute myelogenous leukemia LH-60 cells with IC50 values of 2.7 and 4.1 μM, respectively. On comparing the cytotoxic activity of tetracenoquinocin (127), 5-iminoaranciamycin (S3), and aranciamycin (128), it was found that the ketone functional group at C-5 is essential for the cytotoxicity against the cancer cells [9,22].
Four new γ-pyrones nocapyrones A–D (S65S68) were purified from the Nocardiopsis strain HB383 associated with Halichondria panacea collected from the Baltic Sea (Germany) [9,92]. The new tetronic acid derivatives tetromycins 1-4 (S69, S70, 130, 131) and a known one tetromycin B (S71) were produced by Streptomyces axinellae Pol001T, which was isolated from the Mediterranean sponge Axinella polypoides. Tetromycins 3 and 4 (130, 131) (Figure 6) displayed protease inhibition activities against several cysteine proteases and exhibited pronounced activity against Gram-positive bacteria methicillin-resistant Staphylococcus aureus [9,93].
Three new C-glycosylated benz[a]anthraquinone derivatives, urdamycinone E (132), urdamycinone G (133), and dehydroxyaquayamycin (134) (Figure 6) were identified from Streptomycetes sp. BCC45596 separated from marine sponge Xestospongia sp. collected at Sichang Island, Chonburi, Thailand. Urdamycin E (135) (Figure 6), the possible biosynthetic precursor of 132134, has also been identified from this strain. These compounds (132135) exhibited potent anti-Plasmodium palcifarum K1 strain with IC50 values of 0.0534–2.93 μg/mL and anti-Mycobacterium tuberculosis with minimum inhibition concentrations (MICs) in a range of 3.13–12.50 μg/mL [94,95].
Six dihydroquinone derivatives (136141) (Figure 6) were isolated from sponge-associated Streptomyces sp. CMS JV M18_3 [3]. Chloro-Dihydroquinone 1–4 (136139) has great antibacterial activities against MRSA and vancomycin-resistant Enterococcus faecium (VREF). In addition, compounds 136139 were active against HCT-116 human colon carcinoma [96]. SF2415B3 (141) displayed anti-biofilm activity inhibiting Staphylococcus aureus biofilm formation, and compound 140 exhibited antimicrobial activities against some Gram-positive bacteria [97,98].
In 2014, Min Cheol Kim et al. reported isolation and identification of two novel tetracenedione derivatives nocatriones A (142) and B (143) (Figure 6) from Nocardiopsis sp. KMF-002 was cultivated from an unidentified dark purple marine sponge. Nocatrione A (142) showed a significant protective effect against UVB irradiation in both NHDF cell lines, whereas nocatrione B (143) was active against UVB only in a specific NHDF cell line [99].
Actinokineospora sp. EG49 cultivated from the marine sponge Spheciospongia vagabunda afforded two new actinosporin analogs actinosporins C (144) and D (145) (Figure 6). At 1.25 μM, actinosporins C and D showed a significant antioxidant and protective capacity from the genomic damage induced by hydrogen peroxide in the human promyelocytic (HL-60) cell line. Additionally, two other new antitrypanosomal angucycline-type metabolites actinosporins A (146) and B (147) (Figure 6) were also discovered from strain EG49 [100,101].
Microluside A (148) (Figure 6) is a unique O-glycosylated disubstituted xanthone separated from Micrococcus sp. EG45 obtained from the Red Sea sponge Spheciospongia vagabunda. It exhibited antibacterial potential against Enterococcus faecalis JH212 and Staphylococcus aureus NCTC 8325 with MIC values of 10 and 13 μM, respectively [102]. In 2015, a sponge-derived strain Streptomyces sp. M7_15 led to the isolation of dimethyldehydrorabelomycin (S72) [34].
Three S-bridged pyranonaphthoquinone dimers naquihexcins A (149) (Figure 6) and B (S73), and a related analog (−)-BE-52440A (150) (Figure 6) were produced by sponge-associated Streptomyces sp. HDN-10-293. Among them, (−)-BE-52440A (150) showed cytotoxicity against NB4 and HL-60 cells with IC50 values of 1.7 and 1.8 μM, respectively. Naquihexcin A (149) bears a rare unsaturated hexuronic acid moiety and could inhibit the proliferation of an adriamycin-resistant human breast cancer cell line MCF-7 ADM with IC50 = 16.1 μM, indicating that the unsaturated hexuronic acid moiety could enhance the activity against the cancer cells [103].
Compound S74, identified as 3-hydroxy-2-methyl-4H-pyran-4-one (maltol), was isolated both from the Streptomyces sp. SBT348 and Rhodococcus sp. UA13 [43,45]. Compound S75 was discovered from the co-culture of sponge-derived Saccharomonospora sp. UR22 and Dietzia sp. UR66 [47]. In 2018, Dongbo Xu et al. reported the discovery of three novel angucyclines nocardiopsistins A-C (151153) (Figure 6) separated from Nocardiopsis sp. HB-J378 associated with Theonella sp. Nocardiopsistin B showed the best anti-MRSA activity with the same MIC (3.12 μg/mL) as that of chloramphenicol, whereas nocardiopsistins A and C have a moderate anti-MRSA activity (MIC = 12.5 μg/mL) [104]. On comparing the anti-MRSA activity of nocardiopsistins A–C (151153), it was found that the ketone functional group at C-4 could enhance the anti-MRSA activity, while the hydroxyl group at C-3 weakened activity.
7-methoxy-2,3-dimethylchromone-4-one (S76) was first isolated from Streptomyces rochei MB037 associated with marine sponge Dysidea arenaria collected at Yongxin Island in the South China Sea. And the co-culture with fungus Rhinocladiella similis 35 derived from gorgonian could enhance its production [54]. The new fridamycins H (154), together with three known actinosporins C, D, and G (144, 145, 155) (Figure 6) were obtained from sponge-associated Actinokineospora spheciospongiae sp. nov. Among them, fridamycin H (154) exhibited potent antitrypanosomal activity and growth inhibitory activity towards Trypanosoma brucei strain TC221 [52].
Two new lavandulylated flavonoids (156 and 157) (Figure 6) were produced by the marine sponge-derived Streptomyces sp. G246 had a broad spectrum of antimicrobial activity [56]. Three novel lavandulylated flavonoids 158160 (Figure 6) were isolated along with two known compounds 161 and 162, from the culture broth of sponge-associated Streptomyces sp. G248. Compounds 158160 exhibited remarkable antimicrobial activity. Additionally, two known compounds 161 and 162 (Figure 6) showed inhibitory activity against Mycobacterium tuberculosis H37Rv with MIC values of 6.0 µg/mL and 11.1 µg/mL, respectively [55].
Actinosporins A and C (146, 144) were isolated from the axenic cultures of strain EG49. Actinosporin A showed anti-trypanosomal activity and actinosporin C exhibited antioxidant activity. New actinosporins E–H (163, S77, 155, 164) (Figure 6) were produced by EG49 through the activation of cryptic gene cluster by N-acetyl-D-glucosamine (GluNAc). In addition, tetrangulol (165) (Figure 6) and the same actinosporins E, G, and H were discovered with antimalarial activity from the co-culture of strains Actinokineospora spheciospongiae EG49 and Rhodococcus sp. UR59. Tetrangulol was also reported as an antibiotic in previous research. What’s more, the separate cultivation of strain UR59 leads to the isolation of Kaimonolide B (166) (Figure 6) (plant growth inhibitor) and 8,15-Dideoxylankanolide (S78) [59].

Polyketides Derived from the Coral-Associated Actinomycetes

Isolation of two novel compounds Octalactins A (167) (Figure 7) and B (S79) with fully saturated eight-membered lactone ring were reported in 1991 from Streptomyces sp. PG-19 collected on the surface of Cortez gorgonian octocoral Pacifigorgia sp. Octalactin A demonstrated significant in vitro cytotoxicity toward B-16-F10 murine melanoma (IC50 = 7.2 × 10−3 µg/mL) and HCT-116 human colon tumor (IC50 = 0.5 µg/mL) [105].
A novel analog of jadomycin B, 7b,13-dihydro-7-O-methyl jadomycin B (S80) was produced by a marine actinomycete Micromonospora sp. strain A5-1 obtained from soft coral Scleronephthya sp. in the East China Sea [106]. Streptomyces sp. OUCMDZ-1703 associated with a soft coral collected from the South China Sea led to the discovery of two new chlorinated polyketides strepchloritides A and B (168 and 169) (Figure 7), together with 1-(3,5-dihydroxyphenyl)ethanone (S41). Polyketides 168 and 169 showed moderate cytotoxicity against MCF-7 tumor cells [60].
Aranciamycin K (S81) and two known anthracycline derivatives (170, S82) were produced by coral-associated Streptomyces sp. SCSIO 41399. Compound 170 (Figure 7) exhibited moderate in vitro cytotoxicity against the K562 cell lines [62].
A novel gram-positive antibiotic anthracimycin B (171) and anthracimycin (172) (Figure 7) were isolated from Streptomyces cyaneofuscatus M-169 associated with a gorgonian coral collected in the Avilés Canyon. Anthramycin has significant activity against four Gram-positive bacteria (MSSA, MRSA, vancomycin sensitive Enterococcus faecium, and vancomycin sensitive Enterococcus faecalis) with MIC values less than 0.03 µg/mL. And anthracimycin B was also active against these four Gram-positive bacteria. In addition, anthracimycin displayed anti-tubercular activity against Mycobacterium tuberculosis [107].
1-hydroxy-1-norresistomycin (HNM) (173) (Figure 7) was produced by Streptomyces variabilis obtained from Scleractinia coral Acropora formosa. HNM, and inhibited biofilm formation of E. coli, V. cholerae, and S. aureus with an efficiency of 96%, 92%, and 93%, respectively. Additionally, it also exhibited potent cytotoxic activity against cell lines viz. HMO2 (gastric adenocarcinoma) and HePG2 (hepatic carcinoma) in vitro [108].
Isolation of three novel glycosylated macrolides iseolides A–C (174176) (Figure 7) were reported in 2020 from Streptomyces sp. DC4-5 is associated with a stony coral Dendrophyllia. Iseolides showed potent antifungal activity against a plant pathogen Glomerella cingulata, as well as human pathogens Candida albicans and Trichophyton rubrum with MICs in the range of 0.19–6.25 μg/mL [109].

Polyketides Derived from the Ascidian-Associated Actinomycetes

Ubiquinone Q9 (177) (Figure 8) was isolated from Nocardia sp. strain KMM 3749, a marine actinomycete associated with an unidentified ascidian. This compound inhibited the development of fertilized eggs of Strongylocentrotus intermedius and caused hemolysis of mouse erythrocytes [64,110].
The polyketide griseorhodin A (178) (Figure 8) was found to be biosynthesized by Streptomyces sp. JP95, which is associated with marine ascidian Aplidium lenticulum collected at Heron Island, Queensland, Australia. Griseorhodin A, a member of the rubromycin family, is an inhibitor of human telomerase [64,111].
The isolation of Streptomyces sp. #N1-78-1 from sea squirt Ecteinascidia turbinata in Puerto Rico led to the purification of bisanthraquinones 1 and 2 (179, 180), as well as derivative 3 (181) (Figure 8), the dehydration product of bisanthraquinone 1. Bisanthraquinones 1 and 2 showed potent antimicrobial activities against MRSA (methicillin-resistant Staphylococcus aureus) and VRE (vancomycin-resistant Enterococcus faecalis), and these three compounds displayed cytotoxic activity against HCT-116 cells [64,112].
Four novel anthracyclinones (182, 183, S83, and S84) were produced by a strain of Micromonospora sp. derived from a Brazilian endemic ascidian Eudistoma vannamei. Compounds 182 and 183 (Figure 8) displayed moderate cytotoxic effects on the human colon adenocarcinoma cell line HCT-8 with IC50 values of 12.7 and 6.2 μM, respectively [113].
In 2013, Actinomadura sp. WMMB499 associated with ascidian Ecteinascidia turbinata collected in the Florida Keys led to the discovery of four new halogenated electrophilic polyketides halomadurones A–D (S85, S86, 184, 185). Halomadurones C (184) and D (185) (Figure 8) showed potent nuclear factor E2-related factor antioxidant response element (Nrf2-ARE) activation [64,114]. In addition, a polyether antibiotic ecteinamycin (186) (Figure 8) was also isolated from this strain, acting as an ionophore antibiotic and shows potent antibacterial activity, especially against a wide array of Clostridium difficile strains [115].
Five polyol polyketides containing a decalin ring, including four novel nahuoic acids B−E (187190) and nahuoic acid A (191) (Figure 8) were separated from Streptomyces sp. SCSGAA 0027 and associated with gorgonian coral Melitodes squamata collected at the South China Sea. Nahuoic acid A (191) with an unprecedented carbon skeleton was the first natural product inhibiting the SETD8 lysine methyltransferase and the first selective SETD8 inhibitor. Compounds 187191 showed weak antibiofilm activity against Shewanella onedensis MR-1 biofilm [116]. In 2017, strain SCSGAA 0027 was further investigated which led to the isolation of five new spirocyclic polyketides pteridic acids C–G (S87, S88, 192194). Compounds E-G (192194) (Figure 8) had weak antibacterial activity against B. subtilis [117].
The Streptomyces coelicoflavus strain HQA809, which is isolated from sea squirt Styela clava, produced germicidin (195) and 6-isopropyl group-3-ethyl-4-hydroxy-2-pyrone (196) (Figure 8). Both two compounds were lethal to Artemia salina [64].
The Streptomyces sp. PTY087I2 associated with styela canopus collected from Bastimentos Park, Bocas del Toro, Panama produced three naphthoquinone derivatives granaticin (197), granatomycin D (198), and dihydrogranaticin B (199) (Figure 8). Co-culture of Streptomyces sp. PTY087I2 with human pathogens such as Bacillus subtilis, MSSA, and MRSA, respectively, resulted in increased production of these three antibiotics. In addition, co-culture resulted in greatly enhanced biological activity against the above three Gram-positive human pathogens [64,118].

Polyketides Derived from the Sea Cucumber-Associated Actinomycetes

Five curvularin macrolides (200204) (Figure 9) were isolated from Pseudonocardia sp. HS7, cultivated from obtained from the cloacal aperture of the sea cucumber Holothuria moebii. Compound 202 is a new macrolide with a rare a-D-glucopyranose substituent. Compounds 200203, 204a, and 204c (the acyl products of 204) suppressed the proliferation of cancer cell lines, and metabolites 203 is the most active compound with IC50 values ranging from 0.59 to 3.39 mM. The 11-hydroxycurvularins 200 and 201 also showed antibacterial activity inhibiting the growth of Escherichia coli [119]. Compound 205 (Figure 9), which was isolated from cucumber-associated Streptomyces sp. G278 selectively inhibited Enterococcus faecalis (MIC: 256 μg/mL). And the known metabolite 301 (Figure 9) with antibacterial and antifungal properties was also discovered from strain G278 [82].

Polyketides Derived from the Actinomycetes Associated with Other Marine Invertebrates

Three known analogs X- 14881 E (206), ochromycinone (207), and X-14881 C (208), together with a new angucyclines saccharothrixmicine A (209) (Figure 10), were produced by Saccharothrix espanaensis An 113 obtained from a marine mollusk specimen (Anadara broughtoni), which was collected from Peter the Great Bay, Sea of Japan, Russia. Compounds 206208 showed significant activity against C. albicans, B. subtilis, E. faecium, S. aureus, and Xanthomonas sp. pv. badrii. And Compound 209 exhibited activity towards Candida albicans and Xanthomonas sp. pv. Badrii [120,121]. Macrolactins E (S89) and F (S90), together with gilvocarcins M (S91) and V (S92), have been isolated from an unidentified tunicate-associated Saccharopolyspora sp. SS081219 JE-28 [64,77].
Isolation of two novel 3,4,6-trisubstituted α-pyrone derivatives violapyrones H (210) and I (211) were reported together with known violapyrones B (212) and C (213) (Figure 10), from Streptomyces sp. 112CH148 associated with starfish Acanthaster planci collected at Chuuk, Federated States of Micronesia. Violapyrones (210213) showed growth inhibitory activity against cancer cell lines at concentrations less than 26.12 μg/mL. Wherein compound 210 showed the highest cytotoxic activity against the HCT-15 cell line with a GI50 value of 1.10 μg/mL. Additionally, violapyrones B and C were demonstrated to have antibacterial activities. Therefore, it may be noteworthy that each compound has structural similarities, but showed different activities. The results suggested that the length of the aliphatic side chain and the position of the methyl group affected the activity. Furthermore, violapyrones having an isomethyl group in the alkyl side chain showed better activity than others [122].
In 2015, Marta Pérez et al. reported the discovery of two novel polyhydroxyl macrolides PM100117 (214) and PM100118 (215) (Figure 10) from Streptomyces caniferus GUA-06-05-006A associated with marine polychaete Filograna sp. collected at Guadalupe Island in the Pacific Ocean. Both macrolides 214 and 215 displayed strong cytotoxicity against tumor cells and weak antifungal activity against Candida albicans ATCC10231 [123].
Two new compounds (216, 217) (Figure 10) and five known ones (S93S97) were produced by the sea anemone-derived Streptomyces sp. ZZ406. New compounds 216 and 217 showed potent activity in inhibiting the proliferation of different glioma cells and downregulating the expressions of glioma metabolic regulators. Compound 216 was active against the proliferation of different glioma cells with IC50 values of 4.7 to 8.1 μM, high selectivity index (>12.3 to 21.3), and good stability in human liver microsomes [83]. A novel flavonoid derivative flavoside A (S98) and the known angucyclinone PD116740 (126) were isolated from the sea urchin-derived Streptomyces sp. HDa1 [86].
Two julichrome monomers julichromes Q11 (218) and Q12 (219) (Figure 10), along with five known julichromes Q10, Q3·5, Q3·3, Q6·6, Q6 (220, S99, S100, 221, 222) were separated from the marine gastropod mollusk Batillaria zonalis-associated Streptomyces sampsonii SCSIO 054. In addition, four known anthraquinones chrysophanol (S101), 4-acetylchrysophanol (S102), islandicin (S103), and huanglongmycin A (S104) were also discovered. Julichrome Q12 (219) was found to display antibacterial activity against Micrococcus luteus and Bacillus subtilis with MICs of 2.0 and 8.0 μg/mL. What’s more, compounds 218, 220222 (Figure 10) also showed inhibitory activities against methicillin-resistant Staphylococcus aureus, S. aureus, and S. simulans AKA1 with MIC values ranging from 8 to 64 μg/mL [124].

Polyketides Derived from the Actinomycetes Associated with Marine Vertebrates

Further investigation for metabolites of this strain has led to the isolation of three additional novel cytotoxic metabolites designated as halichoblelide A–C (223225) (Figure 11). Halichoblelide A was active against the murine P-388 cell line (ED50 0.63 μg/mL), and all three novel metabolites exhibited significant cytotoxicity against the 39 human cancer cell lines [125,126].
Ochoa et al. reported the isolation of a new glycosylated polyketide phocoenamicin (226) (Figure 11) in 2017, from a Micromonospora strain obtained from five marine mammals. It showed potent activity against a broad panel of Gram-positive animals and plants such as the intestinal pathogen Clostridium difficile (MIC: 2.6 μM) [127].

3.1.3. Peptides

Most peptides from actinomycetes are circular and contain further rare structural elements, such as chromophore or unusual amino acids.

Peptides Derived from the Sponge-Associated Actinomycetes

Streptomyces sp. DA18 was separated from marine sponge Craniella australiensis collected at Sanya Island in the South China Sea, from which diketopiperazines (DKPs) (227, 228, 229, S105) (Figure 12) were discovered. Compounds 227 and 228 showed antimicrobial activity, of which 228 also had antifouling activity [128].
The cyclic depsipeptide valinomycin (230) (Figure 12) was isolated from Streptomyces sp. strains 22 and 34 associated with the sponges Aplysina aerophoba and Axinella polypoides, respectively. It exhibited significant inhibitory activities against the parasites Leishmania major (IC50 < 0.11 μM) and Trypanosoma brucei brucei (IC50 = 0.0032 μM). Additionally, it was active against 293T kidney epithelial cells (IC50 11.24 µM) and J774.1 macrophages (IC50 < 0.10 µM) [9,21].
A diketopiperazine (231) (Figure 12) with weak cytotoxic activity, known as a synthetic compound before, was first reported to be discovered from Nocardiopsis strain HB383 [92]. The new teleocidin analog JBIR-31 (232) (Figure 12) was purified from the obligate marine Streptomyces sp. NBRC 105896 separated from Haliclona sp. collected offshore Tateyama City (Chiba Prefecture, Japan). The compound displayed weak cytotoxic activity against human cervical carcinoma HeLa cells with an IC50 value of 49 μM [9,129].
In 2010, Motohashi et al. reported the discovery of two new modified indole-containing peptides JBIR-34 (233) and JBIR-35 (234) (Figure 12) from Streptomyces sp. Sp080513GE-23 associated with Haliclona sp. Compounds 233 and 234 exhibited DPPH radical scavenging activity with IC50 values of 1.0 and 2.5 mM [130].
Isolation of two new peptides JBIR-56 (S106) and JBIR-57 (S107) were reported in 2011 from the new isolate Streptomyces sp. SpD081030SC-03, which was obtained from an unidentified sponge collected in Ishigaki, Okinawa, Japan [9,131].
The potent cytotoxic thiodepsipeptide thiocoraline (235) (Figure 12), which was first isolated in 1997 from the mycelia of Micromonospora marina, and five new analogs of thiocoraline (236, S108, S109, 237, 238) were discovered in 2011 from Verrucosispora sp. strain WMMA107, which was separated from Chondrilla caribensis f. caribensis (Florida Keys, USA). 22′-Deoxythiocoraline (236), thiochondrilline C (237), and 12′ -sulfoxythiocoraline (238) (Figure 12) showed significant cytotoxicity against the A549 human cancer cell line with EC50 values of 0.13, 2.86 and 1.26 μM, respectively [9,132].
A sponge-derived Streptomyces sp. strain RV15 was reported to produce four new cyclic lipopeptides cyclodysidins A–D (S110S113) in 2012 [9,133]. Two novel cyclic peptides (239 and 240) (Figure 12), along with the previously reported nocardamine (S114), were isolated from Streptomyces M1087 associated with an unidentified sponge. The new compounds 239240 exhibited weak inhibition against the recombinant enzyme sortase B with EC50 values of 88.3 and 126.4 µg/mL [134].
Streptomyces sp. NIO 10068 derived from a sponge collected from the western coast of India produced linear dipeptides proline–glycine (S115) and N-amido-α- proline (S116) [135]. Kocurin (241) (Figure 12), a new member of the thiazolyl peptide family of antibiotics, was produced by Kocuria palustris. F-276,345 is associated with a marine sponge that was collected in Florida Keys. Kocurin showed potent activity against Gram-positive bacteria (MRSA) with MIC values of 0.25–0.5 μg/mL [136,137].
Xin Zhen et al. reported isolation and identification of two new metabolites quinomycin G (242) (Figure 12) and cyclo-(L-Pro-4-OH-L-Leu) (S117) in 2015, from sponge-associated Streptomyces sp. LS298. Quinomycin G (242) with a terminal double bond in one of the Ser groups exhibited moderate antibacterial activities against Staphylococcuse pidermidis, S. aureus, Enterococcus faecium, and E. faecalis with MIC values ranging from 16 to 64 μg/mL. In addition, it showed potent anti-tumor activities and the highest activity was observed against the Jurkat cell line (human T-cell leukemia) with an IC50 value of 0.414 μM [36].
A new cyclic dipeptide petrocidin A (243) (Figure 12) was discovered from sponge-derived Streptomyces sp. SBT348 exhibited significant cytotoxicity towards the human promyelocytic HL-60 and the human colon adenocarcinoma HT-29 cell lines [45]. New cyclic depsipeptide rakicidin F (244) (Figure 12) and known rakicidin C (S118) were isolated from Streptomyces sp. GKU 220 is associated with a marine sponge sample collected in Andaman sea, Ranong, Thailand. Rakicidin F (244) showed growth inhibitory activity against B. subtilis and E. coli at the dosage of 25 μg per disk [138].
Isolation of a novel antibacterial peptide actinokineosin (245) (Figure 12) was reported in 2016, from sponge-associated Actinokineospora spheciospongiae DSM45935T. Actinokineosin was active against Micrococcus luteus with an inhibition zone diameter of 8.0 mm at 50 μg/disk [139]. Four diketopiperazine compounds S119S122 were purified from the sponge-derived Streptomyces sp. G246 [56]. And three known compounds 246 (Figure 12) and S120, S123 were separated from marine sponge-associated Streptomyces sp. G248 [55].
Four new D-type actinomycin analogs actinomycins D1-D4 (247250) and actinomycin D (251) (Figure 12) were discovered from Streptomyces sp. LHW52447 associated with Phyllospongia foliascens obtained from the Xisha Islands in the South China Sea. Actinomycins D1 (247) and D2 (248) introduced an oxazole unit into the central phenoxazinone chromophore and exhibited more potent activities against three strains of MRSA with MIC values of 0.125–0.25 μg/mL than that of actinomycins D3–D4 (MIC = 0.5–1.0 μg/mL), which indicated that the incorporation of the oxazole unit would enhance the antibacterial activity. In addition, the cytotoxicity evaluation against human lung WI38 embryonal fibroblasts suggested that the incorporation of oxazole unit could decrease the cytotoxicity of actinomycins on human normal cells [140]. The SAR studies had indicated that amino acid substitutions in different positions of the peptides influenced biological potencies. And the oxidation level of the β-ring proline residue influenced both cytotoxic and antibacterial activity. The single most important structure-activity factor in the peptide moieties is, however, the integrity of the cyclic structure. In addition, the configuration of 4-hydroxyl group points has a strong influence on activity [141,142].
Streptomyces sp. Call-36 isolated from sponge Callyspongia sp. collected in the Red Sea was reported to produce a new diketopiperazine actinozine A (252), cyclo(2-OH-D-Pro-L-Leu) (253), cyclo(D-Pro-L-Phe) (254) (Figure 12), and cyclo(L-Pro-L-Phe) (S123). Compounds 252 and 253 displayed potent activity against S. aureus and were moderately active against C. albicans. Compound 254 exhibited moderate and selective activity against HCT-116 with an IC50 of 32.7 µM, while cyclo (L-Pro-L-Phe) (S123) was inactive, indicating that the D/L configuration of Pro had an important effect on the activity [143].
A novel anti-infective molecule nesfactin (255) (Figure 12) was isolated from sponge-associated Nesterenkonia sp. MSA31 and active against multidrug-resistant Pseudomonas aeruginosa by inhibiting the phenotypic expression of virulence factors [144]. Three antifungal cycle peptides Rhodopeptins C1, C2, and B5 (256258) (Figure 12) were obtained from Rhodococcus sp. UR59 is associated with marine sponge Callyspongia sp. [59].

Peptides Derived from the Coral-Associated Actinomycetes

Thiocoraline (235) is a thiodepsipeptide antitumor antibiotic isolated from Micromonospora sp. L-13-ACM2-092 is associated with a soft coral collected in the Indian Ocean off the coast of Mozambique. Thiocoraline had an inhibitory effect on DNA polymerase α. In addition, it displayed potent cytotoxicity and strong activity against Gram-positive bacteria [145,146].

3.1.4. Peptides Derived from the Ascidian-Associated Actinomycetes

In 2012, Wyche et al. reported the discovery of five new lipopeptide peptidolipins B–F (259, 260, S124S126) (Figure 13) from a strain of Nocardia sp., which was cultivated from Trididemnum orbiculatum. Peptidolipins B and E were active against MRSA and MSSA [64,147]. Four known diketopiperazines cyclo (6-OH-D-Pro-L-Phe) (229), bacillusamide B (261) (Figure 13), cyclo (L-Pro-L-Leu) (246), and cyclo (L-Pro-L-Ile) (262) (Figure 13), were separated from a colonial ascidian-associated Streptomyces sp. Did-27 and displayed cytotoxicity against cancer cell lines HCT-116, HepG2, and MCF-7 [64,72].

Peptides Derived from the Fish-Associated Actinomycetes

Isolation of three new 2,5-diketopiperazines (2, 5-DKPs) (263265) (Figure 14) was reported in 2016, from Streptomyces sp. MNU FJ-36 obtained from the intestinal fabric of Katsuwonus sp. All novel compounds displayed weak cytotoxicity against the A-549 cell line, and compounds 264 and 265 also exhibited weak inhibitory activity against HCT-116 cell line [148].

Peptides Derived from the Actinomycetes Associated with Other Marine Animals

Five cyclic depsipeptides with unusual structures salinamides A, B (266, 267) (Figure 15) and salinamides C–E (S127S129) were discovered from Streptomyces sp. CNB-091 obtained from the surface of jellyfish Cassiopeia xamachana collected in the Florida Keys. Salinamides A and B exhibited moderate antibiotic activity against Gram-positive bacteria. Additionally, the results of phorbol ester-induced mouse ear edema assay showed that salinamides A and B displayed significant topical anti-inflammatory activity using [149,150]. In 2014, salinamide F (268) (Figure 15), a new bicyclic depsipeptide, was also reported to be separated from this strain, possessing significant RNAP-inhibitory and antibacterial activity like salinamide A [151].
Three known diketopiperazines 269271 (Figure 15) produced by Saccharothrix espanaensis An 113 were active against Vibrio alginolyticus and Vibrio parahaemolyticus [121]. Thiocoraline (235) was separated from Micromonospora sp. ML1 isolated from a mollusk collected from the Indian Ocean coast of Mozambique [152]. Limazepines G (272) and H (273) (Figure 15) were discovered from Streptomyces seoulensis IFB-A01 cultivated from the gut of shrimp Penasus orientalis. Compounds 272 and 273 showed potent neuraminidase (NA) inhibitory activity with IC50 values of 7.50 and 7.37 μmol/L, respectively [153].
Mengxuan Chen et al. reported isolation and identification of a new compound S130 and the known valinomycin (230) in 2018, from sea anemone-associated Streptomyces sp. ZZ406. Valinomycin (230) was active against the proliferation of different glioma cells and downregulating the expressions of glioma metabolic regulators [83]. In addition, the valinomycin was also isolated from sponge-derived Streptomyces sp. strains 22 and 34 that have been mentioned above. Compound 274 (Figure 15) was discovered from the sea urchin-associated Streptomyces sp. HDa1 displayed weak activity against the Gram-negative bacteria Vibrio harveyi with an inhibition zone of 1.5 mm [85].

3.1.5. Terpenoids

The new diterpene JBIR-65 (275) (Figure 16) was separated from the Actinomadura sp. SpB081030SC-15 associated with an unidentified sponge collected offshore Ishigaki Island, Japan. JBIR-65 showed weak protection of neuronal hybridoma N18-RE-105 cells from L-glutamate toxicity with an EC50 value of 31 μM [9,154].
Five sesquiterpenes (276, 277, S131S133) were isolated from Streptomyces sp. (ZJG1) cultivated from stony corals collected in the South China Sea. Compound 277 (Figure 16) showed great scavenging activity, hemolytic activity, and acetylcholinesterase inhibitory activity. Compound 276 (Figure 16) exhibited well free radical scavenging and acetylcholinesterase inhibitory activity [155].
Isolation of two novel halimane-type diterpenoids micromonohalimanes A (S134) and B (278) (Figure 16) was reported in 2016 from Micromonospora sp. WMMC-218 separated from the ascidian Symplegma brakenhielmi. Micromonohalimane B showed moderate activity against methicillin-resistant Staphylococcus aureus [156].
Guanahanolide A, a meroterpenoid featured with an unprecedented sesterterpene skeleton was obtained from Streptomyces sp. RKBH-B7 is associated with octocoral Eunicea. Guanahanolide A (279) (Figure 16) showed moderate cytotoxicity against human cancer cell lines MCF-7, HTB-26, and HCT-116 [157].
Microeunicellols A (280) (Figure 16) and B (S135), two novel eunicellin diterpenoids, were reported in 2020 from the culture of Streptomyces albogriseolus SY67903 associated with the gorgonian Muricella sibogae collected at the South China Sea. Microeunicellol A exhibited cytotoxicities against several human cancer cell lines [158].

3.1.6. Steroids

3-keto sterols bendigoles D–F (281283) (Figure 17) were discovered from Actinomadura sp. SBMs009 is derived from the sponge Suberites japonicus. They showed anti-inflammatory activity based on NF-kB inhibition and glucocorticoid receptor–protein binding properties. Among them, bendigole F exhibited the highest activity against translocation of GFP-labeled NF-kB into the nucleus of hamster ovary CHO cells in vivo with an IC50 of 71 μM. The three sterols displayed activity against the glucocorticoid receptor translocation and bendigole D was the most potent. Bendigole D showed mild cytotoxicity against the L929 murine aneuploid fibrosarcoma with an IC50 of 30 μM [9,159].
Streptoseolactone (284) (Figure 17), a novel metabolite isolated from Streptomyces seoulensis IFB-A01, showed a significant inhibitory effect on NA in a dose-dependent manner, and the IC50 value was 3.92 μmol/L [153].
Manadoperoxide H (285) (Figure 17) and acanthosterol sulfate F (S136) were isolated from sponge-associated Streptomyces sp. RM66. Manadoperoxide H exhibited antitrypanosomal activity against Trypanosoma brucei rhodesiense with an IC50 value of 0.375 μg/mL [58].

3.1.7. Other Structure Classes

Lutoside (S137) was purified from Micrococcus luteus R-1588-10 associated with sponge Xestospongia sp. collected at Noumea, New Caledonia. Additionally, the previously known synthetic 2,4,4′-trichloro-2′-hydroxydiphenylether (286) (Figure 18) was also discovered from the strain R-1588-10, and it was active against Staphylococcus aureus, Vibrio anguillarum, and Candida albicans [9,160].
Isolation of a new cytotoxic metabolite (287) (Figure 18) characterized as S-methyl-2,4-dihydroxy-6-isopropyl-3,5-dimethylbenzothioate was reported in 2008 from Streptomyces sp. (LA3L2). Two known compounds montagnetol (S138) and erythrin (S139) were also isolated from Streptomyces sp. (LA3L2) that is the first reported actinomycete to produce these lichen-related compounds. In addition, chromomycin A2 (S140), chromomycin A3 (S141), and chromomycin 02-3D (S142) were reported to be separated from Streptomyces sp. (LA3L1) [76].
Three new angucyclinones saccharothrixins A (290), B (291), and C (292), together with two known analogs X-14881 A (288) and X-14881 B (289) (Figure 18), were produced by sponge-associated Saccharothrix espanaensis An 113. In addition, a new angucyclines saccharothrixmicine B (293) (Figure 18) was also reported from this strain in 2010, which displayed activity against Candida albicans and Xanthomonas sp. pv. badrii. Compound 288 exhibited moderate inhibitory activity against C. albicans and weak activity against B. subtilis and E. faecium. Compound 289 possessed significant activity against C. albicans, B. subtilis, E. faecium, S. aureus, and Xanthomonas sp. pv. badrii, whereas activities of compounds 290292 against B. subtilis, E. faecium, and Xanthomonas sp. pv. badrii were only modest [120,121].
Streptomyces sp. strain T03 derived from the sponge Tethya sp. led to the identification of butenolide (294) (Figure 18), which exhibited anti-trypanosoma activity against Trypanosoma brucei brucei (IC50 = 31.77 μM) [9,21]. Ecteinamycin (295) (Figure 18) was obtained from Actinomadura sp associated with a sea squirt Ecteinascidia turbinate. It showed potent activity against Clostridium difficile NAP1/B1/027 [64,115].
New nucleoside derivative (S143) and two known compounds S144 and S145 were isolated from Streptomyces microflavus strain No. HVG29 is associated with the marine sponge Hymeniacidon perlevis collected from the coast of Dalian (China). Compounds S143S145 were the first time to isolate deoxyuridine structures from S. microflavus associated with sponges [161].
Streptomyces sp. NIO 10068 derived from a marine sponge produced cinnamic acid (296) (Figure 18), which was active against QS (quorum sensing)-mediated virulence factors in Pseudomonas aeruginosa. Cinnamic acid was proven to possess previously undescribed QS antagonist properties. In addition, it is demonstrated to display bactericidal activity in the present study [135].
In 2014, Ellis et al. reported isolation and identification of two novel trialkyl-substituted aromatic acids, solwaric acids A and B (297, 298) (Figure 18), together with the known 2,4,6-triphenyl-1-hexene (S146) from Solwaraspora sp. WMMB329 is associated with ascidian Trididemnum orbiculatum. The two novel compounds demonstrated antibacterial activity against MRSA and MSSA [64,162].
Four unusual glycoglycerolipids (299, S147S149) and diphosphatidylglycerol (S150) were produced by Microbacterium sp. HP2 associated with sponge Halichondria panacea collected at the Adriatic coast, Rovinj, Croatia. The major compound 299 (Figure 18) displayed antitumor activity by inhibiting the growth of tumor cell lines HM02 and Hep G2 with GI50 values of 0.38 and 2.7 μg/mL, respectively [9,163].
Isolation of a new oxaphenalene derivative (S151) was reported in 2017 from Streptomyces griseorubens sp. ASMR4 is associated with an unidentified soft coral collected in the Red Sea at the Hurghada coast, Egypt. Additionally, along with metabolite S144, seven other known metabolites S152S158 were also discovered from the strain ASMR4 [164].
Phenylacetic acid methyl ester (S159) and tryptophan (S160) were produced by Rhodococcus sp. UA13 [43]. Two known compounds 1-hydroxy-2-naphthoic acid (S161) and 1,4-dihydroxy-2-naphthoic acid (S162) were isolated from the axenic culture of Saccharomonospora sp. UR22 [47]. In addition, L-tryptophan (S160) and L-phenylalanine (S163) were produced by the sponge-associated Streptomyces sp. G246 [56]. The known metabolite 300 (Figure 18) isolated from sea cucumber-derived Streptomyces sp. G278 exhibited antimicrobial activity [82]. GTRI-02 (S164) was produced by the sea anemone-derived Streptomyces sp. ZZ406 [83].
In 2019, Shaala et al. reported the discovery of two new nucleosides thymidine-3-mercaptocarbamic acid (S165) and thymidine-3-thioamine (S166) from sponge-associated Streptomyces sp. Call-36 [143]. An intriguing cage-like polyhemiketal nesteretal A (302) (Figure 18) was produced by Nesterenkonia halobia E5.1 obtained from a scleractinian coral Platygyra. Nesteretal A is a highly oxygenated compound featuring an unprecedented 5/5/5/5 tetracyclic scaffold. It showed a weak retinoid X receptor-α (RXRα) transcriptional activation effect [165].
A pair of geometrically isomeric unsaturated keto fatty acids (6E,8Z)/(6E, 8E)-5-oxo-6,8-tetradecadienoic acids (303, 304) (Figure 18) were identified from Micrococcus sp. associated with a stony coral Catalaphyllia sp. Compounds 303 and 304 with the unprecedented 2,4-dienone system both showed antibacterial activity against the plant pathogen Rhizobium radiobacter (preferred 303) and the fish pathogen Tenacibaculum maritimum (preferred 304). In addition, compounds 303 and 304 displayed agonistic activity against peroxisome proliferator-activated receptors (PPARs) with an isoform specificity towards PPARα and PPARγ [166].
Isolation of four metabolites ethyl plakortide Z (305) (Figure 18), seco-plakortide Z (S167), actinopolysporin B (S168), and acanthosterol G (S169) was reported in 2021 from sponge-associated Streptomyces sp. RM66. Peroxidessethyl plakortide Z was active against solid tumor and L-1210 leukemia cell lines in vitro [58].

3.2. Natural Products of the Actinomycetes Derived from Marine Plants, Macroalgae, Cyanobacteria, and Lichens

3.2.1. Alkaloids

Alkaloids Derived from the Brown Algae Associated Actinomycetes

Two new anti-inflammatory compounds lobophorins A (Figure 19) and B (306, 109) were produced by an unidentified actinomycete strain CNC-837 obtained from the surface of the Caribbean brown alga Lobophora variegate. The new compounds, distantly related to antibiotics of the kijanimicin class, are potent inhibitors of topical PMA-induced edema in the mouse ear assay when administered either topically or IP [167].
Streptomyces cyaneofuscatus M-27 and Streptomyces carnosus M-40 were associated with diverse intertidal marine brown macroalgae (Phyllum heterokontophyta, Fucus spiralis, and Cystoseira baccata) from the central Cantabrian Sea. Streptomyces cyaneofuscatus M-27 produced several antitumor antibiotics of the anthracycline family, of which two antibiotics were identified as daunomycin (307) and cosmomycin B (308) (Figure 19). And it also led to the isolation of an antifungal macrolactam maltophilin (309) (Figure 19). In addition, lobophorine B (310) (Figure 19) was separated from Streptomyces carnosus M-40 derived from macroalgae Cystoseira baccata with anti-inflammatory and antituberculosis properties [168].
Isolation of a new phenolic acid derivative 4-amino-6-methylsalicylic acid (311) (Figure 19) was reported in 2019 from Nocardiopsis sp. AS23C was obtained from Sargassum arnaudianum collected in the Red Sea at Hurghada coast, Egypt. The extract exhibited antibacterial activity against the Gram-positive Staphylococcus aureus, Bacillus subtilis ATCC6051, and Streptomyces viridochromogenes Tü 57 [169].

Alkaloids Derived from the Green Algae Associated Actinomycetes

A unique indolizinium alkaloid streptopertusacin A (312) (Figure 19) and a novel N-arylpyrazinone derivative streptoarylpyrazinone A (S170) were separated from the seaweed-associated Streptomyces sp. HZP-2216E obtained from sea lettuce Ulva pertusa, a traditional Chinese medicine. They are both existing as zwitterion and streptopertusacin A showed moderate activity against the growth of MRSA [170,171].
Three new compounds (S171S173), together with the known benzamides 2-acetamido-3-hydroxybenzamide (S174), 2-amino-3-hydroxybenzamide (S175), and 2-aminobenzamide (S176), were isolated from the Streptomyces sp. ZZ502 is associated with the seaweed Ulva conglobatea collected in the East China Sea [172].

Alkaloids Derived from the Actinomycetes Associated with Lichens

Isolation of a novel metabolite S177 was reported in 2013 from Streptomyces cavourensis YY01-17 separated from the lichens grown in the Antarctic area [173].

3.2.2. Polyketides

Polyketides Derived from the Brown Algae Associated Actinomycetes

Polyketide 313 (Figure 19) [2-hydroxy-5-((6-hydroxy-4-oxo-4H-pyran-2-yl) methyl)-2-propylchroman -4-one] was isolated as the major metabolite together with three related known phaechromycins B, C, and E (S178S180) from Streptomyces sundarbansensis strain WR1L1S8 associated with brown algae Fucus sp. collected in Bejaia coastline, Algeria. Compound 313 stood out as the most active of the series, showing a selective antibacterial activity against MRSA with a MIC of 6 μΜ [174].
Germicidins A (314) and B (315) (Figure 19), discovered from Streptomyces carnosus M-40, displayed a physiological role in spore germination and hypha elongation. Galtamycin B (S181) produced by Streptomyces cyaneofuscatus M-27 associated with Fucus spiralis was first isolated from Streptomyces or marine environments [168].
In 2018, Kim et al. reported isolation and identification of neomycin B (316) (Figure 19), a 28-membered macrolide containing 19 chiral centers, from Micromonospora sp. CNY-010 associated with Stypopodium zonale collected at the Bahamas Islands. Neaumycin B displayed potent cytotoxicity and showed significant efficacy and selectivity toward U87 human glioblastoma with an LD50 value of 5.6 × 10−5µg/mL [175].

Polyketides Derived from the Green Algae ASSOCIATED Actinomycetes

A new compound of 23-O-butyrylbafilomycin D(317), together with bafilomycin D(320), 9-hydroxybafilomycin D (321), and bafilomycin A1(322) (Figure 19), was isolated from seaweed-derived Streptomyces sp. HZP-2216E separate from traditional Chinese medicine sea lettuce Ulva pertusa. Later, two novel bafilomycins 21,22-en-bafilomycin D (318) and 21,22-en-9-hydroxybafilomycin D (319) (Figure 19), together with bafilomycin A2 (S182) were separated. Compounds 318 and 319 had potent activity against the proliferation of glioma U251 and C6 cells with IC50 values of 0.12–1.08 μM. In addition, they were active against MRSA with MIC values of 12.5 mg/mL. The four bafilomycins of compounds 320322 and 317 showed potent activity in suppressing the proliferation of the four tested glioma cell lines with IC50 values of 0.35 to 2.95 µM. In addition, compounds 320322 were reported to have antibacterial, antifungal, insecticidal, and herbicidal activities. Bafilomycins D and A1 (320, 322) also exhibited potent activity in inhibiting vacuolar-type ATPase [170,171].
Desertomycin G (323) (Figure 19) was separated from Streptomyces althioticus MSM3 associated with intertidal macroalgae Ulva sp. collected at Cantabrian Sea. Desertomycin G displayed potent antibiotic activities against several clinically relevant pathogens and moderate activity against relevant Gram-negative clinical pathogens. Additionally, it affects the viability of tumor cell lines, such as human breast adenocarcinoma (MCF-7) and colon carcinoma (DLD-1), but not normal mammary fibroblasts [176].

Polyketides Derived from the Red Algae-Associated Actinomycetes

Isolation of new α-pyrone polyketides zoumbericins A, B (S183, S184) and Germicidins K, L (S185, S186), together with six previously reported metabolites wailupemycin D (S187), wailupemycin E (S188), enterocin/vulgamycin (324), 5-deoxy-enterocin (325), germicidin A (314) and germicidin B (315) (Figure 19), was reported in 2017 from Streptomyces ambofaciens BI0048 separated from the red alga Laurencia glandulifera. Among them, enterocin showed herbicidal activity and weak antibacterial activity. 5-Deoxy-enterocin was reported to be active against S. lutea, S. aureus, Klebsiella pneumoniae, and Vibrio percolans. Germicidin A displayed weak activity against Streptomyces viridochromogenes and Streptomyces griseus [177].

Polyketides Derived from the Cyanobacteria-Associated Actinomycetes

Isolation of antibiotic Bisanthraquinones (326328) (Figure 19) was reported from Streptomyces sp. N1-78-1 associated with unicellular cyanobacteria parasitized on the tunic surface of Ecteinascidia turbinate collected in La Parguera, Puerto Rico. The metabolites 326 and 327 potently inhibited the growth of MRSA. All compounds were moderately active against HCT-116 human colon tumor cells [112].

Polyketides Derived from the Actinomycetes Associated with Marine Plants

In 2015, Yong-Fu Huang et al. reported the discovery of a new anthraquinone (329) (Figure 19) from Streptomyces sp. FX-58 is associated with Salicornia herbacea collected in Qingdao, Shandong province, China. Compound 329 showed an inhibitory effect on cancer cell HL-60, BCTC-823 and MDA-MB-435 with IC50 values of 6.83, 82.2, 56.59 mg/ mL, respectively [178].

3.2.3. Peptides

Two antifouling diketopiperazines bmDKP (330) and imDKP (331) (Figure 19) were isolated from Streptomyces praecox strain 291-11 separated from the rhizosphere of Undaria pinnatifida. The two compounds inhibited zoospores with a therapeutic ratio (LC50/EC50) of 17.7 and 21 and inhibited diatoms with a therapeutic ratio of 17.7 and 21, respectively [179].

3.2.4. Other Structure Classes

Two known compounds octadecanoic acid (S189) and cholest-4-en-3-one (S190) were discovered from Streptomyces sp. FX-58 is associated with marine plant Salicornia herbacea [178]. Glycoglycerolipids (332335) (Figure 20) produced by Streptomyces coelescens PK206-15 derived from brown algae Laminaria japonica rhizosphere were active against the following fouling animals and plants: zoospores of Ulva pertusa, the diatom Navicula annexa, the mussel Mytilus edulis and fouling bacteria with an EC50 range of 0.005–0.2 μg/mL [180].
Isolation of two antibacterial benzaldehydes 2-hydroxy-5-(3-methylbut-2-enyl) benzaldehyde (336) and 2-hepta-1,5-dienyl-3,6-dihydroxy-5-(3-methylbut-2-enyl) benzaldehyde (337) (Figure 20) were reported from Streptomyces atrovirens PK288-21 separated from the rhizosphere of the brown algae Undaria pinnatifida. The compound 336 is a new benzaldehyde derivative, and metabolite 337 was the first time reported in the genus Streptomyces. The two compounds were active against Edwardsiella tarda and Streptococcus iniae [181].
Shan-Shan Su et al. reported the discovery of a novel compound (S191) in 2013 together with a known compound (S192) from lichen-derived Streptomyces cavourensis YY01-17 [173]. And a novel bioactive antimicrobial kocumarin (338) (Figure 20) was produced by Kocuria marina CMG S2 associated with brown macroalga Pelvetia canaliculata attached to the rocks of Sonmiani Beach and demonstrated prominent and rapid growth inhibition against all tested fungi and pathogenic bacteria [182]. Three known compounds benzoic acid (S158), hydrocinnamic acid (S193), and (E)-cinnamic acid(S194) were separated from Streptomyces ambofaciens BI0048 associated with red alga Laurencia glandulifera [177].
Geosmin (S195), the compound responsible for the “earth smell” and beta-patchoulene (S196) used as a fragrance agent in the perfume industry, were two major metabolites of Streptomyces carnosus M-40 [168]. 5-methylresorcinol (339) and Linoleic acid (340) (Figure 20), which were isolated from Nocardiopsis sp. AS23C associated with brown alga Sargassum arnaudianum exhibited antibacterial activity against the Gram-positive Staphylococcus aureus, Bacillus subtilis ATCC6051, and Streptomyces viridochromogenes Tü 57 [169].

3.3. Data Analysis of the Secondary Metabolites from Actinomycetes Associated to Various Hosts

A total of 536 metabolites have been discovered from 155 actinomycetes associated with various marine hosts belonging to 22 genera. Among them, alkaloids (37%), polyketides (33%), and peptides (15%) comprise the largest proportion of secondary metabolites, while Streptomyces (68%), Micromonospora (6%), and Nocardiopsis (3%) are the dominant producers (Figure 21). Figure 22 showed the distribution of secondary metabolites currently identified in different host-related actinomycetes. The majority of the secondary metabolites were isolated from the actinomycetes associated with sponges (47%), ascidians (11%), and corals (9%), as well as brown algae (5%). Furthermore, the Sankey diagram and histogram were done to show the distribution of secondary metabolites produced by actinomycetes with various genera derived from different hosts, which is convenient for readers to have an overall understanding of the current secondary metabolites from marine organism-associated actinomycetes (Figure 23 and Figure 24).
Approximately 64% of the SMs displayed various biological activities, especially antimicrobial activity and cytotoxicity (Figure 25). Interestingly, some of these active metabolites with multiple biological properties deserve more attention (Table S7); for example, metabolites with cytotoxicity usually have antibacterial or antiparasitic activities, and some metabolites showed antibacterial activity can also act as enzyme inhibitors. This study advances the knowledge of these actinomycetes in respect to the metabolic potential of medicinal lead compounds.

3.4. Clinical Information of the Secondary Metabolites

Molecules with excellent activities, which have been in clinical applications or have entered clinical trials, were listed in Table 1. For example, rifamycin SV (9) is the earliest rifamycin antibiotic used in clinical application. Tetrodotoxin (105) has been widely used as an analgesic, sedative, antispasmodic, and local anesthetic in clinics. And daunomycin (307) has a good effect on acute myeloid leukemia. In addition, for these drugs already in clinical use, more clinical trials are underway for new diseases or new usages. (More detailed information can be found on the website ClinicalTrials.gov accessed on 25 October 2021).

4. Discussion

Metabolites from actinomycetes associated with marine organisms have proven to be an abundant source for the isolation of multiple potent bioactive metabolites with diverse structures. In this review, we attempt to discuss the significance of the special ecological status and genetic factors of these actinomycetes with multiple hosts. The chemical ecology underlying hosts–actinomycetes interactions provide a great opportunity for the discovery of novel drugs. During the co-evolution, these actinomycetes and their specific hosts constructed a coordinated and relatively independent micro-ecological environment, in which SMs can be tolerated by the host and are the active inhibiting specific external invasion. Therefore, actinomycetes associated with various marine hosts play an important ecological role in producing novel medicinal active compounds. Currently, there are relatively few studies on these actinomycetes, but many secondary metabolites have been isolated with excellent bioactivities. Some of these metabolites have been used in clinical applications or have entered clinical trials where they are expected to become new drugs. There is no doubt that further exploration can be a useful strategy for discovering novel marine natural products. These actinomycetes, however, are difficult to be cultured under experimental conditions. Therefore, in-depth exploration of the ecology of these actinomycetes to continuously optimize culture conditions is crucial for further research. Meanwhile, the use of advanced bioinformatics technology for gene detection of uncultured actinomycetes and heterologous expression of the collected biosynthetic gene clusters will be another important pathway for research on SMs of marine organism-associated actinomycetes.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/md19110629/s1, Figures S1–S19: structures of compounds S1S196, Figure S20: original rectangular (Neighbor-joining) tree. Table S1: actinomycetes and their accession number of Figure 1. Tables S2–S6: the data of Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6. Table S7: The summary of all secondary metabolites including information on separation sources, structural types, and biological activities. Table S8: Summarized repetitive compounds identified from multiple actinomycetes.

Author Contributions

J.C. wrote this manuscript; L.X. guided the figure drawing of the manuscript; Y.Z. guided the revision of the manuscript in the biology section; B.H. conceived the idea for this article and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Key Research and Development Program of Zhejiang Province, China (2021C02062), the National Natural Science Foundation of China (81973233) and the Scientific Research Foundation of Zhejiang Sci-Tech University (17042058-Y), the Special Program for Research and Development of Zhejiang Sci-Tech University (2019Y009).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jose, P.A.; Maharshi, A.; Jha, B. Actinobacteria in natural products research: Progress and prospects. Microbiol. Res. 2021, 246, 126708. [Google Scholar] [CrossRef]
  2. Subramani, R.; Aalbersberg, W. Marine actinomycetes: An ongoing source of novel bioactive metabolites. Microbiol. Res. 2012, 167, 571–580. [Google Scholar] [CrossRef]
  3. Vicente, J.; Stewart, A.; Song, B.; Hill, R.T.; Wright, J.L. Biodiversity of Actinomycetes Associated with Caribbean Sponges and Their Potential for Natural Product Discovery. Mar. Biotechnol. 2013, 15, 413–424. [Google Scholar] [CrossRef]
  4. Jagannathan, S.; Manemann, E.; Rowe, S.; Callender, M.; Soto, W. Marine Actinomycetes, New Sources of Biotechnological Products. Mar. Drugs 2021, 19, 365. [Google Scholar] [CrossRef]
  5. Liu, L.; Zheng, Y.-Y.; Shao, C.-L.; Wang, C.-Y. Metabolites from marine invertebrates and their symbiotic microorganisms: Molecular diversity discovery, mining, and application. Mar. Life Sci. Technol. 2019, 1, 60–94. [Google Scholar] [CrossRef] [Green Version]
  6. Jan, R.V. Biodiversity of Actinomycetes Associated with Caribbean Sponges of Puerto Rico, and Their Metabolic Profiles.; University of North Carolina Wilmington: Wilmington, NC, USA, 2010. [Google Scholar]
  7. Mary, T.R.J.; Kannan, R.R.; Iniyan, A.M.; Ramachandran, D.; Vincent, S.G.P. Cell wall distraction and biofilm inhibition of marine Streptomyces derived angucycline in methicillin resistant Staphylococcus aureus. Microb. Pathog. 2021, 150, 104712. [Google Scholar] [CrossRef]
  8. Valliappan, K.; Sun, W.; Li, Z. Marine actinobacteria associated with marine organisms and their potentials in producing pharmaceutical natural products. Appl. Microbiol. Biotechnol. 2014, 98, 7365–7377. [Google Scholar] [CrossRef]
  9. Abdelmohsen, U.R.; Bayer, K.; Hentschel, U. Diversity, abundance and natural products of marine sponge-associated actinomycetes. Nat. Prod. Rep. 2014, 31, 381–399. [Google Scholar] [CrossRef]
  10. Hernández, L.M.C.; Blanco, J.A.D.L.F.; Baz, J.P.; Puentes, J.L.F.; Millán, F.R.; Vázquez, F.E.; Fernández-Chimeno, R.I.; Grávalos, D.G. 4’-N-Methyl-5’-hydroxystaurosporine and 5’-Hydroxystaurosporine, New Indolocarbazole Alkaloids from a Marine Micromonospora sp. Strain. J. Antibiot. 2000, 53, 895–902. [Google Scholar] [CrossRef] [Green Version]
  11. Han, Z.; Pantazis, P.; Lange, T.S.; Wyche, J.H.; Hendrickson, E. The staurosporine analog, Ro-31-8220, induces apoptosis independently of its ability to inhibit protein kinase C. Cell Death Differ. 2000, 7, 521–530. [Google Scholar] [CrossRef] [Green Version]
  12. Schupp, P.; Steube, K.; Meyer, C.; Proksch, P.J.C.L. Anti-proliferative effects of new staurosporine derivatives isolated from a marine ascidian and its predatory flatworm. Cancer Lett. 2001, 174, 165–172. [Google Scholar] [CrossRef]
  13. Liu, R.; Cui, C.B.; Duan, L.; Gu, Q.Q.; Zhu, W.M. Potent in vitro anticancer activity of metacycloprodigiosin and undecylprodigiosin from a sponge-derived actinomycete Saccharopolyspora sp. Nov. Arch. Pharm. Res. 2005, 28, 1341–1344. [Google Scholar] [CrossRef]
  14. Kim, T.K.; Hewavitharana, A.K.; Shaw, N.; Fuerst, J.A. Discovery of a New Source of Rifamycin Antibiotics in Marine Sponge Actinobacteria by Phylogenetic Prediction. Appl. Environ. Microbiol. 2006, 72, 2118–2125. [Google Scholar] [CrossRef] [Green Version]
  15. DiCioccio, R.A.; Srivastava, B.I. Structure-activity relationships and specificity of inhibition of DNA polymerases from normal and leukemia cells of man and from simian sarcoma virus by rifamycin derivatives. J. Natl. Cancer Inst. 1978, 61, 1187–1194. [Google Scholar] [CrossRef]
  16. Luo, Z.M.; Xia, X.D.; Cui, Y.B.; Jiang, X.L.; Chao, M.; Cao, S.H. An update on progress in candidates and structure-activity relationships of the rifamycins. Chin. J. Antibiot. 2012, 37, 308–319. [Google Scholar]
  17. Imamura, N.; Nishijima, M.; Adachi, K.; Sano, H. Novel antimycin antibiotics, urauchimycins A and B, produced by marine actinomycete. J. Antibiot. 1993, 46, 241–246. [Google Scholar] [CrossRef] [Green Version]
  18. Mitova, M.I.; Lang, G.; Wiese, J.; Imhoff, J.F. Subinhibitory Concentrations of Antibiotics Induce Phenazine Production in a Marine Streptomyces sp. J. Nat. Prod. 2008, 71, 824–827. [Google Scholar] [CrossRef]
  19. Bunbamrung, N.; Dramae, A.; Srichomthong, K.; Supothina, S.; Pittayakhajonwut, P. Streptophenazines I–L from Streptomyces sp. BCC21835. Phytochem. Lett. 2014, 10, 91–94. [Google Scholar] [CrossRef]
  20. Pimentel-Elardo, S.M.; Gulder, T.; Hentschel, U.; Bringmann, G. Cebulactams A1 and A2, new macrolactams isolated from Saccharopolyspora cebuensis, the first obligate marine strain of the genus Saccharopolyspora. Cheminform 2009, 49, 6889–6892. [Google Scholar] [CrossRef]
  21. Pimentel-Elardo, S.M.; Kozytska, S.; Bugni, T.S.; Ireland, C.M.; Moll, H.; Hentschel, U. Anti-Parasitic Compounds from Streptomyces sp. Strains Isolated from Mediterranean Sponges. Mar. Drugs 2010, 8, 373–380. [Google Scholar] [CrossRef]
  22. Motohashi, K.; Takagi, M.; Shin-Ya, K. Tetracenoquinocin and 5-iminoaranciamycin from a sponge-derived Streptomyces sp. Sp080513GE-26. J. Nat. Prod. 2010, 73, 755–758. [Google Scholar] [CrossRef]
  23. Izumikawa, M.; Khan, S.; Takagi, M.; Shin-Ya, K. Sponge-Derived Streptomyces Producing Isoprenoids via the Mevalonate Pathway. J. Nat. Prod. 2010, 73, 208–212. [Google Scholar] [CrossRef]
  24. Khan, S.; Izumikawa, M.; Motohashi, K.; Mukai, A.; Takagi, M.; Shin-Ya, K. Distribution of the 3-hydroxyl-3-methylglutaryl coenzyme A reductase gene and isoprenoid production in marine-derived Actinobacteria. FEMS Microbiol. Lett. 2010, 304, 89–96. [Google Scholar] [CrossRef]
  25. Ueda, J.Y.; Khan, S.T.; Takagi, M.; Shin-ya, K. JBIR-58, a new salicylamide derivative, isolated from a marine sponge-derived Streptomyces sp. SpD081030ME-02. J. Antibiot. Tokyo 2010, 63, 267–269. [Google Scholar] [CrossRef]
  26. Huang, X.-L.; Gao, Y.; Xue, D.-Q.; Liu, H.-L.; Peng, C.-S.; Zhang, F.-L.; Li, Z.-Y.; Guo, Y.-W. Streptomycindole, an Indole Alkaloid from a Marine Streptomyces sp. DA22 Associated with South China Sea Sponge Craniella australiensis. Helvetica Chim. Acta 2011, 94, 1838–1842. [Google Scholar] [CrossRef]
  27. Wei, R.-B.; Xi, T.; Li, J.; Wang, P.; Li, F.-C.; Lin, Y.-C.; Qin, S. Lobophorin C and D, New Kijanimicin Derivatives from a Marine Sponge-Associated Actinomycetal Strain AZS17. Mar. Drugs 2011, 9, 359–368. [Google Scholar] [CrossRef]
  28. Abdelmohsen, U.R.; Szesny, M.; Othman, E.M.; Schirmeister, T.; Grond, S.; Stopper, H.; Hentschel, U. Antioxidant and Anti-Protease Activities of Diazepinomicin from the Sponge-Associated Micromonospora Strain RV115. Mar. Drugs 2012, 10, 2208–2221. [Google Scholar] [CrossRef]
  29. Nakayama, O.; Yagi, M.; Tanaka, M.; Kiyoto, S.; Okuhara, M.; Kohsaka, M. WS-9659 A and B, novel testosterone 5.ALPHA.-reductase inhibitors isolated from a Streptomyces. I. Taxonomy, fermentation, isolation, physico-chemical characteristics. J. Antibiot. 1989, 42, 1221–1229. [Google Scholar] [CrossRef] [Green Version]
  30. Hosoya, T.; Hirokawa, T.; Takagi, M.; Shin-ya, K. Trichostatin analogues JBIR-109, JBIR-110, and JBIR-111 from the marine sponge-derived Streptomyces sp. RM72. J. Nat. Prod. 2012, 75, 285–289. [Google Scholar] [CrossRef]
  31. Izumikawa, M.; Kawahara, T.; Hwang, J.-H.; Takagi, M.; Shin-Ya, K. JBIR-107, a New Metabolite from the Marine-Sponge-Derived Actinomycete, Streptomyces tateyamensisNBRC 105047. Biosci. Biotechnol. Biochem. 2013, 77, 663–665. [Google Scholar] [CrossRef]
  32. Sun, D.; Sun, W.; Yu, Y.; Li, Z.; Deng, Z.; Lin, S. A new glutarimide derivative from marine sponge-derived Streptomyces anulatus S71. Nat. Prod. Res. 2014, 28, 1602–1606. [Google Scholar] [CrossRef]
  33. Singh, S.; Prasad, P.; Subramani, R.; Aalbersberg, W. Production and purification of a bioactive substance against multi-drug resistant human pathogens from the marine-sponge-derived Salinispora sp. Asian Pac. J. Trop. Biomed. 2014, 4, 825–831. [Google Scholar] [CrossRef] [Green Version]
  34. Vicente, J.; Stewart, A.K.; Van Wagoner, R.M.; Elliott, E.; Bourdelais, A.J.; Wright, 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] [Green Version]
  35. Reimer, A.; Blohm, A.; Quack, T.; Grevelding, C.G.; Kozjak-Pavlovic, V.; Rudel, T.; Hentschel, U.; Abdelmohsen, U.R. Inhibitory activities of the marine streptomycete-derived compound SF2446A2 against Chlamydia trachomatis and Schistosoma mansoni. J. Antibiot. 2015, 68, 674–679. [Google Scholar] [CrossRef] [Green Version]
  36. Zhen, X.; Gong, T.; Liu, F.; Zhang, P.-C.; Zhou, W.-Q.; Li, Y.; Zhu, P. A New Analogue of Echinomycin and a New Cyclic Dipeptide from a Marine-Derived Streptomyces sp. LS298. Mar. Drugs 2015, 13, 6947–6961. [Google Scholar] [CrossRef] [Green Version]
  37. Espinosa, A.; Socha, A.M.; Ryke, E.; Rowley, D.C. Antiamoebic properties of the actinomycete metabolites echinomycin A and tirandamycin A. Parasitol. Res. 2012, 111, 2473–2477. [Google Scholar] [CrossRef]
  38. Yu, Z.; Vodanovic-Jankovic, S.; Ledeboer, N.; Huang, S.-X.; Rajski, S.R.; Kron, M.; Shen, B. Tirandamycins from Streptomyces sp. 17944 Inhibiting the Parasite Brugia malayi Asparagine tRNA Synthetase. Org. Lett. 2011, 13, 2034–2037. [Google Scholar] [CrossRef] [Green Version]
  39. Tamaoki, T.; Nomoto, H.; Takahashi, I.; Kato, Y.; Morimoto, M.; Tomita, F. Staurosporine, a potent inhibitor of phospholipidCa++dependent protein kinase. Biochem. Biophys. Res. Commun. 1986, 135, 397–402. [Google Scholar] [CrossRef]
  40. Cheng, C.; Othman, E.M.; Fekete, A.; Krischke, M.; Stopper, H.; Edrada-Ebel, U.; Mueller, M.J.; Hentschel, U.; Abdelmohsen, U.R. Strepoxazine A, a new cytotoxic phenoxazin from the marine sponge-derived bacterium Streptomyces sp. SBT345. Tetrahedron Lett. 2016, 57, 4196–4199. [Google Scholar] [CrossRef] [Green Version]
  41. Cheng, C.; Othman, E.M.; Reimer, A.; Grüne, M.; Kozjak-Pavlovic, V.; Stopper, H.; Hentschel, U.; Abdelmohsen, U.R. Ageloline A, new antioxidant and antichlamydial quinolone from the marine sponge-derived bacterium Streptomyces sp SBT345. Tetrahedron Lett. 2016, 57, 2786–2789. [Google Scholar] [CrossRef]
  42. Sekurova, O.N.; Pérez-Victoria, I.; Martín, J.; Degnes, K.F.; Sletta, H.; Reyes, F.; Zotchev, S.B. New Deferoxamine Glycoconjugates Produced upon Overexpression of Pathway-Specific Regulatory Gene in the Marine Sponge-Derived Streptomyces albus PVA94-07. Molecules 2016, 21, 1131. [Google Scholar] [CrossRef] [Green Version]
  43. Elsayed, Y.; Refaat, J.; Abdelmohsen, U.R.; Ahmed, S.; Fouad, M.A. Rhodozepinone, a new antitrypanosomal azepino-diindole alkaloid from the marine sponge-derived bacterium Rhodococcus sp. UA13. Med. Chem. Res. 2017, 26, 2751–2760. [Google Scholar] [CrossRef]
  44. Li, J.L.; Chen, D.; Huang, L.; Ni, M.; Zhao, Y.; Fan, H.; Bao, X. Antichlamydial Dimeric Indole Derivatives from Marine Actinomycete Rubrobacter radiotolerans. Planta Med. 2017, 83, 805–811. [Google Scholar] [CrossRef]
  45. Cheng, C.; Othman, E.M.; Stopper, H.; Edrada-Ebel, R.; Hentschel, U.; Abdelmohsen, U.R. Isolation of Petrocidin A, a New Cytotoxic Cyclic Dipeptide from the Marine Sponge-Derived Bacterium Streptomyces sp. SBT348. Mar. Drugs 2017, 15, 383. [Google Scholar] [CrossRef] [Green Version]
  46. Che, Q.; Liang, Q.; Han, X.; Liu, Y.; Wang, W.; Gu, Q.; Zhu, T.; Li, D. Anthranosides A-C, Anthranilate Derivatives from a Sponge-Derived Streptomyces sp. CMN-62. Org. Lett. 2018, 20, 5466–5469. [Google Scholar] [CrossRef]
  47. El-Hawary, S.S.; Sayed, A.; Mohammed, R.; Khanfar, M.; Rateb, M.E.; Mohammed, T.A.; Hajjar, D.; Hassan, H.; Gulder, T.A.M.; Abdelmohsen, U.R. New Pim-1 Kinase Inhibitor From the Co-culture of Two Sponge-Associated Actinomycetes. Front. Chem. 2018, 6, 538. [Google Scholar] [CrossRef] [Green Version]
  48. Gong, T.; Zhen, X.; Li, X.-L.; Chen, J.-J.; Chen, T.-J.; Yang, J.-L.; Zhu, P. Tetrocarcin Q, a New Spirotetronate with a Unique Glycosyl Group from a Marine-Derived Actinomycete Micromonospora carbonacea LS276. Mar. Drugs 2018, 16, 74. [Google Scholar] [CrossRef] [Green Version]
  49. Huang, X.; Kong, F.; Zhou, S.; Huang, D.; Zheng, J.; Zhu, W. Streptomyces tirandamycinicus sp. nov., a Novel Marine Sponge-Derived Actinobacterium With Antibacterial Potential Against Streptococcus agalactiae. Front. Microbiol. 2019, 10, 482. [Google Scholar] [CrossRef] [Green Version]
  50. Chen, M.-H.; Lian, Y.-Y.; Fang, D.-S.; Chen, L.; Jia, J.; Zhang, W.-L.; Lin, R.; Xie, Y.; Bi, H.-K.; Jiang, H. Identification and antimicrobial properties of a new alkaloid produced by marine-derived Verrucosispora sp. FIM06-0036. Nat. Prod. Res. 2019, 1–7. [Google Scholar] [CrossRef]
  51. Chen, M.-H.; Zhang, W.-L.; Chen, L.; Lin, R.; Xie, Y.; Fang, D.-S.; Lian, Y.-Y.; Jiang, H. Isolation, purification and identification of two new alkaloids metabolites from marine-derived Verrucosispora sp. FIM06025. Nat. Prod. Res. 2018, 33, 2897–2903. [Google Scholar] [CrossRef]
  52. 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]
  53. Dhaneesha, M.; Hasin, O.; Sivakumar, K.C.; Ravinesh, R.; Naman, C.B.; Carmeli, S.; Sajeevan, T.P. DNA Binding and Molecular Dynamic Studies of Polycyclic Tetramate Macrolactams (PTM) with Potential Anticancer Activity Isolated from a Sponge-Associated Streptomyces zhaozhouensis subsp. mycale subsp. nov. Mar. Biotechnol. 2019, 21, 124–137. [Google Scholar] [CrossRef]
  54. Yu, M.; Li, Y.; Banakar, S.P.; Liu, L.; Shao, C.; Li, Z.; Wang, C. New Metabolites From the Co-culture of Marine-Derived Actinomycete Streptomyces rochei MB037 and Fungus Rhinocladiella similis 35. Front. Microbiol. 2019, 10, 915. [Google Scholar] [CrossRef] [Green Version]
  55. Cao, D.D.; Trinh, T.T.V.; Mai, H.D.T.; Vu, V.N.; Le, H.M.; Thi, Q.V.; Nguyen, M.A.; Duong, T.T.; Tran, D.T.; Chau, V.M.; et al. Antimicrobial lavandulylated flavonoids from a sponge-derived Streptomyces sp. G248 in east vietnam sea. Mar. Drugs 2019, 17, 529. [Google Scholar] [CrossRef] [Green Version]
  56. Cao, D.D.; Do, T.Q.; Doan Thi Mai, H.; Vu Thi, Q.; Nguyen, M.A.; Le Thi, H.M.; Tran, D.T.; Chau, V.M.; Cong Thung, D.; Pham, V.C. Antimicrobial lavandulylated flavonoids from a sponge-derived actinomycete. Nat. Prod. Res. 2020, 34, 413–420. [Google Scholar] [CrossRef]
  57. Back, C.; Stennett, H.; Williams, S.; Wang, L.; Gomez, J.O.; Abdulle, O.; Duffy, T.; Neal, C.; Mantell, J.; Jepson, M.; et al. A New Micromonospora Strain with Antibiotic Activity Isolated from the Microbiome of a Mid-Atlantic Deep-Sea Sponge. Mar. Drugs 2021, 19, 105. [Google Scholar] [CrossRef]
  58. Alkhalifah, D.H.M. Sponge-associated sp. RM66 metabolome induction with N-acetylglucosamine: Antibacterial, antifungal and anti-trypanosomal activities. Saudi J. Biol. Sci. 2021, 28, 4691–4698. [Google Scholar] [CrossRef]
  59. Alhadrami, H.; Thissera, B.; Hassan, M.; Behery, F.; Ngwa, C.; Hassan, H.; Pradel, G.; Abdelmohsen, U.; Rateb, M. 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]
  60. Peng, F.; Fandong, K.; Yuanfei, W. Antibiotic Metabolites from the Coral-Associated Actinomycete Streptomyces sp. OUCMDZ-1703. Chin. J. Chem. 2013, 31, 100–104. [Google Scholar]
  61. Braña, A.F.; Sarmiento-Vizcaíno, A.; Osset, M.; Pérez-Victoria, I.; Martín, J.; de Pedro, N.; de la Cruz, M.; Díaz, C.; Vicente, F.; Reyes, F.; et al. Lobophorin K, a New Natural Product with Cytotoxic Activity Produced by Streptomyces sp. M-207 Associated with the Deep-Sea Coral Lophelia pertusa. Mar. Drugs 2017, 15, 144. [Google Scholar] [CrossRef] [Green Version]
  62. Cong, Z.; Huang, X.; Liu, Y.; Liu, Y.; Wang, P.; Liao, S.; Yang, B.; Zhou, X.; Huang, D.; Wang, J. Cytotoxic anthracycline and antibacterial tirandamycin analogues from a marine-derived Streptomyces sp. SCSIO 41399. J. Antibiot. 2019, 72, 45–49. [Google Scholar] [CrossRef] [PubMed]
  63. Fang, Z.; Chen, S.; Zhu, Y.; Li, J.; Khan, I.; Zhang, Q.; Zhang, C. A new uridine derivative and a new indole derivative from the coral-associated actinomycete Pseudonocardia sp. SCSIO 11457. Nat. Prod. Res. 2021, 35, 188–194. [Google Scholar] [CrossRef]
  64. Chen, L.; Hu, J.-S.; Xu, J.-L.; Shao, C.-L.; Wang, G.-Y. Biological and Chemical Diversity of Ascidian-Associated Microorganisms. Mar. Drugs 2018, 16, 362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Janso, J.E.; Haltli, B.; Eustaquio, A.; Kulowski, K.; Waldman, A.J.; Zha, L.; Nakamura, H.; Bernan, V.S.; He, H.; Carter, G.T.; et al. Discovery of the lomaiviticin biosynthetic gene cluster in Salinispora pacifica. Tetrahedron 2014, 70, 4156–4164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. He, H.; Ding, W.D.; Bernan, V.S.; Richardson, A.D.; Ireland, C.M.; Greenstein, M.; Ellestad, G.A.; Carter, G.T. Lomaiviticins A and B, potent antitumor antibiotics from Micromonospora lomaivitiensis. J. Am. Chem. Soc. 2001, 123, 5362–5363. [Google Scholar] [CrossRef]
  67. Charan, R.D.; Schlingmann, G.; Janso, J.; Bernan, V.; Feng, X.; Carter, G.T. Diazepinomicin, a New Antimicrobial Alkaloid from a Marine Micromonospora sp. J. Nat. Prod. 2004, 67, 1431–1433. [Google Scholar] [CrossRef]
  68. Hayakawa, Y.; Shirasaki, S.; Shiba, S.; Kawasaki, T.; Matsuo, Y.; Adachi, K.; Shizuri, Y. Piericidins C7 and C8, New Cytotoxic Antibiotics Produced by a Marine Streptomyces sp. ChemInform 2007, 38, 39. [Google Scholar] [CrossRef]
  69. Asolkar, R.N.; Kirkland, T.N.; Jensen, P.R.; Fenical, W. Arenimycin, an antibiotic effective against rifampin- and methicillin-resistant Staphylococcus aureus from the marine actinomycete Salinispora arenicola. J. Antibiot. Tokyo 2010, 63, 37–39. [Google Scholar] [CrossRef] [Green Version]
  70. Kim, S.H.; Yang, H.O.; Sohn, Y.C.; Kwon, H.C. Aeromicrobium halocynthiae sp. nov., a taurocholic acid-producing bacterium isolated from the marine ascidian Halocynthia roretzi. Int. J. Syst. Evol. Microbiol. 2010, 60, 2793–2798. [Google Scholar] [CrossRef]
  71. Wyche, T.P.; Piotrowski, J.S.; Hou, Y.; Braun, D.; Deshpande, R.; Mcilwain, S.; Ong, I.M.; Myers, C.L.; Guzei, I.A.; Westler, W.M. Forazoline A: Marine-derived polyketide with antifungal in vivo efficacy. Angewandte Chemie 2014, 126, 11767–11770. [Google Scholar] [CrossRef]
  72. Shaala, L.A.; Youssef, D.T.A.; Badr, J.M.; Harakeh, S.M. Bioactive 2(1H)-Pyrazinones and Diketopiperazine Alkaloids from a Tunicate-Derived Actinomycete Streptomyces sp. Molecules 2016, 21, 1116. [Google Scholar] [CrossRef]
  73. Takahashi, C.; Takada, T.; Yamada, T.; Minoura, K.; Uchida, K.; Matsumura, E.; Numata, A. Halichomycin, a new class of potent cytotoxic macrolide produced by an actinomycete from a marine fish. Tetrahedron Lett. 1994, 35, 5013–5014. [Google Scholar] [CrossRef]
  74. López, J.M.S.; Insua, M.M.; Baz, J.P.; Puentes, J.L.F.; Hernández, L.M.C. New Cytotoxic Indolic Metabolites from a MarineStreptomyces. J. Nat. Prod. 2003, 66, 863–864. [Google Scholar] [CrossRef]
  75. Wu, Z.; Xie, L.; Xia, G.; Zhang, J.; Nie, Y.; Hu, J.; Wang, S.; Zhang, R. A new tetrodotoxin-producing actinomycete, Nocardiopsis dassonvillei, isolated from the ovaries of puffer fish Fugu rubripes. Toxicon 2005, 45, 851–859. [Google Scholar] [CrossRef]
  76. Mahyudin, N.A. Actinomycetes and Fungi Associated with Marine Invertebrates: A Potential Source of Bioactive Compounds. Ph.D. Dissertation, University of Canterbury Biological Sciences, Canterbury, UK, 2008. [Google Scholar]
  77. Takagi, M.; Motohashi, K.; Izumikawa, M.; Khan, S.T.; Hwang, J.H.; Shin-Ya, K. JBIR-66, a new metabolite isolated from tunicate-derived Saccharopolyspora sp. SS081219JE-28. Biosci. Biotechnol. Biochem. 2010, 74, 2355–2357. [Google Scholar] [CrossRef] [Green Version]
  78. Lin, Z.; Koch, M.; Pond, C.D.; Mabeza, G.; Seronay, R.A.; Concepcion, G.P.; Barrows, L.R.; Olivera, B.M.; Schmidt, E.W. Structure and activity of lobophorins from a turrid mollusk-associated Streptomyces sp. J. Antibiot. 2013, 67, 121–126. [Google Scholar] [CrossRef] [Green Version]
  79. Kawahara, T.; Itoh, M.; Izumikawa, M.; Kozone, I.; Sakata, N.; Tsuchida, T.; Shin-Ya, K. New hydroxamate metabolite, MBJ-0003, from Micromonospora sp. 29867. J. Antibiot. 2013, 67, 261–263. [Google Scholar] [CrossRef]
  80. Ambavane, V.; Tokdar, P.; Parab, R.; Sreekumar, E.S.; Mahajan, G.; Mishra, P.D.; D’Souza, L.; Ranadive, P. Caerulomycin A—An Antifungal Compound Isolated from Marine Actinomycetes. Adv. Microbiol. 2014, 04, 567–578. [Google Scholar] [CrossRef]
  81. Adnani, N.; Chevrette, M.G.; Adibhatla, S.N.; Zhang, F.; Yu, Q.; Braun, D.R.; Nelson, J.; Simpkins, S.W.; McDonald, B.R.; Myers, C.L.; et al. Coculture of Marine Invertebrate-Associated Bacteria and Interdisciplinary Technologies Enable Biosynthesis and Discovery of a New Antibiotic, Keyicin. ACS Chem. Biol. 2017, 12, 3093–3102. [Google Scholar] [CrossRef]
  82. Cao, D.T.; Tran, V.H.; Vu, V.N.; Mai, H.D.T.; Le, T.H.M.; Vu, T.Q.; Nguyen, H.H.; Chau, V.M.; Pham, V.C. Antimicrobial metabolites from a marine-derived Actinomycete Streptomyces sp. G278. Nat. Prod. Res. 2019, 33, 3223–3230. [Google Scholar] [CrossRef]
  83. Chen, M.; Chai, W.; Song, T.; Ma, M.; Lian, X.-Y.; Zhang, Z. Anti-glioma Natural Products Downregulating Tumor Glycolytic Enzymes from Marine Actinomycete Streptomyces sp. ZZ406. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef] [Green Version]
  84. Zhang, B.; Wang, K.B.; Wang, W.; Bi, S.F.; Mei, Y.N.; Deng, X.Z.; Jiao, R.H.; Tan, R.X.; Ge, H.M. Discovery, Biosynthesis, and Heterologous Production of Streptoseomycin, an Anti-Microaerophilic Bacteria Macrodilactone. Org. Lett. 2018, 20, 2967–2971. [Google Scholar] [CrossRef]
  85. Guo, Z.-K.; Wang, R.; Chen, S.-Q.; Chen, F.-X.; Liu, T.-M.; Yang, M.-Q. Anthocidins A⁻D, New 5-Hydroxyanthranilic Acid Related Metabolites from the Sea Urchin-Associated Actinobacterium, Streptomyces sp. HDa1. Molecules 2018, 23, 1032. [Google Scholar] [CrossRef] [Green Version]
  86. Guo, Z.-K.; Wang, R.; Liu, T.-M.; Chen, F.-X.; Yang, M.-Q. A new flavonoid derivative and a new 5-hydroxyanthranilic acid derivative from the sea urchin-derived Streptomyces sp. HDa1. J. Asian Nat. Prod. Res. 2019, 21, 992–998. [Google Scholar] [CrossRef]
  87. Zhou, Z.; Wu, Q.; Xie, Q.; Ling, C.; Zhang, H.; Sun, C.; Ju, J. New Borrelidins from Onchidium sp. Associated Streptomyces olivaceus SCSIO LO13. Chem. Biodivers. 2020, 17, e1900560. [Google Scholar] [CrossRef]
  88. Zhang, L.; Shi, J.; Liu, C.L.; Xiang, L.; Ma, S.Y.; Li, W.; Jiao, R.H.; Tan, R.X.; Ge, H.M. New borrelidin derivatives from an endophytic Streptomyces sp. Tetrahedron Lett. 2018, 59, 4517–4520. [Google Scholar] [CrossRef]
  89. Wilkinson, B.; Gregory, M.A.; Moss, S.J.; Carletti, I.; Sheridan, R.M.; Kaja, A.; Ward, M.; Olano, C.; Mendez, C.; Salas, J.A.; et al. Separation of anti-angiogenic and cytotoxic activities of borrelidin by modification at the C17 side chain. Bioorg.Med. Chem. Lett. 2006, 16, 5814–5817. [Google Scholar] [CrossRef]
  90. Cañedo, L.M.; Puentes, J.L.F.; Baz, J.P.; Huang, X.-H.; Rinehart, K.L. IB-96212, a Novel Cytotoxic Macrolide Produced by a Marine Micromonospora. II. Physico-chemical Properties and Structure Determination. J. Antibiot. 2000, 53, 479–483. [Google Scholar] [CrossRef] [Green Version]
  91. Perez, M.; Schleissner, C.; Rodriguez, P.; Zuniga, P.; Benedit, G.; Sanchez-Sancho, F.; de la Calle, F. PM070747, a new cytotoxic angucyclinone from the marine-derived Saccharopolyspora taberi PEM-06-F23-019B. J. Antibiot. Tokyo 2009, 62, 167–169. [Google Scholar] [CrossRef]
  92. Schneemann, I.; Ohlendorf, B.; Zinecker, H.; Nagel, K.; Wiese, J.; Imhoff, J.F. Nocapyrones A-D, gamma-pyrones from a Nocardiopsis strain isolated from the marine sponge Halichondria panicea. J. Nat. Prod. 2010, 73, 1444–1447. [Google Scholar] [CrossRef] [Green Version]
  93. Pimentel-Elardo, S.M.; Buback, V.; Gulder, T.A.; Bugni, T.S.; Reppart, J.; Bringmann, G.; Ireland, C.M.; Schirmeister, T.; Hentschel, U. New Tetromycin Derivatives with Anti-Trypanosomal and Protease Inhibitory Activities. Mar. Drugs 2011, 9, 1682–1697. [Google Scholar] [CrossRef]
  94. Wang, C.; Lu, Y.; Cao, S. Antimicrobial compounds from marine actinomycetes. Arch. Pharmacal Res. 2020, 43, 677–704. [Google Scholar] [CrossRef]
  95. 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]
  96. Soria-Mercado, I.E.; Prieto-Davo, A.; Jensen, A.P.R.; Fenical, W. Antibiotic Terpenoid Chloro-Dihydroquinones from a New Marine Actinomycete. J. Nat. Prod. 2005, 68, 904–910. [Google Scholar] [CrossRef]
  97. Bauermeister, A.; Pereira, F.; Grilo, I.; Godinho, C.C.; Paulino, M.; Almeida, V.; Gobbo-Neto, L.; Prieto-Davó, A.; Sobral, R.G.; Lopes, N.P.; et al. Intra-clade metabolomic profiling of MAR4 Streptomyces from the Macaronesia Atlantic region reveals a source of anti-biofilm metabolites. Environ. Microbiol. 2019, 21, 1099–1112. [Google Scholar] [CrossRef]
  98. Shomura, T.; Gomi, S.; Ito, M.; Yoshida, J.; Tanaka, E.; Amano, S.; Watabe, H.-O.; Ohuchi, S.; Itoh, J.; Sezaki, M.; et al. Studies on new antibiotics SF2415. I. Taxonomy, fermentation, isolation, physico-chemical properties and biological activities. J. Antibiot. 1987, 40, 732–739. [Google Scholar] [CrossRef] [Green Version]
  99. Kim, M.C.; Hwang, E.; Kim, T.; Ham, J.; Kim, S.Y.; Kwon, H.C. Nocatriones A and B, Photoprotective Tetracenediones from a Marine-Derived Nocardiopsis sp. J. Nat. Prod. 2014, 77, 2326–2330. [Google Scholar] [CrossRef]
  100. 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] [Green Version]
  101. Grkovic, T.; Abdelmohsen, U.R.; Othman, E.M.; Stopper, H.; Edrada-Ebel, R.; Hentschel, U.; Quinn, R.J. Two new antioxidant actinosporin analogues from the calcium alginate beads culture of sponge-associated Actinokineospora sp. strain EG49. Bioorganic Med. Chem. Lett. 2014, 24, 5089–5092. [Google Scholar] [CrossRef]
  102. Eltamany, E.E.; Abdelmohsen, U.R.; Ibrahim, A.K.; Hassanean, H.A.; Hentschel, U.; Ahmed, S.A. New antibacterial xanthone from the marine sponge-derived Micrococcus sp. EG45. Bioorg. Med. Chem. Lett. 2014, 24, 4939–4942. [Google Scholar] [CrossRef]
  103. Che, Q.; Tan, H.; Han, X.; Zhang, X.; Gu, Q.; Zhu, T.; Li, D. Naquihexcin A, a S-Bridged Pyranonaphthoquinone Dimer Bearing an Unsaturated Hexuronic Acid Moiety from a Sponge-Derived Streptomyces sp. HDN-10-293. Org. Lett. 2016, 18, 3358–3361. [Google Scholar] [CrossRef]
  104. Xu, D.; Nepal, K.K.; Chen, J.; Harmody, D.; Zhu, H.; McCarthy, P.J.; Wright, A.E.; Wang, G. 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]
  105. Tapiolas, D.M.; Roman, M.; Fenical, W.; Stout, T.J.; Clardy, J. Octalactins A and B: Cytotoxic eight-membered-ring lactones from a marine bacterium, Streptomyces sp. J. Am. Chem. Soc. 1991, 113, 4682–4683. [Google Scholar] [CrossRef]
  106. Sun, W.; Peng, C.; Zhao, Y.; Li, Z. Functional Gene-Guided Discovery of Type II Polyketides from Culturable Actinomycetes Associated with Soft Coral Scleronephthya sp. PLoS ONE 2012, 7, e42847. [Google Scholar] [CrossRef] [Green Version]
  107. Rodríguez, V.; Martín, J.; Sarmiento-Vizcaíno, A.; De La Cruz, M.; García, L.A.; Blanco, G.; Reyes, F. Anthracimycin B, a Potent Antibiotic against Gram-Positive Bacteria Isolated from Cultures of the Deep-Sea Actinomycete Streptomyces cyaneofuscatus M-169. Mar. Drugs 2018, 16, 406. [Google Scholar] [CrossRef] [Green Version]
  108. Ramalingam, V.; Mahamuni, D.; Rajaram, R. In vitro and in silico approaches of antibiofilm activity of 1-hydroxy-1-norresistomycin against human clinical pathogens. Microb. Pathog. 2019, 132, 343–354. [Google Scholar] [CrossRef]
  109. Zhang, Z.; Zhou, T.; Harunari, E.; Oku, N.; Igarashi, Y. Iseolides A–C, antifungal macrolides from a coral-derived actinomycete. J. Antibiot. 2020, 73, 534–541. [Google Scholar] [CrossRef]
  110. Kuznetsova, T.A.; Dmitrenok, A.S.; Sobolevskaya, M.P.; Shevchenko, L.S.; Mikhailov, V.V. Ubiquinone Q9 from a marine isolate of an actinobacterium Nocardia sp. Russ. Chem. Bull. 2002, 51, 1951–1953. [Google Scholar] [CrossRef]
  111. Li, A.; Piel, J. A Gene Cluster from a Marine Streptomyces Encoding the Biosynthesis of the Aromatic Spiroketal Polyketide Griseorhodin A. Chem. Biol. 2002, 9, 1017–1026. [Google Scholar] [CrossRef] [Green Version]
  112. Socha, A.M.; Garcia, D.; Sheffer, R.; Rowley, D.C. Antibiotic Bisanthraquinones Produced by a Streptomycete Isolated from a Cyanobacterium Associated with Ecteinascidia turbinata. J. Nat. Prod. 2006, 69, 1070–1073. [Google Scholar] [CrossRef]
  113. Sousa, T.D.S.; Jimenez, P.C.; Ferreira, E.G.; Silveira, E.R.; Filho, R.B.; Pessoa, O.D.L.; Costa-Lotufo, L.V. Anthracyclinones from Micromonospora sp. J. Nat. Prod. 2012, 75, 489–493. [Google Scholar] [CrossRef]
  114. Wyche, T.P.; Standiford, M.; Hou, Y.; Braun, D.; Johnson, D.A.; Johnson, J.A.; Bugni, T.S. Activation of the Nuclear Factor E2-Related Factor 2 Pathway by Novel Natural Products Halomadurones A–D and a Synthetic Analogue. Mar. Drugs 2013, 11, 5089–5099. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wyche, T.P.; Alvarenga, R.F.R.; Piotrowski, J.S.; Duster, M.N.; Warrack, S.R.; Cornilescu, G.; De Wolfe, T.J.; Hou, Y.; Braun, D.R.; Ellis, G.A.; et al. Chemical Genomics, Structure Elucidation, and in Vivo Studies of the Marine-Derived Anticlostridial Ecteinamycin. ACS Chem. Biol. 2017, 12, 2287–2295. [Google Scholar] [CrossRef] [Green Version]
  116. Nong, X.-H.; Zhang, X.-Y.; Xu, X.-Y.; Wang, J.; Qi, S.-H. Nahuoic Acids B–E, Polyhydroxy Polyketides from the Marine-Derived Streptomyces sp. SCSGAA 0027. J. Nat. Prod. 2015, 79, 141–148. [Google Scholar] [CrossRef]
  117. Nong, X.-H.; Wei, X.-Y.; Qi, S.-H. Pteridic acids C–G spirocyclic polyketides from the marine-derived Streptomyces sp. SCSGAA 0027. J. Antibiot. 2017, 70, 1047–1052. [Google Scholar] [CrossRef] [PubMed]
  118. Sung, A.A.; Gromek, S.M.; Balunas, M.J. Upregulation and Identification of Antibiotic Activity of a Marine-Derived Streptomyces sp. via Co-Cultures with Human Pathogens. Mar. Drugs 2017, 15, 250. [Google Scholar] [CrossRef] [Green Version]
  119. Ye, X.; Anjum, K.; Song, T.; Wang, W.; Yu, S.; Huang, H.; Lian, X.-Y.; Zhang, Z. A new curvularin glycoside and its cytotoxic and antibacterial analogues from marine actinomycete Pseudonocardia sp. HS7. HS7. Nat. Prod. Res. 2016, 30, 1156–1161. [Google Scholar] [CrossRef] [PubMed]
  120. 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, 10. [Google Scholar] [CrossRef] [Green Version]
  121. Kalinovskaya, N.I.; Kalinovsky, A.I.; Romanenko, L.A.; Dmitrenok, P.S.; Kuznetsova, T.A. New Angucyclines and Antimicrobial Diketopiperazines from the Marine Mollusk-Derived Actinomycete Saccharothrix espanaensis An 113. Nat. Prod. Commun. 2010, 5, 597–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Shin, H.J.; Lee, H.-S.; Lee, J.S.; Shin, J.; Lee, M.A.; Lee, H.-S.; Lee, Y.-J.; Yun, J.; Kang, J.S. Violapyrones H and I, New Cytotoxic Compounds Isolated from Streptomyces sp. Associated with the Marine Starfish Acanthaster planci. Mar. Drugs 2014, 12, 3283–3291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Pérez, M.; Schleissner, C.; Fernandez, R.; Rodriguez, P.; Reyes, F.; Zuniga, P.; De La Calle, F.; Cuevas, C. PM100117 and PM100118, new antitumor macrolides produced by a marine Streptomyces caniferus GUA-06-05-006A. J. Antibiot. 2015, 69, 388–394. [Google Scholar] [CrossRef]
  124. Dong, Y.; Ding, W.; Sun, C.; Ji, X.; Ling, C.; Zhou, Z.; Chen, Z.; Chen, X.; Ju, J. Julichrome Monomers from Marine Gastropod Mollusk-Associated Streptomyces and Stereochemical Revision of Julichromes Q 3 5 and Q 3 3. Chem. Biodivers. 2020, 17, e2000057. [Google Scholar] [CrossRef]
  125. Yamada, T.; Minoura, K.; Numata, A. Halichoblelide, a Potent Cytotoxic Macrolide from a Streptomyces Species Separated from a Marine Fish. Tetrahedron Lett. 2002, 43, 1721–1724. [Google Scholar] [CrossRef]
  126. Yamada, T.; Kikuchi, T.; Tanaka, R.; Numata, A. Halichoblelides B and C, potent cytotoxic macrolides from a Streptomyces species separated from a marine fish. Tetrahedron Lett. 2012, 53, 2842–2846. [Google Scholar] [CrossRef]
  127. Ochoa, J.L.; Sanchez, L.M.; Koo, B.-M.; Doherty, J.S.; Rajendram, M.; Huang, K.C.; Gross, C.A.; Linington, R.G. Marine Mammal Microbiota Yields Novel Antibiotic with Potent Activity Against Clostridium difficile. ACS Infect. Dis. 2018, 4, 59–67. [Google Scholar] [CrossRef]
  128. Gao, Y.; Yu, L.; Peng, C.; Li, Z.; Guo, Y. Diketopiperazines from two strains of South China Sea sponge-associated microorganisms. Biochem. Syst. Ecol. 2010, 38, 931–934. [Google Scholar] [CrossRef]
  129. Izumikawa, M.; Khan, S.; Komaki, H.; Takagi, M.; Shin-Ya, K. JBIR-31, a new teleocidin analog, produced by salt-requiring Streptomyces sp. NBRC 105896 isolated from a marine sponge. J. Antibiot. 2009, 63, 33–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Motohashi, K.; Takagi, M.; Shin-Ya, K. Tetrapeptides possessing a unique skeleton, JBIR-34 and JBIR-35, isolated from a sponge-derived actinomycete, Streptomyces sp. Sp080513GE-23. J. Nat. Prod. 2010, 73, 226–228. [Google Scholar] [CrossRef] [PubMed]
  131. Motohashi, K.; Inaba, K.; Fuse, S.; Doi, T.; Izumikawa, M.; Khan, S.T.; Takagi, M.; Takahashi, T.; Shin-ya, K. JBIR-56 and JBIR-57, 2(1H)-pyrazinones from a marine sponge-derived Streptomyces sp. SpD081030SC-03. J. Nat. Prod. 2011, 74, 1630–1635. [Google Scholar] [CrossRef] [PubMed]
  132. Wyche, T.P.; Hou, Y.; Braun, D.; Cohen, H.C.; Xiong, M.P.; Bugni, T.S. First Natural Analogs of the Cytotoxic Thiodepsipeptide Thiocoraline A from a Marine Verrucosispora sp. J. Org. Chem. 2011, 76, 6542–6547. [Google Scholar] [CrossRef] [Green Version]
  133. Abdelmohsen, U.R.; Zhang, G.; Philippe, A.; Schmitz, W.; Pimentel-Elardo, S.M.; Hertlein-Amslinger, B.; Hentschel, U.; Bringmann, G. Cyclodysidins A–D, cyclic lipopeptides from the marine sponge-derived Streptomyces strain RV15. Tetrahedron Lett. 2012, 53, 23–29. [Google Scholar] [CrossRef]
  134. Lee, H.-S.; Shin, H.J.; Jang, K.H.; Kim, T.S.; Oh, K.-B.; Shin, J. Cyclic Peptides of the Nocardamine Class from a Marine-Derived Bacterium of the Genus Streptomyces. J. Nat. Prod. 2005, 68, 623–625. [Google Scholar] [CrossRef] [PubMed]
  135. Naik, D.; Wahidullah, S.; Meena, R. Attenuation of Pseudomonas aeruginosa virulence by marine invertebrate-derived Streptomyces sp. Lett. Appl. Microbiol. 2013, 56, 197–207. [Google Scholar] [CrossRef] [PubMed]
  136. Palomo, S.; González, I.; De La Cruz, M.; Martínez, I.G.; Tormo, J.R.; Anderson, M.; Hill, R.; Vicente, F.; Reyes, F.; Genilloud, O. Sponge-Derived Kocuria and Micrococcus spp. as Sources of the New Thiazolyl Peptide Antibiotic Kocurin. Mar. Drugs 2013, 11, 1071–1086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Martin, J.; Sousa, T.D.S.; Crespo, G.; Palomo, S.; González, I.; Tormo, J.R.; De La Cruz, M.; Anderson, M.; Hill, R.T.; Vicente, F.; et al. Kocurin, the True Structure of PM181104, an Anti-Methicillin-Resistant Staphylococcus aureus (MRSA) Thiazolyl Peptide from the Marine-Derived Bacterium Kocuria palustris. Mar. Drugs 2013, 11, 387–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Kitani, S.; Ueguchi, T.; Igarashi, Y.; Leetanasaksakul, K.; Thamchaipenet, A.; Nihira, T. Rakicidin F, a new antibacterial cyclic depsipeptide from a marine sponge-derived Streptomyces sp. J. Antibiot. 2017, 71, 139–141. [Google Scholar] [CrossRef]
  139. Takasaka, N.; Kaweewan, I.; Ohnishi-Kameyama, M.; Kodani, S. Isolation of a new antibacterial peptide actinokineosin from Actinokineospora spheciospongiae based on genome mining. Lett. Appl. Microbiol. 2017, 64, 150–157. [Google Scholar] [CrossRef]
  140. Jiao, W.H.; Yuan, W.; Li, Z.Y.; Li, J.; Li, L.; Sun, J.B.; Gui, Y.H.; Wang, J.; Ye, B.P.; Lin, H.W. Anti-MRSA actinomycins D1-D4 from the marine sponge-associated Streptomyces sp. LHW52447—ScienceDirect. Tetrahedron 2018, 74, 5914–5919. [Google Scholar] [CrossRef]
  141. Bitzer, J.; Streibel, M.; Langer, H.-J.; Grond, S. First Y-type actinomycins from Streptomyces with divergent structure-activity relationships for antibacterial and cytotoxic properties. Org. Biomol. Chem. 2009, 7, 444–450. [Google Scholar] [CrossRef]
  142. Meienhofer, J.; Atherton, E. Structure—Activity Relationships in the Actinomycins. Int. Rev. Cytol. 1973, 16, 203–300. [Google Scholar] [CrossRef]
  143. Shaala, L.A.; Youssef, D.T.A.; Badr, J.M.; Harakeh, S.M.; Genta-Jouve, G. Bioactive Diketopiperazines and Nucleoside Derivatives from a Sponge-Derived Streptomyces Species. Mar. Drugs 2019, 17, 584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Kiran, G.S.; Sajayan, A.; Priyadharshini, G.; Balakrishnan, A.; Prathiviraj, R.; Sabu, A.; Selvin, J. A novel anti-infective molecule nesfactin identified from sponge associated bacteria Nesterenkonia sp. MSA31 against multidrug resistant Pseudomonas aeruginosa. Microb. Pathog. 2021, 157, 104923. [Google Scholar] [CrossRef]
  145. Boger, D.L.; Ichikawa, S. Total Syntheses of Thiocoraline and BE-22179: Establishment of Relative and Absolute Stereochemistry. J. Am. Chem. Soc. 2000, 122, 2956–2957. [Google Scholar] [CrossRef]
  146. Qi, S.; Gui, M.; Li, H.; Yu, C.; Li, H.; Zeng, Z.; Sun, P. Secondary Metabolites from Marine Micromonospora: Chemistry and Bioactivities. Chem. Biodivers. 2020, 17, 2000024. [Google Scholar] [CrossRef]
  147. Wyche, T.P.; Hou, Y.; Vazquez-Rivera, E.; Braun, D.; Bugni, T.S. Peptidolipins B–F, Antibacterial Lipopeptides from an Ascidian-Derived Nocardia sp. J. Nat. Prod. 2012, 75, 735–740. [Google Scholar] [CrossRef] [Green Version]
  148. Ou, Y.-X.; Huang, J.-F.; Li, X.-M.; Kang, Q.-J.; Pan, Y.-T. Three new 2,5-diketopiperazines from the fish intestinal Streptomyces sp. MNU FJ-36. Nat. Prod. Res. 2016, 30, 1771–1775. [Google Scholar] [CrossRef]
  149. Trischman, J.A.; Tapiolas, D.M.; Jensen, P.R.; Dwight, R.; Fenical, W.; Mckee, T.C.; Ireland, C.M.; Stout, T.J.; Clardy, J. Salinamides A and B: Anti-inflammatory depsipeptides from a marine streptomycete. Cheminform 1994, 25, 757–758. [Google Scholar] [CrossRef]
  150. Moore, B.S.; Trischman, J.A.; Seng, D.; Kho, D.; Jensen, P.R.; Fenical, W. Salinamides, Antiinflammatory Depsipeptides from a Marine Streptomycete. J. Org. Chem. 1999, 64, 1145–1150. [Google Scholar] [CrossRef]
  151. Hassan, H.M.; Degen, D.; Jang, K.H.; Ebright, R.H.; Fenical, W. Salinamide F, new depsipeptide antibiotic and inhibitor of bacterial RNA polymerase from a marine-derived Streptomyces sp. J. Antibiot. 2014, 68, 206–209. [Google Scholar] [CrossRef] [Green Version]
  152. Lombó, F.; Velasco, A.; Castro, A.; Calle, F.; Braña, A.F.; Sánchez-Puelles, J.M.; Méndez, C.; Salas, J.A. Deciphering the Biosynthesis Pathway of the Antitumor Thiocoraline from a Marine Actinomycete and Its Expression in Two Streptomyces Species. ChemBioChem 2006, 7, 366–376. [Google Scholar] [CrossRef]
  153. Jiao, R.; Xu, H.; Cui, J.; Ge, H.; Tan, R. Neuraminidase Inhibitors from marine-derived actinomycete Streptomyces seoulensis. J. Appl. Microbiol. 2013, 114, 1046–1053. [Google Scholar] [CrossRef]
  154. Takagi, M.; Motohashi, K.; Khan, S.T.; Hashimoto, J.; Shin-Ya, K. JBIR-65, a new diterpene, isolated from a sponge-derived Actinomadura sp. SpB081030SC-15. J. Antibiot. Tokyo 2010, 63, 401–403. [Google Scholar] [CrossRef] [Green Version]
  155. Wen, L.; Chen, G.; Zhang, S.; You, T.; Liu, F. In Vitro Antioxidant and Acetylcholinesterase Inhibitory Activities of the Sesquiterpenes of a Symbiotic Actinomycete Streptomyces sp. from South China Sea. Asian J. Chem. 2013, 25, 6865–6869. [Google Scholar] [CrossRef]
  156. Zhang, Y.; Adnani, N.; Braun, D.R.; Ellis, G.A.; Barns, K.J.; Parker-Nance, S.; Guzei, I.A.; Bugni, T.S. Micromonohalimanes A and B: Antibacterial Halimane-Type Diterpenoids from a Marine Micromonospora Species. J. Nat. Prod. 2016, 79, 2968–2972. [Google Scholar] [CrossRef] [Green Version]
  157. Marchbank, D.H.; Ptycia-Lamky, V.C.; Decken, A.; Haltli, B.A.; Kerr, R.G. Guanahanolide A, a Meroterpenoid with a Sesterterpene Skeleton from Coral-Derived Streptomyces sp. Org. Lett. 2020, 22, 6399–6403. [Google Scholar] [CrossRef] [PubMed]
  158. Ma, L.-F.; Chen, M.-J.; Liang, D.-E.; Shi, L.-M.; Ying, Y.-M.; Shan, W.-G.; Li, G.-Q.; Zhan, Z.-J. Streptomyces albogriseolus SY67903 Produces Eunicellin Diterpenoids Structurally Similar to Terpenes of the Gorgonian Muricella sibogae, the Bacterial Source. J. Nat. Prod. 2020, 83, 1641–1645. [Google Scholar] [CrossRef] [PubMed]
  159. Simmons, L.; Kaufmann, K.; Garcia, R.; Schwär, G.; Huch, V.; Müller, R. Bendigoles D–F, bioactive sterols from the marine sponge-derived Actinomadura sp. SBMs009. Bioorg Med. Chem. 2011, 19, 6570–6575. [Google Scholar] [CrossRef]
  160. Bultel-Poncé, V.; Debitus, C.; Bergé, J.; Cerceau, C.; Guyot, M. Metabolites from the sponge-associated bacterium Micrococcus luteus. J. Mar. Biotechnol. 1998, 6, 233. [Google Scholar]
  161. Li, K.; Li, Q.-L.; Ji, N.-Y.; Liu, B.; Zhang, W.; Cao, X.-P. Deoxyuridines from the Marine Sponge Associated Actinomycete Streptomyces microflavus. Mar. Drugs 2011, 9, 690–695. [Google Scholar] [CrossRef] [PubMed]
  162. Ellis, G.; Wyche, T.P.; Fry, C.G.; Braun, D.R.; Bugni, T.S. Solwaric Acids A and B, Antibacterial Aromatic Acids from a Marine Solwaraspora sp. Mar. Drugs 2014, 12, 1013–1022. [Google Scholar] [CrossRef] [Green Version]
  163. Wicke, C.; Hüners, M.; Wray, V.; Nimtz, M.; Bilitewski, U.; Lang, S. Production and Structure Elucidation of Glycoglycerolipids from a Marine Sponge-Associated Microbacterium Species. J. Nat. Prod. 2000, 63, 621–626. [Google Scholar] [CrossRef] [PubMed]
  164. Hamed, A.; Abdel-Razek, A.; Frese, M.; Wibberg, D.; El-Haddad, A.F.; Ibrahim, T.M.A.; Kalinowski, J.; Sewald, N.; Shaaban, M. New oxaphenalene derivative from marine-derived Streptomyces griseorubens sp. ASMR4. Z. Für. Nat. B 2017, 72, 53–62. [Google Scholar] [CrossRef]
  165. Xie, C.-L.; Chen, R.; Yang, S.; Xia, J.-M.; Zhang, G.-Y.; Chen, C.-H.; Zhang, Y.; Yang, X.-W. Nesteretal A, A Novel Class of Cage-Like Polyketide from Marine-Derived Actinomycete Nesterenkonia halobia. Org. Lett. 2019, 21, 8174–8177. [Google Scholar] [CrossRef]
  166. Sharma, A.R.; Harunari, E.; Oku, N.; Matsuura, N.; Trianto, A.; Igarashi, Y. Two antibacterial and PPARalpha/gamma-agonistic unsaturated keto fatty acids from a coral-associated actinomycete of the genus Micrococcus. Beilstein. J. Org. Chem. 2020, 16, 297–304. [Google Scholar] [CrossRef] [PubMed]
  167. Jiang, Z.D.; Jensen, P.R.; Fenical, W. Lobophorins A and B, new antiinflammatory macrolides produced by a tropical marine bacterium. Bioorg. Med. Chem. Lett. 1999, 9, 2003–2006. [Google Scholar] [CrossRef]
  168. Brana, A.F.; Fiedler, H.P.; Nava, H.; Gonzalez, V.; Sarmiento-Vizcaino, A.; Molina, A.; Acuna, J.L.; Garcia, L.A.; Blanco, G. Two Streptomyces species producing antibiotic, antitumor, and anti-inflammatory compounds are widespread among intertidal macroalgae and deep-sea coral reef invertebrates from the central Cantabrian Sea. Microb. Ecol. 2015, 69, 512–524. [Google Scholar] [CrossRef] [PubMed]
  169. Eliwa, E.M.; Abdel-Razek, A.S.; Frese, M.; Halawa, A.H.; El-Agrody, A.M.; Bedair, A.H.; Sewald, N.; Shaaban, M. New naturally occurring phenolic derivatives from marine Nocardiopsis sp. AS23C: Structural elucidation and in silico computational studies. Vietnam. J. Chem. 2019, 57, 164–174. [Google Scholar] [CrossRef] [Green Version]
  170. Zhang, X.; Chen, L.; Chai, W.; Lian, X.-Y.; Zhang, Z. A unique indolizinium alkaloid streptopertusacin A and bioactive bafilomycins from marine-derived Streptomyces sp. HZP-2216E. Phytochemistry 2017, 144, 119–126. [Google Scholar] [CrossRef] [PubMed]
  171. Zhang, Z.; Chen, L.; Zhang, X.; Liang, Y.; Anjum, K.; Chen, L.; Lian, X.-Y. Bioactive Bafilomycins and a New N-Arylpyrazinone Derivative from Marine-derived Streptomyces sp. HZP-2216E. Planta Med. 2017, 83, 1405–1411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Zhang, X.; Shu, C.; Li, Q.; Lian, X.-Y.; Zhang, Z. Novel cyclohexene and benzamide derivatives from marine-associated Streptomyces sp. ZZ502. Nat. Prod. Res. 2019, 33, 2151–2159. [Google Scholar] [CrossRef] [PubMed]
  173. Su, S.-S.; Tian, L.; Chen, G.; Li, Z.-Q.; Xu, W.-F.; Pei, Y.-H. Two new compounds from the metabolites of a marine-derived actinomycete Streptomyces cavourensis YY01-17. J. Asian Nat. Prod. Res. 2013, 15, 265–269. [Google Scholar] [CrossRef]
  174. Djinni, I.; Defant, A.; Kecha, M.; Mancini, I. Antibacterial Polyketides from the Marine Alga-Derived Endophitic Streptomyces sundarbansensis: A Study on Hydroxypyrone Tautomerism. Mar. Drugs 2013, 11, 124–135. [Google Scholar] [CrossRef] [Green Version]
  175. Kim, M.C.; Machado, H.; Jang, K.H.; Trzoss, L.; Jensen, P.R.; Fenical, W. Integration of Genomic Data with NMR Analysis Enables Assignment of the Full Stereostructure of Neaumycin B, a Potent Inhibitor of Glioblastoma from a Marine-Derived Micromonospora. J. Am. Chem. Soc. 2018, 140, 10775–10784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Braña, A.F.; Sarmiento-Vizcaíno, A.; Pérez-Victoria, I.; Martín, J.; Otero, L.; Palacios-Gutiérrez, J.-J.; Fernández, J.; Mohamedi, Y.; Fontanil, T.; Salmón, M.; et al. Desertomycin G, a New Antibiotic with Activity against Mycobacterium tuberculosis and Human Breast Tumor Cell Lines Produced by Streptomyces althioticus MSM3, Isolated from the Cantabrian Sea Intertidal Macroalgae Ulva sp. Mar. Drugs 2019, 17, 114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Rab, E.; Kekos, D.; Roussis, V.; Ioannou, E. α-Pyrone Polyketides from Streptomyces ambofaciens BI0048, an Endophytic Actinobacterial Strain Isolated from the Red Alga Laurencia glandulifera. Mar. Drugs 2017, 15, 389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Huang, Y.-F.; Tian, L.; Fu, H.-W.; Hua, H.-M.; Pei, Y.-H. One new anthraquinone from marine Streptomyces sp. FX-58. Nat. Prod. Res. 2006, 20, 1207–1210. [Google Scholar] [CrossRef]
  179. Cho, J.Y.; Kang, J.Y.; Hong, Y.K.; Baek, H.H.; Shin, H.W.; Kim, M.S. Isolation and Structural Determination of the Antifouling Diketopiperazines from Marine-Derived Streptomyces praecox 291-11. Biosci. Biotechnol. Biochem. 2012, 76, 1116–1121. [Google Scholar] [CrossRef] [Green Version]
  180. Cho, J.Y. Glycoglycerolipids Isolated from Marine Derived Streptomyces coelescens PK206-15. Biosci. Biotechnol. Biochem. 2012, 76, 1746–1751. [Google Scholar] [CrossRef] [Green Version]
  181. Ji, Y.C.; Kim, M.S. Antibacterial benzaldehydes produced by seaweed-derived Streptomyces atrovirens PK288-21. Fish. Sci. 2012, 78, 1065–1073. [Google Scholar]
  182. Uzair, B.; Menaa, F.; Khan, B.A.; Mohammad, F.V.; Ahmad, V.U.; Djeribi, R.; Menaa, B. Isolation, purification, structural elucidation and antimicrobial activities of kocumarin, a novel antibiotic isolated from actinobacterium Kocuria marina CMG S2 associated with the brown seaweed Pelvetia canaliculata. Microbiol. Res. 2018, 206, 186–197. [Google Scholar] [CrossRef]
Figure 1. Neighbor-Joining phylogenetic tree based on 16S rRNA gene sequences of natural product-producing actinomycetes. The aligned sequences were analyzed by the bootstrap method with a bootstrap number of 10,000. The colored labels indicate the classification of actinomycetes; the outer color strip shows their host sources.
Figure 1. Neighbor-Joining phylogenetic tree based on 16S rRNA gene sequences of natural product-producing actinomycetes. The aligned sequences were analyzed by the bootstrap method with a bootstrap number of 10,000. The colored labels indicate the classification of actinomycetes; the outer color strip shows their host sources.
Marinedrugs 19 00629 g001
Figure 2. Structures of compounds 178.
Figure 2. Structures of compounds 178.
Marinedrugs 19 00629 g002aMarinedrugs 19 00629 g002bMarinedrugs 19 00629 g002cMarinedrugs 19 00629 g002d
Figure 3. Structures of compounds 7984.
Figure 3. Structures of compounds 7984.
Marinedrugs 19 00629 g003
Figure 4. Structures of compounds 85100.
Figure 4. Structures of compounds 85100.
Marinedrugs 19 00629 g004
Figure 5. Structures of compounds 101123.
Figure 5. Structures of compounds 101123.
Marinedrugs 19 00629 g005aMarinedrugs 19 00629 g005bMarinedrugs 19 00629 g005c
Figure 6. Structures of compounds 124166.
Figure 6. Structures of compounds 124166.
Marinedrugs 19 00629 g006aMarinedrugs 19 00629 g006bMarinedrugs 19 00629 g006c
Figure 7. Structures of compounds 167176.
Figure 7. Structures of compounds 167176.
Marinedrugs 19 00629 g007
Figure 8. Structures of compounds 177199.
Figure 8. Structures of compounds 177199.
Marinedrugs 19 00629 g008aMarinedrugs 19 00629 g008b
Figure 9. Structures of compounds 200205, 301.
Figure 9. Structures of compounds 200205, 301.
Marinedrugs 19 00629 g009
Figure 10. Structures of compounds 206222.
Figure 10. Structures of compounds 206222.
Marinedrugs 19 00629 g010
Figure 11. Structures of compounds 223226.
Figure 11. Structures of compounds 223226.
Marinedrugs 19 00629 g011
Figure 12. Structures of compounds 227258.
Figure 12. Structures of compounds 227258.
Marinedrugs 19 00629 g012aMarinedrugs 19 00629 g012bMarinedrugs 19 00629 g012cMarinedrugs 19 00629 g012d
Figure 13. Structures of compounds 259262.
Figure 13. Structures of compounds 259262.
Marinedrugs 19 00629 g013
Figure 14. Structures of compounds 263265.
Figure 14. Structures of compounds 263265.
Marinedrugs 19 00629 g014
Figure 15. Structures of compounds 266274.
Figure 15. Structures of compounds 266274.
Marinedrugs 19 00629 g015aMarinedrugs 19 00629 g015b
Figure 16. Structures of compounds 275280.
Figure 16. Structures of compounds 275280.
Marinedrugs 19 00629 g016
Figure 17. Structures of compounds 281285.
Figure 17. Structures of compounds 281285.
Marinedrugs 19 00629 g017
Figure 18. Structures of compounds 286300, 302305.
Figure 18. Structures of compounds 286300, 302305.
Marinedrugs 19 00629 g018
Figure 19. Structures of compounds 306331.
Figure 19. Structures of compounds 306331.
Marinedrugs 19 00629 g019aMarinedrugs 19 00629 g019b
Figure 20. Structures of compounds 332340.
Figure 20. Structures of compounds 332340.
Marinedrugs 19 00629 g020
Figure 21. The structural distribution of metabolites from the actinomycetes is divided by genera.
Figure 21. The structural distribution of metabolites from the actinomycetes is divided by genera.
Marinedrugs 19 00629 g021
Figure 22. The structural distribution of metabolites from -actinomycetes associated with various hosts. The dots in that figure represent the number of compounds and the color darkens when the dots overlap.
Figure 22. The structural distribution of metabolites from -actinomycetes associated with various hosts. The dots in that figure represent the number of compounds and the color darkens when the dots overlap.
Marinedrugs 19 00629 g022
Figure 23. The distribution of secondary metabolites produced by actinomycetes with various genera derived from different hosts. The width of the extended branches in the figure corresponds to the number of secondary metabolites.
Figure 23. The distribution of secondary metabolites produced by actinomycetes with various genera derived from different hosts. The width of the extended branches in the figure corresponds to the number of secondary metabolites.
Marinedrugs 19 00629 g023
Figure 24. The structural distribution of metabolites from three dominant actinomycetes in the main host’s sponge and ascidian. The x-axis labels represent the host-genus-structure classes of actinomycetes: (A) Ascidian; (S) Sponge; (Str) Streptomyces; (Mic) Micromonospora; (Sac) Saccharopolyspora; (Act) Actinomadura; (al) alkaloid; (po) polyketide.
Figure 24. The structural distribution of metabolites from three dominant actinomycetes in the main host’s sponge and ascidian. The x-axis labels represent the host-genus-structure classes of actinomycetes: (A) Ascidian; (S) Sponge; (Str) Streptomyces; (Mic) Micromonospora; (Sac) Saccharopolyspora; (Act) Actinomadura; (al) alkaloid; (po) polyketide.
Marinedrugs 19 00629 g024
Figure 25. The diversity distribution of biological activity with different structures.
Figure 25. The diversity distribution of biological activity with different structures.
Marinedrugs 19 00629 g025
Table 1. Clinical information of the secondary metabolites.
Table 1. Clinical information of the secondary metabolites.
CompoundsStudy TitleConditionsRelated Compounds for InterventionsPhaseNCT Number
Rifamycin SV (9)Rifamycin SV-MMX® 400 mg b.i.d. vs. Rifamycin SV-MMX® 600 mg t.i.d. vs. Placebo in Acute Uncomplicated DiverticulitisUncomplicated DiverticulitisRifamycin SV-MMX® Phase 2NCT01847664
Rifamycin SV-MMX® 600 mg Tablets Administered Three or Two Times Daily to Patients With IBS-DDiarrhea-predominant Irritable Bowel SyndromeRifamycin SV Phase 2NCT03099785
Study to Evaluate Safety and Efficacy of Rifamycin SV Multi-Matrix System (MMX) for the Treatment of Traveler’s Diarrhea (TD)Traveler’s DiarrheaRifamycin SV MMXPhase 3NCT01142089
Rifamycin SV-MMX® Tablets Versus Ciprofloxacin Capsules in Acute Traveller’s DiarrhoeaTraveler’s DiarrheaRifamycin SV-MMX®Phase 3NCT01208922
Diazepinomicin (28)A Phase I Study of ECO-4601 in Patients With Advanced CancerTumors
Glioma
Colorectal Cancer
ECO-4601Phase 1NCT00338026
Efficacy Study of TLN-4601 in Patients With Recurring Glioblastoma MultiformeGlioblastoma MultiformeTLN-4601Phase 2NCT00730262
Staurosporine (3)A Phase I Trial of Continuous Infusion UCN-01 in Patients With Refractory NeoplasmsBreast Cancer
Lymphoma
Neoplasm
Prostatic Neoplasm
7-hydroxystaurosporine (UCN-01)Phase 1NCT00001444
PK and Safety of Midostaurin in Subjects With Impaired Hepatic Function and Subjects With Normal Hepatic FunctionHepatic ImpairmentMidostaurinPhase 1NCT01429337
Phase I Combination of Midostaurin, Bortezomib, and Chemo in Relapsed/Refractory Acute Myeloid LeukemiaAcute Myeloid Leukemia
AML With Multilineage Dysplasia Following
Myelodysplastic Syndrome
MidostaurinPhase 1NCT01174888
Azacitidine With or Without Nivolumab or Midostaurin, or Decitabine and Cytarabine Alone in Treating Older Patients With Newly Diagnosed Acute Myeloid Leukemia or High-Risk Myelodysplastic SyndromeAcute Myeloid Leukemia
Myelodysplastic Syndrome
Myelodysplastic Syndrome With Excess Blasts-2
MidostaurinPhase 2
Phase 3
NCT03092674
Tetrodotoxin (105)Tetrodotoxin Open-label Efficacy and Safety Continuation StudyPain
Cancer
TetrodotoxinPhase 3NCT00726011
Safety & Efficacy Study of Subcutaneous Tetrodotoxin for Moderate to Severe Inadequately Controlled Cancer-related PainPain
Cancer
TetrodotoxinPhase 3NCT00725114
Daunomycin (307)Pilot Study Efficacy and Tolerance Fish Oil Emulsion Daunorubicin and Cytarabine Treatment of AML Younger PatientsAcute Myeloid Leukemia (AML)DaunorubicinPhase 2NCT01999413
A Randomized Study of Gemtuzumab Ozogamicin (GO) With Daunorubicine and Cytarabine in Untreated Acute Myeloid Leukemia (AML) Aged of 50–70 Years OldAcute Myeloid LeukemiaDaunorubicinPhase 3NCT00927498
Linoleic acid (340)Proof of Principle Trial to Determine if Nutritional Supplement Conjugated Linoleic Acid (CLA) Can Modulate the Lipogenic Pathway in Breast Cancer TissueBreast CancerConjugated Linoleic Acid (CLA)Early Phase 1NCT00908791
Conjugated Linoleic Acid / Leucine Versus Metformin on Visceral Fat in Metabolic SyndromeMetabolic SyndromeConjugated linoleic acid/LeucinePhase 2NCT02629627
Conjugated Linoleic Acid and AtherosclerosisAtherosclerosisCis9, trans11 conjugated linoleic acidPhase 3NCT00706745
Actinomycin D (251)Dactinomycin in Treating Patients With Persistent or Recurrent Gestational Trophoblastic NeoplasiaGestational Trophoblastic TumorDactinomycinPhase 2NCT00003688
Addition of Ipilimumab (MDX-010) To Isolated Limb Infusion (ILI) With Standard Melphalan and Dactinomycin In The Treatment of Advanced Unresectable Melanoma of The ExtremityMelanomaDactinomycinPhase 2NCT01323517
Methotrexate Compared With Dactinomycin in Treating Patients With Gestational Trophoblastic NeoplasiaGestational Trophoblastic NeoplasiaDactinomycinPhase 3NCT00003702
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, J.; Xu, L.; Zhou, Y.; Han, B. Natural Products from Actinomycetes Associated with Marine Organisms. Mar. Drugs 2021, 19, 629. https://doi.org/10.3390/md19110629

AMA Style

Chen J, Xu L, Zhou Y, Han B. Natural Products from Actinomycetes Associated with Marine Organisms. Marine Drugs. 2021; 19(11):629. https://doi.org/10.3390/md19110629

Chicago/Turabian Style

Chen, Jianing, Lin Xu, Yanrong Zhou, and Bingnan Han. 2021. "Natural Products from Actinomycetes Associated with Marine Organisms" Marine Drugs 19, no. 11: 629. https://doi.org/10.3390/md19110629

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