Next Article in Journal
Checklist of Odonata of Sidi Bel Abbes, Northwest Algeria
Previous Article in Journal
Satellite-Based Quantification of Methane Emissions from Wetlands and Rice Paddies Ecosystems in North and Northeast India
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hydrobiological Aspects of Fatty Acids: Unique, Rare, and Unusual Fatty Acids Incorporated into Linear and Cyclic Lipopeptides and Their Biological Activity

by
Valery M. Dembitsky
Centre for Applied Research, Innovation and Entrepreneurship, Lethbridge College, 3000 College Drive South, Lethbridge, AB T1K 1L6, Canada
Hydrobiology 2022, 1(3), 331-432; https://doi.org/10.3390/hydrobiology1030024
Submission received: 8 August 2022 / Revised: 1 September 2022 / Accepted: 5 September 2022 / Published: 14 September 2022

Abstract

:
The study of lipopeptides and their related compounds produced by various living organisms from bacteria to marine invertebrates is of fundamental interest for medicinal chemistry, pharmacology, and practical clinical medicine. Using the principles of retrosynthetic analysis of linear and cyclic peptides, the pharmacological activity of unique, unusual, and rare fatty acids (FA) that are part of natural lipopeptides was investigated. To search for new biologically active natural metabolites from natural sources, more than 350 FA incorporated into linear and cyclic peptides isolated from bacteria, cyanobacteria, microalgae, marine invertebrates, fungal endophytes, and microorganisms isolated from sediments are presented. Biological activities have been studied experimentally in various laboratories, as well as data obtained using QSAR (Quantitative Structure-Activity Relationships) algorithms. According to the data obtained, several FA were identified that demonstrated strong antibacterial, antimicrobial, antifungal, or antitumor activity. Along with this, FA have been found that have shown rare properties such as antiviral, antidiabetic, anti-helmintic, anti-inflammatory, anti-psoriatic, anti-ischemic, and anti-infective activities. In addition, FA have been found as potential regulators of lipid metabolism, as well as agents for the treatment of acute neurological disorders, as well as in the treatment of atherosclerosis and multiple sclerosis. For 36 FA, 3D graphs are presented, which demonstrate their predicted and calculated activities.

1. Introduction

It is known that peptides are natural biological molecules, are found in all living organisms living on planet Earth and apparently play a key role in all types of biological activity [1,2,3,4,5,6,7,8,9]. Linear and/or cyclic lipopeptides and depsipeptides of bacteria, cyanobacteria, marine invertebrates, and fungi and fungal endophytes are of great interest, both from the point of view of academic research and from a purely practical point of view, as well as their use in pharmacology and medicine [1,2,10,11,12,13,14,15,16,17,18,19]. Both types of peptides have shown different biological activities, which include antiviral, anti-inflammatory, antibacterial, antifungal, anti-tumor, and other activities [3,5,12,15,17]. Lipopeptides are molecules consisting of a lipid fragment connected to peptides with an ester bond, while depsipeptides are peptides in which one or more amide groups are replaced by a corresponding ester [20,21,22,23,24,25].
When scanning more than 30,000 peptide structures, especially lipopeptides and depsipeptides isolated from marine and terrestrial organisms, it was found that these compounds in the absolute majority (over 80%) contain fragments of saturated FA (C6:0–C26:0), about 15% contain iso-, anteiso- and methyl-branched chain FA (C6:0–C24:0), and about 4–5% contain unsaturated FA. A few exceptions to FA not included in this review are amino fatty (carboxylic) acids and those mentioned above. Unique, rare, and unusual FA make up less than one percent of the total peptides screening.
This review focuses on this rare group of FA that are part of lipopeptides or depsipeptides and are covalently linked. The lipopeptides used in this review were found and isolated from bacteria, microalgae, cyanobacteria, marine invertebrates, fungal endophytes, and phytopathogenic microorganisms. The biological activity of many lipopeptides has not been studied. The biological activity of FA, as well as some amino acids that make up the peptides, was studied using the PASS software [26,27,28]. In addition, 3D graphs of the most interesting FA from the point of view of their pharmacological activity are presented.

2. Bacterial and Cyanobacterial Linear and Cyclic Peptides and Their Fatty Acids

It is known that cyanobacteria (or blue–green algae) belong to the division of Gram-negative bacteria capable of photosynthesis, accompanied by the release of oxygen, which have a blue–green color, and they are called cyanobacteria [29,30,31,32,33]. From the Archean rocks of Western Australia, fossils of cyanobacteria have been identified, the age of which is determined at 3.5 billion years [34,35].
Marine cyanobacteria have been attracting increasing attention for probe and drug discovery due to the high incidence of structurally novel bioactive secondary metabolites that complement those known from terrestrial sources [36,37,38]. These natural products are predominantly modified peptides and depsipeptides, polyketides, and peptide–polyketide hybrids, many of which are cyclic and oftentimes halogenated [39,40,41]. Cyanobacteria produce many bioactive compounds of various chemical structures, with about 40% of them being lipopeptides [36,37,38,39,40,41,42,43,44,45].
Using cytotoxic cyclic depsipeptides as an example, we want to show which FA are incorporated into peptides and are of interest to academic science for their unusual chemical structures. All the discovered cryptophycins can be retrosynthetically divided into four sub-fragments or subunits, namely A–D. Unit A represents the most exotic fragment, (5S,6S,E)-5-hydroxy-6-(2R,3R)-3-phenyloxiran-2-yl)hept-2-enoic acid. Fragment B can be derived from D-O-methyltyrosine and represents the (R)-2-amino-3-(3-chloro-4-methoxyphenyl)-propanoic acid. Unit C represents 3-aminopropanoic acid (known as β-alanine). Finally, (S)-2-hydroxy-4-methyl-pentanoic acid, also known as L-leucic acid (or 2-hydroxyisocaproic acid), constitutes the fragment D (Figure 1).
Cryptophycin-1 is the most important member of the cryptophycins family. It was first isolated from Nostoc sp. ATCC 53787 by Merck scientists as an antifungal agent [46]. However, a detailed study of cryptophycin-1 showed that it was too toxic to be of practical use, at least as an antifungal agent. Subsequent studies have shown that cryptophycin-1 is an active microtubule depolymerization agent, showing excellent activity against a wide range of solid tumors implanted in mice, including multiple drug-resistant tumors [47]. In addition, cryptophycin-1 suppresses tubulin dynamics and induces apoptosis [48,49].
Considering the above experimental data, we tested the biological activity of cryptophycin-1 and its subunits A, B, C and D included in this depsipeptide. The data of the PASS analysis are shown in Table 1. The data obtained by various groups of scientists are fully confirmed by the PASS program. The dominant activity was antifungal with a confidence level of 84.5%, in addition, antitumor activity was found with a confidence level of 77%, antineoplastic (solid tumors) activity with a reliability of 63%, and apoptosis agonist-62%. Thus, PASS fully confirmed the experimental data, and thus the program showed the correctness and quality of its work.
Interestingly, the antifungal activity that is characteristic of cryptophycin-1 was found only in subunit A and was not found in other subunits B, C and D. From this it can be inferred that subunit A makes a significant contribution to the overall pool of activity of this depsipeptide (Table 1).
Cryptophycin-1, similar to other cryptophycins, is a class of 16-membered highly cytotoxic macrocyclic depsipeptides produced by the cyanobacterium from the strain Nostoc [46]. So, (5S,6S,E)-5-hydroxy-6-(2R,3R)-3-phenyloxiran-2-yl)hept-2-enoic acid (1, structure shown in Figure 2) is a fragment of the cryptophycins 1, 2, 16, 21, 23, 24, 28, 31, 38, 50, 52, 54, 176 and 326 [50,51], and (2E,5S,6R,7E)-5-hydroxy-6-methyl-8-phenylocta-2,7-dienoic acid (2) was detected in cryptophycins 3, 4, 17, 18, 19, 29, 31, 43, 45, 46, 49, 175 and 327 (Figure 3). (S,E)-5-Hydroxy-6-(2R,3R)-3-phenyloxiran-2-yl)hex-2-enoic acid (3) was found in the structure of cryptophycin 28, (R,2E,7E)-5-hydroxy-8-phenylocta-2,7-dienoic acid (4) was detected in cryptophycin 40, and (6S,7S,Z)-6-hydroxy-7-((2R,3R)-3-phenyloxiran-2-yl)oct-3-enoic acid (5) was isolated from cryptophycin 327.
Cryptophycin-38, -326, and -327 were isolated from the terrestrial cyanobacterium Nostoc sp. GSV 224 by Chaganty and co-workers [52], and cryptophycin-2, -21, -46, -175, and -176 have been identified from the MeCN-CH2Cl2 extract of the same blue–green algae [50,53,54].
(5S,6S,7R,8S,E)-8-Chloro-5,7-dihydroxy-6-methyl-8-phenyloct-2-enoic acid (6) has been detected in the cryptophycins 8 and 55 which were isolated from Nostoc sp. GSV, and (5S,6S,7R,8S,E)-5,7-dihydroxy-8-methoxy-6-methyl-8-phenyloct-2-enoic acid (7) was found in the cryptophycins 9 and 10 [55]. 4-(2S,3R,4R,5S)-4-Hydroxy-3-methyl-5-phenyltetrahydrofuran-2-yl)-butanoic acid (8) has been found in the linear cryptophycins 6 and 7, and (5S,6S,E)-5-hydroxy-6-((2S,3S)-3-phenyloxiran-2-yl)hept-2-enoic acid (9) was detected in cryptophycin-101. The anticancer activity of natural, semi-synthetic and synthetic cryptophycins has been studied in detail and summarized in several review articles [50,51,52,53,54,55,56,57]. The biological activity of the fatty acids incorporated into cryptophycins is shown in Table 2.
The analysis of the predicted biological activity of subunit A incorporated into cryptophycins, which is presented in Table 2, shows that all these fatty acids have three dominant properties such as antiviral (Arbovirus), antifungal and moderate antineoplastic activities with some variations in different fatty acids. The 3D graph of the activities of cryptophycin-1 and its four subunits A, B, C and D is shown in Figure 4.
It is known that natural compounds containing an acetylenic (triple) bond have been isolated from many species of plants, fungi, fungal endophytes, and various marine invertebrates [58,59,60,61,62,63]. Numerous studies have shown that many of these metabolites exhibit various biological activities, such as antibacterial, antimicrobial, antifungal, antitumor, and other medicinal properties [58,64,65,66]. Various species of freshwater and marine plants, macrophytes, microalgae, cyanobacteria and some other aquatic organisms produce a wide variety of different bioactive molecules containing acetylenic bonds [58,64,65,66].
The genus Lyngbya is the most abundant and available cyanobacterial species and is distributed throughout the world in tropical and subtropical regions. The species L. majuscule (see Figure 5), L. martensiana, L. aestuarii and L. wollei are currently the most important species of their genus Lyngbya and synthesize many secondary metabolites including lipopeptides.
Through numerous studies, it has been established that the widespread tropical cyanobacterium Lyngbya majuscula synthesizes more than 30% of all natural products derived from all marine cyanobacteria [67], which exhibit various activities, including antiproliferative, antifeedant, anti-inflammatory, molluscicidal, and immunosuppressive properties. It has been established that more than half of the known secondary metabolites are either linear or cyclic lipopeptides, some of which contain an acetylene fragment [67]. The linear lipopeptides called apramides A, B, and G were found in the cytotoxic fraction of L. majuscula collected at Apra Harbor (Guam) [68], and apramide G showed strong cytotoxic activity against KB and LoVo cells, respectively [69]. (R)-2-methyloct-7-ynoic acid (10, see Figure 6 for structure, and biological activity shown in Table 3) was found in apramide A and G, and apramide B contained oct-7-ynoic acid (11).
Dragonamide, dysidenamide, nordysidenin, and pseudodysidenin were isolated from L. majuscula collected from the beach at Boca del Drago Beach, Bocas del Toro, Panama. Dragonamide containing (S)-2-methyloct-7-ynoic acid (12) showed strong cytotoxic activity against P-388, A-549, HT-29, and MEL-28 cancer cells [70], and was synthesized four years later after its discovery [71]. The n-BuOH extract of L. majuscula contained a cytotoxic, linear lipotetrapeptide named carmabin A, which contains 2,4-dimethyldec-9-ynoic acid (13) [72]. The depsipeptides named antanapeptins A–D, of which antanapeptin A and D contain 3-hydroxy-2-methyloct-7-ynoic acid (14), were isolated from L. majuscula of the Antany Mora collection (Madagascar), and both depsipeptides showed moderate cytotoxic activity against neuroblastoma-2A cells in mice [73].
Pitipeptolide A, which belongs to the cyclodepsipeptides, was isolated from extracts of the cyanobacterium L. majuscula, which lives around the Piti Bomb Holes (Guam Reefs), where marine blooms caused by these cyanobacteria occur with a certain periodicity. This lipopeptide containing a 2,2-dimethyl-3-hydroxy-7-octynoic acid (15) [74] exhibited weak cytotoxicity against LoVo cancer cells but possessed moderate antimycobacterial activity and stimulated elastase activity. Other pitipeptolides D, E, and F also contained this fatty acid. Two depsipeptides named yanucamides A and B were found in lipid extracts of two cyanobacteria, L. majuscula and Schizothrix sp. collected at Yanuca Island (Fiji) [75]. Both lipopeptides contained the same fatty acid (15), which has previously been described as a major component of kulolide-1 and kulokainalide-1 isolated from the marine mollusk Philinopsis speciosa [76].
This acetylenic acid (15) was also found in the depsipeptides called wewakpeptins A and C, and both peptides showed cytotoxic activity against H460 human lung tumor and the mouse neuroblastoma-2A cell lines [77,78,79]. Additionally, another cyclic depsipeptide called georgamide was isolated from a non-identified Australian cyanobacterium [79], containing two hydroxy carboxylic acids, 2(S)-hydroxy-3(R)-methyl-pentanoic acid and FA (15), which are also present in wewakpeptins A and B [79].
Cytotoxic depsipeptide, which was named onchidin B, was isolated from extracts of the pulmonate mollusk Onchidium sp. and contained the 3-hydroxy-2-methyl-7-octynoic acid (14) [80,81,82].
Cyclic depsipeptide, kulolide-1 was isolated from a cephalaspidean mollusk, Philinopsis speciosa, and contains two carboxylic acids, L-3-phenyllacetic acid and the unprecedented (R)-3-hydroxy-2,2-dimethyl-7-octynoic acid (15). The isolated depsipeptide showed activity against L-1210 leukemia cells and P388 murine leukemia cells and caused a morphological change in 3Y1 rat fibroblast cells [83,84]. In addition, this mollusk yielded a linear peptide, pupukeamide, and an unprecedented macrolide, tolytoxin-23-acetate, which contained this acetylenic acid [84]. Interestingly, the lipopeptides kulolide-2, kulolide-3, kulokainalide-1, kulomoopunalide-1, kulomoopunalide-2, and tolytoxin 23-acetate were found in combined extracts of ((EtOH and CHCl3/MeOH (1:1))) mollusk Ph. speciosa. Kulokainalide-1 contains acetylenic acid (15), and kulomoopunalide-1 and kulomoopunalide-2 contains (R)-3-hydroxy-2-methyl-7-octynoic acid (14) [84]. Widely present in lipopeptides, acetylenic acid (15) was also found in structures such as in the depsipeptides mantillamide, and dudawalamide A, isolated from extracts of the marine cyanobacterium Lyngbya sp. Both peptides show anticancer and antimalarial activity [85]. A cyclic depsipeptide, guineamide G with FA (15), is produced by the marine cyanobacterium L. majuscula, collected from Papua New Guinea, and exhibits potent brine shrimp toxicity and moderate cytotoxicity to a mouse neuroblastoma cell line [86]. Additionally, the cocosamides A and B from the lipophilic extract of a collection of L. majuscula from Cocos Lagoon (Guam), demonstrated activity against HT-29 cells, and also contained FA (15) [87]. FA (15) was present in cyclic depsipeptides, and the viequeamides A–F, which were discovered from a shallow subtidal collection of Rivularia sp.; viequeamide A is active against the H460 human lung cancer cell line [88]. Other cyclic depsipeptides, named dudawalamides A−D, were isolated from Papua New Guinea from the cyanobacterium Moorea sp., and FA (15) was found in dudawalamide A as recently reported [89]. The 3D graph of the activity of fatty acids (15) is shown in Figure 7.
The Luesch group from the Florida University reported the isolation of cytotoxic cyclodepsipeptides, veraguamides A–F from a cyanobacterium Symploca cf. hydnoides at Cetti Bay, Guam [90], and veraguamides H, I–L from the marine cyanobacterium Oscillatoria margaritifera at the Coiba National Park, Panama [91], which contained 3-hydroxy-2-methyl-7-octynoic acid (14). Among them, veraguamides A and B are 8-bromo-3-hydroxy-2-methyl-7-octynoic acid (16)-containing cyclic peptides, while veraguamides K and L are acid (16)-containing linear peptides. The Okinawan marine cyanobacterium Oceania sp. produces a cytotoxic depsipeptide called odobromoamide containing an alkynyl bromide (16), and this demonstrated cytotoxic activity against HeLa S3 cells and broad-spectrum cytotoxicity against a panel of human cancer cell lines [92].
An alkynyl-containing cyclic depsipeptide, palauamide, containing 5,7-dihydroxy-2,6-dimethyldodec-2-en-11-ynoic acid (17), was extracted from a Lyngbya sp. from Palau. Palauamide showed strong cytotoxicity against KB cells with an IC50 value of 13 nM [93]. Sitachitta and co-workers reported the isolation and identification of cyclic peptides, trungapeptins A, B, and C, containing 3-hydroxy-2-methyl-7-octynoic (18), 3-hydroxy-2-methyl-7-octenoic (19), and 3-hydroxy-2-methyl-7-octanoic acid (20), respectively [94]. Trungapeptin A exhibited potent brine shrimp toxicity and ichthyotoxicity at 10 ppm and 6.2 ppm, respectively.
In 2009, a hmoya-containing analog of hantupeptin A (19, 3-hydroxy-2-methyl-7-octynoic acid) was discovered from the marine cyanobacterium Lyngbya majuscula from Pulau Hantu Besar, Singapore [95]. Further, hantupeptin A afforded both brine shrimp toxicity at 10 ppm and strong cytotoxicity against the leukemia cell line MOLT-4 with an IC50 value of 32 nM. The same FA (19) was detected in trungapeptin A, which was detected in the marine cyanobacterium L. majuscula collected from Trung Province, Thailand [96].
A sample of brown Lyngbya polychroa from Hollywood Beach, Fort Lauderdale, Florida yielded an impressive array of structurally diverse cytotoxic linear tetrapeptide–octynoates, the dragonamides A, B, C, D and E [97,98,99]. The dragonamides C and D showed weak activity, with GI50 values of 56 and 59 µM against U2OS osteosarcoma cells, 22 and 32 µM against HT29 colon adenocarcinoma cells, and 49 and 51 µM against IMR-32 neuroblastoma cells, respectively. From isolated linear tetrapeptides, the dragonamides A and B contain FA (12), dragonamide C contains FA (21), dragonamide D contains FA (22) and dragonamide E contains FA (23).
Several linear alkynoic lipopeptides have been isolated from a Panamanian strain of the marine cyanobacterium L. majuscula, including carmabin A, dragomabin, and dragonamide A, which showed good antimalarial activity (IC50 4, 6, and 7.7 μM, respectively), whereas the non-aromatic analog, dragonamide B, was inactive [98]. The isolated linear lipopeptides dragomabin and dragonamide A and B contained fatty acid (12), while fatty acid (13) were determined in carmabin A.
A marine cyanobacterium Oscillatoria nigro-viridis from Panama area led to the isolation of two linear alkynoic lipopeptides, viridamides A and B with 5-methoxydec-9-ynoic acid (24). Viridamide A showed anti-trypanosomal activity with an IC50 = 1.1 µM, and anti-leishmanial activity with an IC50 = 1.5 µM [100]. An acetylene-containing lipopeptide, kurahyne with FA (23), was isolated from a cyanobacterial assemblage that mostly consisted of Lyngbya sp. Kurahyne inhibited the growth of human cancer cells and induced apoptosis in HeLa cells, and it seemed to localize in mitochondria [101].
Cyclic depsipeptides, the cocosamides A and B, containing FA (15) and (25), respectively, have been detected in the lipophilic extract of a collection of L. majuscula from Cocos Lagoon, Guam. Both metabolites showed weak cytotoxicity against MCF7 breast cancer and HT-29 colon cancer cells [102].
The marine benthic cyanobacteria Oscillatoria nigroviridis from the Colombian Caribbean Sea produces lipopeptides named almiramides B, D, E, F, H, and G with FA (12). Almiramides B and D show a strong activity against human tumor cell lines A549, MDA-MB231, MCF-7, HeLa and PC3 [103]. Pitipeptolides D, E and F with (15, FA) showing antimycobacterial cyclodepsipeptides were detected in the marine cyanobacterium L. majuscula from Piti Bomb Holes, Guam. Obtained compounds showed weak cytotoxicity against HT-29 colon adenocarcinoma and MCF7 breast cancer cells [86].
Cytotoxic cyclic depsipeptides, hantupeptins A and B, have been derived from the marine cyanobacterium L. majuscula from Pulau Hantu Besar, Singapore. Hantupeptin A with FA (12) showed strong cytotoxicity against leukemia cells and breast cancer MCF-7 cells (IC50 values of 32 and 4.0 µM, respectively), while hantupeptin B with FA (19) displayed moderate cytotoxicity against MOLT-4 (leukemia) and MCF-7 cell lines [95,104].
The cyclic depsipeptide guineamide G with FA (15) has been extracted from Lyngbya semiplena and L. majuscula. The isolated lipopeptide exhibited brine shrimp toxicity and showed potent cytotoxicity against a mouse neuroblastoma cell line with an LC50 value of 2.7 µM [105,106].
A slightly halophilic myxobacterial strain, SMH-27-4, was isolated from nearshore soil. This slowly-growing myxobacterium produced the novel antibiotic depsipeptides named miuraenamides A and B, which both contain (R,E) -9-hydroxy-6-methyldec-5-enoic acid (26, structure see Figure 8, and Table 4). Miuraenamide A exhibited potent and selective inhibition against a phytopathogenic microorganism, Phytophthora sp., And moderate inhibition against some fungi and yeasts. Both metabolites inhibited NADH oxidase at IC50 values of 50 μM [107].
Cyanobacteria are known to require iron to grow, and they often inhabit iron-restricted habitats and produce several siderophores, including an unusual FA (27). Iron starvation triggered the synthesis of β-OH-Asp lipopeptides in cyanobacteria Rivularia sp. strain PCC 7116, Leptolyngbya sp. strain NIES-3755, and Rubidibacter lacunae strain KORDI 51-2 [108].
Unusual (2E,6E,8E)-12-hydroxy-14-methoxy-2,4,9-trimethylpentadeca-2,6,8-trienoic acid (28) was detected in cyclic peptide named alotamide A, which is produced by the marine cyanobacterium Lyngbya bouillonii. Alotamide A displays an unusual calcium influx activation profile in murine cerebrocortical neurons with an EC50 of 4.18 μM [109]. The cyclic peptides apratoxins A–C possess the unprecedented (E)-3-((4S)-2-((2S,3R,5S)-3,8-dihydroxy-5,9,9-trimethyldecan-2-yl)-4,5-dihydrothiazol-4-yl)-2-methylacrylic acid (29) as the polyketide moiety, and apratoxin D with FA (30) showed potent in vitro cytotoxicity against H-460 human lung cancer cells with an IC50 value of 2.6 nM. Apratoxin A possesses IC50 values for in vitro cytotoxicity against human tumor cell lines, ranging from 0.3 to 0.5 nM; however, it was only marginally active in vivo against a colon tumor and ineffective against a mammary tumor. Apratoxins A–D have been isolated from the marine cyanobacteria Lyngbya majuscula and Lyngbya sordida [110,111].
The nNeuroactive cyclic depsipeptide hoiamide A was originally isolated from a consortium of two different filamentous cyanobacteria identified as Lyngbya majuscula and Phormidium gracile, and two related peptide metabolites, one a cyclic depsipeptide, hoiamide B, and the other a linear lipopeptide, hoiamide C, were isolated from two different collections of marine cyanobacteria obtained in Papua New Guinea. All the isolated hoiamides A–C contain unusual FA (31) [112].
The 3D graph demonstrating the predicted and calculated FA (31) activity is shown in Figure 9. A collection of the marine cyanobacterium Lyngbya bouillonii from Guam afforded the cytotoxic apratoxin E, which displayed stronger cytotoxicity than its closest analog, semisynthetic E-dehydroapratoxin A, against several cancer cell lines derived from colon, cervix, and bone, ranging from 21 to 72 nM. Both cyclic peptides contained FA (32) and (33), respectively [113]. A halogenated metabolite, (S)-7,7-dichloro-3-hydroxy-2,2-dimethyloctanoic acid (34) is incorporated into many lipopeptides that are produced predominantly by the marine cyanobacterium of the genus Lyngbya, as well as other cyanobacterial species, L. bouillonii and the Fijian marine cyanobacterium Moorea producens [114,115]. Thus, lyngbyabellin A, a significantly cytotoxic compound with unusual structural features, was isolated from a Guamanian strain of the marine cyanobacterium Lyngbya majuscula [116]. Lyngbyabellin A was shown to be a potent disrupter of the cellular microfilament network. In addition, lyngbyabellin A showed anticancer activity against HT29, HeLA, KB, LoVo cancer cell lines [114,115]. Antitumor lipopeptides lyngbyabellin B, C, J and 27-deoxylyngbyabellin A have shown activity against HT29, HeLA, and MCF7 cancer cell lines [115,117,118].
A halogenated FA (34, see Figure 10) has also been found in the cytotoxic lipopeptide named hectochlorin which was isolated from marine isolates of Lyngbya majuscula collected from Hector Bay, Panama [119]. Another chlorine-containing metabolite, (2S,3S)-7,7-dichloro-3-hydroxy-2-methyl-octanoic acid (35) was detected in bioactive lipopeptides such as the lyngbyabellins E, F, G, H, I, K, M, N, O and P, which exhibit antimalarial, anticancer, and antifouling activities [117,118,120,121,122]. The above lyngbyabellins have been isolated from the marine cyanobacteria Lyngbya majuscula, Lyngbya sp., L. bouillonii, Okeania sp., Moorea bouillonii, and M. producens. (2S,3S,7R)-7-Chloro-3-hydroxy-2-methyloctanoic acid (36) was isolated from lipopeptides named lyngbyabellins K, L and 7-epi-lyngbyabellin L, which showed antitumor activity against H-460 cancer cell lines [123].
The cyclic lipopeptides named antillatoxins, ATx-A (41) and ATx-B (42) have been isolated from a marine cyanobacteria Lyngbya majuscula [124], and both toxins contained (4S,5S,6E,8E)-5-hydroxy-4,6,8,10,10-pentamethyl-3-methyleneundeca-6,8-dienoic acid (37), and acid (40). Antillatoxin B exhibited significant sodium channel-activating (EC50 = 1.8 μM) and ichthyotoxic (LC50 = 1 μM) properties [125]. It was shown that the natural product antillatoxin B (42) is 10 times less active than antillatoxin A (41), and synthetic stereoisomers (43 and 44) of the cyclic depsipeptide of antillatoxin A were 20–55 times less active than the natural isomer [125,126].
Both synthetic stereoisomers (43 and 44) contain different lipophilic fragments, (4S,5S,6E,8E)-5-hydroxy-4,6,10,10-tetramethyl-3-methyleneundeca-6,8-dienoic acid (38) and (4S,5S),(E)-5-hydroxy-4,6,10,10-tetramethyl-3-methyleneundec-6-enoic acid (39), respectively, although the cytotoxicity (44) for Neuro-2a cells was shown to be 10 times more effective than the cytotoxicity of the molecule (43) as shown by Okura and co-workers [127].
The biological activities of FA 34–40 and the cyclic lipopeptides named antillatoxins 41–44 are shown in Table 5. Undoubtedly, (S)-7,7-dichloro-3-hydroxy-2,2-dimethyloctanoic acid (34) containing two chlorine atoms is of great interest since it demonstrates strong hypolipemic activity with a confidence level of more than 92%. The 3D graph of this acid is shown in Figure 11.
It seems very interesting to compare the activities of (4S,5S,6E,8E)-5-hydroxy-4,6,8,10,10-pentamethyl-3-methyleneundeca-6,8-dienoic acid (37), two cyclic lipopeptides named antillatoxin A (41) and B (42), and a synthetic analogue (43). As shown by the PASS analysis, for all samples the dominant property is moderate antitumor activity with a reliability of about 90%. The 3D graph of this acid (37) and cyclic lipopeptides is shown in Figure 12.
Lyngbyabellin N is a bioactive lipopeptide (structure shown in Figure 10) that is produced by the filamentous marine cyanobacterium Moorea bouillonii. This metabolite possesses a leucine statin residue and showed strong cytotoxic activity against HCT116 colon cancer cell lines (IC50 = 41 nM) and is highly cytotoxic to H-460 human lung cancer cells [117,118,119,120,121,122,123]. It was interesting to determine which of the four fragments (A, B, C and D) of lyngbyabellin N contributes to its overall activity. The actuality of these fragments is shown in Table 6. As can be seen from these PASS data, fragment D demonstrates antineoplastic (liver cancer) activity with a confidence level of 92.3%.
Neo FA, neo alkanes and their analogs and derivatives are quite rare lipid molecules that are found in marine invertebrates, algae, fungi, plants, and microorganisms, but they are not found in free form [128]. Several neo fatty (carboxylic) acids (37–41, 45–48) have been incorporated into the lipopeptides produced by cyanobacteria. Two cytotoxic peptides with pivalic acid (45, activity see in Table 5), named bisebromoamide and norbisebromoamide, have been identified from the marine cyanobacterium Lyngbya sp. Bisebromoamide exhibit cytotoxicity against HeLa S3 cells (IC50 = 0.04 µg/mL) and inhibit the phosphorylation of extracellular signal-regulated protein kinases in NRK cells, showing potent and selective inhibitory effects on protein kinases [129,130]. The antibiotic bottromycin B with pivalic acid (45), produced by Streptomyces sp. strain No. 3668-L2, Kitasatoa purpurea strain KA-281, and Micromonospora chalcea strain FERM-P 1823 [131,132,133,134]. Bottromycin has shown antibacterial activity against six antibiotic-resistant strains of Staphylococcus aureus, Streptococcus pyogenes, Micrococcus flavus, Bacillus subtilis, B. cereus, B. megaterium, B. anthracis, Corynebacterium xerosis, and Mycobacterium phlei in concentrations of 0.03–3 µg/mL.
Pyrrolinone-containing lipopeptides named ypaoamide with (Z)- and (E)- double bonds, herbivore antifeedant metabolites, were isolated from the extract of a mixed cyanobacterial assemblage that was composed of Schizothrix calcicola and Lyngbya majuscula [135,136,137]. More than 20 years later, similar pyrrolinone-containing lipopeptides named ypaoamides B and C were isolated from marine cyanobacterium Okeania sp. collected in Okinawa. Both ypaoamides B and C stimulated glucose uptake in cultured rat L6 myotubes, and ypaoamide B showed potent activity and activated AMP-activated protein kinase. All four lipopeptides contained the rare 6,6-dimethylheptanoic acid (46) [138].
The unique polytheonamides A and B with 5,5-dimethyl-2-oxohexanoic acid (47) are highly cytotoxic polypeptides with 48 amino acid residues have been isolated from the marine sponge Theonella swinhoei [139].
Janadolide, obtained from cyanobacteria Okeania sp. (Janado, Japan) and a cyclic peptide polyketide hybrid possessing a rare tert-butyl moiety, (2R,7S,E)-7-hydroxy-2,5,8,8-tetramethylnon-5-enoic acid (48), showed potent activity towards Trypanosoma brucei (IC50 47 nM) which was superior to the commonly used therapeutic drug suramin. Furthermore, significant selectivity towards the trypanosome parasite was identified, since no in vitro cytotoxicity towards the human cell lines MRC-5, HL60 and HeLa cells was noted at 10 mM [140].
Compounds containing an aromatic ring in a molecule are widespread in nature. They are found in marine invertebrates, algae, fungi, microorganisms, and other living organisms [141,142,143,144,145,146,147,148,149,150,151,152,153]. Acyclic lipopeptides named hoshinoamides A and B with 4-(4-hydroxyphenyl)-butanoic acid (49) have been isolated from the marine cyanobacterium Caldora penicillata. Both compounds inhibited the in vitro growth of the malarial parasite Plasmodium falciparum (IC50 = 0.5 and 1.0 μM, respectively) [154]. Cytotoxic depsipeptides, anaenamides A and B with 2-methoxy-6-pentylbenzoic acid (50), were discovered from a green filamentous cyanobacterium Hormoscilla sp. from Guam [155].
A linear lipopeptide aldehyde with 2-hydroxy-3-(4-hydroxyphenyl)-propanoic acid (51) was detected in a hydrophilic extract of the Nostoc sp. [156]. Two trypsin inhibitors called nostosin A and B with (S)-2-hydroxy-4-(4-hydroxyphenyl)-butanoic acid (52) and (R)-2-hydroxy-4-(4-hydroxyphenyl)-butanoic acid (53), respectively, were isolated from a hydrophilic extract of Nostoc sp. strain FSN, which was collected from a paddy field in the Golestan Province, Iran. Nostosins A and B exhibited IC50 values of 0.35 and 55 μM against porcine trypsin, respectively, suggesting that the argininal aldehyde group plays a crucial role in the efficient inhibition of trypsin [157].
Nannocystin A, a lipopeptide with epoxy-containing FA (54), was isolated from a myxobacterium Nannocystis sp. [158,159]. The isolated compound has a strong antifungal effect against C. albicans and displays potent cell proliferation inhibitive properties by inducing apoptosis early in tested cell lines. Besides, nannocystin A has shown antiproliferative properties against 472 cancer cell lines in the nanomolar concentration range (IC50 values ranging from 0.5 µM to 5 nM).
A cytotoxic and linear peptide was isolated from the marine cyanobacterium Geitlerinema sp. The structure of mitsoamide contains an unusual polyketide unit (3,7-dimethoxy-5-methyl-nonanedioic acid, 55) incorporated into a homolysine residue, and possesses a highly unusual piperidine amino moiety. This peptide showed antitumor activity against NCI-H460 human lung tumor cells, IC50 460 nM [160].
A lipopeptide, minnamide A with unique (3S,5S,7R,9S,11R,13S,15R)-3,7,11,15-tetrahydroxy-5,9,13-trimethyloctadecanoic acid (56) from the marine cyanobacterium Okeania hirsute showed growth-inhibitory activity toward HeLa cells with an IC50 value of 0.17 μM, and rapidly induced cell death at a concentration of 2 μM [161].
Rare (2E,4E,10E)-15-hydroxy-7-methoxy-2-methylhexadeca-2,4,10-trienoic acid (57) was incorporated into cyclic lipopeptide named malevamide E which was found in extracts of the marine cyanobacterium Symploca laeteviridis (see Figure 13) and showed store-operated Ca 2+ entry in thapsigargin-treated human embryonic kidney (HEK) cells with a dose-dependent inhibition (2–45 µM) [162].
The tropical marine cyanobacterium, Moorea bouillonii from New Britain, Papua New Guinea, yielded a cytotoxic cyclic depsipeptide, bouillonamide [163]. The obtained metabolite exhibited mild toxicity with an IC50 of 6.0 µM against the neuron 2a mouse neuroblastoma cells. In addition, the cyclopeptide contained two unique polyketide-derived moieties, namely a 2-methyl-6-methylamino-hex-5-enoic acid (58) and 3-methyl-5-hydroxy-heptanoic acid (59).
The antifungal glycosylated lipopeptide, hassallidin A, was isolated from an epilithic cyanobacterium Tolypothrix basionym collected in Bellano, Italy. The isolated lipopeptide with 2,3-dihydroxytetradecanoic acid (60, see Figure 14) exhibited antifungal activity against Aspergillus fumigatus and Candida albicans [164]. The 3D graph demonstrating the predicted and calculated activity of FA (51, 52, 53 and 54) is shown in Figure 15.
Yu and co-workers [165] isolated nine linear lipopeptides of the named microcolins E–M from the marine cyanobacteria Moorea produns, which showed significant cytotoxic activity against lung carcinoma, and they all contained (2S,4S)-2,4-dimethyloctanoic acid (61, for structure see Figure 16, and activity see in Table 7). Two linear lipopeptides, gageostatins B and C belonging to the heptapeptides were obtained from the fermentation broth of a marine-derived bacterium Bacillus subtilis. The isolated lipopeptides contain (3R)-3-hydroxy-9,11-dimethyltridecanoic (62) and (3S,E)-3-hydroxy-9,11-dimethyl-tridec-4-enoic (63) acids, respectively. The gageostatins exhibited good antifungal activities with MIC values of 4–32 µg/mL when tested against pathogenic fungi (Rhizoctonia solani, Bacillus cinerea and Colletotrichum acutatum) and both compounds shown moderate antibacterial activity against bacteria (Bacillus subtilis, Staphylococcus aeureus, Salmonella typhi and Pseudomonas aeruginosa) with MIC values of 8–64 µg/mL. Furthermore, gageostatins displayed cytotoxicity against six human cancer cell lines with GI50 values of 4.6–19.6 µg/mL [166].
An antimalarial lipopeptide, ikoamide with (3S,5R)-3,5-dimethoxyoctanoic acid (64), was isolated from an Okeania sp. marine cyanobacterium. Ikoamide showed strong antimalarial activity with an IC50 value of 0.14 μM without cytotoxicity against human cancer cell lines at 10 μM [167]. A malyngamide with (E)-7-hydroxytetradec-4-enoic acid (65) was isolated from the marine cyanobacterium Moorea producens collected in Hawaii. The compound showed cytotoxicity against the L1210 cell line at an IC50 value of 2.9 mM and lethal toxicity against the shrimp Palaemon paucidens at an LD100 value of 33.3 mg/kg [168].
The marine bacterium Saccharomonospora sp. CNQ-490 produced the chlorinated lipopeptide taromycin A, and taromycin B was detected in ethyl acetate extracts of S. coelicolor M1146-M1 cultures [169,170]. Both taromycins A and B display potent activity against methicillin-resistant Staphylococcus aureus and vancomycin-resistant Enterococcus faecium clinical isolates. Both lipopeptides contain a rare (2E,4E)-octa-2,4-dienoic (66) and (2E,4E)-6-methylocta-2,4-dienoic (67) acids, respectively.
Cyclic lipopeptides named bananamides D-G were detected in the crude extract of Pseudomonas sp. COW3. Both bananamides D and G contain (Z)-3-hydroxydodec-4-enoic acid (68). COW3 displayed antagonistic activity and mycophagy against Pythium myriotylum, while it mainly showed mycophagy on Pyricularia oryzae. Purified bananamides D-G inhibited the growth of P. myriotylum and P. oryzae and caused hyphal distortion [171].
The cyclic lipopeptide named gageopeptin A with (3S)-3-hydroxy-12,14-dimethylhexadecanoic acid (69) was obtained from the ethyl acetate extract of the fermentation broth of a marine-derived strain Bacillus sp. 109GGC020 and it exhibited moderate antibacterial and good antifungal activities [172], and two linear lipopeptides, gageopeptides C and D with (3S)-3-hydroxy-8,10-dimethyldodecanoic (70) and (3R)-3-hydroxy-9,11-dimethyltridecanoic (71) acids, were discovered from a marine Bacillus subtilis strain 109GGC020 [173].
A lipopeptide antibiotic, stalobacin I was discovered from a culture broth of an unidentified Gram-negative bacterium. Stalobacin I had a unique chemical architecture composed of an upper and a lower half peptide sequence, which were linked via a hemiaminal methylene moiety. The sequence of one contained an unusual amino acid, carnosadine, 3,4-dihydroxyariginine, 3-hydroxy-isoleucine, and 3-hydroxyaspartic acid, and a novel cyclopropyl FA, (E)-2-hydroxy-4-((1R,2R)-2-((Z)-tridec-6-en-1-yl)-cyclopropyl)-but-3-enoic acid (72). This compound showed antibacterial activity against a broad range of drug-resistant Gram-positive bacteria and was much stronger than those of “last resort” antibiotics such as vancomycin, linezolid, and telavancin (MIC 0.004–0.016 μg/mL) [174].
Lyngbya majuscula from Papua New Guinea yielded the guineamides B and C with (S)-3-hydroxy-2,2-dimethylhexanoic acid (73), which possess moderate cytotoxicity to a mouse neuroblastoma cell line with IC50 values of 15 and 16 μM, respectively [40]. The same acid (73) was detected in a cyclic depsipeptide palmyramide A that was isolated from a Lyngbya majuscula–Centroceras sp. association [175]. Pure palmyramide A showed sodium channel-blocking activity in neuro-2a cells and cytotoxic activity in H-460 human lung carcinoma cells. The 3D graph demonstrating the predicted and calculated activity of FA (66) is shown in Figure 17, and the 3D graph on Figure 18 shows the activity of FA (62), (63), (68), and (70).
A lipopeptide with (Z)-hexadec-9-enoic acid (74) named pseudoalteropeptide A was isolated from the marine bacterium Pseudoalteromonas piscicida SWA4_PA4. It showed moderate iron-chelating activity as well as cytotoxic activity against Jurkat human T lymphocyte cells [176].
Bioactive compounds of marine cyanobacteria from a Jamaican collection of Lyngbya majuscula led to the isolation of jamaicamides A–C. Two jamaicamides A and B are highly functionalized lipopeptides containing an alkynyl, vinyl chloride, α-methoxy eneone system, and pyrrolinone ring fragments [176A]. Three halogenated FA, (4E,9E)-14-bromo-9-(chloromethylene)-6-methyltetradec-4-en-13-ynoic acid (75, for structure see Figure 19, and activity see in Table 8), (4E,9E)-9-(chloromethylene)-6-methyltetradec-4-en-13-ynoic acid (76) and (4E,9E)-9-(chloromethylene)-6-methyltetradeca-4,13-dienoic acid (77) were found in the neurotoxins, jamaicamides A–C, respectively. Halogenated linear lipopeptides called vatiamides F and E have been found in the cyanobacterium Moorea producens. Both lipopeptides contained same FA (75) and (76), respectively [177].
The cyanobacteria genus Lyngbya is an amazing source of chlorine-containing metabolites, and Lyngbya majuscula from Grenada has identified depsipeptides named itralamides A and B, which contain 4,4-dichloro-3-methylbutanoic acid (78) [178]. The 3D graph demonstrating the predicted and calculated activity of FA (78) is shown in Figure 20.
Cyanobacterial field collections from American Samoa and Palmyra Atoll yielded three cyclic peptides called tutuilamides A–C. Tutuilamides A–C show potent elastase inhibitory activity together with moderate potency in H-460 lung cancer cell cytotoxicity assays. The tutuilamides A and B contain (E)-4-chloro-3-methylbut-3-enoic acid (79) and (E)-3-(chloromethylene)-pentanoic acid (80) and were detected in tutuilamide [179].
The Floridian marine cyanobacterium, Symploca sp., produces a cytotoxic depsipeptide named largazole and contains a 4-methyl-thiazoline unit (81) and an unusual a 3-hydroxy-7-mercaptohept-4-enoic acid (82). This thioether-functional depsipeptide is a potent inhibitor of the growth of transformed human mammary epithelial cells (MDA-MB-231) and is less susceptible to non-transformed mouse mammary epithelial cells. In addition, largazole showed exceptional antiproliferative activity against transformed U2OS fibroblast osteosarcoma cells [180].
Three lipopetides called lyngbyastatins 4–6 have been identified from the marine cyanobacterium Lyngbya confervoides from the Florida Atlantic coast and South Florida. Lyngbyastatin 4 shows potent and selective inhibitory effects on elastase as well as chymotrypsin in vitro over other serine proteases with IC50 values of 0.03 and 0.30 μM, respectively. Rare (R)-2-hydroxy-3-(sulfooxy)-propanoic acid (83) was detected in lyngbyastatin 4 and sodium (R)-2-carboxy-2-hydroxyethyl sulfate (84) was found in lyngbyastatin 6 [181,182].
A lipopeptide antibiotic called amphomycin A with (E)-10-methyldodec-3-enoic acid (85) was first isolated from extracts of the bacterium Streptococcus canis demonstrating antibacterial activity against Gram-positive pathogens [183,184]. A lipopeptide antibiotic called tsushimycin which was isolated from Bacillus subtilis contains three different FA in varying proportions: (E)-11-methyldodec-3-enoic (86), (E)-12-methyltridec-3-enoic (87) and (E)-11-methyltridec-3-enoic (88) acids [185].
Several lipopeptide antibiotics, friulimicins A, B, C, D, and the acidic lipopeptides of the amphomycin type that were also present in the culture fluid, compounds A-1437 A, B, E, and G, were isolated from cultures of Actinoplanes friuliensis HAG01 0964 after fermentation in different nutrient media. All eight lipopeptides possess an identical peptide macrocycle as their central element, linked via a diaminobutyric acid N-terminal either to an acylated asparagine residue or, in the case of the amphomycin series, to an acylated aspartic acid residue. Friulimicin A and A-1437 A contains cis-3-iso-13:1 acid (87), friulimicin B and A-1437 B contains cis-3-isotetradecenoic acid (88), friulimicin C and A-1437 E contains fatty acid (89), and friulimicin D and A-1437 G contains cis-3-anteisopentadecenoic acid (90, for structure see Figure 21, and activity in Table 9) [186,187].
Four cyclolipopeptides, glycinocins A to D, were isolated from the fermentation broth of an unidentified terrestrial Actinomycete species. The glycinocin antibiotics are structurally related to amphomycin that was originally reported as a linear lipopeptide with a C-terminal diketopiperazine moiety [188]. All isolated glycinocins contain rare double-bond FA in the second position. So, glycinocins A and D contain (E)-14-methylpentadec-2-enoic acid (91), B-(E)-15-methylhexadec-2-enoic acid (92) and C-(E)-13-methyltetradec-2-enoic acid (93), respectively.
A series of different lipopeptides containing the same FA fragment (94–100) were isolated from marine cyanobacteria living in different regions of the world’s oceans. Thus, Lyngbya majuscula from Papua New Guinea led to the isolation of two lipopeptides, aurilides B and C. Both compounds with (2E,5S,6S,7S,8E)-5,7-dihydroxy-2,6,8-trimethylundeca-2,8-dienoic acid (94) showed in vitro cytotoxicity toward NCI-H460 human lung tumor and the neuro-2a mouse neuroblastoma cell lines, with LC50 values between 0.01 and 0.13 μM, and aurilide B exhibited a high level of cytotoxicity against leukemia, renal, and prostate cancer cell lines [189].
Cyclic depsipeptides, lagunamides A and B were isolated from the marine cyanobacterium L. majuscula obtained from Pulau Hantu Besar (Singapore). Both lagunamides displayed significant antimalarial properties against Plasmodium falciparum, with IC50 values of 0.2 and 0.9 μM, respectively. Lagunamides A and B contained (5S,6S,7R,8R,E)-5,7-dihydroxy-2,6,8-trimethyldec-2-enoic (98, for predicted activity see Figure 22) and (2E,5S,6S,7S,8E)-5,7-dihydroxy-2,6,8-trimethyldeca-2,8-dienoic (99) acids and possessed potent cytotoxic activity against P388 murine leukemia cell lines, with IC50 values of 6.4 and 20.5 nM, respectively [190]. The same cyanobacterium from Singapore produced a cytotoxic cyclodepsipeptide named lagunamide C, which displayed potent cytotoxic activity against a panel of cancer cell lines, such as P388, A549, PC3, HCT8, and SK-OV3 cell lines, with IC50 values ranging from 2.1 to 24.4 nM. This compound with (5S,6R,8R,9S,E)-5,8-dihydroxy-2,6,9-trimethylundec-2-enoic acid (95) also displayed significant antimalarial activity with an IC50 value of 0.29 μM when tested against Plasmodium falciparum. In addition, lagunamide C exhibited weak anti-swarming activity when tested at 100 ppm against the Gram-negative bacterial strain, Pseudomonas aeruginosa PA01 [191].
Cytotoxic macrocyclic depsipeptides, lagunamide D and D, were discovered from a mixture containing marine cyanobacteria Dichothrix sp., Lyngbya sp. and Rivularia sp. from Florida. Both depsipeptides contain (5S,6R,7R,E)-5,7-dihydroxy-2,6-dimethylundec-2-enoic acid (96) [192]. In addition, lagunamide A and D exhibited antiproliferative activity even in the low-nanomolar range against A549 human lung adenocarcinoma cells with an IC50 value of 6.7 and 7.1 nM, respectively.
Okinawan marine cyanobacterium Okeania sp. led to the isolation of the cyclodepsipeptide named odoamide. Notably, this compound containing (5S,6S,7R,8S,E)-5,7-dihydroxy-2,6,8-trimethylundec-2-enoic acid (97) showed potent cytotoxicity against HeLa S3 human cervical cancer cells with an IC50 value of 26.3 nM [193].
The extract of a species of Lyngbya sp. from Palau has yielded the cyclodepsipeptide palauamide with (5S,6R,7R,E)-5,7-dihydroxy-2,6-dimethyldodec-2-en-11-ynoic acid (100), which had an IC50 value of 13 nM against the KB tumor cell line [93].
The Baltic Sea cyanobacterium Anabaena cylindrica Bio33, cultivated in the laboratory, has provided the antifungal lipopeptides balticidins A–D. An unusual, chlorinated FA (101, for structure see Figure 23, and activity see in Table 10) was detected in balticidins A and B, and dechlorinated acid (102) was found in balticidins C and D. Antifungal activity with these compounds is also observed against Candida albicans, C. krusei, Aspergillus fumigatus, Microsporum gypseum, Mucor sp., and Microsporum canis [194,195]. Antibiotic lipopeptides from Pseudomonads (see Figure 24), brabantamides A–C, were isolated from plant-associated Pseudomonas sp. SH-C52. Brabantamides A–C displayed moderate to high in vitro activities against Gram-positive bacterial pathogens. Brabantamide B contains unsaturated fatty acid (103), and brabantamide A and C contains saturated FA (104) and (105), respectively [196].
The cyanobacterium Hassallia sp. produces a family of bioactive compounds which exhibits a broad spectrum of antifungal activities. One of the bioactive glycolipopeptides is hassallidin B, which contains glycosidic FA (106, activity of this glycosidic FA sees in Figure 25) [197], and hassallidin D, which contains glycosidic FA (107) [198].
Characterizing the fatty acids of lipopeptides of bacteria and cyanobacteria, some conclusions can be drawn. So, these lipopeptides are characterized by fatty acids containing an aromatic ring, as measured by cryptophycins. Acetylene-containing FA are a hallmark of bacterial lipopeptides produced by Nostoc species. For bacterial lipopeptides, fatty acids with oxazole and thiazole rings, as well as sulfur- and chlorine-containing FA, are an interesting feature.

3. Linear and Cyclic Peptides Derived from Seaweeds and Invertebrates

Many algae and invertebrate species have long been used as human food, animal fodder and sources of valuable substances, including lipids. Marine seaweeds and invertebrates are rich in unusual lipids, steroids, triterpenoids, phospholipids, glycolipids, and polyunsaturated FA and are of potential value as sources of essential FA, important in the nutrition of humans and animals [199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218]. In addition, proteins of marine algae and invertebrates, which are natural reservoirs of bioactive peptides, are of great interest [219,220,221,222,223,224,225].

3.1. Fatty Acids Derived from Seaweed Lipopeptides

Marine and freshwater algae are a phylogenetically heterogeneous group of aquatic plants that belong to three main taxonomic groups: green (Chlorophyta), brown (Phaeophyta), and red (Rhodophyta) [226]. Since ancient times, seaweeds have been of great practical interest since they contain bioactive elements such as iodine, bromine or chlorine, and metabolites, steroids, carotenoids, fatty acids, lipopeptides, alkaloids, and other organic molecules that have antimicrobial, antiviral, anti-inflammatory and immunotropic properties [227,228,229,230,231].
Two cyclic lipopeptides, mebamamide A and B, containing FA (108, for structure see Figure 26, and activity see in Table 11) and (3R,8S)-3,8-dihydroxy-9-methyldecanoic acid (109), respectively, were isolated from the green alga Derbesia marina [232].
The sacoglossan mollusc Elysia rufescens is known to use the green algae Bryopsis pennata and B. plumosa as its main diet [233,234,235,236,237]. Analysis of lipid extracts from molluscs and algae showed that they contain biologically active cyclic depsipeptides, (kahalalides A–F, iso-KF, 5-OHKF, K, O–S, R′, S′, W, and Y) and five linear depsipeptides (kahalalides G, H, J, V, and X), which exhibit cytotoxic, antitumour, antimicrobial, antileishmanial and immunosuppressive activities [233].
The (R)-2-methylbutanoic acid (110, for structures see Figure 27, and biological activity is shown in Table 11) was found in kahalalide A and 5-methylhexanoic acid (111) was included in the structure of the kahalalides B, F, G, O, R2 and S2. 3-Hydroxy-9-methyldecanoic acid (112) was present in kahalalides E, H, J, K, and Y. (R)-4-Methylhexanoic acid (113), 5-hydroxy-5-methylhexanoic acid (114), (S)-2-hydroxy-9-methyldecanoic acid (115), 5-hydroxy-7-methyloctanoic acid (116), and (R)-3-hydroxy-7-methyloctanoic acid (117) were incorporated into the lipopeptides iso-kahalalide F, 5-OH-kahalalide F, kahalalide P and Q, kahalalide R1 and S1, and kahalalide V, respectively. As specimens, Figure 28 shows the green alga Bryopsis pennata, B. plumosa and the sacoglossan mollusc Elysia rufescens.
According to the PASS data, FA (115) showed properties as a cerebral anti-ischemic agent with a confidence level of more than 94%. This FA is found in cyclic depsipeptides, kahalalides P and Q, and a 3D graph of its predicted and calculated cerebral anti-ischemic activity is shown in Figure 29.

3.2. Fatty Acids Incorporated into Lipopeptides of Marine Sponges

Marine and freshwater sponges (class Demospongiae) are known to be home to many symbiotic microorganisms, including fungal endophytes, bacteria, and some unicellular organisms. Sponges, including their symbiotic microorganisms, synthesize many secondary metabolites such as steroids, terpenoids, carotenoids, halogenated and unusual fatty acids, alkaloids, and of course cyclic and linear lipopeptides [1,4,6,17,23,58,60,61,65,66,128,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215].
The marine sponges of the genus Theonella synthesize a wide variety of lipopeptides, and in the sponge Theonella aff. mirabilis, a pentapeptide was found that contained a rare (2R,3R)-aziridine-2,3-dicarboxylic acid (118, see Figure 30) [238]. The isolated lipopeptide inhibits the proteolytic activity of trypsin-like serine proteases, papain-like cysteine proteases, and pepsin-like aspartyl proteases [239]. Previously, this FA (118) was found and isolated from the ascomycete Streptomyces sp. MD 398-A1 [240]. A similar lipopeptide was isolated from the Red Sea sponge Theonella swinhoei (order Lithistida, see Figure 31) and is a potent inhibitor of cathepsin B, protease, and HIV [241]. The aziridine-containing compounds are powerful immuno-modulatory and anticancer agents and are of practical interest to pharmacologists [242,243].

3.2.1. Saturated, Methyl-Branched, and Unsaturated Fatty Acids

We have previously mentioned that neo fatty (carboxylic) acids have been isolated from cyanobacteria, microalgae, and some marine invertebrates [128]. Highly cytotoxic polypeptides named polytheonamides A and B are found in extracts of the marine sponge Theonella swinhoei [244,245,246,247]. Both polypeptides are quite unusual in that one peptide molecule contains nine amino acids with tert-butyl units, and both linear polypeptides contain a rare neo-FA, 5,5-dimethyl-2-oxohexanoic acid (119, for structure see Figure 32, and activity see in Table 12).
Linear peptides named yakuamides A and B were found and isolated from extracts of the Japanese sponge Ceratopsia sp. and they showed activity against P388 murine leukemia cells [245] and both contained 2,2,4,6-tetramethyl-3-oxoheptanoic acid (120).
The marine sponge Poecillastra sp. (Bahamas) yielded the potently cytotoxic poecillastrins A–C, which are related to the chondropsin D, and the closely related cytotoxic poecillastrin D was isolated from Jaspis serpentina (Oshimashinsone, Japan) [248]. All the above mentioned lipopeptides contain (E)-7-hydroxy-4,4,6,8-tetramethyl-5-oxonon-2-enoic acid (121).
Rare cyclic lipodepsipeptides, lipodiscamides A and C, were found in extracts of the marine sponge Discodermia kiiensis. Lipodiscamides A and C contain (3S,5R,6E,8E,11Z)-3-hydroxy-5-methoxy-2,2,15-trimethylhexadeca-6,8,11-trienoic acid (122), and lipodiscamide B contains FA (123) [249].
Two “head-to-side-chain” depsiundecapeptides named stellatolide A and B were present in lipid extracts from the marine sponge Ecionemia acervus. Both compounds showed strong antiproliferative activity against three human cancer cell lines (Lung-NSCLC A549, Colon HT-29 and Breast MDA-MB-231). (3S,6S,Z)-3-Hydroxy-6,8-dimethylnon-4-enoic acid (124) was isolated from stellatolide A, and (3S,6S,Z)-3-hydroxy-6-methylnon-4- enoic acid (125) was found in stellatolide B [250].
HIV-inhibitory cyclic depsipeptides known as neamphamide A–C were isolated from Papua New Guinea in the marine sponge Neamphius huxleyi. All lipopeptides contain 2R,3R,4R)-3-hydroxy-2,4,6-trimethylheptanoic acid (126) [251].
Cyclic depsipeptides named halipeptin A and B were found in extracts from the marine sponge Haliclona sp. (see Figure 33). (3R,4R,7S)-3,7-Dihydroxy-2,2,4-trimethyldecanoic acid (127) was present in halipeptin B and C, and (3R,4R,7S)-3-hydroxy-7-methoxy-2,2,4-trimethyldecanoic acid (128) was found in halipeptin A and D [252].
Cytotoxic depsipeptides, seragamides A–F, containing (2R,6S,8R,E)-8-Hydroxy-2,4,6-trimethylnon-4-enoic acid (129) were detected in the lipid extracts in the Okinawan sponge Suberites japonicus. The same FA has been found in jasplakinolide D, M, Q, and R1, as well as in cyclic depsipeptides named geodiamolides J, P, and R, which have been isolated from the marine sponge Cymbastela sp. and found in geodiamolides A and B from the sponge Geodia sp., and geodiamolide D from the sponge Pseudoaxinyssa sp. [253].
A cyclic depsipeptide, Jaspamide (jasplakinolide), containing (2R,6S,8S,E)-8-hydroxy-2,4,6-trimethylnon-4-enoic acid (130) was found in the lipid fraction of Fijian sponges of the genus Jaspis [254], and a similar peptide was found in other types of sponges [255]. The cytotoxic peptides, jaspamide and geodiamolide TA containing (E)-8-hydroxy-2,4,6-trimethylnon-4-enoic acid (131), found in the lipid fraction of the sponge Hemiasterella, while lipopeptides the geodiamolides J, K, and jaspamide B containing (2R,6S,8R)-8-hydroxy-2,6-dimethyl-4-methylene-5-oxononanoic acid (132) were isolated as minor metabolites from the sponge Cymbastela sp. [256].
Homophymines A–E and A1–E1 are a series of cyclodepsipeptides isolated from Homophymia sp. collected from shallow waters off the east coast of New Caledonia [257,258]. They are similar in structure to the previously published antiviral marine cyclodepsipeptides callipeltin A, neamphamide A, papuamides, theopapuamides, and mirabamides [259,260,261,262,263,264,265]. Homophymine A was cytotoxic against uninfected PBMC cells with an IC50 of 1.19 μM, but it was almost sixteen times more effective against infected cells and exhibited potent cytotoxicity with IC50 values ranging from 2 to 100 nM. These compounds were the most potent against the PC3 human prostate adenocarcinoma and the SK-OV3 human ovarian adenocarcinoma cell lines [257,258,259,260,261,262,263,264,265]. (2S,3S,4S,6S)-3-Hydroxy-2,4,6-trimethyloctanoic acid (134, for structure see Figure 34, and activity is shown in Table 13) was isolated from the homophymines A and A1, (2S,3S,4S)-3-hydroxy-2,4,6-trimethylheptanoic acid (135) from the homophymines B and B1, (2S,3S,4S,6S)-3-hydroxy-2,4,6-trimethylnonanoic acid (136) from the homophymines C and C1, (2S,3S,4S,6S)-3-hydroxy-2,4,6,8-tetramethylnonanoic acid (137) from the homophymines D and D1, and (2S,3S,4S,6S,8S)-3-hydroxy-2,4,6,8-tetramethyldecanoic acid (138) was isolated from the homophymines E and E1.
According to PASS data, among methyl-branched FA (119–133), of particular interest is FA (127). A rare feature of this acid that has been shown to be anti-psoriatic and anti-eczematic under the general concept of dermatologic activity with a high certainty of over 90%. The 3D graph of this methyl-branched FA (127) is shown in Figure 35.
(4E,6E)-2,3-Dihydroxy-2,6,8-trimethyldeca-4,6-dienoic acid (139) has been found in the depsipeptides papuamide A–D, which are produced by the sponge Theonella [262,263]. HIV-inhibitory depsipeptides, mirabamides A–D, contain (2R,3R,4R)-3-hydroxy-2,4,6-trimethyloctanoic acid (140) and were extracted from Siliquariaspongia mirabilis, while the cyclic depsipeptide neamphamide D also contains this FA and was found in the Australian marine sponge Neamphius huxleyi [266].
It is known that depsipeptides called didemnins, which are cytotoxins and immunosuppressive agents, were first isolated over 40 years ago from the Caribbean tunicate Trididemnum solidum, and contain (2S,4S)-4-hydroxy-2,5-dimethyl-3-oxohexanoic acid (141) [267]. However, recent data indicate that didemnins do not synthesize tunicate, but rather the symbiotic bacteria Tistrella mobilis [268]. These the symbiotic bacteria of the genus Tistrella have been found in marine sponges and appear to synthesize depsipeptides like didemnins [269].
Cytotoxic undecapeptides, theopapuamides and celebesides A–C from the sponge Theonella swinhoei, showed anticancer activity against HCT–116 cells (colon cancer) [270]. (2R,3R)–3–Hydroxy–2,4,6–trimethyloctanoic acid (142) was present in the undecapeptides theopapuamide A–D, (2E,4E,7S,8R,9S,10R)–7,9–dihydroxy–8,10-dimethyltrideca–2,4–dienoic acid (143) was isolated from celebeside A and C, and (2E,4E,7S,8R,9S,10R)–7,9–dihydroxy–8,10–dimethyldodeca–2,4–dienoic acid (144) was found in celebeside B [264]. Cytotoxic cyclic peptides, aciculitins A–C, were found in the active lipid fraction of the lithistid sponge Aciculites orientalis. Aciculitin A, containing FA (145) and FA (146), was present in aciculitin B, and aciculitin C contains FA (147) [271].
According to PASS data, among the group of fatty acids (134–147), glycosidic FA (145, 146 and 147) are of the greatest interest, which demonstrate anti-infective and antineoplastic activities with a high degree of confidence, more than 93%. Figure 36 demonstrates the 3D graph of the activities of these acids.
A potent cytotoxin, psymberin, also known as irciniastatin A, is found in the sponge Psammocinia sp. with (2S)-2-hydroxy-3-methoxy-5-methylhex-5-enoic acid (148, for structure see Figure 37, and predicted activity is shown in Table 14) [272,273,274,275] and the keto analogue irciniastatin B was isolated from Ircinia ramosa (Borneo) and contains 2-hydroxy-3-methoxy-5-methylhex-5-enoic acid (149) [276,277].
Marine sponges belonging to the Jaspidae family produce related bioactive lipopeptides, bengamides [278,279,280,281,282]. Thus, bengamides AE, G, H, J, L, M, O, Y, and Z contain (2R,3R,4S,5R,E)-3,4,5-trihydroxy-2-methoxy-8-methylnon-6-enoic acid (150), bengamides E’ and F’ contain (2R,3R,4S,5R,E)-3,4,5-trihydroxy-2-methoxy-8-methyldec-6-enoic acid (151), bengamides P and Q contain (2R,3R,4R,5R,E)-3,4-dihydroxy-2-methoxy-8-methyl-5-(tetradecanoyloxy)non-6-enoic acid (152), and (2R,3R,4R,5R,E)-3,4-dihydroxy-2-methoxy-8-methyl-5-(palmitoyloxy)non-6-enoic acid (153, activity see in Figure 38) was found in bengamide R [278,279,280,281,282,283].
(6S,7S,E)-7-Hydroxy-4,4,6,8-tetramethyl-5-oxonon-2-enoic acid (154) was incorporated into a macrocyclic lactam, mirabalin, which was found and isolated from extracts of Siliquariaspongia mirabilis. Mirabalin is known to inhibit the growth of the HCT-116 cell line [284,285,286]. Poecillastrin C and D are isolated from the deep-sea sponge, Japsis serpentine [287,288]. These compounds showed potent cytotoxicity against various tumor cell lines, and both poecillastrins contain (E)-7-hydroxy-4,4,6,8-tetramethyl-5-oxonon-2-enoic acid (153) [287].
An aqueous extract of the marine sponge Chondropsis sp. contains several macrolides called chondropsins [289]. Chondropsin A, B, and D and deoxychondropsin A contain (E)-7-hydroxy-9-methoxy-4,4,6,8,8-pentamethyl-5,9-dioxonon-2-enoic acid (156), and (E)-7-hydroxy-4,4,6,8-tetramethyl-5-oxonon-2-enoic acid (155) was found in chondropsin C [289].
Cytotoxic peptides, theopapuamides A–D, which contain (2R,3R)-3-hydroxy-2,4,6-trimethyloctanoic acid (157) were obtained from Theonella swinhoei sponge extracts [290], while geodiamolide TA was isolated from the marine sponge Hemiasterella minor, which also contained the same FA [291].
The cyclodepsipeptides named homophymines A–E and A1–E1 were obtained from lipid extracts of the marine sponge Homophymia sp. living in the island of Barneo [257,258]. All the members described so far exhibit potent cytotoxic activity. (2R,3R,4R,6R)-3-Hydroxy-2,4,6,8-tetramethylnonanoic acid (158) is found in homophymines B and B1, (2R,3R,4R,6R,8R)-3-hydroxy-2,4,6,8-tetramethyldecanoic acid (159) in homophymines A and A1. Homophymines C and C1 contain (2R,3R,4R,6R)-3-hydroxy-2,4,6-trimethylnonanoic acid (160), homophymines D and D1 contain (2R,3R,4R,6R)-3-hydroxy- 2,4,6,8-tetramethylnonanoic acid (161), and homophymines E and E1 contain (2R,3R,4R,6R)-3-hydroxy-2,4,6,9-tetramethyldecanoic acid (162) [257,258].
The sponge Homophymia lamellosa from the coast of Madagascar yielded cytotoxic cyclic depsipeptides, pipecolidepsins. Both pipecolidepsins A and B contain FA (134), and 3-hydroxy-2,4,6,8-tetramethylnonanoic acid (163) is found in pipecolidepsin C [292].
(2E,4E,7R,8S,9S,10S)-9-Hydroxy-7-methoxy-8,10-dimethyltrideca-2,4-dienoic acid (164) is incorporated into depsipeptide nagahamide A, which demonstrated antibacterial properties and was found in a water–methanol extract of the marine sponge Theonella swinhoei [293].

3.2.2. Chlorinated Fatty Acids Derived from Sponge Lipopeptides

It is known that chlorinated fatty acids are widely distributed in nature and are part of neutral lipids, phospholipids, and glycolipids, and are also found in natural lipopeptides of marine invertebrates [203]. Polychlorinated peptides from Lamellodysidea herbacea such as dysidin and dysidenin contain (S)-4,4,4-trichloro-3-methylbutanoic acid (165, for structure see Figure 39, and activity is shown in Table 15) [294]. Five dysideaprolines A–F contain FA (166 and 167), and barbaleucamides A and B, which contain (E)-6,6,6-trichloro-3-methoxy-5-methylhex-2-enoic acid (168), were obtained from the Philippines sponge Dysidea sp. (E)-6,6,6-Trichloro-5-methylhex-2-enoic acid, 169) or herbacic acid is the major trichloroleucine metabolite of herbaceamide A in the sponge Dysidea herbacea [294,295] and FA (170) was isolated from chlorinated lipopeptides found in the marine sponge Dysidea sp.
Analyzing the PASS data, all chlorine-containing FA (165–170) show the dominant property of anticancer activity with varying degrees of reliability. Strong antitumor activity is characteristic of 165 and 170 acids. Figure 40 demonstrates the 3D graph which shows the predicted and calculated antitumor activity of fatty acid (170).

3.2.3. Miscellaneous Fatty Acids Incorporated into Sponge Lipopeptides

A cytotoxic cyclic didepsipeptide named arenastatin A containing (5S,6S,7S,E)-6-hydroxy-5,6-dimethyl-7-(3-phenyloxiran-2-yl)oct-2-enoic acid (171, for structure see Figure 41, and activity is shown in Table 16) was found in a chloroform–methanol extract of the Okinawan marine sponge Dysidea arenaria [296,297,298,299]. This cyclodepsipeptide has an extremely strong cytotoxic activity against KB 3-1 cells (human epidermoid carcinoma cell line) [300,301,302,303].
The cytotoxic compounds onnamide A, B, and C were obtained from marine sponge Theonella sp. Onnamide A contains (S)-2-hydroxy-2-((2R,5R,6R)-2-methoxy-5,6-dimethyl-4-methylene-tetrahydro-2H-pyran-2-yl)acetic acid (172), onnamide B contains (S)-2-hydroxy-2-((2S,5R,6R)-2-methoxy-5,6-dimethyl-4-methylenetetrahydro-2H-pyran-2-yl)acetic acid (173), and onnamide C contains (R)-2-hydroxy-2-((2S,5R,6R)-2-methoxy-5,6-dimethyl-4-methylenetetrahydro-2H-pyran-2-yl)acetic acid (174). Onnamide A analogues, 21,22-dihydroxyonnamides A1–A4, containing FA (172), were isolated from an Okinawan collection of Theonella swinhoei [304,305,306]. A cyclic peptide oriamide with 2,5-dihydroxybenzoic acid (175), was detected in the marine sponge Theonella sp. collected in Sodwana Bay [307].
It is known that dysinosin A is an inhibitor of Factor VIIa and thrombin and is produced by the Australian sponge of the family Dysideidae, and contains a sulfated glyceric acid, (R)-2-methoxy-3-(sulfooxy)-propanoic acid (176), as its analogues dysinosins B and C contain this FA (176) [308,309].
The cyclic peptide, scleritodermin A with sodium (S)-(1-carboxy-2-methoxyethyl)-sulfamate (177) inhibited tubulin polymerization and showed significant in vitro cytotoxicity against human tumor cell lines [310] and was isolated from the lithistid sponge Scleritoderma nodosum. The bioactive hexapeptide, keramamide A, from the Okinawan marine sponge Theonella sp. contains (R)-3-formamido-2-hydroxypropanoic acid (178), which was also found in keramamides A, J, K, H, and G [311].
(4R,5S,6E,11E)-5-Hydroxy-4,7,9,11-tetramethyl-12-(oxazol-5-yl)-3-oxododeca-6,11-dienoic acid (179) is incorporated into a cytotoxic lipodepsipeptide named taumycin A, which was obtained from the Madagascar sponge Fascaplysinopsis sp. [256], and the sponge Discodermia kiiensis yielded the cyclic depsipeptides, discokiolide A–C, with (E)-3-hydroxy-2-methyl-3-(2-(4-phenylbut-3-en-2-yl)-oxazol-4-yl)-propanoic acid (180) [312].
According to PASS data, among FA (171–180), tetrahydro-2H-pyran-containing FA (172, 173 and 174) are of the greatest interest, which demonstrate antineoplastic activity with a high degree of certainty, more than 96%. Figure 42 shows the 3D graph of FA (173) activity, and a single peak in the red area corresponds to strong antitumor activity.
Marine sponges, like their freshwater relatives, often contain dense and diverse microbial communities or symbionts, with many microorganisms specific to sponge hosts. Symbiont microorganisms can include bacteria, archaea, and unicellular eukaryotes (fungi and microalgae), and account for up to 40% of the volume of the sponge. These symbionts synthesize a wide variety of organic molecules, including lipopeptides, which can have a profound effect on the biology of the host sponge. To date, there is no definite answer as to who the true producer of certain organic molecules isolated from the body of sponges is. Therefore, when we say that lipopeptides are isolated from sea sponges, this does not mean that these lipopeptides were synthesized by the sponge; they can be synthesized, for example, by fungal endophytes, microalgae, or bacteria.
It is very difficult to characterize the FA composition of lipopeptides in algae and invertebrates, and especially in marine and freshwater sponges. This is since sponges and other invertebrates contain a huge pool of various symbiotic bacteria and fungi. And it is not easy to determine what contribution the symbionts of invertebrates make. It should be noted that most invertebrate lipopeptide FA contain similar fragments of bacterial FA, such as chlorine-containing, oxirane, polymethyl- or phenyl-containing FA. But there are fragments of FA that are not found in bacterial lipopeptides, such as aziridine and tetrahydro-2H-pyran-containing FA.

4. Fatty Acids Derived from Mollusca Lipopeptides

Numerous scientific publications have shown that gastropods are a rich source of bioactive compounds that include steroids, terpenoids, polyketides, FA, and lipopeptides. Many of the drugs found demonstrate anticancer, antibacterial, and antifungal properties [313,314,315,316,317,318,319].
(2E,5S,6S,7S,8E)-5,7-Dihydroxy-2,6,8-trimethyldeca-2,8-dienoic acid (181, for structure see Figure 43, activity is shown in Table 17, and picture of this mollusc is shown in Figure 43) is found in a cytotoxic depsipeptide, kulokekahilide-2, which is derived from a cephalaspidean mollusc, Philinopsis speciosa [320]. Kulolide-1 contains (S)-3-hydroxy-2,2-dimethyloct-7-enoic acid (182), kulolide-2 contains (S)-3-hydroxy-2,2-dimethyloct-7-enoic acid (183) [321,322] and kulomoopunalide-2 contains (2R,3S)-3-hydroxy-2-methyloct-7-ynoic acid (184) [323]. All depsipeptides were obtained from Ph. Speciosa [321,322,323]. Cytotoxic depsipcptides, onchidin A and B, were found in extracts of the ulmonated mollusc Onchidium sp. and contain (S)-2-hydroxy-3-methylbutanoic acid (185) and 3-hydroxy-2-methyloct-7-ynoic acid (186), respectively [324,325].
Cyclic depsipeptides, kahalalides R, S and kahalalides F, D, were isolated from the mollusc Elysia grandifolia [326], and kahalalide F was also found in the molluscs Elysia rufescens, the bivalve mollusc Spisula polynyma, and from the green alga Bryopsis sp. [327]. Kahalalide S contains 5-hydroxy-7-methyloctanoic acid (187), kahalalide F contains 5-methylhexanoic acid, and kahalalide D contains 3-hydroxy-7-methyloctanoic acid (188).
A 26-membered cyclodepsipeptide, aurilide, with (2E,5R,6R,7S,8E)-5,7-dihydroxy-2,6,8-trimethylundeca-2,8-dienoic acid (189), has been isolated from the Japanese sea hare Dolabella auricularia (Figure 44) [328], and an antineoplastic agent, dolastatin 13 with 3-hydroxy-2-methoxypropanoic acid (190) was found in the same sea hare [329,330].
(2E,4Z,10E)-15-Hydroxy-7-methoxy-2-methylhexadeca-2,4,10-trienoic acid (191) was incorporated into a cytostatic depsipeptide, dolastatin 14 from the Indian Ocean shell-less mollusc Dolabella auricularia [331], while dolastaine C was found in the Japanese sea hare D. auricularia and contains (2S,3R)-2-(dimethylamino)-3-methylpentanoic acid (192) [332]. Dolastatin H, isodolastatin H, and dolastatin 10 contain (S)-2-(dimethylamino)-3-methylbutanoic acid (193) [333]. Two 35-membered depsipeptides, dolastatin G and nordolastatin G, demonstrated strong cytotoxicity against HeLa cells. Both the depsipeptides contain (2Z,4E,7R,8S)-8-hydroxy-3-methoxy-4,7-dimethylnona-2,4-dienoic (194) and (2R,3R,7S)-3,7-dihydroxy-2,8-dimethyl-nonanoic (195) acids [334]. Two antineoplastic cyclic depsipeptides, designated dolastatin 11 and dolastatin 12, were isolated from the Indian Ocean D. auricularia, and both depsipeptides contain (3S)-2-hydroxy-3-methylpentanoic acid (196) [330,335,336], dolastatin 18 contain 2,2-dimethyl-3-oxohexanoic acid (197) [337,338,339], and the cytotoxic cyclic depsipeptide, (-)-doliculide, similar to jasplakinolide, contains (2S,3S,5S,6S,8S)-6,8-dihydroxy-2,3,5,9-tetramethyl-decanoic acid (198) [340].
According to the PASS data, among FA (181–198) isolated from mollusc lipopeptides, the most interesting is FA (183), which demonstrates anti-psoriatic activity with a high degree of certainty, more than 92%. Figure 45 shows the 3D graph of the activity of acid 183; two peaks in the red area correspond to strong anti-psoriatic and anti-eczematic activities.
Lipopeptide FA isolated from marine or freshwater molluscs are not very diverse. It is known that molluscs tend to eat microalgae, macrophytes, and algae residues where there may be various bacterial communities. The mollusc lipopeptides found in their bodies are likely to be ingested, since FA do not fundamentally differ in structure from bacterial FA.

5. Fatty Acids Derived from Tunicate Lipopeptides

Marine ascidians are considered one of the richest sources of biologically active substances, including lipopeptides [341,342,343,344,345,346,347,348,349,350]. Many biologically active compounds of marine ascidians are already at different stages of clinical and preclinical research. Various lipopeptides have antitumor, antihypertensive, antioxidant and antimicrobial properties [58,351,352,353,354]. Below we present some data on lipopeptides that are the most interesting from the point of view of medicine and pharmacology.
Sagittamide A and B have been isolated from a tropical tunicate, Dolabella auricularia (Micronesia) [355,356]; other minor congeners, sagittamides C–F were isolated from Didemnidae ascidia [355,356,357]. According to published data, sagittamides A-F each have different fatty acids (199–204, for structures see Figure 46, activity see in Table 18, and the same samples of tunicate are shown in Figure 47).
Cyclic polyether lipopeptides named bistramides B, C, D, and K, which were related to bistramide A from the ascidian Lissoclinum bistratum, were found in the extracts and their structures were determined. Bistramides A–C contain FA (205), bistramide D contains FA (206), and bistramide K contains FA (207) [358,359,360,361].
Cyclic depsipeptide from a tunicate of the genus Trididemnum, didemnins A, B and D, contain (2R,4R)-4-hydroxy-2,5-dimethyl-3-oxohexanoic (208), (2S,4R)-4-hydroxy-2,5-dimethyl-3-oxohexanoic [209], and (2R,4S)-4-hydroxy-2,5-dimethyl-3-oxohexanoic (210) acids, respectively [362,363], and two other cyclic peptides named Eudistomides A and B, which are derived from a Fijian ascidian Eudistoma sp. contain identical FAs (211) [364].
Sea squirts are sedentary and filter water containing plankton, microalgae, and bacteria. For sea squirts, the obligate cyanobacterial symbionts are Prochloron spp. Apparently, ascidian metabolites are organic molecules, including lipopeptides, which are synthesized by microorganisms associated with ascidians. Therefore, if we are talking about ascidian lipopeptides, then we mean organic metabolites isolated from the body of ascidians and nothing more.

6. Fatty Acids Incorporated into Actinomycete and Fungal Lipopeptides

Marine fungi, fungal endophytes and fungi growing in other ecosystems synthesize linear and/or cyclic lipopeptides. These organisms are an inexhaustible source of new biologically active compounds. These compounds are unique because the aquatic environment requires many specific and potent biologically active molecules. Various lipopeptides have been discovered with a wide spectrum of biological activity, including antimicrobial, antitumor and antiviral activity, and toxins [365,366,367,368].
An Australian marine-derived fungus, Acremonium sp. (MST-MF588a), yielded a family of lipodepsipeptides, acremolides A–D [369]. 3,5,11-trihydroxy-2,6-dimethyl-dodecanoic FA (212, for structures see Figure 48, activity is shown in Table 19) was incorporated into acremolides A, C, and D, and 3,5-dihydroxy-2,6-dimethyl-11-oxododecanoic FA (213) was found in the structure of acremolide B.
As inhibitors of topoisomerases, the cyclic lipopeptides fusaristatins A and B were isolated from rice cultures of endophytic fungus Fusarium sp. YG-45. An unusual (8E,10E)-3-hydroxy-2,6,10,14-tetramethyl-7-oxoicosa-8,10-dienoic acid (214) was found in cyclic lipopeptides of fusaristatin A and B, as well as the linear lipopeptide YM 170320. Figure 49 shows the 3D graph of the predicted and calculated activity of this FA [370,371].
Rakicidins A–D are the 15-membered cytotoxic depsipeptides produced by the actinomycetes Micromonospora and Streptomyces [372,373], while rakicidin D was isolated from the culture broth of an actinomycete strain of the genus Streptomyces sp. MWW064 [374]. 3-Hydroxy-2,4,15-trimethylhexadecanoic acid (215) was found in rakicidin A, 3-hydroxy-2,4,16-trimethylheptadecanoic acid (216)—in rakicidin B, 3-hydroxy-2,4,6,8-tetramethyl-nonanoic acid (217)—in rakicidin C, and 3-hydroxy-2,4-dimethyldecanoic acid (218) was detected in rakicidin D.
The cyclic depsipeptides tumescenamides A and B were isolated from the fermentation broth of a marine bacterium, Streptomyces tumescens YM23-260 [375], and tumescenamide C was detected in a culture broth of an actinomycete Streptomyces sp. KUSC F05 [376]. (2S,4S)-2,4-dimethylheptanoic acid (219) was found in tumescenamide A and C, and (2S,4S)-2,4,6-trimethylnonanoic acid (220) was isolated from tumescenamide B.
In a mixture of cultures of fungal Emericella sp. and the actinomycete Salinispora arenicola, two cyclic depsipeptides, emericellamides A and B, were found. It is known that Salinispora arenicola was isolated from marine sediments and exhibited weak cytotoxicity against human colon cancer cells HCT-116 [377]. (2R,3R,4S)-3-hydroxy-2,4-dimethyl-decanoic acid (221) found in emericellamide A, and (2R,3R,4S,6S)-3-hydroxy-2,4,6-trimethyldodecanoic acid (222)—in emericellamides B. The cyclic hexadepsipeptides arenamides A and B are characterized by a 19-membered macrocycle with six subunits—Phe, Ala, Val, Gly, Leu, were isolated from the marine actinomycete S. arenicola. These two compounds, having an aromatic amino acid phenyl alanine in the molecule, inhibited NO production in a dose-dependent manner (2–10 μM), besides displaying weak activity against HCT116 cells [378]. The arenamide A and B blocked TNF-induced activation in a dose- and t-dependent manner with IC50 values of 3.7 and 1.7 μM respectively. Two different FA, (3R,4R)-3-hydroxy-4-methyldecanoic (223), and (3S,4S)-3-hydroxy-4-methyldecanoic (224), were isolated from arenamide A (A1) and arenamide B (A2), respectively.
Bicyclic depsipeptide antibiotics, salinamides A and B, with anti-inflammatory properties, were produced by fermentation of a specific marine actinomycete, a Streptomyces sp. (CNB-091) in saltwater-based media [379]. (2S,3S)-3-hydroxy-2,4-dimethylpentanoic acid (225) was present in both the bicyclic depsipeptides salinamides A and B.
Vinylamycin, a depsipeptide antibiotic, was isolated from the culture broth of a Streptomyces sp. It has showed antimicrobial activities against Gram-positive bacteria including methicillin-resistant Staphylococcus aureus [380]. 3-hydroxy-2-(2-hydroxyethyl)-4-methyl-decanoic acid (226) was found in vinylamycin. A cyclic depsipeptide antibiotic NA30851A useful for insecticides or microbicides, is manufactured by culturing Streptomyces sp. NA30851A (FERM P-16214) [381,382]. Two 2-hydroxy-3-methylpentanoic (227) and 3,4-dihydroxy-2,2-dimethyl-5-phenyl-pentanoic (228) FA were present in cyclic depsipeptide antibiotic NA 30851A.
A metabolite of antimycin family, JBIR-06, was isolated from Streptomyces sp. ML55, and it inhibited the expression of GRP78 induced by 2-deoxyglucose at the IC50 value of 250 nM [383]. An actinomycete, Streptomyces sp. ML55, produced the antibiotic JBIR-52 and containing the FA 4-Hydroxy-2,2-dimethyl-3-oxo-nonanoic acid (229, for structure see Figure 50, and activity see in Table 20) was found in JBIR-06, and 4-hydroxy-2-methyl-3-oxo-nonanoic acid (230) was present in JBIR-52 [384].
The depsipeptides WS9326A, WS9326C, WS9326D, and WS9326E were detected in a culture of Streptomyces sp. 9078 [385]. Phenyl-containing (E)-3-(2-((Z)-pent-1-en-1-yl)-phenyl)-acrylic acid (231) was found in all depsipeptides, and other antibiotics with unsaturated fatty acids were obtained from Streptomyces hawaiiensis culture extracts [386]. Thus, antibiotic A 54556A contains (2E,4E)-hexa-2,4-dienoic acid (232), antibitotic A 54556B contains (2E,4E,6E)-octa-2,4,6-trienoic acid (233), and four same acyl depsipeptides contain (2E,4E)-octa-2,4-dienoic (234), (2E,4E)-7-hydroxyhepta-2,4-dienoic (235), (2E,4E)-6-oxohexa-2,4-dienoic (236), and (2E,4E)-7-hydroxyocta-2,4-dienoic acids (237), respectively.
Structurally and functionally similar lipodepsipeptide antibiotics, enduracidin and ramoplanin, have been found in cultivated cultures of Streptomyces fungicidicus B5477 [387,388,389,390]. (2Z,4E)-10-Methylundeca-2,4-dienoic acid (238) was detected in the structures of enduracidin A, C, D, F and (2Z,4E)-10-methyldodeca-2,4-dienoic acid (239) was found in enduracidin B, E, and G.
Streptomyces macrosporeus ATCC 21 strain produces janiemycin as the main peptide antibiotic, which serves as a bactericidal ointment against Streptococcus pyogenes C203 and Diplococcus pneumoniae [388,389,390,391]. Janiemycin and ramoplanin A1 contain (2E,4E)-octa-2,4-dienoic acid (234), ramoplanin contains A2-(2Z,4E)-7-methylocta-2,4-dienoic acid (240), and ramoplanin A3 contains (2Z,4E)-8-methylnona-2,4-dienoic acid (241) [390].
It is known that the endophytic fungus Pestalotiopsis sp. produces the cyclopeptolide antibiotic HUN-7293 pesthivin and anti-HIV agent pesthivin DM, which are used to treat chronic inflammatory diseases and inhibit VCAM-1 expression on activated endothelial cells [392,393,394]. All antibiotics contain the same unusual (R)-4-cyano-2-hydroxybutanoic acid (242).
The lipopeptides rotihibin A and B, known as plant growth inhibitors, were obtained from Streptomyces graminofaciens 3C02 filtrate [395], and both contain (Z)-dec-2-enoic acid (243). The predicted and calculated antiviral activity of unsaturated FA (234, 237 and 243) is shown in Figure 51.
A cultivated endophytic fungus from Copris tripartitus is a producer of coprisamides A and B, which have significant activity for the induction of quinone reductase [396]. Both cyclic peptides contain (Z)-3-(2-((1Z,3E,5E)-hepta-1,3,5-trien-1-yl)-phenyl)-acrylic acid (244). A lipopeptide antibiotic related to amphomycin and named laspartomycin has a side chain in the form of (E)-13-methyltetradec-2-enoic acid (245) [397].
The Floridian marine sediment-derived fungus Microascus sp. EGM-556 produces a hybrid biosynthetic cyclodepsipeptide [398], which was structurally identical to turnagainolide A. It is known that turnagainolide A is produced in culture by a Bacillus sp. [399]. These cyclodepsipeptides, EGM-556 and turnagainolide A, contain (E)-3-hydroxy-5-phenylpent-4-enoic acid (246).
The actinomycete Streptomyces pristinaespiralis synthesizes a mixture of antibiotics called pristinamycin IA and IIA in a ratio of 30:70 [400,401]. Both antibiotics are known to inhibit protein synthesis elongation, exhibiting powerful bacteriostatic activity [402,403]. (4R,5R,E)-5-hydroxy-4,6-dimethylhept-2-enoic (247) FA was incorporated into both antibiotics IA and IIA.
Unusual guanidine-containing fatty acids (248–251) are produced by some fungus species. Monoamidocin, N-[(S)-5-guanidino-2-hydroxypentanoyl]-l-phenylalanine, is a dipeptide analogue has been isolated from Streptomyces sp. NR 0637. Monoamidocin inhibits the binding of fibrinogen to GP IIb/IIIa receptors [404]. This compound contains (S)-5-guanidino-2-hydroxypentanoic acid (248, for structure see Figure 52, and activity is shown in Table 21). The monoamidocin analogue was shown to have a 10-fold increase in activity and contains (R)-5-guanidino-2-hydroxypentanoic acid (249). The fusaricidins A, B, C and D, depsipeptide antibiotics, have been isolated as minor components from the culture broth of Bacillus polymyxa KT-8 which was obtained from the rhizosphere of garlic suffering from basal rot caused by Fusarium oxysporum. The fusaricidins B, C and D are active against fungi and Gram-positive bacteria, as well as fusaricidin A [405]. All the fusaricidins contain 15-guanidino-3-hydroxypentadecanoic acid (250).
Eulicin is a potent antibiotic against a broad range of Gram-positive and Gram-negative bacteria which was isolated from a Streptomyces sp. [406]. More recently, eulicin and its related analogues, as a muscarinic receptor antagonist, have also been isolated from a Streptomyces strain SCC 2268 [407]. Recently, it has been shown that eulicin inhibits human immunodeficiency virus infection and replication in a dose-dependent manner [408]. Both compounds contain 9-guanidinononanoic acid (251).
It is known that depsipeptides and streptogramins are potent drugs against numerous highly resistant pathogens and are used for human treatment, and one of them is virginiamycin [409,410]. This depsipeptide was obtained by fermentation of a culture of Streptomyces sp. G-89 [411], and contains (4R,5S,E)-5-hydroxy-4,6-dimethylhept-2-enoic acid (252).
Pterulamides I-VI are linear lipopeptides that were isolated from the fruiting bodies of a Malaysian fungus Pterula sp. Pterulamides I and IV are cytotoxic against the P388 cell line with IC50 values of 0.55 and 0.95 μM/mL (0.79 and 1.33 μM), respectively [412]. Two sulfur-containing acids, (E)-3-(methylsulfinyl)-acrylic (253) and (E)-3-(methylthio)-acrylic (254) acids were present in pterulamides I–VI.
The arylomycins are lipopeptide antibiotics, and they were detected in the culture filtrate and mycelium extracts of Streptomyces sp. Tu6075b. The isolated antibiotics demonstrate antimicrobial activity against Staphylococcus aureus, Escherichia coli, and Pseudomonas aeruginosa [413]. Two monoenoic fatty acids, (E)-2,4-dimethylhept-2-enoic acid (255) and (E)-11-hydroxy-14-methylpentadec-2-enoic acid (256), were isolated from arylomycin D.
Analysis of a fermentation broth of Streptomyces RK-1051 showed that the solution contains two depsipeptides that inhibit the growth of Staphylococcus pyogenes [414]. Both depsipeptides contain (2E,4E,6E,8E,10E)-dodeca-2,4,6,8,10-pentaenedioic acid (257). This fatty acid also was found in enopeptin A, which was isolated from Streptomyces griseus and showed antimicrobial activity against several microorganisms, and enopeptin A exhibited anti-bacteriophage activity [415,416]. A cyclic depsilipopeptide, colisporifungin, was isolated from a fungal isolate of Fusarium sp. and contained 3,5,11-trihydroxy-2,6-dimethyl-dodecanoic acid (258) [417].
The actinomycete Streptomyces fungicidicus has yielded three enduracidin analogues. (2Z,4E)-10-methylundeca-2,4-dienoic acid (259) was found in enduracidin and monodeschloro-enduracidin A, and (2Z,4E)-10-methyldodeca-2,4-dienoic acid (260) was present in mono-deschloroenduracidin B [418]. Enamidonin, a cyclic lipopeptide antibiotic, has been isolated from a culture broth of Streptomyces sp. 91–75 [419]. (2E,4E,9E)-13-hydroxytetradeca-2,4,9-trienoic acid (261) was found in structure of enamidonin.
Strain YL-03706F, a mutant of Candida tropicalis pK233, has produced a lipopeptide antibiotic designated YM-170320 (for structure see Figure 52) [420,421]. (8E,10E)-3-Hydroxy-2,6,10,14-tetramethyl-7-oxoicosa-8,10-dienoic acid (262) was included into the lipopeptide structure. Comparative analysis of the biological activity of this linear lipopeptide YM-170320 and FA (262) included in its structure showed that this lipopeptide demonstrates antifungal and antibacterial activity, and the same activity is demonstrated by FA (see Table 21).
Analyzing the activity of FA (248–262), which was obtained using PASS, acid (253) showed antihypertensive properties with a confidence level of more than 96%. Figure 53 presents the 3D graph which shows the predicted and calculated activity of this FA.
The fermentation broth actinomycete Streptomyces sp. contains two cyclic lipopeptides K97-0239 A and B, both having (2E,4E)-13-hydroxytetradeca-2,4-dienoic acid (263, for structure see Figure 54, and biological activity see in Table 22) [422].
Ulleungamides A and B, cyclic depsipeptides, were obtained from cultures of Streptomyces sp. Ulleungamide A with (S)-2-isopropylsuccinic acid (264) displayed growth inhibitory activity against Staphylococcus aureus and Salmonella typhimurium without cytotoxicity [423].
More than 40 years ago, Eli Lilly published the isolation of a similar depsipeptide antibiotic A54556, a complex of eight depsipeptide factors A-H, which was produced by aerobic fermentation of Streptomyces hawaiiensis NRRL 15010 [424]. The depsipeptides had promising in vitro activity against enterococci and streptococci but only moderate in vitro potency against staphylococci. A complex antibiotic A54556 depsipeptide factor contains several acids: A and F a (2E,4E,6E)-octa-2,4,6-trienoic acid (265), B-(2E,4E)-6-hydroxyhexa-2,4-dienoic acid (266), C and D-(2E,4E)-hexa-2,4-dienoic acid (267), and H-(2E,4E)-octa-2,4-dienoic acid (268).
The peptide–polyketide glycoside totopotensamide A and its aglycone totopotensamide B were detected in the fermentation broth of Streptomyces sp. 1053U. This Actinomycete was isolated from the gastropod mollusc, Lienardia totopotens, collected in the Philippines (Mactan Is., Cebu) [425]. Similar to the glycolipid, (269) and (8S,9R,10S,11R,12S)-9,11,12-trihydroxy-4,6,8,10-tetramethyl-3-oxotri-decanoic (270) FA were found in totopotensamide A and B, respectively.
(E)-6-Hydroxy-4-methylhex-2-enoic acid (271) was detected in several lipopeptides and cyclic peptides, called asperchrome A, B1, B2, B3, C, D1, D2, D3, and rerrirubin. These antibiotics are produced by Aspergillus ochraceous; fusarinine C, and N,N’N’’-triacetylfusarinin C are produced by Aspergillus fumigatus, A.nidulans and Fusarium cubense; coprogen and dimerumic acid are produced by Aspergillus terreus; fusarinine C and asperchrome F1 is produced by fungus Aureobasidium pullulans and Penicillium chrysogenum; basidiochrome is produced by Ceratobasidium cornigerum; basidiochrome, ferrirhodin, and Ceratobasidium globisporum [426].
The unique cyclic peptide thioviridamide is an apoptosis inducer found in the fermentation broth of Streptomyces olivoviridis with 2-hydroxy-2-methyl-4-oxopentanoic acid (272) [427]. The fermentation broths of the fungal strains Chalara sp. no. 22210 and Tolypocladium parasiticum 16616 contain the antifungal lipopeptides FR227673 and FR190293, with 10,12-dimethyltetradecanoic (273), and 12,14-dimethylhexadecanoic (274) FA, respectively [428].
A tripeptide, pre-sclerotiotide F, was isolated from a marine sediment-derived fungus, Aspergillus insulicola, and has showed cytotoxicity against selected cancer cells in vitro. The effects of pre-sclerotiotide F and sclerotiotide F on LPS-induced NF-κB and iNOS expression were also reported [429]. (2E,4E)-6-Oxohexa-2,4-dienoic acid (275) was present in pre-sclerotiotide F and sclerotiotide F.
The marine mangrove endophytic unidentified fungus from the South China Sea produced xyloallenoide A with (E)-3-(4-(buta-2,3-dien-1-yloxy)-phenyl)-acrylic acid (276) [430].
The marine fungus Hypoxylon oceanicum LL-15G256 is a producer of several lipodepsipeptides with antifungal activity. Compound 15G256γ contains (2S,3S,4R)-3-hydroxy-2,4-dimethyl-dodecanoic (277), 15G256d-(2R,3R,4R)-3-hydroxy-4-(hydroxymethyl)-2-methyl-dodecanoic (278), and 15G256e-(2S,3S,4R)-3-hydroxy-2,4-dimethyldecanoic (279) acids, respectively [431,432], and butanone extracts of another marine-derived fungus Beauveria feline with cytotoxic and anti-tuberculosis activity contains destruxin E chlorohydrin and pseudodestruxin C with (2S,4R)-5-chloro-2,4-dihydroxypentanoic acid (280) [433].
The fungal endophyte Metarhizium anisopliae produces several destruxins A, B, and E (DA, DB and DE). Destruxin-A4 chlorohydrin and destruxin Ed1 contain 2-hydroxy-3-((S)-oxiran-2-yl)-propanoic acid (281) [434,435,436].
The antibiotic E-64 is a thiol protease inhibitor, and its producer is Aspergillus japonicus TPR-64, which contains (2R,3R)-oxirane-2,3-dicarboxylic acid (282) [437,438] and a lipopeptide called cystargamide has been identified in the fermentation broth of the actinomycete Kitasatospora cystarginea and contains 3-heptyloxirane-2-carboxylic acid (283) [439].
The acidic lipopeptides produced by Streptomyces sp. such as CDA1b, CDA2a, CDA2b, CDA3a, CDA3b, CDA4a and CDA4b contain 3-pentyloxirane-2-carboxylic acid (284), and CDA1 and CDA2 contain 3-propyloxirane-2-carboxylic acid (285) [440,441,442]. The antitumor antibiotic carzinophilin A was derived from Streptomyces sahachiroi and azinomycin B was found in Streptomyces griseofuscus S42227, and both lipopeptides contain a 2-hydroxy-2-((S)-2-methyloxiran-2-yl)-acetic acid (286) [443]. The Streptomyces sp. strain associated with fungus-growing termites is a producer of microtermolide A, which contains a (2S,3R,4R)-3-hydroxy-2-(2-hydroxyethyl)-4-methylheptanoic acid (287) [444]. Oxirane-2,3-dicarboxylic acid (288) is part of many antibiotics that produce different fungal species, and it found in: antibiotic TMC 52A, TMC 52B, TMC 52C, rexostatine, cathestatin A, B, C, antibiotic AM 4299B, antibiotics WF 14861A, 14865A, 14865B, and antibiotic 460B [445].
Very interesting data were obtained in the analysis of the biological activity of FA (263–288). Acids containing the epoxy group 282, 283, 284, 285 and 288 have been shown to treat Myasthenia Gravis with a confidence level of 78 to 96%. This is a rare property that epoxy FA exhibit (282, 283 and 284), and their 3D activity is shown in Figure 55.
Pyrane-containing FA (289–301, for structures see Figure 56, and biological activity see in Table 23) have been found in the structures of many lipopeptides. Thus, two hexadepsipeptides were detected in a fermentation broth of Streptomyces nobilis JCM4274. Isolated hexadepsipeptides have shown an IC50 of 30 nM against human lung cancer NCI-H358 cells [446], and both compounds contained (289) FA. A fermentation broth of Streptomyces sp. contained a cyclic hexadepsipeptide antibiotic GE3 A and a linear lipopeptide GE3 B, which had the same acid (290, see 3D graph in Figure 57) [447]. Polyoxypeptins A and B with FA (291) are potent apoptosis-inducing peptides and were detected in the culture broth of Streptomyces sp. [448].
The hexadepsipeptide antibiotics named aurantimycins A, B, and C with FA (292) were detected in the mycelium of Streptomyces aurantiacus JA4570 [449,450], and another actinomycete, Streptomyces flavidovirens, produced a cyclic hexadepsipeptide antibiotic, citropeptin, and contained pyrane-containing acid (293).
The hexadepsipeptide antibiotic, azinothricin with (294) FA was detected in extracts of the culture filtrate of Streptomyces sp. X-14950 [451], and a culture of the Streptomyces karnatakensis contained a cyclic hexadepsipeptide antibiotic designated A83586C with (295) FA [452].
Glyco-hexadepsipeptide-polyketide with FA (296) named mollemycin A is produced by a marine-derived Streptomyces sp. (CMB-M0244), which was isolated from a sediment collected off South Molle Island (Queensland) [453]. The cyclic hexadepsipeptide antibiotics which contained FA (297, 3D graph of activity sees in Figure 57) were obtained from a fermentation broth of Streptomyces species (PM0895172/MTCC 684) and showed antitumor activity [454].
The cyclic hexadepsipeptide named pipalamycin with the rare FA (298) was isolated from a culture filtrate of Streptomyces sp. ML297-90F8 as an apoptosis-inducing agent [455].
Variapeptin and citropeptin were found to be hexadepsipeptide antibiotics produced by Streptomyces variabilis and Streptomyces flavidovirens, respectively. Both antibiotics were structurally related to azinothricin and A83586C, respectively. A culture of Streptomyces variabilis was also to produce a variapeptin. This antibiotic was active against Gram-positive bacteria and showed cytotoxic activity against mammalian cells [456,457], and acid (299) was present in variapeptin. Two antibacterial cyclic hexadepsipeptides named oleamycin A and B were detected in Streptomyces sp. [458] and contain pyrane-containing FA (300, see 3D graph in Figure 58)).
Verucopeptin with FA (301) is an antitumor antibiotic and was found in the culture broth of Actinomadura verrucosospora Q886-2 [459,460]. Urauchimycins A and B are antimycin antibiotics and were isolated from a fermentation broth of a Streptomyces sp. Ni-80. Both antibiotics showed inhibitory activity against of Candida albicans, and contain different FA, 2-(1,2-dihydroxypropyl)-4-methylhexanoic (302) and 2-(1,2-dihydroxypropyl)-5-methylhexanoic acid (303) [461]. Two cyclic lipopeptides, K97-0239A and B, are produced by Actinomycete Streptomyces sp., and both compounds contain (2E,4E)-13-hydroxytetradeca-2,4-dienoic acid (304) [462].
A depsipeptide SCH 58149 containing 3-hydroxy-4-methyloctanoic acid (305) was found in the organic extract of the fermentation broth of a fungus of Acremonium sp. SCH 58149, which exhibited weak activity against cholesterol ester transfer protein (CETP) with an IC50 of 50 mM [463]. Tachykinin (NK2) receptor inhibitors named SCH 378161, SCH 217048, SCH 378199, and SCH 378167 with 2-hydroxy-3-methylhexanoic (306) FA were detected in the fermentation broth of a taxonomically unidentified fungus [464].
Chlorinated polyketide peptides named peritoxins were produced only by a pathogenic fungus, Periconia circinata. For both compounds A and B, biologically inactive intermediates, N-3-(E-pentenyl)-glutaroyl-aspartate, circinatin with (E)-3-(pent-1-en-1-yl)-pentanedioic acid (307), and 7-chlorocircinatin with (2S,3R)-2-chloro-3-((E)-pent-1-en-1-yl)-pentanedioic acid (308) were detected only in the culture fluids of the Tox(+) strains, and peritoxin B contains FA (309); other toxins, peritoxin A, periconin A and B, contain the same FA (310) [465,466].
Tetrahydro-2H-pyran containing FA (289–301) isolated from fungal and bacterial lipopeptides are of considerable interest, since the lipopeptides themselves show antitumor activity, many of the acids incorporated into these lipopeptides also demonstrate antitumor activity, and some acids are both inhibitors of glycopeptide antibiotics. This is a rather rare function for FA. Figure 56 and Figure 57 show 3D graphs of some biologically active FA.
The mycoparasitic fungus Acremonium domschii (NRRL 39465) was obtained from a basidioma of Rigidoporus microsporus found on a dead branch in a Hawaiian forest. The crude EtOAc extract of solid-substrate fermentation cultures of A. domschii showed significant antiinsectan and antifungal activities, and contained four depsipeptides, named domschisins A–D. Domschisin A exhibited significant antiinsectan activity against Spodoptera frugiperda. All isolated compounds contained (2R,3S,5S,6S,11S)-3,5,11-trihydroxy-2,6-dimethyldodecanoic acid (311, for structures see Figure 59, and biological activity see in Table 24) [467]. Depsipeptide β-D-glucosyl-hydroxydestruxin B is produced by the fungus Alternaria alternata f. sp. mali, belongs to the phytotoxins and contains FA (312, 3D graph of activity is shown in Figure 60) [468].
Linear lipopeptides named curmenins contain an α-substituted β-methoxyacrylate, and fatty acids, (2E,4Z)-2,11-dimethyldodeca-2,4-dienoic (313), and (2E,4Z)-2,10-dimethyl-undeca-2,4-dienoic (314) FA, and have been isolated by several higher fungi. Both peptides were inhibitors of the mitochondrial respiratory energy metabolism [469]. Liposidomycines, complex molecules, and two derivatives such as liposomycin A contain (S,7Z,10Z)-3-hydroxyhexadeca-7,10-dienoic acid (315, for structure see Figure 61) and liposomycin K contains (S,9Z,12Z)-3-hydroxyoctadeca-9,12-dienoic acid (316) [470]. Liposidomycins A, B and C strongly inhibited peptidoglycan synthetase prepared from Escherichia coli, and these lipopeptides are synthesized by the fungus Streptomyces griseosporeus.
Several lipopeptide antibiotics, friulimicin A–D and lipopeptides A1437 A, A1437 B, A1437 E, A1437 G, were detected in extracts of Actinoplanes friuliensis. These compounds showed activity against Gram-positive bacteria, such as methicillin-resistant Staphylococcus epidermidis and Staphylococcus aureus strains [469]. Friulimicin A and A1437 A contain (Z)-11-methyldodec-3-enoic (317), friulimicin B and A1437 B-(Z)-12-methyltridec-3-enoic (318), friulimicin C and A1437 E-(Z)-10-methyldodec-3-enoic (319), and friulimicin D and A1437 G-(Z)-12-methyltetradec-3-enoic (320) FA [468]. FA (317–320) also found in amphomycin-type lipopeptide antibiotics include: amphomycin (glumamycin) [184,471,472,473,474,475,476,477,478,479,480,481,482,483,484]. Antibiotic F contains (E)-10-methyldodec-3-enoic acid (321), antibiotics G and H-(E)-12-methyl-tetradec-3-enoic acid (322) [470].
Glycinocins A–D, types of cyclolipopeptides, were isolated from the fermentation broth of an unidentified Actinomycete species [485]. Glycinocins A and D contain FA (245), glycinocin B contains (E)-14-methylpentadec-2-enoic (323), and glycinocin C contains (E)-12-methyltridec-2-enoic acid (324). A rare (R)-2-hydroxypent-4-enoic acid (325) was detected in the toxic cyclodepsipeptides named roseotoxin B and destroxin A, which are produced by the fungus Trichothecium roseum [486,487].
The marine and endophytic fungus Calcarisporium sp. strain KF525 produced calcaripeptides A, B, and C from the German Wadden Sea [488]. The calcaripeptides A and B contain (6R,9S,E)-9-hydroxy-4,6-dimethyl-3-oxodec-4-enoic acid (326), and calcaripeptide C contains (2S,4R,7S)-7-hydroxy-2,4-dimethyl-3-oxooctanoic acid (327). The cyclodepsipeptide trichomide A with (2R,4R)-2,5-dihydroxy-4-methylpentanoic acid (328) was isolated from the fermentation products of the fungus Trichothecium roseum [489].
The marine-derived fungus Penicillium purpurogenum G59 from the unidentified sponge produced antitumor lipopeptides, penicimutanin A, and penicimutanin B, and these drugs contain (E)-4,6-dimethyldodec-2-enoic acid (329) [490].
The lipopeptide topostatin with (8E,10E)-3-hydroxy-2,6,10,13-tetramethyl-7-oxoicosa-8,10-dienoic acid (330) is an inhibitor of topoisomerases and was isolated from the culture filtrate of Thermomonospora alba strain No. 1520 [491]. The cyclic lipopeptides pneumocandin A and pneumocandin B produced by the fungus Glarea lozoyensis contain (10S,12R)-10,12-dimethyltetradecanoic acid (331). [492,493,494]. The liquid culture broth of Pseudomonas sp. MF381-IODS yielded two antimicrobial peptides named pseudotrienic acid A and B, and both compounds contain (3E,5E)-7-hydroxy-4-methylhexadeca-3,5-dienoic (332) and (3E,5E)-7-hydroxy-4-methyltetradeca-3,5-dienoic (333) FA [495].
Cyclodepsipeptide derivatives named emericellamides A and B were produced by the marine-derived fungus Emericella sp. strain CNL-878 [496]. Emericellamides A, C, D, E, and F were also found in Aspergillus nidulans. Emericellamide A contains (2R,3R,4S)-3-hydroxy-2,4-dimethyldecanoic acid (334, for structures see Figure 62, and activity is shown in Table 25), and emericellamide B contains (2R,3R,4S,6S)-3-hydroxy-2,4,6-trimethyldodecanoic acid (335).
The depsipeptides named chondramides A–D were produced by several myxobacteria from the genus Chondromyces, and isolated compounds contain (E)-7-hydroxy-2,4,6-trimethyloct-4-enoic acid (336) [497]. The fungus Schizosaccharomyces pombe produced the antitumor antibiotic depsipeptide, FK228, and this drug contained an unusual (R)-3-hydroxy-7-mercaptoheptanoic acid (337) and (R)-3-hydroxy-7-(methylthio)-heptanoic acid (338) [498].
(S,E)-3-hydroxy-7-mercaptohept-4-enoic acid (339) was present in an anticancer depsipeptide named romidepsin, which was isolated from a culture of a Gram-negative, facultative anaerobic, coccobacillus known as Chromobacterium violaceum [499].
The macrocyclic depsipeptide named ngercheumicin D with 3-hydroxy-4-(methylthio)-butanoic acid (340) produced by Photobacterium strains was active against the non-pathogenic Pseudovibrio denitrificans [500,501]. Aetherobacter yielded the cyclic peptides aetheramides A and B. Aetheramides showed cytostatic activity against human colon carcinoma (HCT-116) cells with IC50 values of 0.11 μM. Both peptides contained (4E,10E,12E,14S,15S)-14,15-dihydroxy-9-methoxy-2,4,10-trimethyl-3-oxo-15-phenyl-pentadeca-4,10,12-trienoic (341) acid and (4E,10E,12E,14R,15R)-14,15-dihydroxy-9-methoxy-2,4,10-trimethyl-3-oxo-15-phenyl-pentadeca-4,10,12-trienoic acid (342), respectively [502].
The depsipeptides known as miuraenamides A–D are produced by a slightly halophilic myxobacterial strain, SMH-27-4 [503]. All miuraenamides contain (S,E)-9-hydroxy-6-methyldec-5-enoic acid (343). Antifungal metabolites, cyrmenin A with (2E,4Z)-2,11-dimethyldodeca-2,4-dienoic acid (344) and cyrmenin B with (2E,4Z)-2-methyldodeca-2,4-dienoic acid (345), have been isolated from Cystobacter armeniaca and Archangium gephyra, respectively [504,505].
Antitumor agents BU-2867T A, B, and C with (2E,4E)-dodeca-2,4-dienoic acid (346, 3D graph sees in Figure 63), (2E,4E,8E)-tetradeca-2,4,8-trienoic acid (347), and (2E,4E)-tetradeca-2,4-dienoic acid (348) were produced by Polyangium brachysporum sp. nov [506,507]. Peptide antibiotics designated herein as BU-2867T F with (2E,4E)-deca-2,4-dienoic (349) and G with (346) FA were produced by fermentation of the Polyangium brachysporum strain K481-B101 [507].
A rare class of bicyclic depsipeptide antibiotics, desmethylsalinamide C and salinamide A, were derived from the marine Streptomyces sp. CNB-091 [508] (specimens of Streptomyces found in various ecosystems, see in Figure 64). Desmethylsalinamide C contains (2S,3R)-3-hydroxy-2,4-dimethylpentanoic acid (350) and salinamide A contains (2Z,4E)-4-methylhexa-2,4-dienoic acid (351).
Myxobacterial species such as Myxococcus xanthus and Stigmatella aurantiaca produce cyclic depsipeptides, myxochromides with an unsaturated polyketide side chain [509]. Thus, myxochromide S contains (352) acid, myxochromides A2 and S2 contain FA (353), A3 and S3—(354), and A4—(355) FA.
In this group, we included lipopeptides that are produced by Actinomycetes, fungal endophytes and fungi. A very interesting group, since these microorganisms themselves synthesize many biologically active substances. Many FAs are similar in structure and diversity to bacterial FA. Apparently, many of these microorganisms are symbionts in more complex biological structures.

7. Conclusions

The present review is devoted to an interesting topic of studying the biological activity of FAs that are part of linear and cyclic peptides produced by organisms living in both marine and freshwater habitats. The most extensively studied lipopeptides are marine and, to a lesser extent, freshwater invertebrates.
The study of FAs in various types of complex molecules such as neutral lipids, glyco- and phospholipids isolated from various organisms is of great interest to biochemists and molecular biologists due to the high biological activity of FAs. In recent years, more and more attention has been paid to lipopeptides by scientists since these complex molecules have pronounced and specific biological activities. Given such a great interest in this group of natural complex molecules, we tried to combine known and published data on the activity of lipopeptides and their constituent fragments. This review presents the biological activity of both individual lipopeptides, and more than 350 FAs incorporated into these molecules.
Based on the presented data on the biological activity of lipopeptides and individual fragments such as FA, we can draw a preliminary conclusion that the study of the activity of such complex molecules as lipopeptides is already a fait accompli, however, in the future, an important area of research is to determine the biological activity of individual fragments, in particular FA, as well as amino-containing fatty acids.
Of undoubted interest are activities that demonstrate FAs incorporated into lipopeptides. So, several fatty acids showed pronounced antibacterial, antineurotic, antimicrobial, antitoxic, antifungal, or antitumor activity. In addition, FAs were found that exhibited rare beneficial properties, such as antiparasitic, antidiabetic, anthelmintic, anti-inflammatory, and anti-psoriasis effects. Separate fatty acids were stimulants of leukopoiesis, natural killer cells, and had anti-infective action.
We must be aware that the activities of FA incorporated into lipopeptides presented in this review are computer simulations. In the pool of this program (QSAR) there are about 1,000,000 natural and synthesized complex molecules that are associated with more than 10,000 experimentally obtained biological activities. The prospects for this predictive activity (QSAR) in medicine and pharmacology are developing very rapidly, and the probability of predicting the activity of simple and complex molecules currently reaches about 90% in many cases. Currently, about 20 million organic and inorganic compounds have been synthesized, but their biological activity has not yet been determined, and using the QSAR method, several new useful active molecules can be identified to combat numerous human diseases.
For readers and researchers who are interested in PASS, they can go to the website of this program [510]. In addition, those who wish can use this program to determine any organic molecule, both of natural origin and synthetic type, may find it useful. The site also describes all the details of the PASS.

Funding

This work did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author is grateful to Tatyana A. Gloriozova (Institute of Biomedical Chemistry, Moscow, No.122030100170-5) for prompt help in determining the biological activity of lipopeptides and FA presented in this article.

Conflicts of Interest

The author declares that he has no known competing financial interest or personal relationships that could affect the work described in this article.

References

  1. Cheung, R.C.F.; Ng, T.B.; Wong, J.H. Marine peptides: Bioactivities and applications. Mar. Drugs 2015, 13, 4006–4043. [Google Scholar] [CrossRef]
  2. Youssef, F.S.; Ashour, M.L.; Singab, A.N.B.; Wink, M. A comprehensive review of bioactive peptides from marine fungi and their biological significance. Mar. Drugs 2019, 17, 559. [Google Scholar] [CrossRef] [PubMed]
  3. Ngo, D.H.; Vo, T.S.; Ngo, D.N.; Wijesekara, I.; Kim, S.K. Biological activities and potential health benefits of bioactive peptides derived from marine organisms. Int. J. Biol. Macromol. 2012, 51, 378–383. [Google Scholar] [CrossRef] [PubMed]
  4. Anjum, K.; Abbas, S.Q.; Akhter, N.; Shagufta, B.I.; Shah, S.A.A.; Hassan, S.S.U. Emerging biopharmaceuticals from bioactive peptides derived from marine organisms. Chem. Biol. Drug Des. 2017, 90, 12–30. [Google Scholar] [CrossRef] [PubMed]
  5. Ruiz-Ruiz, F.; Mancera-Andrade, E.I.; Iqbal, H.M. Marine-derived bioactive peptides for biomedical sectors: A review. Protein Pept. Lett. 2017, 24, 109–117. [Google Scholar] [CrossRef] [PubMed]
  6. Kang, H.K.; Seo, C.H.; Park, Y. Marine peptides and their anti-infective activities. Mar. Drugs 2015, 13, 618–654. [Google Scholar] [CrossRef]
  7. Kim, S.K.; Kang, K.H. Medicinal effects of peptides from marine microalgae. Adv. Food Nutr. Res. 2011, 64, 313–323. [Google Scholar]
  8. Harnedy, P.A.; Fitz Gerald, R.J. Bioactive peptides from marine processing waste and shellfish: A review. J. Functional Food. 2012, 4, 6–24. [Google Scholar] [CrossRef]
  9. Eghtedari, M.; Jafari, S.; Nowruzi, P.B. Anticancer potential of natural peptides from terrestrial and marine environments: A review. Phytochem. Lett. 2021, 42, 87–103. [Google Scholar] [CrossRef]
  10. Ahmed, S.; Mirzaei, H.; Aschner, M.; Khan, A.; Al-Harrasi, A.; Khan, H. Marine peptides in breast cancer: Therapeutic and mechanistic understanding. Biomed. Pharmacother. 2021, 142, 112038. [Google Scholar] [CrossRef]
  11. Gogineni, V.; Hamann, M.T. Marine natural product peptides with therapeutic potential: Chemistry, biosynthesis, and pharmacology. Biochim. Biophys. Acta 2018, 1862, 81–196. [Google Scholar] [CrossRef] [PubMed]
  12. Tardón, M.C.; Allard, M.; Dutoit, V.; Dietrich, P.Y.; Dietrich, P.Y.; Walker, P.R. Peptides as cancer vaccines. Curr. Opin. Pharmacol. 2019, 47, 20–26. [Google Scholar] [CrossRef] [PubMed]
  13. Klimpel, A.; Lützenburg, T.; Neundorf, I. Recent advances of anti-cancer therapies including the use of cell-penetrating peptides. Curr. Opin. Pharmacol. 2019, 47, 8–13. [Google Scholar] [CrossRef]
  14. Kurrikoff, K.; Aphkhazava, D.; Lange, Ü. The future of peptides in cancer treatment. Curr. Opin. Pharmacol. 2019, 47, 27–32. [Google Scholar] [CrossRef]
  15. Skjånes, K.; Aesoy, R.; Herfindal, L.; Skomedal, H. Bioactive peptides from microalgae. Focus on anti-cancer and immunomodulating activity. Physiol. Plantarum 2021, 173, 612–623. [Google Scholar] [CrossRef]
  16. Shannon, E.; Abu-Ghannam, N. Antibacterial derivatives of marine algae: An overview of pharmacological mechanisms and applications. Mar. Drugs 2016, 14, 81. [Google Scholar] [CrossRef]
  17. Kastin, A. Handbook of Biologically Active Peptides, 2nd ed.; Academic Press: San Diego, CA, USA, 2013. [Google Scholar]
  18. Mayer, A.M.S.; Gustafson, K.R. Marine pharmacology in 2003–2004: Antitumour and cytotoxic compounds. Eur. J. Cancer 2006, 42, 2241–2270. [Google Scholar] [CrossRef]
  19. Jin, L.; Quan, C.; Hou, X.; Fan, S. Potential pharmacological resources: Natural bioactive compounds from marine-derived fungi. Mar. Drugs 2016, 14, 76. [Google Scholar] [CrossRef]
  20. Biniarz, P.; Łukaszewicz, M.; Janek, T. Screening concepts, characterization, and structural analysis of microbial-derived bioactive lipopeptides: A review. Crit. Rev. Biotechnol. 2017, 37, 393–410. [Google Scholar] [CrossRef]
  21. Mnif, I.; Ghribi, D. Review lipopeptides biosurfactants: Mean classes and new insights for industrial, biomedical, and environmental applications. Pept. Sci. 2015, 104, 129–147. [Google Scholar] [CrossRef]
  22. Hutchinson, J.A.; Burholt, S. Peptide hormones and lipopeptides: From self-assembly to therapeutic applications. J. Pept. Sci. 2017, 23, 82–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Meena, K.R.; Sharma, A.; Kanwar, S.S. Microbial lipopeptides and their medical applications. Ann. Pharmacol. Pharm. 2017, 2, 1126. [Google Scholar]
  24. Anke, H.; Laatsch, H. Cyclic peptides and depsipeptides from fungi. In Physiology and Genetics. The Mycota (A Comprehensive Treatise on Fungi as Experimental Systems for Basic and Applied Research); Anke, T., Schüffler, A., Eds.; Springer: Cham, Switzerland, 2018; Volume 15. [Google Scholar]
  25. Rangel, M.; José Correia de Santana, C.; Pinheiro, C.; dos Anjos, L.A. Marine depsipeptides as promising pharmacotherapeutic agents. Curr. Protein Pept. Sci. 2017, 18, 72–91. [Google Scholar] [CrossRef] [PubMed]
  26. Poroikov, V.V. Computer-aided drug design: From discovery of novel pharmaceutical agents to systems pharmacology. Biochemistry 2020, 14, 216–227. [Google Scholar]
  27. Muratov, E.N.; Bajorath, J.; Sheridan, R.P.; Tetko, I.V.; Filimonov, D.; Poroikov, V.V. QSAR without borders. Chem. Soc. Rev. 2020, 49, 3525–3564. [Google Scholar] [CrossRef]
  28. Dembitsky, V.M.; Gloriozova, T.A.; Poroikov, V.V. Antitumor profile of carbon-bridged steroids (CBS) and triterpenoids. Mar. Drugs 2021, 19, 324. [Google Scholar] [CrossRef]
  29. Fogg, G.E. The comparative physiology and biochemistry of the blue-green algae. Bacteriol. Rev. 1956, 20, 148–156. [Google Scholar] [CrossRef]
  30. Echlin, P.; Morris, I. The relationship between blue-green algae and bacteria. Biol. Rev. 1965, 40, 143–184. [Google Scholar] [CrossRef]
  31. Wolk, C.P. Physiology and cytological chemistry blue-green algae. Bacteriol. Rev. 1973, 37, 32–101. [Google Scholar] [CrossRef]
  32. Whitton, B.A.; Potts, M. Introduction to the Cyanobacteria. In Ecology of Cyanobacteria II: Their Diversity in Space and Time; Whitton, B.A., Ed.; Springer Science & Business Media, B.V.: Berlin/Heidelberg, Germany, 2012; pp. 1–13. [Google Scholar]
  33. Soo, R.M.; Hemp, J.; Hugenholtz, P. Evolution of photosynthesis and aerobic respiration in the cyanobacteria. Free Radic. Biol. Med. 2019, 140, 200–205. [Google Scholar] [CrossRef]
  34. Allaby, M. The Concise Dictionary of Botany; Oxford University Press: Oxford, UK, 1992. [Google Scholar]
  35. Sánchez-Baracaldo, R.; Bianchini, G.; Wilson, J.D.; Knoll, A.H. Cyanobacteria, and biogeochemical cycles through Earth history. Trends Microbiol. 2022, 30, 143–157. [Google Scholar] [CrossRef] [PubMed]
  36. Khalifa, S.A.M.; Shedid, E.S.; Saied, E.M.; Jassbi, A.R.; Jamebozorgi, F.H.; Rateb, M.E.; Du, M.; Abdel-Daim, M.M.; Kai, G.-Y.; Al-Hammady, M.A.M. Cyanobacteria—From the oceans to the potential biotechnological and biomedical applications. Mar. Drugs 2021, 19, 241. [Google Scholar] [CrossRef] [PubMed]
  37. Mondal, A.; Bose, S.; Banerjee, S.; Patra, J.K. Marine cyanobacteria and microalgae metabolites—A rich source of potential anticancer drugs. Mar. Drugs 2020, 18, 476. [Google Scholar] [CrossRef] [PubMed]
  38. Namikoshi, M.; Rinehart, K. Bioactive compounds produced by cyanobacteria. J. Ind. Microbiol. Biotechnol. 1996, 17, 373–384. [Google Scholar] [CrossRef]
  39. Burja, A.M.; Banaigs, E.B.; Burgess, J.G.; Wright, P.C. Marine cyanobacteria—A prolific source of natural products. Tetrahedron 2001, 57, 9347–9377. [Google Scholar] [CrossRef]
  40. Tan, L.T.; Sitachitta, N.; Gerwick, W.H. The guineamides, novel cyclic depsipeptides from a Papua New Guinea collection of the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2003, 66, 764–771. [Google Scholar] [CrossRef]
  41. Singh, S.; Bhushan, N.; Kate, U.C. Banerjee. Bioactive compounds from cyanobacteria and microalgae: An Overview. Crit. Rev. Biotechnol. 2005, 25, 73–95. [Google Scholar] [CrossRef]
  42. Dembitsky, V.M.; Řezanka, T. Metabolites produced by nitrogen fixing Nostoc species. Folia Microbiol. 2005, 50, 363–391. [Google Scholar] [CrossRef]
  43. Dembitsky, V.M.; Shkrob, I.; Dor, I. Separation and identification of hydrocarbons and other volatile compounds from cultured blue-green alga Nostoc sp. by gas chromatography—Mass spectrometry using serially coupled capillary columns with consecutive nonpolar and semipolar stationary phases. J. Chromatogr. A 1999, 862, 221–229. [Google Scholar] [CrossRef]
  44. Řezanka, T.; Dembitsky, V.M. Metabolites produced by cyanobacteria belonging to several species of the family Nostocaceae. Folia Microbiol. 2006, 51, 159–182. [Google Scholar] [CrossRef]
  45. Temina, M.; Rezankova, H.; Rezanka, T.; Dembitsky, V.M. Diversity of the fatty acids of the Nostoc species and their statistical analysis. Microbiol. Res. 2007, 162, 308–321. [Google Scholar] [CrossRef] [PubMed]
  46. Schwartz, R.E.; Hirsch, C.F.; Sesin, D.F.; Flor, J.E.; Chartrain, M. Pharmaceuticals from cultured algae. J. Ind. Microbiol. 1990, 5, 113–124. [Google Scholar] [CrossRef]
  47. Trimurtulu, G.; Ohtani, I.; Patterson, G.M.L.; Moore, R.E.; Corbett, T.H.; Valeriote, F.A.; Demchik, L. Total structures of cryptophycins, potent antitumor depsipeptides from the blue-green alga Nostoc sp. strain GSV 224. J. Am. Chem. Soc. 1994, 116, 4729–4737. [Google Scholar] [CrossRef]
  48. Mooberry, S.L.; Busquets, L.; Tien, G. Induction of apoptosis by cryptophycin 1, a new antimicrotubule agent. Int. J. Cancer 1997, 73, 440–448. [Google Scholar] [CrossRef]
  49. Hirsch, C.F.; Liesch, J.M.; Salvatore, M.J.; Schwartz, R.E.; Sesin, D.F. Antifungal Fermentation Product and Method. U.S. Patent 4946835, 18 August 1990. [Google Scholar]
  50. Weiss, C.; Figueras, E.; Borbely, A.N.; Sewald, N. Cryptophycins: Cytotoxic cyclodepsipeptides with potential for tumor targeting. J. Pept. Sci. 2017, 23, 514–531. [Google Scholar] [CrossRef]
  51. Subbaraju, G.V.; Golakoti, T.; Patterson, G.M.L.; Moore, R.E. Three new cryptophycins from Nostoc sp. GSV 224. J. Nat. Prod. 1997, 60, 302–305. [Google Scholar] [CrossRef]
  52. Chaganty, S.; Golakoti, T.; Heltzel, C.; Moore, R.E.; Yoshida, W.Y. Isolation and structure determination of cryptophycins 38, 326, and 327 from the terrestrial cyanobacterium Nostoc sp. GSV 224. J. Nat. Prod. 2004, 67, 1403–1406. [Google Scholar] [CrossRef]
  53. Golakoti, T.; Ogino, J.; Heltzel, C.E.; Husebo, T.L.; Jensen, C.M.; Larsen, L.K.; Patterson, G.M.L.; Moore, R.E.; Mooberry, S.L.; Corbett, T.H.; et al. Structure determination, conformational analysis, chemical stability studies, and antitumor evaluation of the cryptophycins. Isolation of 18 new analogs from Nostoc sp. strain GSV 224. J. Am. Chem. Soc. 1995, 117, 12030–12049. [Google Scholar] [CrossRef]
  54. Shih, C.; Teicher, B.A. Cryptophycins: A novel class of potent antimitotic antitumor depsipeptides. Curr. Pharm. Des. 2001, 7, 1259–1276. [Google Scholar] [CrossRef]
  55. Rohr, J. Cryptophycin anticancer drugs revisited. ACS Chem. Biol. 2006, 1, 747–752. [Google Scholar] [CrossRef]
  56. Eggen, M.J.; Georg, G.I. The cryptophycins: Their synthesis and anticancer activity. Med. Res. Rev. 2002, 22, 85–101. [Google Scholar] [CrossRef] [PubMed]
  57. Costa, M.; Costa-Rodrigues, J.; Fernandes, M.H.; Barros, P.; Vasconcelos, V.; Martins, R. Marine cyanobacteria compounds with anticancer properties: A review on the implication of apoptosis. Mar. Drugs 2012, 10, 2181–2207. [Google Scholar] [CrossRef] [PubMed]
  58. Siddiq, A.; Dembitsky, V. Acetylenic anticancer agents. Anti-Cancer Agents Med. Chem. 2008, 8, 132–170. [Google Scholar] [CrossRef] [PubMed]
  59. Avato, P.; Tava, A. Rare fatty acids and lipids in plant oilseeds: Occurrence and bioactivity. Phytochem. Rev. 2022, 21, 401–428. [Google Scholar] [CrossRef]
  60. Dembitsky, V.M. Anticancer activity of natural and synthetic acetylenic lipids. Lipids 2006, 41, 883–924. [Google Scholar] [CrossRef]
  61. Dembitsky, V.M.; Levitsky, D.O. Acetylenic terrestrial anticancer agents. Nat. Prod. Commun. 2006, 1, 405. [Google Scholar] [CrossRef]
  62. Dembitsky, V.M.; Rezanka, T. Distribution of acetylenic acids and polar lipids in some aquatic bryophytes. Phytochemistry 1995, 40, 93–97. [Google Scholar] [CrossRef]
  63. Dembitsky, V.M.; Rezanka, T. Acetylenic fatty acids of the Dicranaceae. Phytochemistry 1994, 36, 685–689. [Google Scholar] [CrossRef]
  64. Cahoon, E.B.; Li-Beisson, Y. Plant unusual fatty acids: Learning from the less common. Curr. Opin. Plant Biol. 2020, 55, 66–73. [Google Scholar] [CrossRef]
  65. Dembitsky, V.M.; Levitsky, D.O.; Gloriozova, T.A.; Poroikov, V.V. Acetylenic aquatic anticancer agents and related compounds. Nat. Prod. Commun. 2006, 1, 773. [Google Scholar] [CrossRef]
  66. Kuklev, D.V.; Domb, A.J.; Dembitsky, V.M. Bioactive acetylenic metabolites. Phytomedicine 2013, 20, 1145–1159. [Google Scholar] [CrossRef] [PubMed]
  67. Shimizu, Y. Microalgal metabolites. Curr. Opin. Microbiol. 2003, 6, 236. [Google Scholar] [CrossRef]
  68. Luesch, H.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J. Apramides A−G, novel lipopeptides from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2000, 63, 1106. [Google Scholar] [CrossRef] [PubMed]
  69. Williams, P.G.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J. Micromide and guamamide:  Cytotoxic alkaloids from a species of the marine cyanobacterium Symploca. J. Nat. Prod. 2004, 67, 49. [Google Scholar] [CrossRef]
  70. Jimenez, J.I.; Scheuer, P.J. New lipopeptides from the Caribbean cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2001, 64, 200. [Google Scholar] [CrossRef]
  71. Chen, H.; Feng, Y.; Xu, Z.; Ye, T. The total synthesis and reassignment of stereochemistry of dragonamide. Tetrahedron 2005, 61, 11132. [Google Scholar] [CrossRef]
  72. Hooper, G.J.; Orjala, J.; Schatzman, R.C.; Gerwick, W.H. Carmabins A and B, new lipopeptides from the caribbean cyanobacterium Lyngbya majuscula. J. Nat. Prod. 1998, 61, 529. [Google Scholar] [CrossRef]
  73. Nogle, L.M.; Gerwick, W.H. Isolation of four new cyclic depsipeptides, antanapeptins A−D, and dolastatin 16 from a Madagascan collection of Lyngbya majuscula. J. Nat. Prod. 2002, 65, 21. [Google Scholar] [CrossRef]
  74. Luesch, H.; Pangilinan, R.Y.; Moore, R.E.; Paul, V.J. Pitipeptolides A and B, new cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2001, 64, 304. [Google Scholar] [CrossRef]
  75. Sitachitta, N.; Williamson, R.T.; Gerwick, W.H. Yanucamides A and B, two new depsipeptides from an assemblage of the marine cyanobacteria Lyngbya majuscula and Schizothrix species. J. Nat. Prod. 2000, 63, 197. [Google Scholar] [CrossRef]
  76. Xu, Z.; Peng, Y.; Ye, T. The total synthesis and stereochemical revision of yanucamide A. Org. Lett. 2003, 5, 2821. [Google Scholar] [CrossRef] [PubMed]
  77. Williams, P.G.; Yoshida, W.Y.; Quon, M.K.; Moore, R.E.; Paul, V.J. Ulongapeptin, a cytotoxic cyclic depsipeptide from a Palauan marine cyanobacterium Lyngbya sp. J. Nat. Prod. 2003, 66, 651. [Google Scholar] [CrossRef] [PubMed]
  78. Han, B.; Goeger, D.; Maier, C.S.; Gerwick, W.H. The Wewakpeptins, Cyclic depsipeptides from a Papua New Guinea collection of the marine cyanobacterium Lyngbya semiplena. J. Org. Chem. 2005, 70, 3133. [Google Scholar] [CrossRef] [PubMed]
  79. Blunt, J.W.; Copp, B.R.; Munro, M.H.G.; Northcote, P.T.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2005, 22, 15–61. [Google Scholar] [CrossRef] [PubMed]
  80. Wan, F.; Erickson, K.L. Georgamide, a new cyclic depsipeptide with an alkynoic acid residue from an Australian cyanobacterium. J. Nat. Prod. 2001, 64, 143. [Google Scholar] [CrossRef]
  81. Rodriguez, J.; Fernandez, R.; Quinoa, E.; Riguera, R.; Debitus, C.; Bouchet, P. Onchidin: A cytotoxic depsipeptide with C2 symmetry from a marine mollusc. Tetrahedron Lett. 1994, 35, 9239. [Google Scholar] [CrossRef]
  82. Fernandez, R.; Rodriguez, J.; Quiñoa, E.; Riguera, R.; Muñoz, L.; Fernandez-Suarez, M.; Debitus, C. Onchidin B:  A new cyclodepsipeptide from the mollusc Onchidium sp. J. Am. Chem. Soc. 1996, 118, 11635–11643. [Google Scholar] [CrossRef]
  83. Reese, M.T.; Gulavita, N.K.; Nakao, Y.; Hamann, M.T.; Yoshida, W.Y.; Coval, S.J.; Scheuer, P.J. Kulolide:  A cytotoxic depsipeptide from a cephalaspidean mollusk, Philinopsis speciosa. J. Am. Chem. Soc. 1996, 118, 11081–11084. [Google Scholar] [CrossRef]
  84. Nakao, Y.; Yoshida, W.Y.; Szabo, C.M.; Baker, B.J.; Scheuer, P.J. More peptides and other diverse constituents of the marine mollusk Philinopsis speciosa. J. Org. Chem. 1998, 63, 3272–3280. [Google Scholar] [CrossRef]
  85. Liu, W.-T.; Ng, J.; Meluzzi, D.; Bandeira, N.; Gutierrez, M.; Simmons, T.L.; Schultz, A.W.; Linington, R.G.; Moore, B.S.; Gerwick, W.H. Interpretation of tandem mass spectra obtained from cyclic nonribosomal peptides. Anal. Chem. 2009, 81, 4200–4209. [Google Scholar] [CrossRef]
  86. Montaser, R.; Paul, V.J.; Luesch, H. Pitipeptolides C-F, antimycobacterial cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula from Guam. Phytochemistry 2011, 72, 2068–2074. [Google Scholar] [CrossRef] [PubMed]
  87. Gunasekera, S.P.; Owle, C.S.; Montaser, R.; Luesch, H.; Paul, V.J. Malyngamide 3 and cocosamides A and B from the marine cyanobacterium Lyngbya majuscula from Cocos lagoon, Guam. J. Nat. Prod. 2011, 74, 871–876. [Google Scholar] [CrossRef] [PubMed]
  88. Boudreau, P.D.; Byrum, T.; Liu, W.T.; Dorrestein, P.C.; Gerwick, W.H. Viequeamide A, a cytotoxic member of the kulolide superfamily of cyclic depsipeptides from a marine button cyanobacterium. J. Nat. Prod. 2012, 75, 1560–1570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Almaliti, J.; Malloy, K.L.; Glukhov, E.; Spadafora, C.; Gutierrez, M.; Gerwick, W.H. Dudawalamides A−D, antiparasitic cyclic depsipeptides from the marine cyanobacterium Moorea producers. J. Nat. Prod. 2017, 80, 1827–1836. [Google Scholar] [CrossRef] [PubMed]
  90. Salvador, L.A.; Biggs, J.S.; Paul, V.J.; Luesch, H. Veraguamides A–G, cyclic hexadepsipeptides from a dolastatin 16-producing cyanobacterium Symploca cf. hydnoides from Guam. J. Nat. Prod. 2011, 74, 917–927. [Google Scholar] [CrossRef]
  91. Mevers, E.; Liu, W.T.; Engene, N.; Mohimani, H.; Byrum, T.; Pevzner, P.A.; Dorrestein, P.C.; Spadafora, C.; Gerwick, W.H. Cytotoxic veraguamides, alkynyl bromide-containing cyclic depsipeptides from the marine cyanobacterium cf. Oscillatoria margaritifera. J. Nat. Prod. 2011, 74, 928–936. [Google Scholar] [CrossRef]
  92. Sueyoshi, K.; Kudo, T.; Yamano, A.; Sumimoto, S. Odobromoamide, a terminal alkynyl bromide-containing cyclodepsipeptide from the marine cyanobacterium Okeania sp. Bull. Chem. Soc. Jpn. 2017, 90, 436–440. [Google Scholar] [CrossRef]
  93. Williams, P.G.; Yoshida, W.Y.; Quon, M.K.; Moore, R.E.; Paul, V.J. The structure of palauamide, a potent cytotoxin from a species of the marine cyanobacterium Lyngbya. J. Nat. Prod. 2003, 66, 1545–1549. [Google Scholar] [CrossRef]
  94. Bunyajetpong, S.; Yoshida, W.Y.; Sitachitta, N.; Kaya, K. Trungapeptins A−C, cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2006, 69, 1539–1542. [Google Scholar] [CrossRef]
  95. Tripathi, A.; Puddick, J.; Prinsep, M.R.; Lee, P.P.F.; Tan, L.T. Hantupeptin A, a cytotoxic cyclic depsipeptide from a Singapore collection of Lyngbya majuscula. J. Nat. Prod. 2009, 72, 29–32. [Google Scholar] [CrossRef]
  96. Arif, J.M.; Farooqui, A.; Siddiqui, M.H.; Al-Karrawi, M. Novel bioactive peptides from cyanobacteria: Functional, biochemical, and biomedical significance. Bioactive Nat. Prod. 2012, 36, 111–161. [Google Scholar]
  97. Gunasekera, S.P.; Ross, C.; Paul, V.J.; Matthew, S.; Luesch, H. Dragonamides C and D, linear lipopeptides from the marine cyanobacterium brown Lyngbya polychroa. J. Nat. Prod. 2008, 71, 887–890. [Google Scholar] [CrossRef] [PubMed]
  98. McPhail, K.L.; Correa, J.; Linington, R.G.; González, J.; Ortega-Barría, E.; Capson, T.L.; Gerwick, W.H. Antimalarial linear lipopeptides from a Panamanian strain of the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2007, 70, 984–988. [Google Scholar] [CrossRef] [PubMed]
  99. Balunas, M.J.; Linington, R.G.; Tidgewell, K.; Fenner, A.M.; Ureña, L.D.; Togna, G.D.; Kyle, D.E.; Gerwick, W.H. Dragonamide E, a modified linear lipopeptide from Lyngbya majuscula with antileishmanial activity. J. Nat. Prod. 2010, 73, 60–66. [Google Scholar] [CrossRef] [Green Version]
  100. Simmons, T.L.; Engene, N.; Ureña, L.D.; Romero, L.I.; Ortega-Barría, E.; Gerwick, L.; Gerwick, W.H. Viridamides A and B, lipodepsipeptides with anti-protozoal activity from the marine cyanobacterium Oscillatoria nigro-viridis. J. Nat. Prod. 2008, 71, 1544–1550. [Google Scholar] [CrossRef]
  101. Iwasaki, A.; Ohno, O.; Sumimoto, S.; Sud, S.; Suenaga, K. Kurahyne, an acetylene-containing lipopeptide from a marine cyanobacterial assemblage of Lyngbya sp. Commun. RSC Adv. 2014, 4, 12840–12843. [Google Scholar] [CrossRef]
  102. Dixit, R.B.; Suseela, M.R. Cyanobacteria: Potential candidates for drug discovery. Antonie van Leeuwenhoek 2013, 103, 947–961. [Google Scholar] [CrossRef]
  103. Quintana, J.; Bayona, L.M.; Castellanos, L.; Puyana, M.; Camargo, P.; Aristizábal, F.; Edwards, C.; Tabudravu, J.N.; Jaspars, M.; Ramos, F.A. Almiramide D, cytotoxic peptide from the marine cyanobacterium Oscillatoria nigroviridis. Bioorg. Med. Chem. 2014, 22, 6789–6795. [Google Scholar] [CrossRef]
  104. Tripathi, A.; Puddick, J.; Prinsep, M.R.; Lee, P.P.F.; Tan, L.T. Hantupeptins B and C, cytotoxic cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula. Phytochemistry 2010, 71, 307–311. [Google Scholar] [CrossRef]
  105. Bingnan, H.; Gross, H.; McPhail, K.L.; Goeger, D.; Maier, C.S.; Gerwick, W.H. Wewakamide A and guineamide G, cyclic depsipeptides from the marine cyanobacteria Lyngbya semiplena and Lyngbya majuscula. J. Microbiol. Biotechnol. 2011, 21, 930–936. [Google Scholar]
  106. Chai, Q.Y.; Yang, Z.; Lin, H.W.; Han, B.N. Alkynyl-containing peptides of marine origin: A review. Mar. Drugs 2016, 14, 216. [Google Scholar] [CrossRef] [PubMed]
  107. Iizuka, T.; Fudou, R.; Jojima, Y.; Ogawa, S.; Yamanaka, S.; Inukai, Y.; Ojika, M. Miuraenamides A and B, novel antimicrobial cyclic depsipeptides from a new slightly halophilic myxobacterium: Taxonomy, production, and biological properties. J. Antibiot. 2006, 59, 385–391. [Google Scholar] [CrossRef] [PubMed]
  108. Galica, T.; Borbone, N.; Mareš, J.; Kust, A.; Caso, A.; Esposito, G.; Saurav, K. Cyanochelins, an overlooked class of widely distributed cyanobacterial siderophores, discovered by silent gene cluster awakening. Appl. Environ. Microbiol. 2021, 87, e0312820. [Google Scholar] [CrossRef] [PubMed]
  109. Soria-Mercado, I.E.; Pereira, A.; Cao, Z.; Murray, T.F.; Gerwick, W.H. Alotamide A, a novel neuropharmacological agent from the marine cyanobacterium Lyngbya bouillonii. Org. Lett. 2009, 11, 4704–4707. [Google Scholar] [CrossRef]
  110. Luesch, H.L.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J.; Corbett, T.H. Total structure determination of apratoxin A, a potent novel cytotoxin from the marine cyanobacterium Lyngbya majuscula. J. Am. Chem. Soc. 2001, 123, 5418–5423. [Google Scholar] [CrossRef]
  111. Gutiérrez, M.; Suyama, T.L.; Engene, N.; Wingerd, J.S.; Matainaho, T.; Gerwick, W.H. Apratoxin D, a potent cytotoxic cyclodepsipeptide from Papua New Guinea collections of the marine cyanobacteria Lyngbya majuscula and Lyngbya sordida. J. Nat. Prod. 2008, 71, 1099–1103. [Google Scholar] [CrossRef]
  112. Choi, H.; Pereira, A.R.; Cao, Z.; Shuman, C.F.; Engene, N.; Byrum, T.; Matainaho, T.; Murray, T.F.; Mangoni, A.; Gerwick, W.H. The hoiamides, structurally intriguing neurotoxic lipopeptides from Papua New Guinea marine cyanobacteria. J. Nat. Prod. 2010, 73, 1411–1421. [Google Scholar] [CrossRef]
  113. Matthew, S.; Schupp, P.J.; Luesch, H. Apratoxin E, a cytotoxic peptolide from a Guamanian collection of the marine cyanobacterium Lyngbya bouillonii. J. Nat. Prod. 2008, 71, 1113–1116. [Google Scholar] [CrossRef]
  114. Sweeney-Jones, A.M.; Gagaring, K.; Antonova-Koch, J.; Zhou, H.; Mojib, N.; Soapi, K.; Skolnick, J.; McNamara, C.W.; Kubanek, J. Antimalarial peptide and polyketide natural products from the Fijian marine cyanobacterium Moorea producens. Mar. Drugs 2020, 18, 167. [Google Scholar] [CrossRef]
  115. Matthew, S.; Salvador, L.A.; Schupp, P.J.; Paul, V.J.; Luesch, V.J. Cytotoxic halogenated macrolides and modified peptides from the apratoxin-producing marine cyanobacterium Lyngbya bouillonii from Guam. J. Nat. Prod. 2010, 73, 1544–1552. [Google Scholar] [CrossRef]
  116. Luesch, H.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J.; Mooberry, S.L. Isolation, structure determination, and biological activity of lyngbyabellin A from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2000, 63, 611–615. [Google Scholar] [CrossRef] [PubMed]
  117. Fathoni, I.; Petitbois, J.G.; Alarif, W.M.; Abdel-Lateff, A.; Al-Lihaibi, S.S.; Yoshimura, E.; Nogata, Y.; Vairappan, C.S.; Sholikhah, E.N.; Okino, T. Bioactivities of lyngbyabellins from cyanobacteria of Moorea and Okeania genera. Molecules 2020, 25, 3986. [Google Scholar] [CrossRef] [PubMed]
  118. Han, B.; McPhail, K.L.; Gross, H.; Goeger, D.E.; Mooberry, S.L.; Gerwick, W.H. Isolation and structure of five lyngbyabellin derivatives from a Papua New Guinea collection of the marine cyanobacterium Lyngbya majuscula. Tetrahedron 2005, 61, 11723–11729. [Google Scholar] [CrossRef]
  119. Marquez, B.L.; Watts, K.S.; Yokochi, A.; Roberts, M.A.; Verdier-Pinard, P.; Jimenez, J.I.; Hamel, E.; Scheuer, P.J.; Gerwick, W.H. Structure and absolute stereochemistry of hectochlorin, a potent stimulator of actin assembly. J. Nat. Prod. 2002, 65, 866–871. [Google Scholar] [CrossRef]
  120. Petitbois, J.G.; Casalme, L.O.; Lopez, J.A.V.; Alarif, W.M.; Abdel-Late, A.; Al-Lihaibi, S.S.; Yoshimura, E.; Nogata, Y.; Umezawa, T.; Matsuda, F. Serinolamides and lyngbyabellins from an Okeania sp. cyanobacterium collected from the Red Sea. J. Nat. Prod. 2017, 80, 2708–2715. [Google Scholar] [CrossRef]
  121. Luesch, H.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J. Structurally diverse new alkaloids from Palauan collections of the apratoxin-producing marine cyanobacterium Lyngbya sp. Tetrahedron 2002, 58, 7958–7959. [Google Scholar] [CrossRef]
  122. Williams, P.G.; Luesch, H.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J. Continuing studies on the cyanobacterium Lyngbya sp.: Isolation and structure determination of 15-norlyngbyapeptin A and lyngbyabellin D. J. Nat. Prod. 2003, 66, 595–598. [Google Scholar] [CrossRef]
  123. Choi, H.; Mevers, E.; Byrun, T.; Valeriote, F.A.; Gerwick, W.H. Lyngbyabellins K–N from two Palmyra atoll collections of the marine cyanobacterium Moorea bouillonii. Eur. J. Org. Chem. 2012, 27, 5141–5150. [Google Scholar] [CrossRef]
  124. Li, W.I.; Berman, F.W.; Okino, T.; Yokokawa, F.; Shioiri, T.; Gerwick, W.H.; Murray, T.F. Antillatoxin is a marine cyanobacterial toxin that potently activates voltage-gated sodium channels. Proc. Natl. Acad. Sci. USA 2001, 98, 7599–7604. [Google Scholar] [CrossRef]
  125. Nogle, L.M.; Okino, T.; Gerwick, W.H. Antillatoxin B, a neurotoxic lipopeptide from the Marine Cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2001, 64, 983–985. [Google Scholar] [CrossRef]
  126. Li, W.I.; Marquez, B.L.; Okino, T.; Yokokawa, F.; Shioiri, T.; Gerwick, W.H.; Murray, T.F. Characterization of the preferred stereochemistry for the neuropharmacologic actions of antillatoxin. J. Nat. Prod. 2004, 67, 559–568. [Google Scholar] [CrossRef] [PubMed]
  127. Okura, K.; Matsuoka, S.; Goto, R.; Inoue, M. The twisted side chain of antillatoxin is important for potent toxicity: Total synthesis and biological evaluation of antillatoxin and analogues. Angew. Chem. Int. Ed. Engl. 2010, 49, 329–332. [Google Scholar] [CrossRef] [PubMed]
  128. Dembitsky, V.M. Natural neo acids and neo alkanes: Their analogs and derivatives. Lipids 2006, 41, 309–340. [Google Scholar] [CrossRef]
  129. Teruya, T.; Sasaki, H.; Fukazawa, H.; Suenaga, K. Bisebromoamide, a potent cytotoxic peptide from the marine cyanobacterium Lyngbya sp.: Isolation, stereostructure, and biological activity. Org. Lett. 2009, 11, 5062–5065. [Google Scholar] [CrossRef]
  130. Sasaki, H.; Teruya, T.; Fukazawa, H.; Suenaga, K. Revised structure and structureeactivity relationshipof bisebromoamide and structure of norbisebromoamide from the marine cyanobacterium Lyngbya sp. Tetrahedron 2011, 67, 990–994. [Google Scholar] [CrossRef]
  131. Nakamura, S.; Chikaike, T.; Karasawa, K.; Tanaka, N.; Yonehara, H.; Umezawa, H. Isolation and characterization of bottromycins A and B. J. Antibiot. 1965, 18, 47–52. [Google Scholar]
  132. Nakamura, S.; Tanaka, N.; Umezawa, H. Bottromycins A1 and A2 and their structures. J. Antibiot. 1966, 19, 10–12. [Google Scholar]
  133. Hata, F.; Matsumae, A.; Abe, K.; Sano, Y.; Otani, M.; Omura, S. Fermentative preparation of bottromycin, Japan Tokkyo Koho. Japanese Patent JP 47010036, 22 May 1972. [Google Scholar]
  134. Mizuno, K.; Muto, N.; Kamata, S.; Asano, K. Antibiotic, bottromycin, Japan Kokai Tokkyo Koho. Japanese Patent JP 49116297 19741106/JP 73–28677 19730312, 12 December 1974. [Google Scholar]
  135. Nagle, D.G.; Paul, V.J.; Roberts, M.A. Ypaoamide, a new broadly acting feeding deterrent from the marine cyanobacterium Lyngbya majuscula. Tetrahedron Lett. 1996, 37, 6263–6266. [Google Scholar] [CrossRef]
  136. Nagle, D.G.; Paul, V.J. Chemical defense of a marine cyanobacterial bloom. J. Exp. Mar. Biol. Ecol. 1998, 225, 29–38. [Google Scholar] [CrossRef]
  137. Nagle, D.G.; Paul, V.J. Production of secondary metabolites by filamentous tropical marine cyanobacteria: Ecological functions of the compounds. J. Phycol. 1999, 35, 1412–1421. [Google Scholar]
  138. Sueyoshi, K.; Yamada, M.; Yamano, A.; Ozaki, K.; Sumimoto, S.; Iwasaki, A.; Suenaga, K.; Teruya, T. Ypaoamides B and C, linear lipopeptides from an Okeania sp. marine cyanobacterium. J. Nat. Prod. 2018, 81, 1103–1107. [Google Scholar] [CrossRef] [PubMed]
  139. Hamada, T.; Matsunaga, S.; Yano, G.; Fusetani, N. Polytheonamides A and B, highly cytotoxic, linear polypeptides with unprecedented structural features, from the marine sponge Theonella swinhoei. J. Am. Chem. Soc. 2005, 127, 110–118. [Google Scholar] [CrossRef] [PubMed]
  140. Ogawa, H.; Iwasaki, A.; Sumimoto, S.; Kanamori, Y.; Ohno, O.; Iwatsuki, M. Janadolide, a cyclic polyketide–peptide hybrid possessing a tert-butyl group from an Okeania sp. marine cyanobacterium. J. Nat. Prod. 2016, 79, 1862–1866. [Google Scholar] [CrossRef] [PubMed]
  141. Dembitsky, V.M.; Ermolenko, E.; Savidov, N.; Gloriozova, T.A.; Proroikov, V.V. Antiprotozoal and antitumor activity of natural polycyclic endoperoxides: Origin, structures, and biological activity. Molecules 2021, 26, 686. [Google Scholar] [CrossRef]
  142. Sikorsky, T.V.; Ermolenko, E.V.; Gloriozova, T.A.; Dembitsky, V.M. Mini Review: Anticancer activity of diterpenoid peroxides. Vietnam, J. Chem. 2020, 58, 273–280. [Google Scholar] [CrossRef]
  143. Ermolenko, E.V.; Imbs, A.B.; Gloriozova, T.A.; Poroikov, V.V.; Sikorskaya, T.V.; Dembitsky, V.M. Chemical diversity of soft coral steroids and their pharmacological activities. Mar. Drugs 2020, 18, 613. [Google Scholar] [CrossRef]
  144. Dembitsky, V.M. Antitumor and hepatoprotective activity of natural and synthetic neo steroids. Prog. Lipid Res. 2020, 79, 101048. [Google Scholar] [CrossRef]
  145. Dembitsky, V.M.; Yaremenko, I.A. Stable and unstable 1,2-dioxolanes: Origin, synthesis, and biological activities. Sci. Synth. 2020, 38, 277–321. [Google Scholar]
  146. Dembitsky, V.M.; Savidov, N.; Poroikov, V.V.; Gloriozova, T.A.; Imbs, A.B. Naturally occurring aromatic steroids and their biological activities. Appl. Microbiol. Biotechnol. 2018, 102, 4663–4674. [Google Scholar] [CrossRef]
  147. Dembitsky, V.M.; Gloriozova, T.A.; Savidov, N. Steroid phosphate esters and phosphonosteroids and their biological activities. Appl. Microbiol. Biotechnol. 2018, 102, 7679–7692. [Google Scholar] [CrossRef]
  148. Vil, V.A.; Gloriozova, T.A.; Poroikov, V.V.; Terent’ev, A.O.; Savidov, N.; Dembitsky, V.M. Peroxy steroids derived from plant and fungi and their biological activities. Appl. Microbiol. Biotechnol. 2018, 102, 7657–7667. [Google Scholar] [CrossRef] [PubMed]
  149. Savidov, N.; Gloriozova, T.A.; Poroikov, V.V.; Dembitsky, V.M. Highly oxygenated isoprenoid lipids derived from fungi and fungal endophytes: Origin and biological activities. Steroids 2018, 140, 114–124. [Google Scholar] [CrossRef] [PubMed]
  150. Vil, V.; Terentev, A.O.; Al Quntar, A.A.A.; Gloriozova, T.A.; Savidov, N.; Dembitsky, V.M. Oxetane-containing metabolites: Origin, structures, and biological activities. Appl. Microbiol. Biotechnol. 2019, 103, 2449–2467. [Google Scholar] [CrossRef] [PubMed]
  151. Vil, V.A.; Gloriozova, T.A.; Terentev, A.O.; Savidov, N.; Dembitsky, V.M. Hydroperoxides derived from marine sources: Origin and biological activities. Appl. Microbiol. Biotechnol. 2019, 103, 1627–1642. [Google Scholar] [CrossRef] [PubMed]
  152. Vil, V.A.; Gloriozova, T.A.; Poroikov, V.V.; Terentev, A.O.; Savidov, N.; Dembitsky, V.M. Naturally occurring of α,β-diepoxy-containing compounds: Origin, structures, and biological activities. Appl. Microbiol. Biotechnol. 2019, 103, 3249–3264. [Google Scholar] [CrossRef] [PubMed]
  153. Vil, V.A.; Terent’ev, A.O.; Savidov, N.; Gloriozova, T.A.; Poroikov, V.V.; Pounina, T.A.; Dembitsky, V.M. Hydroperoxy steroids and triterpenoids derived from plant and fungi: Origin, structures, and biological activities. J. Steroid Biochem. Mol. Biol. 2019, 190, 76–87. [Google Scholar] [CrossRef]
  154. Iwasaki, A.; Tadenuma, T.; Sumimoto, S.; Shiota, I.; Matsubara, T. Hoshinoamides A and B, acyclic lipopeptides from the marine cyanobacterium Caldora penicillata. J. Nat. Prod. 2018, 81, 2545–2552. [Google Scholar] [CrossRef]
  155. Brumley, D.A.; Gunasekera, S.P.; Chen, Q.Y.; Paul, V.J.; Luesch, H. Discovery, total synthesis, and SAR of anaenamides A and B: Anticancer cyanobacterial depsipeptides with a chlorinated pharmacophore. Org. Lett. 2020, 22, 4235–4239. [Google Scholar] [CrossRef]
  156. Nowruzi, B.; Wahlsten, M.; Jokela, J. A report on finding a new peptide aldehyde from cyanobacterium Nostoc sp. Iran, J. Biotechnol. 2019, 17, e1853. [Google Scholar] [CrossRef]
  157. Liu, L.; Jokela, J.; Wahlsten, M.; Nowruzi, B.; Permi, P.; Zhang, Y.Z. Nostosins, trypsin inhibitors isolated from the terrestrial cyanobacterium Nostoc sp. strain FSN. J. Nat. Prod. 2014, 77, 1784–1790. [Google Scholar] [CrossRef]
  158. Hoffmann, H.; Kogler, H.; Heyse, W.; Matter, H.; Caspers, M.; Schummer, D.; Klemke-Jahn, C.; Bauer, A.; Penarier, G.; Debussche, L. Discovery, structure elucidation, and biological characterization of nannocystin A, a macrocyclic myxobacterial metabolite with potent antiproliferative properties. Angew. Chem. 2015, 54, 10145–10148. [Google Scholar] [CrossRef] [PubMed]
  159. Krastel, P.; Roggo, S.; Schirle, M.; Ross, N.T.; Perruccio, F.; Aspesi, P., Jr.; Aust, T.; Buntin, K.; Estoppey, D.; Liechty, B. Nannocystin A: An elongation factor 1 inhibitor from Myxobacteria with differential anti-cancer properties. Angew. Chem. 2015, 54, 10149–10154. [Google Scholar] [CrossRef] [PubMed]
  160. Andrianasolo, E.H.; Goeger, D.; Gerwick, W.H. Mitsoamide: A cytotoxic linear lipopeptide from the Madagascar marine cyanobacterium Geitlerinema sp. Pure Appl. Chem. 2007, 79, 593–602. [Google Scholar] [CrossRef]
  161. Sumimoto, S.; Kobayashi, M.; Sato, R.; Shinomiya, S.; Iwasaki, A.; Suda, S.; Teruya, T.; Inuzuka, T.; Ohno, O.; Suenaga, K. Minnamide A, a linear lipopeptide from the marine cyanobacterium Okeania hirsuta. Org. Lett. 2019, 21, 1187–1901. [Google Scholar] [CrossRef]
  162. Pereira, A.; Cao, Z.Y.; Murray, T.F.; Gerwick, W.H. Hoiamide A, a sodium channel activator of unusual architecture from a consortium of two Papua New Guinea cyanobacteria. Chem. Biol. 2009, 16, 893–906. [Google Scholar] [CrossRef]
  163. Tan, L.T.; Tatsufumi Okino, T.; Gerwick, W.H. Bouillonamide: A mixed polyketide–peptide cytotoxin from the marine cyanobacterium Moorea bouillonii. Mar. Drugs 2013, 11, 3015–3024. [Google Scholar] [CrossRef]
  164. Neuhof, T.; Schmieder, P.; Seibold, M.; Preussel, K.; Döhrena, H. Hassallidin B—Second antifungal member of the Hassallidin family. Bioorg. Med. Chem. Lett. 2006, 16, 4220–4222. [Google Scholar] [CrossRef]
  165. Yu, H.-B.; Glukhov, E.; Li, Y.; Iwasaki, A.; Gerwick, L.; Dorrestein, P.C.; Jiao, B.-H.; Gerwick, W.H. Cytotoxic microcolin lipopeptides from the marine cyanobacterium Moorea producens. J. Nat. Prod. 2019, 82, 2608–2619. [Google Scholar] [CrossRef]
  166. Tareq, F.S.; Lee, M.A.; Lee, H.S.; Lee, J.S.; Lee, Y.J.; Shin, H.J. Gageostatins A–C, antimicrobial linear lipopeptides from a marine Bacillus subtilis. Mar. Drugs 2014, 12, 871–885. [Google Scholar] [CrossRef] [Green Version]
  167. Iwasaki, K.; Iwasaki, A.; Sumimoto, S.; Matsubara, T.; Sato, T.; Nozaki, T.; Saito-Nakano, Y.; Suenaga, K. Ikoamide, an antimalarial lipopeptide from an Okeania sp. marine cyanobacterium. J. Nat. Prod. 2020, 83, 481–488. [Google Scholar] [CrossRef]
  168. Jiang, W.; Zhou, W.; Othman, R.; Uchida, H.; Watanabe, R.; Suzuki, T. A new malyngamide from the marine cyanobacterium Moorea producens. Nat. Prod. Res. 2018, 32, 97–104. [Google Scholar] [CrossRef]
  169. Yamanaka, K.; Reynolds, K.A.; Kerstena, R.D.; Ryan, K.S.; Gonzalez, D.J. Direct cloning and refactoring of a silent lipopeptide biosynthetic gene cluster yields the antibiotic taromycin A. Proc. Natl. Acad. Sci. USA 2014, 111, 1957–1962. [Google Scholar] [CrossRef] [PubMed]
  170. Reynolds, K.A.; Luhavaya, H.; Li, J.; Dahesh, S.; Nizet, V.; Yamanaka, K.; Moore, B.S. Isolation and structure elucidation of lipopeptide antibiotic taromycin B from the activated taromycin biosynthetic gene cluster. J. Antibiot. 2018, 71, 333–338. [Google Scholar] [CrossRef] [PubMed]
  171. Omoboye, O.O.; Geudens, N.; Duban, M.; Chevalier, M.; Flahaut, C.; Martins, J.C.; Leclère, V.; Oni, F.E.; Höfte, M. Pseudomonas sp. COW3 produces new bananamide type cyclic lipopeptides with antimicrobial activity against Pythium myriotylum and Pyricularia oryzae. Molecules 2019, 24, 4170. [Google Scholar] [CrossRef]
  172. Tareq, F.S.; Hasan, C.M.; Lee, H.S.; Lee, Y.J.; Lee, J.S.; Surovy, M.Z.; Islam, M.T.; Shin, H.J. Gageopeptins A and B, new inhibitors of zoospore motility of the phytopathogen Phytophthora capsici from a marine-derived bacterium Bacillus sp. 109GGC020. Bioorg. Med. Chem. Lett. 2015, 28, 3325–3329. [Google Scholar] [CrossRef]
  173. Chakraborty, M.; Mahmud, N.U.; Gupta, D.R.; Tareq, F.S.; Shin, H.J.; Islam, T. Inhibitory effects of linear lipopeptides from a marine Bacillus subtilis on the wheat blast fungus Magnaporthe oryzae triticum. Front. Microbiol. 2020, 11, 665. [Google Scholar] [CrossRef]
  174. Matsui, K.; Kan, Y.; Kikuchi, J.; Matsushima, K.; Takemura, M.; Maki, H.; Kozono, I.; Ueda, T.; Minagawa, K. Stalobacin: Discovery of novel lipopeptide antibiotics with potent antibacterial activity against multidrug-resistant bacteria. J. Med. Chem. 2020, 63, 6090–6095. [Google Scholar] [CrossRef]
  175. Taniguchi, M.; Nunnery, J.K.; Engene, N.; Esquenazi, E.; Byrum, T.; Dorrestein, P.C.; Gerwick, W.H. Palmyramide A, a cyclic depsipeptide from a Palmyra atoll collection of the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2010, 73, 393–398. [Google Scholar] [CrossRef]
  176. Ueoka, R.; Shinzato, N.; Kagaya, N. Pseudoalteropeptide A, a novel lipopeptide from the marine bacterium Pseudoalteromonas piscicida SWA4_PA4 isolated from marine seaweed. J. Antibiot. 2021, 74, 105–110. [Google Scholar] [CrossRef]
  177. Moss, N.A.; Seiler, G.; Leão, T.F.; Castro-Falcón, G.; Gerwick, L.; Hughes, C.C.; Gerwick, W.H. Nature’s combinatorial biosynthesis produces vatiamides A–F. Angew. Chem. Int. Ed. 2019, 58, 9027–9031. [Google Scholar] [CrossRef]
  178. Jiménez, J.I.; Vansach, T.; Yoshida, W.Y.; Sakamoto, B.; Pörzgen, P.; Horgen, F.D. Halogenated fatty acid amides and cyclic depsipeptides from an eastern Caribbean collection of the cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2009, 72, 1573–1578. [Google Scholar] [CrossRef] [PubMed]
  179. Keller, L.; Canuto, K.M.; Liu, C.; Suzuki, B.M.; Almaliti, J. Tutuilamides A–C: Vinyl-chloride-containing cyclodepsipeptides from marine cyanobacteria with potent elastase inhibitory properties. ACS Chem. Biol. 2020, 15, 751–757. [Google Scholar] [CrossRef] [PubMed]
  180. Taori, K.; Paul, V.J.; Luesch, H. Structure and activity of largazole, a potent antiproliferative agent from the Floridian marine cyanobacterium Symploca sp. J. Am. Chem. Soc. 2008, 130, 1806–1807. [Google Scholar] [CrossRef] [PubMed]
  181. Taori, K.; Matthew, S.; Rocca, J.R.; Paul, V.J.; Luesch, H. Lyngbyastatins 5–7, potent elastase inhibitors from Floridian marine cyanobacteria, Lyngbya spp. J. Nat. Prod. 2007, 70, 1593–1600. [Google Scholar] [CrossRef] [PubMed]
  182. Matthew, S.; Ross, C.; Rocca, J.R.; Paul, V.J.; Luesch, H. Lyngbyastatin 4, a dolastatin 13 analog with elastase and chymotrypsin inhibitory activity from the marine cyanobacterium Lyngbya confervoides. J. Nat. Prod. 2007, 70, 124–127. [Google Scholar] [CrossRef] [PubMed]
  183. Bodanszky, M.; Sigler, G.F.; Bodanszky, A. Structure of the peptide antibiotic amphomycin. J. Am. Chem. Soc. 1973, 95, 2352–2357. [Google Scholar] [CrossRef] [PubMed]
  184. Heinemann, B.; Kaplan, M.A.; Muir, R.D.; Hooper, I.R. Amphomycin, a new antibiotic. Antibiot. Chemother. 1953, 3, 1239–1242. [Google Scholar]
  185. Shoji, J.; Otsuka, H. Studies on tsushimycin. II. J. Antibiot. 1969, 22, 473–479. [Google Scholar] [CrossRef]
  186. Heinzelmann, E.; Berger, S.; Muller, C.; Hartner, T.; Poralla, K.; Wohlleben, W.; Schwartz, D. An acyl-CoA dehydrogenase is involved in the formation of the Dcis3 double bond in the acyl residue of the lipopeptide antibiotic friulimicin in Actinoplanes friuliensis. Microbiology 2005, 151, 1963–1974. [Google Scholar] [CrossRef]
  187. Vertesy, L.; Ehlers, E.; Kogler, H.; Kurz, M.; Meiwes, J.; Seibert, G.; Vogel, M.; Hammann, P. Friulimicins: Novel lipopeptide antibiotics with peptidoglycan synthesis inhibiting activity from Actinoplanes friuliensis sp. nov. II. Isolation and structural characterization. J. Antibiot. 2000, 53, 816–827. [Google Scholar] [CrossRef] [Green Version]
  188. Kong, F.; Carter, G.T. Structure determination of glycinocins A to D, further evidence for the cyclic structure of the amphomycin antibiotics. J. Antibiot. 2003, 56, 557–564. [Google Scholar] [CrossRef] [PubMed]
  189. Han, B.; Gross, H.; Goeger, D.E.; Mooberry, S.L.; Gerwick, W.H. Aurilides B and C, cancer cell toxins from a Papua New Guinea collection of the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2006, 69, 572–575. [Google Scholar] [CrossRef] [PubMed]
  190. Tripathi, A.; Puddick, J.; Prinsep, M.R.; Rottmann, M.; Tan, L.T. Lagunamides A and B: Cytotoxic and antimalarial cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2010, 73, 1810–1814. [Google Scholar] [CrossRef] [PubMed]
  191. Tripathi, A.; Puddick, J.; Prinsep, M.R.; Rottmann, M.; Chan, K.P.; Chen, D.K.K.; Tan, L.T. Lagunamide C, a cytotoxic cyclodepsipeptide from the marine cyanobacterium Lyngbya majuscula. Phytochemistry 2011, 72, 2369–2375. [Google Scholar] [CrossRef]
  192. Luo, D.; Putra, M.Y.; Ye, T.; Paul, V.J.; Luesch, H. Isolation, structure elucidation and biological evaluation of lagunamide D: A new cytotoxic macrocyclic depsipeptide from marine cyanobacteria. Mar. Drugs 2019, 17, 83. [Google Scholar] [CrossRef]
  193. Sueyoshi, K.; Kaneda, M.; Sumimoto, S.; Oishi, S.; Fujii, N.; Suenaga, K.; Teruya, T. Odoamide, a cytotoxic cyclodepsipeptide from the marine cyanobacterium Okeania sp. Tetrahedron 2016, 72, 5472–5478. [Google Scholar] [CrossRef]
  194. Bui, T.-H.; Wray, V.; Nimtz, M.; Fossen, T.; Preisitsch, M.; Schröder, G.; Wende, K.; Heiden, S.E.; Mundt, S. Balticidins A–D, antifungal hassallidin-like lipopeptides from the Baltic Sea cyanobacterium Anabaena cylindrica Bio33. J. Nat. Prod. 2014, 77, 1287–1296. [Google Scholar] [CrossRef]
  195. Shah, S.A.A.; Akhter, N.; Auckloo, B.N.; Khan, I.; Lu, Y.; Wang, K.; Wu, B.; Guo, Y.-W. Structural diversity, biological properties and applications of natural products from cyanobacteria. A Review. Mar. Drugs 2017, 15, 354. [Google Scholar] [CrossRef]
  196. Schmidt, Y.; van der Voort, M.; Crsemann, M.; Piel, J. Biosynthetic origin of the antibiotic cyclocarbamate brabantamide A (SB-253514) in plant-associated Pseudomonas. ChemBioChem. 2014, 15, 259–266. [Google Scholar] [CrossRef]
  197. Neuhof, T.; Schmieder, P.; Preussel, K.; Dieckmann, R.; Pham, H.; Bartl, F.; von Dohren, H. Hassallidin A, a glycosylated lipopeptide with antifungal activity from the cyanobacterium Hassallia sp. J. Nat. Prod. 2005, 68, 695–700. [Google Scholar] [CrossRef]
  198. Humisto, A.; Jokela, J.; Teigen, K.; Wahlsten, M.; Permi, P.; Sivonen, K.; Herfindal, L. Characterization of the interaction of the antifungal and cytotoxic cyclic glycolipopeptide hassallidin with sterol-containing lipid membranes. BBA–Biomembr. 2019, 1861, 1510–1521. [Google Scholar] [CrossRef] [PubMed]
  199. Ackman, R.A. Marine Biogenic Lipids, Fats & Oils; CRC Press: Boca Raton, FL, USA, 1989; Volume I. [Google Scholar]
  200. Dembitsky, V.M. Lipids of Bryophytes. Prog. Lipid Res. 1993, 32, 281–356. [Google Scholar] [CrossRef]
  201. Dembitsky, V.M. Betaine ether-linked glycerolipids: Chemistry and biology. Prog. Lipid Res. 1996, 35, 1–51. [Google Scholar] [CrossRef]
  202. Dembitsky, V.M.; Levitsky, D.O. Arsenolipids. Prog. Lipid Res. 2004, 43, 403–448. [Google Scholar] [CrossRef] [PubMed]
  203. Dembitsky, V.M.; Srebnik, M. Natural halogenated fatty acids: Their analogues and derivatives. Prog. Lipid Res. 2002, 41, 315–367. [Google Scholar] [CrossRef]
  204. Dembitsky, V.M.; Maoka, T. Allenic and cumulenic lipids. Prog. Lipid Res. 2007, 46, 328–375. [Google Scholar] [CrossRef]
  205. Kuklev, D.V.; Dembitsky, V.M. Epoxy acetylenic lipids: Their analogs and derivatives. Prog. Lipid Res. 2014, 56, 67–91. [Google Scholar] [CrossRef]
  206. Hashizume, H.; Nishimura, Y. Cyclic lipopeptide antibiotics. Stud. Nat. Prod. Chem. 2008, 35, 693–751. [Google Scholar]
  207. Dembitsky, V.M. Astonishing diversity of natural surfactants. 1. Glycosides of fatty acids and alcohols. Lipids 2004, 39, 933–953. [Google Scholar] [CrossRef]
  208. Dembitsky, V.M. Astonishing diversity of natural surfactants. 2. Polyether glycosidic ionophores and macrocyclic glycosides. Lipids 2005, 40, 219–248. [Google Scholar] [CrossRef]
  209. Dembitsky, V.M. Astonishing diversity of natural surfactants. 3. Carotenoid glycosides and isoprenoid glycolipids. Lipids 2005, 40, 535–557. [Google Scholar] [CrossRef] [PubMed]
  210. Dembitsky, V.M. Astonishing diversity of natural surfactants. 4. Fatty acid amide glycosides, their analogues, and derivatives. Lipids 2005, 40, 641–660. [Google Scholar] [CrossRef] [PubMed]
  211. Dembitsky, V.M. Astonishing diversity of natural surfactants. 5. Biological active glycosides of aromatic metabolites. Lipids 2005, 40, 869–900. [Google Scholar] [CrossRef] [PubMed]
  212. Dembitsky, V.M. Astonishing diversity of natural surfactants. 6. Biological active marine and terrestrial alkaloid glycosides. Lipids 2005, 40, 1081–1105. [Google Scholar] [CrossRef] [PubMed]
  213. Dembitsky, V.M. Astonishing diversity of natural surfactants. 7. Biological active hemi- and monoterpenoid glycosides. Lipids 2006, 41, 1–27. [Google Scholar] [CrossRef]
  214. Guschina, I.A.; Harwood, J.L. Lipids, and lipid metabolism in eukaryotic algae. Prog. Lipid Res. 2006, 45, 160–186. [Google Scholar] [CrossRef]
  215. Dembitsky, V.M.; Quntar, A.; Srebnik, M. Natural and synthetic small boron-containing molecules as potential inhibitors of bacterial and fungal quorum sensing. Chem. Rev. 2011, 111, 112–143. [Google Scholar] [CrossRef]
  216. Dembitsky, V.M. In silico prediction of steroids and triterpenoids as potential regulators of lipid metabolism. Mar. Drugs 2021, 19, 650. [Google Scholar] [CrossRef]
  217. Miyashita, K.; Mikami, N.; Hosokawa, H. Chemical, and nutritional characteristics of brown seaweed lipids: A review. J. Funct. Foods 2013, 5, 1507–1517. [Google Scholar] [CrossRef]
  218. Pounina, T.A.; Gloriozova, T.A.; Savidov, N.; Dembitsky, V.M. Sulfated and sulfur-containing steroids and their pharmacological profile. Mar. Drugs 2021, 19, 240. [Google Scholar] [CrossRef]
  219. Admassu, H.; Gasmalla, M.A.A.; Yang, R.; Zhao, W. Bioactive peptides derived from seaweed protein and their health benefits: Antihypertensive, antioxidant, and antidiabetic properties. J. Food Sci. 2018, 83, 6–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Lafarga, T.; Acién-Fernández, F.G.; Garcia-Vaquero, M. Bioactive peptides and carbohydrates from seaweed for food applications: Natural occurrence, isolation, purification, and identification. Algal Res. 2020, 48, 101909. [Google Scholar] [CrossRef]
  221. Mehdi, A.; Ali, H.; Mehdi, A. Seaweed proteins as a source of bioactive peptides. Curr. Pharm. Des. 2021, 27, 1342–1352. [Google Scholar]
  222. Bizzaro, G.; Vatland, A.K.; Pampanin, D.M. The One-Health approach in seaweed food production. Environ. Internat. 2022, 158, 106948. [Google Scholar] [CrossRef]
  223. Fusetani, N.; Matsunaga, S. Bioactive sponge peptides. Chem. Rev. 1993, 93, 1793–1806. [Google Scholar] [CrossRef]
  224. Agrawal, S.; Adholeya, A.; Deshmukh, S.K.; Agrawal, S. The pharmacological potential of non-ribosomal peptides from marine sponge and tunicates. Front. Pharmacol. 2016, 7, 333. [Google Scholar] [CrossRef]
  225. Matsunaga, S.; Fusetani, N. Nonribosomal peptides from marine sponges. Curr. Org. Chem. 2003, 7, 945–966. [Google Scholar] [CrossRef]
  226. Williams, S.L.; Smith, J.E. A global review of the distribution, taxonomy, and impacts of introduced seaweeds. Ann. Rev. Ecol. Syst. 2007, 38, 327–359. [Google Scholar] [CrossRef]
  227. Pérez, M.J.; Falqué, E.; Domínguez, H. Antimicrobial action of compounds from marine seaweed. Mar. Drugs 2016, 14, E52. [Google Scholar] [CrossRef]
  228. Maciel, E.; Costa Leal, M.; Lillebø, A.I.; Domingues, P.; Domingues, M.R.; Calado, R. Bioprospecting of marine macrophytes using MS based lipidomics as a new approach. Mar. Drugs 2016, 14, E49. [Google Scholar] [CrossRef]
  229. Plouguerné, E.; da Gama, B.A.; Pereira, R.C.; Barreto Bergter, E. Glycolipids from seaweeds and their potential biotechnological applications. Front. Cell Infect. Microbiol. 2014, 4, 174–189. [Google Scholar] [CrossRef] [PubMed]
  230. Moussavou, G.; Kwak, D.H.; Obiang Obonou, B.W. Anticancer effects of different seaweeds on human colon and breast cancers. Mar. Drugs 2014, 12, 4898–4911. [Google Scholar] [CrossRef] [PubMed]
  231. Yende, S.R.; Harle, U.N.; Chaugule, B.B. Therapeutic potential and health benefits of Sargassum species. Pharmacogn. Rev. 2014, 8, 1–7. [Google Scholar] [CrossRef] [PubMed]
  232. Iwasaki, A.; Ohno, O.; Sumimoto, S. Mebamamides A and B, cyclic lipopeptides isolated from the green alga Derbesia marina. J. Nat. Prod. 2015, 78, 901–908. [Google Scholar] [CrossRef] [PubMed]
  233. Pelay-Gimeno, M.; Tulla-Puche, J.; Albericio, F. “Head-to-Side-Chain” cyclodepsipeptides of marine origin. Mar. Drugs 2013, 11, 1693–1717. [Google Scholar] [CrossRef]
  234. Hamann, M.T.; Scheuer, P.J. Kahalalide F: A bioactive depsipeptide from the sacoglossan mollusk Elysia rufescens and the green alga Bryopsis sp. J. Am. Chem. Soc. 1993, 115, 5825–5826. [Google Scholar] [CrossRef]
  235. Hamann, M.T.; Otto, C.S.; Scheuer, P.J.; Dunbar, D.C. Kahalalides: Bioactive peptides from a marine mollusk Elysia rufescens and its algal diet Bryopsis sp. J. Org. Chem. 1996, 61, 6594–6600. [Google Scholar] [CrossRef]
  236. Goetz, G.; Nakao, Y.; Scheuer, P.J. Two acyclic kahalalides from the sacoglossan mollusk Elysia rufescens. J. Nat. Prod. 1997, 60, 562–567. [Google Scholar] [CrossRef]
  237. Dmitrenok, A.; Iwashita, T.; Nakajima, T.; Sakamoto, B.; Namikoshi, M.; Nagai, H. New cyclic depsipeptides from the green alga Bryopsis species; application of a carboxypeptidase hydrolysis reaction to the structure determination. Tetrahedron 2006, 62, 1301–1308. [Google Scholar] [CrossRef]
  238. Nakao, Y.; Fujita, M.; Warabi, K. Miraziridine A, a novel cysteine protease inhibitor from the marine sponge Theonella aff. mirabilis. J. Am. Chem. Soc. 2000, 122, 10462–10463. [Google Scholar] [CrossRef]
  239. Schaschke, N. Miraziridine A: Natures blueprint towards protease class-spanning inhibitors. Bioorg. Med. Chem. Lett. 2004, 14, 855–857. [Google Scholar] [CrossRef] [PubMed]
  240. Naganawa, H.; Usui, N.; Takita, T. S-2,3-dicarboxy-aziridine: A new metabolite from a Streptomyces. J. Antibiot. 1975, 28, 828–829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Tabares, P.; Degel, B.; Schaschke, N. Identification of the protease inhibitor miraziridine A in the Red Sea sponge Theonella swinhoei. Pharmacogn. Res. 2012, 4, 63–66. [Google Scholar]
  242. Ismail, F.M.D.; Levitsky, D.O.; Dembitsky, V.M. Aziridine alkaloids as potential therapeutic agents. Eur. J. Med. Chem. 2009, 44, 3373–3387. [Google Scholar] [CrossRef]
  243. Oettel, H.; Wilhelm, G. Wege zur chemotherapic des krebses. Arzneim.-Forsch. 1954, 4, 681–703. [Google Scholar]
  244. Jermy, A. Ribosomal origin for polytheonamides. Nat. Rev. Microbiol. 2012, 10, 802. [Google Scholar] [CrossRef]
  245. Ueoka, R.; Ise, Y.; Ohtsuka, S. Yaku’amides A and B, cytotoxic linear peptides rich in dehydroamino acids from the marine sponge Ceratopsion sp. J. Am. Chem. Soc. 2010, 132, 17692–17694. [Google Scholar] [CrossRef]
  246. Takada, K.; Choi, B.W.; Rashid, M.A. Structural assignment of poecillastrins B and C, macrolide lactams from the deep-water Caribbean sponge Poecillastra species. J. Nat. Prod. 2007, 70, 428–431. [Google Scholar] [CrossRef]
  247. Rashid, M.A.; Cantrell, C.L.; Gustafson, K.R.; Boyd, M.R. Chondropsin D, a new 37-membered-ring macrolide lactam from the marine sponge Chondropsis species. J. Nat. Prod. 2001, 64, 1341–1344. [Google Scholar] [CrossRef]
  248. Takemoto, D.; Takekawa, Y.; van Soest, R.W.M. Poecillastrin D: A new cytotoxin of the chondropsin class from marine sponge Jaspis serpentina. Biosci. Biotechnol. Biochem. 2007, 71, 2697–2700. [Google Scholar] [CrossRef]
  249. Tan, K.C.; Wakimoto, T.; Abe, I. Lipodiscamides A–C, new cytotoxic lipopeptides from Discodermia kiiensis. Org. Lett. 2014, 16, 3256–3259. [Google Scholar] [CrossRef] [PubMed]
  250. Martín, M.J.; Rodríguez-Acebes, R.; García-Ramos, Y. Stellatolides, a new cyclodepsipeptide family from the sponge Ecionemia acervus: Isolation, solid-phase total synthesis, and full structural assignment of stellatolide A. J. Am. Chem. Soc. 2014, 136, 6754–6762. [Google Scholar] [CrossRef] [PubMed]
  251. Oku, N.; Gustafson, K.R.; Cartner, L.K. Neamphamide A, a new HIV-inhibitory depsipeptide from the Papua New Guinea marine sponge Neamphius huxleyi. J. Nat. Prod. 2004, 67, 1407–1411. [Google Scholar] [CrossRef] [PubMed]
  252. Randazzo, A.; Bifulco, G.; Giannini, C. Halipeptins A and B: Two novel potent anti-inflammatory cyclic depsipeptides from the Vanuatu marine sponge Haliclona species. J. Am. Chem. Soc. 2001, 123, 10870–10876. [Google Scholar] [CrossRef] [PubMed]
  253. Tanaka, C.; Tanaka, J.; Bolland, R.F.; Marriott, G.; Higa, T. Seragamides A–F, new actin-targeting depsipeptides from the sponge Suberites japonicus Thiele. Tetrahedron 2006, 62, 3536–3542. [Google Scholar] [CrossRef]
  254. Scott, V.R.; Boehme, R.; Matthews, T.R. New class of antifungal agents: Jasplakinolide, a cyclodepsipeptide from the marine sponge, Jaspis species. Antimicrob. Agents Chemother. 1988, 32, 1154–1157. [Google Scholar] [CrossRef]
  255. Molinski, T.F. Anti-infective agents. Curr. Med. Chem. 2004, 3, 197–220. [Google Scholar]
  256. Bishara, A.; Rudi, A.; Aknin, M. Taumycins A and B, two bioactive lipodepsipeptides from the Madagascar sponge Fascaplysinopsis sp. Org. Lett. 2008, 10, 4307–4309. [Google Scholar] [CrossRef]
  257. Zampella, A.; Sepe, V.; Bellotta, F.; Luciano, P. Homophymines B E and A1 E1, a family of bioactive cyclodepsipeptides from the sponge Homophymia sp. Org. Biomol. Chem. 2009, 7, 4037–4044. [Google Scholar] [CrossRef]
  258. Zampella, A.; Sepe, V.; Luciano, P.; Bellotta, F. Homophymine A, an anti-HIV cyclodepsipeptide from the sponge Homophymia sp. J. Org. Chem. 2008, 73, 5319–5327. [Google Scholar] [CrossRef]
  259. Trevisi, L.; Bova, S.; Cargnelli, G.; Danieli Betto, D. Callipeltin A, a cyclic depsipeptide inhibitor of the cardiac sodium calcium exchanger and positive inotropic agent. Biochem. Biophys. Res. Commun. 2000, 279, 219–222. [Google Scholar] [CrossRef] [PubMed]
  260. Trevisi, L.; Cargnelli, G.; Ceolotto, G.; Papparella, I. Callipeltin A: Sodium ionophore effect and tension development in vascular smooth muscle. Biochem. Pharmacol. 2004, 68, 1331–1338. [Google Scholar] [CrossRef] [PubMed]
  261. Noro, J.C.; Kalaitzis, J.A.; Neilan, B.A. Bioactive natural products from Papua New Guinea marine sponges. Chem. Biodivers. 2012, 9, 2077–2095. [Google Scholar] [CrossRef]
  262. Ford, P.W.; Gustafson, K.R.; McKee, T.C.; Shigematsu, N. Papuamides A−D, HIV inhibitory and cytotoxic depsipeptides from the sponges Theonella mirabilis and Theonella swinhoei collected in Papua New Guinea. J. Am. Chem. Soc. 1999, 121, 5899–5909. [Google Scholar] [CrossRef]
  263. Prasad, P.; Aalbersberg, W.; Feussner, K.D.; Van Wagoner, R.M. Papuamides E and F, cytotoxic depsipeptides from the marine sponge Melophlus sp. Tetrahedron 2011, 67, 8529–8531. [Google Scholar] [CrossRef] [Green Version]
  264. Plaza, A.; Bifulco, G.; Keffer, J.L.; Lloyd, J.R. Celebesides A C and theopapuamides B D, depsipeptides from an Indonesian sponge that inhibit HIV 1 entry. J. Org. Chem. 2009, 74, 504–512. [Google Scholar] [CrossRef] [PubMed]
  265. Plaza, A.; Gustchina, E.; Baker, H.L.; Kelly, M.; Bewley, C.A. Mirabamides A D, depsipeptides from the sponge Siliquariaspongia mirabilis that inhibit HIV 1 fusion. J. Nat. Prod. 2007, 70, 1753–1760. [Google Scholar] [CrossRef]
  266. Tran, T.D.; Pham, N.B.; Fechner, G.; Zencak, D. Cytotoxic cyclic depsipeptides from the Australian marine sponge Neamphius huxleyi. J. Nat. Prod. 2012, 75, 2200–2208. [Google Scholar] [CrossRef]
  267. Rinehart, K.L.; Kishore, V.; Bible, K.C.; Sakai, R. Didemnins and tunichlorin: Novel natural products from the marine tunicate Trididemnum solidum. J. Nat. Prod. 1988, 51, 1–21. [Google Scholar] [CrossRef]
  268. Tsukimoto, M.; Nagaoka, M.; Shishido, Y.; Junji Fujimoto, J.; Fukiko Nishisaka, F. Bacterial production of the tunicate-derived antitumor cyclic depsipeptide didemnin B. J. Nat. Prod. 2011, 74, 2329–2331. [Google Scholar] [CrossRef]
  269. Ahir, U.N.; Vyas, T.K.; Gandhi, K.D.; Faldu, P.R.; Patel, K.G. In vitro efficacy for chlorpyrifos degradation by novel isolate Tistrella sp. AUC10 isolated from chlorpyrifos contaminated field. Curr. Microbiol. 2020, 77, 2226–2232. [Google Scholar] [CrossRef] [PubMed]
  270. Ratnayake, A.S.; Bugni, T.S.; Feng, X.; Harper, M.K.; Skalicky, J.J.; Mohammed, K.A.; Andjelic, C.D.; Barrows, L.R.; Ireland, C.M. Theopapuamide, a cyclic depsipeptide from a Papua New Guinea lithistid sponge Theonella swinhoei. J. Nat. Prod. 2006, 69, 1582–1586. [Google Scholar] [CrossRef] [PubMed]
  271. Bewley, C.A.; He, H.; Williams, D.H.; Faulkner, D.J. Aciculitins A−C: Cytotoxic and antifungal cyclic peptides from the lithistid sponge Aciculites orientalis. J. Am. Chem. Soc. 1996, 118, 4314–4321. [Google Scholar] [CrossRef]
  272. Uesugi, S.; Watanabe, T.; Imaizumi, T.; Ota, Y.; Yoshida, K. Total synthesis and biological evaluation of irciniastatin A (a.k.a. Psymberin) and irciniastatin B. Org. Chem. 2015, 80, 12333–123350. [Google Scholar] [CrossRef] [PubMed]
  273. Crimmins, M.T.; Stevens, J.M.; Schaaf, G.M. Total synthesis of irciniastatin A (Psymberin). Org. Lett. 2009, 11, 3990–3993. [Google Scholar] [CrossRef] [Green Version]
  274. Fisch, K.M.; Gurgui, C.; Heycke, N.; van der Sar, S.A. Polyketide assembly lines of uncultivated sponge symbionts from structure-based gene targeting. Nat. Chem. Biol. 2009, 5, 494–501. [Google Scholar] [CrossRef]
  275. Cardani, C.; Ghiringhelli, D.; Mondelli, R.; Quilico, A. The structure of pederin. Tetrahedron Lett. 1965, 6, 2537–2545. [Google Scholar] [CrossRef]
  276. Cichewicz, R.H.; Valeriote, F.A.; Crews, P. Psymberin, a potent sponge derived cytotoxin from Psammocinia distantly related to the pederin family. Org. Lett. 2004, 6, 1951–1954. [Google Scholar] [CrossRef]
  277. An, C.; Jurica, J.A.; Walsh, S.P.; Hoye, A.T.; Smith, A.B., 3rd. Total synthesis of (+) irciniastatin A (a.k.a. psymberin) and irciniastatin B. J. Org. Chem. 2013, 78, 4278–4296. [Google Scholar] [CrossRef]
  278. García Ruiz, C.; Sarabia, F. Chemistry and biology of bengamides and bengazoles, bioactive natural products from Jaspis sponges. Mar. Drugs 2014, 12, 1580–1622. [Google Scholar] [CrossRef]
  279. Quinoa, E.; Adamczeski, M.; Crews, P.; Bakus, G.J. Bengamides, heterocyclic anthelmintics from a Jaspidae marine sponge. J. Org. Chem. 1986, 51, 4494–4497. [Google Scholar] [CrossRef]
  280. White, K.N.; Tenney, K.; Crews, P. The bengamides: A mini-review of natural sources, analogues, biological properties, biosynthetic origins, and future prospects. J. Nat. Prod. 2017, 80, 740–755. [Google Scholar] [CrossRef] [PubMed]
  281. Groweiss, A.; Newcomer, J.J.; O’Keefe, B.R. Cytotoxic metabolites from an Australian collection of the sponge Jaspis species. J. Nat. Prod. 1999, 62, 1691–1693. [Google Scholar] [CrossRef]
  282. Galitz, A.; Nakao, Y.; Schupp, P.J.; Wörheide, G.; Erpenbeck, D. A soft spot for chemistry—Current taxonomic and evolutionary implications of sponge secondary metabolite distribution. Mar. Drugs 2021, 19, 448. [Google Scholar] [CrossRef] [PubMed]
  283. Porras-Alcalá, C.; Moya-Utrera, F.; García-Castro, M.; Sánchez-Ruiz, A.; López-Romero, J.M.; Pino-González, M.S.; Díaz-Morilla, A.; Kitamura, S.; Wolan, D.W.; Prados, J.; et al. The development of the bengamides as new antibiotics against drug-resistant bacteria. Mar. Drugs 2022, 20, 373. [Google Scholar] [CrossRef]
  284. Plaza, A.; Baker, H.L.; Bewley, C.A. Mirabalin, an antitumor macrolide lactam from the marine sponge Siliquariaspongia mirabilis. J. Nat. Prod. 2008, 71, 473–477. [Google Scholar] [CrossRef]
  285. Winder, P.L.; Pomponi, S.A.; Wright, A.E. Natural products from the Lithistida: A review of the literature since 2000. Mar. Drugs 2011, 9, 2643–2682. [Google Scholar] [CrossRef]
  286. Cornil, J.; Echeverria, P.G.; Reymond, S. Synthetic studies toward the C14–C29 fragment of mirabalin. Org. Lett. 2016, 18, 4534–4537. [Google Scholar] [CrossRef]
  287. Chhetri, B.K.; Lavoie, S.; Sweeney-Jones, A.M.; Kubanek, J. Recent trends in the structural revision of natural products. Nat. Prod. Rep. 2018, 35, 514–531. [Google Scholar] [CrossRef]
  288. Huss, M.; Wieczorek, H. Inhibitors of V-ATPases: Old and new players. J. Exp. Biol. 2009, 212, 341–346. [Google Scholar] [CrossRef]
  289. Cantrell, C.L.; Gustafson, K.R.; Cecere, M.R. Chondropsins A and B:  Novel tumor cell growth-inhibitory macrolide lactams from the marine sponge Chondropsis sp. J. Am. Chem. Soc. 2000, 122, 8825–8829. [Google Scholar] [CrossRef]
  290. Molinski, T.F. Cyclic azole-homologated peptides from marine sponges. Org. Biomol. Chem. 2018, 16, 21–29. [Google Scholar] [CrossRef] [PubMed]
  291. Talpir, R.; Benayahu, Y.; Kashman, Y. Hemiasterlin and geodiamolide TA.; two new cytotoxic peptides from the marine sponge Hemiasterella minor (Kirkpatrick). Tetrahedron Lett. 1994, 35, 4453–4456. [Google Scholar] [CrossRef]
  292. Coello, L.; Reyes, F.; Martín, M.J. Isolation and structures of pipecolidepsins A and B, cytotoxic cyclic depsipeptides from the Madagascan sponge Homophymia lamellose. J. Nat. Prod. 2014, 77, 298–303. [Google Scholar] [CrossRef] [PubMed]
  293. Okada, Y.; Matsunaga, S.; van Soest, R.W.; Fusetani, N. Nagahamide A, an antibacterial depsipeptide from the marine sponge Theonella swinhoei. Org. Lett. 2002, 4, 3039–3042. [Google Scholar] [CrossRef]
  294. MacMillan, J.B.; Molinski, T.F. Herbacic acid, a simple prototype of 5,5,5-trichloroleucine metabolites from the sponge Dysidea herbacea. J. Nat. Prod. 2000, 63, 155–157. [Google Scholar] [CrossRef]
  295. Lee, G.M.; Molinski, T.F. Herbaceamide, a chlorinated N-acyl amino ester from the marine sponge, Dysidea herbacea. Tetrahedron Lett. 1992, 33, 7671–7674. [Google Scholar] [CrossRef]
  296. Kobayashi, M.; Aoki, S.; Ohyabu, N. Arenastatin A, a potent cytotoxic depsipeptide from the okinawan marine sponge Dysidea arenaria. Tetrahedron Lett. 1994, 35, 7969–7972. [Google Scholar] [CrossRef]
  297. Kobayashi, M.; Wang, W.; Ohyabu, N. Improved total synthesis and structure-activity relationship of arenastatin A, a potent cytotoxic spongean depsipeptide. Chem. Pharm. Bull. 1995, 43, 1598–1600. [Google Scholar] [CrossRef]
  298. Koiso, Y.; Morita, K.; Kobayashi, M. Effects of arenastatin A and its synthetic analogs on microtubule assembly. Chem. Biol. Interact. 1996, 102, 183–191. [Google Scholar] [CrossRef]
  299. Morita, K.; Koiso, Y.; Hashimoto, Y. Interaction of arenastatin A with porcine brain tubulin. Biol. Pharm. Bull. 1997, 20, 171–174. [Google Scholar] [CrossRef] [PubMed]
  300. Murakami, N.; Tamura, S.; Koyama, K. New analogue of arenastatin A, a potent cytotoxic spongean depsipeptide, with anti-tumor activity. Bioorg. Med. Chem. Lett. 2004, 14, 2597–2601. [Google Scholar] [CrossRef] [PubMed]
  301. Murakami, N.; Wang, W.; Tamura, S.; Kobayashi, M. Synthesis, and biological property of carba and 20-deoxo analogues of arenastatin A. Bioorg. Med. Chem. Lett. 2000, 10, 1823–1826. [Google Scholar] [CrossRef]
  302. Andavan, G.S.; Lemmens Gruber, R. Cyclodepsipeptides from marine sponges: Natural agents for drug research. Mar. Drugs 2010, 8, 810–834. [Google Scholar] [CrossRef]
  303. Singh, R.; Sharma, M.; Joshi, P.; Rawat, D.S. Clinical status of anti-cancer agents derived from mar one sources. Anticancer Agents Med. Chem. 2008, 8, 603–617. [Google Scholar] [CrossRef]
  304. Sakemi, S.; Ichiba, T.; Kohmoto, S.; Saucy, G.; Higa, T. Isolation and structure elucidation of onnamide A, a new bioactive metabolite of a marine sponge, Theonella sp. J. Am. Chem. Soc. 1988, 110, 4851–4853. [Google Scholar] [CrossRef]
  305. Matsunaga, S.; Fusetani, N.; Nakao, Y. Eight new cytotoxic metabolites closely related to onnamide A from two marine sponges of the genus Theonella. Tetrahedron 1992, 48, 8369–8376. [Google Scholar] [CrossRef]
  306. Lee, K.H.; Nishimura, S.; Matsunaga, S.; Fusetani, N. Inhibition of protein synthesis and activation of stress activated protein kinases by onnamide A and theopederin B, antitumor marine natural products. Cancer Sci. 2005, 96, 357–364. [Google Scholar] [CrossRef]
  307. Chill, L.; Kashman, Y.; Schleyer, M. Oriamide, a new cytotoxic cyclic peptide containing a novel amino acid from the marine sponge Theonella sp. Tetrahedron 1997, 53, 16147–16152. [Google Scholar] [CrossRef]
  308. Carroll, A.R.; Pierens, G.K.; Fechner, G. Dysinosin A: A novel inhibitor of Factor VIIa and thrombin from a new genus and species of Australian sponge of the family Dysideidae. J. Am. Chem. Soc. 2002, 124, 13340–13341. [Google Scholar] [CrossRef]
  309. Carroll, A.R.; Buchanan, M.S.; Edser, A. Dysinosins B-D, inhibitors of factor VIIa and thrombin from the Australian sponge Lamellodysidea chlorea. J. Nat. Prod. 2004, 67, 1291–1294. [Google Scholar] [CrossRef] [PubMed]
  310. Schmidt, E.W.; Raventos-Suarez, C.; Bifano, M. Scleritodermin A, a cytotoxic cyclic peptide from the lithistid sponge Scleritoderma nodosum. J. Nat. Prod. 2004, 67, 475–478. [Google Scholar] [CrossRef] [PubMed]
  311. Kobayashi, J.; Sato, M.; Ishibashi, M. Keramamide A, a novel peptide from the Okinawan marine sponge Theonella sp. J. Chem. Soc. Perkin Trans 1 1991, 21, 2609–2611. [Google Scholar] [CrossRef]
  312. Toda, H.; Tozyo, T.; Terui, Y.; Hayashi, F. Discokiolides. Cytotoxic cyclic depsipeptides from the marine sponge Discodermia kiiensis. Chem. Lett. 1992, 2, 431–434. [Google Scholar] [CrossRef]
  313. Odeleye, T.; White, W.L.; Lu, J. Extraction techniques and potential health benefits of bioactive compounds from marine molluscs: A review. Food Funct. 2019, 10, 2278–2289. [Google Scholar] [CrossRef]
  314. Russo, P.; Nastrucci, C.; Cesario, A. From the sea to anticancer therapy. Curr. Med. Chem. 2011, 18, 3551–3562. [Google Scholar] [CrossRef] [PubMed]
  315. Avila, C.; Angulo-Preckler, C. Bioactive compounds from marine heterobranchs. Mar. Drugs 2020, 18, 657. [Google Scholar] [CrossRef]
  316. Bingham, J.P.; Andrews, E.A.; Kiyabu, S.M. Drugs from slugs. Part II—Conopeptide bioengineering. Chem. Biolog. Interact. 2012, 200, 92–113. [Google Scholar] [CrossRef]
  317. Ribeiro, R.; Pinto, E.; Fernandes, C.; Sousa, E. Marine cyclic peptides: Antimicrobial activity and synthetic strategies. Mar. Drugs 2022, 20, 397. [Google Scholar] [CrossRef]
  318. Ahmed, S.; Alam, W.; Jeandet, P.; Aschner, M.; Alsharif, K.F.; Saso, L.; Khan, H. Therapeutic potential of marine peptides in prostate cancer: Mechanistic insights. Mar. Drugs 2022, 20, 466. [Google Scholar] [CrossRef]
  319. Katz, L.; Baltz, R.H. Natural product discovery: Past, present, and future. J. Ind. Microbiol. Biotechnol. 2016, 43, 155–176. [Google Scholar] [CrossRef] [PubMed]
  320. Nakao, Y.; Yoshida, W.Y.; Takada, Y. Kulokekahilide-2, a cytotoxic depsipeptide from a cephalaspidean mollusk Philinopsis speciose. J. Nat. Prod. 2004, 67, 1332–1340. [Google Scholar] [CrossRef] [PubMed]
  321. Zheng, L.-H.; Wang, Y.-J.; Sheng, J.; Wang, F.; Zheng, Y.; Lin, X.-K.; Sun, M. Antitumor peptides from marine organisms. Mar. Drugs 2011, 9, 1840–1859. [Google Scholar] [CrossRef] [PubMed]
  322. Hilla, R.A. Marine natural products. Annu. Rep. Prog. Chem. Sect. B: Org. Chem. 2005, 101, 124–136. [Google Scholar] [CrossRef]
  323. Liu, Y. Mechanism of Action of Seven Marine-Derived Natural Products. Ph.D. Thesis, University of Hawaii, Honolulu, HI, USA, 2014; p. 47. [Google Scholar]
  324. Singh, K.S. Pyrone-derived marine natural products: A review on isolation, bioactivities, and synthesis. Curr. Org. Chem. 2020, 24, 354–401. [Google Scholar]
  325. Wang, B.; Chen, C.; Yu, M.; Liu, Y.; Liu, P.; Zhang, X. A review on metabolites from Onchidium genus: Chemistry and bioactivity. Chem. Biodiver. 2021, 18, e2000580. [Google Scholar] [CrossRef]
  326. Tilvi, S.; Naik, C.G. Tandem mass spectrometry of kahalalides: Identification of two new cyclic depsipeptides, kahalalide R and S from Elysia grandifolia. J. Mass. Spectr. 2007, 42, 70–80. [Google Scholar] [CrossRef]
  327. Faircloth, G.; del Carmen Cuevas Marchante, M. Kahalalide F and ES285: Potent anticancer agents from marine molluscs. Molluscs 2006, 43, 363–379. [Google Scholar]
  328. Suenaga, K.; Mutou, T.; Shibata, T. Isolation and stereostructure of aurilide, a novel cyclodepsipeptide from the Japanese sea hare Dolabella Auricularia. Tetrahedron Lett. 1996, 37, 6771–6774. [Google Scholar] [CrossRef]
  329. Pettit, G.R.; Kamano, Y.; Herald, C.L. Antineoplastic agent. 174. Isolation and structure of the cytostatic depsipeptide dolastatin 13 from the sea hare Dolabella Auricularia. J. Am. Chem. Soc. 1989, 111, 5015–5017. [Google Scholar] [CrossRef]
  330. Pettit, G.R.; Kamano, Y.; Kizu, H. Isolation and structure of the cell growth inhibitory depsipeptides dolastatins 11 and 12. Heterocycles 1989, 28, 553–558. [Google Scholar] [CrossRef]
  331. Pettit, G.R.; Kamano, Y.; Herald, C.L. Antineoplastic agents. 190. Isolation and structure of the cyclodepsipeptide dolastatin 14. J. Org. Chem. 1990, 55, 2989–2990. [Google Scholar] [CrossRef]
  332. Sone, H.; Nemoto, T.; Ojika, M.; Yamada, K. Isolation, structure, and synthesis of dolastatin C, a new depsipeptide from the sea hare Dolabella auricularia. Tetrahedron Lett. 1993, 34, 8445–8448. [Google Scholar] [CrossRef]
  333. Sone, H.; Shibata, T.; Fujita, T.; Ojika, M.; Yamada, K. Dolastatin H and isodolastatin H, potent cytotoxic peptides from the sea hare Dolabella auricularia:  Isolation, stereostructures, and synthesis. J. Am. Chem. Soc. 1996, 118, 1874–1880. [Google Scholar] [CrossRef]
  334. Mutou, T.; Kondo, T.; Ojika, M.; Yamada, K. Isolation and stereostructures of dolastatin G and nordolastatin G, cytotoxic 35-membered cyclodepsipeptides from the Japanese sea hare Dolabella auricularia. J. Org. Chem. 1996, 61, 6340–6345. [Google Scholar] [CrossRef]
  335. Oda, T.; Crane, Z.D.; Dicus, C.W. Dolastatin 11 connects two long-pitch strands in F-actin to stabilize microfilaments. J. Mol. Biol. 2003, 328, 319–324. [Google Scholar] [CrossRef]
  336. Bai, R.; Verdier-Pinard, P.; Gangwar, S. Dolastatin 11, a marine depsipeptide, arrests cells at cytokinesis and induces hyperpolymerization of purified actin. Mol. Pharmacol. 2001, 59, 462–469. [Google Scholar] [CrossRef]
  337. Pettit, G.R.; Xu, J.-P.; Hogan, F.; Schmidt, J.M. Antineoplastic agents 370. Isolation and structure of dolastatin 18. Bioorg. Med. Chem. Lett. 1997, 7, 827–832. [Google Scholar] [CrossRef]
  338. Ishiwata, H.; Nemoto, T.; Ojika, M.; Yamada, K. Isolation and stereostructure of doliculide, a cytotoxic cyclodepsipeptide from the Japanese sea hare Dolabella auricularia. J. Org. Chem. 1994, 59, 4710–4711. [Google Scholar] [CrossRef]
  339. Ishiwata, H.; Sone, H.; Kigoshi, H.; Yamada, K. Total synthesis of doliculide, a potent cytotoxic cyclodepsipeptide from the Japanese sea hare Dolabella auricularia. J. Org. Chem. 1994, 59, 4712–4713. [Google Scholar] [CrossRef]
  340. Bai, R.; Covell, D.G.; Liu, C. (−)-Doliculide, a new macrocyclic depsipeptide enhancer of actin assembly. J. Biol. Chem. 2002, 277, 32165–32171. [Google Scholar] [CrossRef]
  341. Casertano, M.; Menna, M.; Imperatore, C. The ascidian-derived metabolites with antimicrobial properties. Antibiotics 2020, 9, 510. [Google Scholar] [CrossRef]
  342. Arumugam, V.; Venkatesan, M.; Ramachandran, S. Bioactive peptides from marine ascidians and future drug development—A review. Int. J. Pept. Res. Ther. 2018, 24, 13–18. [Google Scholar] [CrossRef]
  343. Phyo, Y.Z.; Ribeiro, J.; Fernandes, C.; Kijjoa, A.; Pinto, M.M.M. Marine natural peptides: Determination of absolute configuration using liquid chromatography methods and evaluation of bioactivities. Molecules 2018, 23, 306. [Google Scholar] [CrossRef]
  344. Laurent, D.; Pietra, F. Natural-product diversity of the New Caledonian marine ecosystem compared to other ecosystems: A pharmacologically oriented view. Chem. Biodiver. 2004, 1, 539–594. [Google Scholar] [CrossRef]
  345. Tincu, J.A.; Taylor, S.W. Antimicrobial peptides from marine invertebrates. Antimicrob. Agents Chemother. 2004, 48, 3645–3654. [Google Scholar] [CrossRef]
  346. Crews, P.; Manes, L.V.; Boehler, M. Jasplakinolide, a cyclodepsipeptide from the marine sponge, Jaspis sp. Tetrahedron Lett. 1986, 27, 2797–2800. [Google Scholar] [CrossRef]
  347. Rinehart, K.L. Antitumor compounds from tunicates. Med. Res. Rev. 2000, 20, 1–27. [Google Scholar] [CrossRef]
  348. Chun, H.G.; Davies, B.; Hoth, D.; Suffness, M.; Plowman, J.; Didemnin, B. The first marine compound entering clinical trials as an antineoplastic agent. Investig. New Drugs 1986, 4, 279–284. [Google Scholar]
  349. Weiss, G.R.; Arteaga, C.L.; Brown, T.D.; Craig, J.B.; Harman, G.S. New anticancer agents. Cancer Chemother. Biol. Response Modif. 1988, 10, 85–116. [Google Scholar]
  350. Dembitsky, V.M.; Gloriozova, T.A.; Poroikov, V.V. Chlorinated plant steroids and their biological activities. Int. J. Curr. Res. Biosci. Plant Biol. 2017, 4, 70–85. [Google Scholar] [CrossRef]
  351. Le, V.H.; Inai, M.; Williams, R.M.; Kan, T. Ecteinascidins. A review of the chemistry, biology and clinical utility of potent tetrahydroisoquinoline antitumor antibiotics. Nat. Prod. Rep. 2015, 32, 328–347. [Google Scholar] [CrossRef]
  352. Dembitsky, V.M.; Gloriozova, T.A.; Poroikov, V.V. Naturally occurring plant isoquinoline N-oxide alkaloids: Their pharmacological and SAR activities. Phytomedicine 2015, 22, 183–202. [Google Scholar] [CrossRef]
  353. Zewail Foote, M.; Hurley, L.H. Ecteinascidin 743: A minor groove alkylator that bends DNA toward the major groove. J. Med. Chem. 1999, 42, 2493–2497. [Google Scholar] [CrossRef]
  354. Takebayashi, Y.; Pourquier, P.; Zimonjic, D.B.; Nakayama, K. Antiproliferative activity of ecteinascidin 743 is dependent upon transcription coupled nucleotide excision repair. Nat. Med. 2001, 7, 961–966. [Google Scholar] [CrossRef]
  355. Lievens, S.C.; Molinski, T.F. Sagittamides A and B. Polyacetoxy long chain acyl amino acids from a Didemnid ascidian. Org. Lett. 2005, 7, 2281–2284. [Google Scholar] [CrossRef]
  356. Schuetz, A.; Junker, J.; Leonov, A.; Lange, O.F.; Molinski, T.F.; Griesinger, C. Stereochemistry of sagittamide A from residual dipolar coupling enhanced NMR. J. Am. Chem. Soc. 2007, 129, 15114–15115. [Google Scholar] [CrossRef]
  357. Humbert, A.; Plé, K.; Harakat, D.; Martinez, A.; Haudrechy, A. A further contribution to the study of sagittamide: Synthesis of a pivotal intermediate belonging to a rare L series. Molecules 2012, 17, 7709–7721. [Google Scholar] [CrossRef]
  358. Biard, J.F.; Roussakis, C.; Kornprobst, J.M.; Gouiffes Barbin, D.; Verbist, J.F. Bistramides A, B, C, D, and K: A new class of bioactive cyclic polyethers from Lissoclinum bistratum. J. Nat. Prod. 1994, 57, 1336–1345. [Google Scholar]
  359. Riou, D.; Roussakis, C.; Robillard, N.; Biard, J.F.; Verbist, J.F. Bistramide A induced irreversible arrest of cell proli feration in a non-small cell bronchopulmonary carcinoma is similar to induction of terminal maturation. Biol. Cell 1993, 77, 261–264. [Google Scholar] [CrossRef]
  360. Statsuk, A.V.; Bai, R.; Baryza, J.L.; Verma, V.A.; Hamel, E.; Wender, P.A.; Kozmin, S.A. Actin is the primary cellular receptor of bistramide A. Nat. Chem. Biol. 2005, 1, 383–388. [Google Scholar] [CrossRef]
  361. Rinehart, K.L., Jr.; Gloer, J.B.; Hughes, R.G., Jr.; Renis, H.E. Didemnins: Antiviral and antitumor depsipeptides from a caribbean tunicate. Science 1981, 212, 933–935. [Google Scholar] [CrossRef]
  362. Crampton, S.L.; Adams, E.G.; Kuentzel, S.L.; Li, L.H. Biochemical and cellular effects of didemnins A and B. Cancer Res. 1984, 44, 1796–1801. [Google Scholar]
  363. Lee, J.; Currano, J.N.; Carroll, P.J.; Joullié, M.M. Didemnins, tamandarins and related natural products. Nat. Prod. Rep. 2012, 29, 404–424. [Google Scholar] [CrossRef]
  364. Whitson, E.L.; Ratnayake, A.S.; Bugni, T.S.; Harper, M.K.; Ireland, C.M. Isolation, structure elucidation and synthesis of eudistomides A and B, lipopeptides from a Fijian ascidian Eudistoma sp. J. Org. Chem. 2009, 74, 1156–1162. [Google Scholar] [CrossRef]
  365. Gutiérrez-Chávez, C.; Benaud, N.; Ferrari, B.C. The ecological roles of microbial lipopeptides: Where are we going? Comp. Struct. Biotechnol. J. 2021, 19, 1400–1413. [Google Scholar] [CrossRef]
  366. Zhao, P.; Xue, Y.; Li, X.; Li, J.; Zhao, Z.; Quan, C.; Gao, W.; Zu, X. Fungi-derived lipopeptide antibiotics developed since 2000. Peptides 2019, 113, 52–65. [Google Scholar] [CrossRef]
  367. Kashif, A.; Rehman, R.; Fuwad, A.; Shahid, M.K. Current advances in the classification, production, properties and applications of microbial biosurfactants – A critical review. Adv. Colloid Interface Sci. 2022, 306, 102718. [Google Scholar] [CrossRef]
  368. Inès, M.; Dhouha, G. Lipopeptide surfactants: Production, recovery and pore forming capacity. Peptides 2015, 71, 100–112. [Google Scholar] [CrossRef]
  369. Ratnayake, R.; Fremlin, L.J.; Lacey, E.; Gill, J.H.; Capon, R.J. Acremolides A–D, lipodepsipeptides from an Australian marine-derived fungus, Acremonium sp. J. Nat. Prod. 2008, 71, 403–408. [Google Scholar] [CrossRef]
  370. Shiono, Y.; Tsuchinari, M.; Shimanuki, K.; Miyajima, T.; Murayama, T.; Koseki, T.; Laatsch, H. Fusaristatins A and B, 2 new cyclic lipopeptides from an endophytic Fusarium sp. J. Antibiot. 2007, 60, 309–316. [Google Scholar] [CrossRef] [Green Version]
  371. Sugawara, T.; Tanaka, A.; Tanaka, K.; Nagai, K.; Suzuki, K.; Suzuki, T. YM-170320, a novel lipopeptide antibiotic inducing morphological change of colonies in a mutant of Candida tropicalis pK233. J. Antibiot. 1998, 51, 435–438. [Google Scholar] [CrossRef]
  372. McBrien, K.D.; Berry, R.L.; Lowe, S.E.; Neddermann, K.M.; Bursuker, I. Rakicidins, new cytotoxic lipopeptides from Micromonospora sp. Fermentation, isolation and characterization. J. Antibiot. 1995, 48, 1446–1452. [Google Scholar] [CrossRef]
  373. Hu, J.-F.; Wunderlich, D.; Sattler, I.; Feng, X.Z.; Grabley, S.; Thiericke, R. Rakicidin C, a new cyclic depsipeptide from Streptomyces sp. Eur. J. Org. Chem. 2000, 28, 3353–3356. [Google Scholar] [CrossRef]
  374. Igarashi, Y.; Shimasaki, R.; Miyanaga, S.; Oku, N.; Onaka, H.; Sakurai, H. Rakicidin D, an inhibitor of tumor cell invasion from marine-derived Streptomyces sp. J. Antibiot. 2010, 63, 563–565. [Google Scholar] [CrossRef]
  375. Motohashi, K.; Toda, T.; Sue, M.; Furihata, K.; Shizuri, Y.; Matsuo, Y.; Kasai, H. Isolation and structure elucidation of tumesce amides A and B, two peptides produced by Streptomyces tumescens YM23–260. J. Antibiot. 2010, 63, 549–552. [Google Scholar] [CrossRef]
  376. Kishimoto, S.; Tsunematsu, Y.; Nishimura, S.; Hayashi, Y.; Hattori, A.; Kakeya, H. Tumescenamide C, an antimicrobial cyclic lipodepsipeptide from Streptomyces sp. Tetrahedron 2012, 68, 5572–5578. [Google Scholar] [CrossRef]
  377. Oh, D.C.; Kauffman, C.A.; Jensen, P.R.; Fenical, W. Induced production of emericellamides A and B from the marine-derived fungus Emericella sp. in competing co-culture. J. Nat. Prod. 2007, 70, 515–520. [Google Scholar] [CrossRef]
  378. Asolkar, R.N.; Freel, K.C.; Jensen, P.R.; Fenical, W. Arenamides A–C, cytotoxic NFκB inhibitors from the marine actinomycete Salinispora arenicola. J. Nat. Prod. 2009, 72, 396–402. [Google Scholar] [CrossRef]
  379. 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. J. Am. Chem. Soc. 1994, 116, 757–758. [Google Scholar] [CrossRef]
  380. Igarashi, M.; Shida, T.; Sasaki, Y.; Kinoshita, N.; Naganawa, H.; Hamada, M.; Takeuchi, T. Vinylamycin, a new depsipeptide antibiotic, from Streptomyces sp. J. Antibiot. 1999, 52, 873–879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  381. Yoshida, H.; Nishide, M.; Kurotaki, M.; Watanabe, Y. New Antibiotic na30851a, Its Production and Use Thereof. Japan Patent 1999049768, 23 February 1999. [Google Scholar]
  382. Watanabe, Y.; Kurotaki, M.; Nishide, M.; Yoshida, H. Antibiotic NA30851A, Its Manufacture with Streptomyces, Insecticides or Microbicides Containing It, and Streptomyces sp. NA30851A. Kokai Tokkyo Koho. Japanese Patent JP 11049768 A 19990223, 18 May 1999. [Google Scholar]
  383. Ueda, J.Y.; Nagai, A.; Izumikawa, M.; Chijiwa, S.; Takagi, M.; Shin-ya, K. A novel antimycin-like compound, JBIR-06, from Streptomyces sp. ML55. J. Antibiot. 2008, 61, 241–244. [Google Scholar] [CrossRef] [PubMed]
  384. Kozone, I.; Ueda, J.Y.; Takagi, M.; Shin-Ya, K. JBIR-52, a new antimycin-like compound, from Streptomyces sp. ML55. J. Antibiot. 2009, 62, 593–595. [Google Scholar] [CrossRef] [PubMed]
  385. Yu, Z.; Vodanovic-Jankovic, S.; Kron, M.; Shen, B. New WS9326A congeners from Streptomyces sp. 9078 Inhibiting Brugia malayi asparaginyl-tRNA synthetase. Org. Lett. 2012, 14, 4946–4949. [Google Scholar] [CrossRef]
  386. Mayer, C.; Sass, P.; Brötz-Oesterhelt, H. Consequences of dosing and timing on the antibacterial effects of ADEP antibiotics. Int. J. Med. Microbiol. 2019, 309, 151329. [Google Scholar] [CrossRef]
  387. Yin, X.; Chen, Y.; Zhang, L.; Wang, Y.; Zabriskie, T.M. Enduracidin analogues with altered halogenation patterns produced by genetically engineered strains of Streptomyces fungicidicus. J. Nat. Prod. 2010, 73, 583–589. [Google Scholar] [CrossRef]
  388. Cavalleri, B.; Pagani, H.; Volpe, G.; Selva, E.; Parenti, F. A-16686, a new antibiotic from Actinoplanes, I. Fermentation, isolation and preliminary physico-chemical characteristics. J. Antibiot. 1984, 37, 309–317. [Google Scholar] [CrossRef]
  389. Pallanza, R.; Berti, M.; Scotti, R.; Randisi, E.; Arioli, E.V. A-16686, a new antibiotic from Actinoplanes II. Biological properties. J. Antibiot. 1984, 37, 318–324. [Google Scholar] [CrossRef]
  390. Nicolaou, K.C.; Chen, J.S.; Edmonds, D.J.; Estrada, A.A. Recent advances in the chemistry and biology of naturally occurring antibiotics. Angew. Chem. Int. Ed. Engl. 2009, 48, 660–719. [Google Scholar] [CrossRef]
  391. Meyers, E.; Weisenborn, F.L.; Pansy, F.E.; Slusarchyk, D.S.; Von Saltza, M.H.; Rathnum, M.L.; Parker, W.L. Janiemycin, a new peptide antibiotic. J. Antibiot. 1970, 23, 502–507. [Google Scholar] [CrossRef]
  392. Hommel, U.; Weber, H.P.; Oberer, L.; Naegeli, H.U.; Oberhauser, B.; Foster, C.A. The 3D-structure of a natural inhibitor of cell adhesion molecule expression. FEBS Lett. 1996, 379, 69–73. [Google Scholar] [CrossRef] [Green Version]
  393. Foster, C.A.; Dreyfuss, M.; Mandak, B.; Merngassner, J.G.; Naegeli, H.U.; Nussbaumer, A.; Oberer, L.; Schell, G.; Swoboda, E.M. Pharmacological modulation of endothelial cell-associated adhesion molecule expression: Implications for future treatment of dermatological diseases. J. Dermatol. 1994, 21, 847–854. [Google Scholar] [CrossRef] [PubMed]
  394. Oberhauser, B.; Grohmann, B.; Sperner, H. Selective epimerisation of a fungal cyclopeptolide via an 2-amino-oxazole intermediate—Conformational consequences. In Proceedings of the 2nd International Electronic Conference on Synthetic Organic Chemistry (ECSOC-2), Basel, Switzerland, 1–30 September 1998; MDPI: Basel, Switzerland, 1999. Available online: https://www.mdpi.org/ecsoc/ (accessed on 7 August 2022).
  395. Fukuchi, N.; Furihata, K.; Nakayama, J.; Goudo, T.; Takayama, S.; Isogai, A.; Suzuki, A. Rotihibins, novel plant growth regulators from Streptomyces graminofaciens. J. Antibiot. 1995, 48, 1004–1010. [Google Scholar] [CrossRef]
  396. Um, S.; Park, S.H.; Kim, J.; Park, H.J.; Ko, K.; Bang, H.S.; Lee, S.K.; Shin, J.; Oh, D.C. Coprisamides A and B, new branched cyclic peptides from a gut bacterium of the dung beetle Copris tripartitus. Org. Lett. 2015, 17, 1272–1275. [Google Scholar] [CrossRef]
  397. Borders, D.B.; Leese, R.A.; Jarolmen, H.; Francis, N.D.; Fantini, A.A.; Falla, T.; Fiddes, J.C.; Aumelas, A. Laspartomycin, an acidic lipopeptide antibiotic with a unique peptide core. J. Nat. Prod. 2007, 70, 443–446. [Google Scholar] [CrossRef]
  398. Vervoort, H.C.; Drašković, M.; Crews, P. Histone deacetylase inhibitors as a tool to up-regulate new fungal biosynthetic products: Isolation of EGM-556, a cyclodepsipeptide, from Microascus sp. Org. Lett. 2011, 13, 410–413. [Google Scholar] [CrossRef]
  399. Li, D.; Carr, G.; Zhang, Y.; Williams, D.E.; Amlani, A.; Bottriell, H.; Mui, A.L.; Andersen, R.J. Turnagainolides A and B, cyclic depsipeptides produced in culture by a Bacillus sp.: Isolation, structure elucidation, and synthesis. J. Nat. Prod. 2011, 74, 1093–1099. [Google Scholar] [CrossRef] [PubMed]
  400. Cooper, E.C.; Curtis, N.; Cranswick, N.; Gwee, A. Pristinamycin: Old drug, new tricks? J. Antimicrob. Chemother. 2014, 69, 2319–2325. [Google Scholar] [CrossRef]
  401. Weber, P. Streptococcus pneumoniae: Lack of emergence of pristinamycin resistance. Pathol. Biol. 2001, 49, 840–845. [Google Scholar] [CrossRef]
  402. De Crecy-Lagard, V.; Blanc, V.; Gil, P.; Naudin, L.; Lorenzon, S.; Famechon, A.; Bamas-Jacques, N.; Crouzet, J.; Thibaut, D. Pristinamycin I biosynthesis in Streptomyces pristinaespiralis: Molecular characterization of the first two structural peptide synthetase genes. J. Bacteriol. 1997, 179, 705–713. [Google Scholar] [CrossRef]
  403. Zhao, Y.; Feng, R.; Zheng, G.; Tian, J.; Ruan, L.; Ge, M.; Jiang, W.; Lu, Y. Involvement of the TetR-type regulator PaaR in the regulation of pristinamycin I biosynthesis through an effect on precursor supply in Streptomyces pristinaespiralis. J. Bacteriol. 2015, 197, 2062–2071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  404. Kamiyama, T.; Satoh, T.; Umino, T.; Katoh, H.; Uemura, F.; Nakamura, Y.; Yokose, K. Monamidocin, a novel fibrinogen receptor antagonist. J. Antibiot. 1995, 48, 1226–1229. [Google Scholar] [CrossRef] [PubMed]
  405. Kajimura, Y.; Kaneda, M. Fusaricidins B, C and D, new depsipeptide antibiotics produced by Bacillus polymyxa KT-8: Isolation, structure elucidation and biological activity. J. Antibiot. 1997, 50, 220–228. [Google Scholar] [CrossRef]
  406. Harman, R.E.; Ham, E.A.; Bolhofer, W.A.; Brink, N.G. The structure of eulicin, a new antifungal agent. J. Am. Chem. Soc. 1958, 80, 5173–5178. [Google Scholar] [CrossRef]
  407. Hedge, V.R.; Silver, J.E.; Patel, M.G.; Gullo, V.P.; Das, P.R.; Puar, M.S. Isolation and structure of two novel muscarinic receptor antagonists. J. Nat. Prod. 1995, 58, 843–848. [Google Scholar] [PubMed]
  408. Nakamura, M.; Yoshida, K.; Yoshida, M.; Kunimoto, S.; Takeuchi, T.; Ohno, T. Eulicin inhibits human immunodeficiency virus infection and replication. J. Antibiot. 1995, 48, 1362–1363. [Google Scholar] [CrossRef]
  409. Mast, Y.; Wohlleben, W. Streptogramins—Two are better than one! Int. J. Med. Microbiol. 2014, 304, 44–50. [Google Scholar] [CrossRef]
  410. Patel, R.; Gallagher, J.C. Vancomycin-resistant enterococcal bacteremia pharmacotherapy. Ann. Pharmacother. 2015, 49, 69–85. [Google Scholar] [CrossRef]
  411. Nomura, H.; Kimura, K.; Sasao, K.; Okabe, M.; Ishikura, T. A Method for Enhancing the Yield of Depsipeptide Antibiotics by Fermentation. Jpn. Kokai Tokkyo Koho. Japanese Patent 62198617 A 19870902, 6 December 1987. [Google Scholar]
  412. Lang, G.; Mitova, M.I.; Cole, A.L.; Din, L.B.; Vikineswary, S.; Abdullah, N.; Blunt, J.W.; Munro, M.H. Pterulamides I-VI, linear peptides from a Malaysian Pterula sp. J. Nat. Prod. 2006, 69, 1389–1393. [Google Scholar] [CrossRef]
  413. Smith, P.A.; Roberts, T.C.; Romesberg, F.E. Broad-spectrum antibiotic activity of the arylomycin natural products is masked by natural target mutations. Chem. Biol. 2010, 17, 1223–1231. [Google Scholar] [CrossRef]
  414. Isono, K.; Osada, H.; Etsuno, H. Manufacture of Depsipeptide A and B with Streptomyces as Antibacterial and Antiviral Agents. Jpn. Kokai Tokkyo Koho. Japanese Patent 05117298 A 19930514, 9 April 1993. [Google Scholar]
  415. Osada, H.; Yano, T.; Koshino, H.; Isono, K. Enopeptin A, a novel depsipeptide antibiotic with anti-bacteriophage activity. J. Antibiot. 1991, 44, 1463–1466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  416. Ortíz-López, F.J.; Monteiro, M.C.; González-Menéndez, V.; Tormo, J.R.; Genilloud, O.; Bills, G.F. Cyclic colisporifungin and linear cavinafungins, antifungal lipopeptides isolated from Colispora cavincola. J. Nat. Prod. 2015, 78, 468–475. [Google Scholar] [CrossRef] [PubMed]
  417. Feng, Y.; Blunt, J.W.; Cole, A.L.; Munro, M.H. A novel cyclodepsipeptide, HA23, from a Fusarium sp. Org. Lett. 2002, 4, 2095–2096. [Google Scholar] [CrossRef] [PubMed]
  418. Gribble, G.W. Recently discovered naturally occurring heterocyclic organohalogen compounds. Heterocycles 2012, 84, 157–207. [Google Scholar] [CrossRef]
  419. Koshino, S.; Koshino, H.; Matsuura, N.; Kobinata, K.; Onose, R.; Isono, K.; Osada, H. A new cyclic lipopeptide antibiotic, enamidonin. J. Antibiot. 1995, 48, 185–187. [Google Scholar] [CrossRef]
  420. Lelais, G.; Seebach, D. α-Amino acids—Syntheses, occurrence in natural products, and components of β-peptides. Biopolymers 2004, 76, 206–243. [Google Scholar] [CrossRef]
  421. Tomoda, H.; Ōmura, S. Potential therapeutics for obesity and atherosclerosis: Inhibitors of neutral lipid metabolism from microorganisms. Pharmacol. Therapeut. 2007, 115, 375–389. [Google Scholar] [CrossRef]
  422. Namatame, I.; Tomoda, H.; Matsuda, D.; Tabata, N.; Kobayashi, S.; Omura, S. K97–0239 A and B, new inhibitors of macrophage foam cell formation, produced by Streptomyces sp. K97–0239. Proc. Jpn. Acad. Ser. B. 2002, 78, 45–50. [Google Scholar] [CrossRef]
  423. Son, S.; Ko, S.K.; Jang, M.; Lee, J.K.; Ryoo, I.J.; Lee, J.S.; Lee, K.H.; Soung, N.K.; Oh, H. Ulleungamides A and B, modified α,β-dehydropipecolic acid containing cyclic depsipeptides from Streptomyces sp. KCB13F003. Org. Lett. 2015, 17, 4046–4069. [Google Scholar] [CrossRef]
  424. Michel, K.H.; Kastner, R.E. AS4556 Antibiotics and Process for Production Thereof. U.S. Patent 4492650, 28 December 1985. [Google Scholar]
  425. Lin, Z.; Flores, M.; Forteza, I.; Henriksen, N.; Concepcion, G.P.; Rosenberg, G.; Haygood, M.G. Totopotensamides, polyketide-cyclic peptide hybrids from a mollusk-associated bacterium Streptomyces sp. J. Nat. Prod. 2012, 75, 644–649. [Google Scholar] [CrossRef]
  426. Jalal, M.A.; Hossain, M.B.; van der Helm, D.; Barnes, C.L. Structure of ferrichrome-type siderophores with dissimilar N delta-acyl groups: Asperchrome B1, B2, B3, D1, D2 and D3. Biol. Met. 1988, 1, 77–89. [Google Scholar] [CrossRef] [PubMed]
  427. Hayakawa, Y.; Sasaki, K.; Nagai, K.; Shin-ya, K.; Furihata, K. Structure of thioviridamide, a novel apoptosis inducer from Streptomyces olivoviridis. J. Antibiot. 2006, 59, 6–10. [Google Scholar] [CrossRef] [PubMed]
  428. Kanasaki, R.; Kobayashi, M.; Fujine, K.; Sato, I.; Hashimoto, M.; Takase, S. FR227673 and FR190293, novel antifungal lipopeptides from Chalara sp. No. 22210 and Tolypocladium parasiticum No. 16616. J. Antibiot. 2006, 59, 158–167. [Google Scholar] [CrossRef] [PubMed]
  429. Wu, Q.X.; Jin, X.J.; Draskovic, M.; Crews, M.S.; Tenney, K.; Valeriote, F.A.; Yao, X.J.; Crews, P. Unraveling the numerous biosynthetic products of the marine sediment-derived fungus, Aspergillus insulicola. Phytochem. Lett. 2012, 5, 114–117. [Google Scholar] [CrossRef]
  430. Wang, S.Y.; Xu, Z.L.; Wang, H.; Li, C.R.; Fu, L.W. Total synthesis, absolute configuration, and biological activity of xyloallenoide A. Helvetica Chim. Acta 2012, 95, 973–982. [Google Scholar] [CrossRef]
  431. Abbanat, D.; Leighton, M.; Maiese, W.; Jones, E.B.; Pearce, C.; Greenstein, M. Cell wall active antifungal compounds produced by the marine fungus Hypoxylon oceanicum LL-15G256. I. Taxonomy and fermentation. J. Antibiot. 1998, 51, 296–302. [Google Scholar] [CrossRef]
  432. Schlingmann, G.; Milne, L.; Williams, D.R.; Carter, G.T. Cell wall active antifungal compounds produced by the marine fungus Hypoxylon oceanicum LL-15G256. II. Isolation and structure determination. J. Antibiot. 1998, 51, 303–316. [Google Scholar] [CrossRef]
  433. Lira, S.P.; Vita-Marques, A.M.; Seleghim, M.H.; Bugni, T.S.; La Barbera, D.V.; Sette, L.D. New destruxins from the marine-derived fungus Beauveria felina. J. Antibiot. 2006, 59, 553–563. [Google Scholar] [CrossRef]
  434. Cai, P.; Smith, D.; Katz, B.; Pearce, C.; Venables, D.; Houck, D. Destruxin-A4 chlorohydrin, a novel destruxin from fungus OS-F68576: Isolation, structure determination, and biological activity as an inducer of erythropoietin. J. Nat. Prod. 1998, 61, 290–293. [Google Scholar] [CrossRef]
  435. Odier, F.; Vey, A.; Bureau, J.P. In vitro effect of fungal cyclodepsipeptides on leukemic cells: Study of destruxins A, B and E. Biol. Cell. 1992, 74, 267–271. [Google Scholar] [CrossRef]
  436. Jegorov, A.; Sedmera, P.; Havlicek, V.; Mat’ha, V. Destruxin Ed1 a cyclopeptide from the fungus Metarhizium anisopliae. Phytochemistry 1999, 49, 1815–1817. [Google Scholar] [CrossRef]
  437. Hanada, K.; Tamai, M.; Yamagishi, M.; Ohmura, S.; Sawada, J.; Tanaka, I. Isolation and characterization of E–64, a new thiol protease inhibitor. Agric. Biol. Chem. 1978, 42, 523–528. [Google Scholar]
  438. Barrett, A.; Kembhavi, A.; Brown, M.; Kirschke, H.; Knight, C.; Tamai, M.; Hanada, K. L-trans-Epoxysuccinyl-leucylamido-(4-guanidino)butane (E-64) and its analogues as inhibitors of cysteine proteinases including cathepsins B, H and L. J. Biochem. 1982, 201, 189–198. [Google Scholar] [CrossRef] [PubMed]
  439. Gill, K.A.; Berrué, F.; Arens, J.C.; Kerr, R.G. Isolation and structure elucidation of cystargamide, a lipopeptide from Kitasatospora cystarginea. J. Nat. Prod. 2014, 77, 1372–1376. [Google Scholar] [CrossRef]
  440. Vilhena, C.; Bettencourt, A. Daptomycin: A review of properties, clinical use, drug delivery and resistance. Mini Rev. Med. Chem. 2012, 12, 202–209. [Google Scholar] [CrossRef]
  441. Baltz, R.H.; Miao, V.; Wrigley, S.K. Natural products to drugs: Daptomycin and related lipopeptide antibiotics. Nat. Prod. Rep. 2005, 22, 717–741. [Google Scholar] [CrossRef]
  442. Mahlert, C.; Kopp, F.; Thirlway, J.; Micklefield, J.; Marahiel, M.A. Stereospecific enzymatic transformation of α-ketoglutarate to (2S,3R)-3-methyl glutamate during acidic lipopeptide biosynthesis. J. Am. Chem. Soc. 2007, 129, 12011–12018. [Google Scholar] [CrossRef]
  443. Hata, T.; Koga, F.; Sano, Y.; Kanamori, K.; Matsumae, A.; Sugawara, R. Carzinophilin, a new tumor inhibitory substance produced by Streptomyces. I. J Antibiot. 1954, 7, 107–112. [Google Scholar]
  444. Carr, G.; Poulsen, M.; Klassen, J.L.; Hou, Y.; Wyche, T.P.; Bugni, T.S.; Currie, C.R.; Clardy, J. Microtermolides A and B from termite-associated Streptomyces sp. and structural revision of vinylamycin. Org. Lett. 2012, 14, 2822–2825. [Google Scholar] [CrossRef]
  445. Bycroft, B.W. Dictionary of Antibiotics and Related Substances; Chapman and Hall: London, UK, 1988; p. 962. [Google Scholar]
  446. Sohda, K.; Takebayashi, Y.; Nagai, K.; Yokoi, T.; Nanya, T.; Kuromitsu, S.; Shindo, N. Novel Antitumor Hexadepsipeptide Compounds of Streptomyces nobilis. WO 9845320 A1 19981015, 4 April 1998. [Google Scholar]
  447. Sakai, Y.; Yoshida, T.; Tsujita, T.; Ochiai, K.; Agatsuma, T.; Saitoh, Y.; Tanaka, F.; Akiyama, T.; Akinaga, S.; Mizukami, T. GE3, a novel hexadepsipeptide antitumor antibiotic, produced by Streptomyces sp. I. Taxonomy, production, isolation, physico-chemical properties, and biological activities. J. Antibiot. 1997, 50, 659–664. [Google Scholar] [CrossRef]
  448. Umezawa, K.; Nakazawa, K.; Ikeda, Y.; Naganawa, H.; Kondo, S. Polyoxypeptins A and B produced by Streptomyces: Apoptosis-inducing cyclic depsipeptides containing the novel amino acid (2S,3R)-3-hydroxy-3-methylproline. J. Org. Chem. 1999, 64, 3034–3038. [Google Scholar] [CrossRef] [PubMed]
  449. Graefe, U.; Schlegel, R.; Ritzau, M.; Ihn, W.; Dornberger, K.; Stengel, C.; Fleck, W.F.; Gutsche, W.; Haertl, A.; Paulus, E.F. Aurantimycins, new depsipeptide antibiotics from Streptomyces aurantiacus IMET 43917: Production, isolation, structure elucidation, and biological activity. J. Antibiot. 1995, 48, 119–125. [Google Scholar] [CrossRef] [PubMed]
  450. Hayakawa, Y.; Nakagawa, M.; Toda, Y.; Seto, H. A new depsipeptide antibiotic, citropeptin. Agric. Biol. Chem. 1990, 54, 1007–1011. [Google Scholar] [PubMed] [Green Version]
  451. Maehr, H.; Liu, C.M.; Palleroni, N.J.; Smallheer, J.; Todaro, L.; Williams, T.H.; Blount, J.F. Microbial products. VIII. Azinothricin, a novel hexadepsipeptide antibiotic. J. Antibiot. 1986, 39, 17–25. [Google Scholar] [CrossRef] [PubMed]
  452. Smitka, T.A.; Deeter, J.B.; Hunt, A.H.; Mertz, F.P.; Ellis, R.M.; Boeck, L.D.; Yao, R.C. A83586C, a new depsipeptide antibiotic. J. Antibiot. 1988, 41, 726–733. [Google Scholar] [CrossRef]
  453. Raju, R.; Zeinab, G.; Piggott, A.M.; Blumenthal, A.; Gardiner, D.L.; Skinner-Adams, T.S.; Capon, R.J. Mollemycin A: An antimalarial and antibacterial glyco-hexadepsipeptide-polyketide from an Australian marine-derived Streptomyces sp. (CMB-M0244). Org. Lett. 2014, 16, 1716–1719. [Google Scholar] [CrossRef]
  454. Mishra, P.D.; Eyyammadichiyil, S.S.; George, S.D.; Sonawane, S.; Chakor, N.S.; Roychowdhury, A.; Sharma, R. Hexadepsipeptide Analogues as Anticancer Compounds. Wipo Patent WO2013168075A1, 14 November 2013. [Google Scholar]
  455. Uchihata, Y.; Ando, N.; Ikeda, Y.; Kondo, S.; Hamada, M.; Umezawa, K. Isolation of a novel cyclic hexadepsipeptide pipalamycin from Streptomyces as an apoptosis-inducing agent. J. Antibiot. 2002, 55, 1–5. [Google Scholar] [CrossRef]
  456. Nakagawa, M.; Hayakawa, Y.; Adachi, K.; Seto, H. A new depsipeptide antibiotic, variapeptin. Agric. Biol. Chem. 1990, 54, 791–794. [Google Scholar]
  457. Nakagawa, M.; Hayakawa, Y.; Furihata, K.; Seto, H. Structural studies on new depsipeptide antibiotics, variapeptin and citropeptin. J. Antibiot. 1990, 43, 477–484. [Google Scholar] [CrossRef]
  458. Raju, R.; Gromyko, O.; Fedorenko, A.B.; Luzhetsky, A.V.; Müller, R. Oleamycins A and B: New antibacterial cyclic hexadepsipeptides isolated from a terrestrial Streptomyces sp. J. Antibiot. 2014, 67, 339–343. [Google Scholar] [CrossRef]
  459. Nishiyama, Y.; Sugawara, K.; Tomita, K.; Yamamoto, H.; Kamei, H.; Oki, T. Verucopeptin, a new antitumor antibiotic active against B16 melanoma. I. Taxonomy, production, isolation, physico-chemical properties and biological activity. J. Antibiot. 1993, 46, 921–927. [Google Scholar] [CrossRef] [PubMed]
  460. Yoshimura, A.; Nishimura, S.; Otsuka, S.; Hattori, A.; Kakeya, H. Structure elucidation of verucopeptin, a HIF-1 inhibitory polyketide-hexapeptide hybrid metabolite from an Actinomycete. Org. Lett. 2015, 17, 5364–5367. [Google Scholar] [CrossRef] [PubMed]
  461. 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] [PubMed]
  462. Chen, H.; Wang, Q. Regulatory mechanisms of lipid biosynthesis in microalgae. Biol. Rev. 2021, 96, 2373–2391. [Google Scholar] [CrossRef] [PubMed]
  463. Hegde, V.R.; Dai, P.; Patel, M.; Das, P.R.; Wang, S.; Puar, M.S. A depsipeptide fungal metabolite inhibitor of cholesteryl ester transfer protein. Bioorg. Med. Chem. Lett. 1998, 8, 1277–1280. [Google Scholar] [CrossRef]
  464. Hedge, V.R.; Puar, M.S.; Dai, P.; Pu, H.; Patel, M.; Anthes, J.C. A family of depsi-peptide fungal metabolites, as selective and competitive human tachykinin receptor (NK2) antagonists: Fermentation, isolation, physico-chemical properties, and biological activity. J. Antibiot. 2001, 54, 125–135. [Google Scholar]
  465. Churchill, A.C.; Dunkle, L.D.; Silbert, W.; Kennedy, K.J.; Macko, V. Differential synthesis of peritoxins and precursors by pathogenic strains of the fungus Periconia circinata. Appl. Environ. Microbiol. 2001, 67, 5721–5728. [Google Scholar] [CrossRef]
  466. Macko, V.; Stimmel, M.B.; Wolpert, T.J.; Dunkle, L.D.; Acklin, W.; Bänteli, R.; Jaun, B.; Arigoni, D. Structure of the host-specific toxins produced by the fungal pathogen Periconia circinata. Proc. Natl. Acad. Sci. USA 1992, 89, 9574–9578. [Google Scholar] [CrossRef]
  467. Jiao, P. Chemical Investigations of Freshwater and Fungicolous Fungi. PhD Thesis, University of Iowa, Iowa City, IA, USA, 2006. [Google Scholar]
  468. Donadio, S.; Maffioli, S.; Paolo Monciardini, P.; Sosio, M.; Jabes, D. Sources of novel antibiotics—aside the common roads. Appl. Microbiol. Biotechnol. 2010, 88, 1261–1267. [Google Scholar] [CrossRef]
  469. Yount, N.Y.; Yeaman, M.R. Peptide antimicrobials: Cell wall as a bacterial target. Ann. N. Y. Acad. 2013, 1277, 127–138. [Google Scholar] [CrossRef]
  470. Hammann, P.; Babenhausen, J.M.; Seibert, G.; Vertesy, L.; Wink, J.; Markus, A. Lipopeptides from Actinoplanes sp. with Pharmacological Action, Process for Their Production and The Use Thereof. U.S. Patent 6,194,383 B1, 27 February 2001. [Google Scholar]
  471. Dini, C. MraY inhibitors as novel antibacterial agents. Current Topics Med. Chem. 2005, 5, 1221–1236. [Google Scholar] [CrossRef] [PubMed]
  472. Boeck, L.D.; Fukuda, D.S.; Abbott, B.J.; Debono, M. Deacylation of A21978C, an acidic lipopeptide antibiotic complex, by Actinoplanes utahensis. J. Antibiot. 1988, 41, 1085–1092. [Google Scholar] [CrossRef] [PubMed]
  473. Debono, M.; Abbott, B.J.; Molloy, R.M.; Fukuda, D.S.; Hunt, A.H.; Daupert, V.M. Enzymatic and chemical modifications of lipopeptide antibiotic A21978C: The synthesis and evaluation of daptomycim (LY146032). J. Antibiot. 1988, 41, 1093–1105. [Google Scholar] [CrossRef] [PubMed]
  474. Fujino, M. On glumamycin, a new antibiotic. VI. An approach to the amino acid sequence. Bull. Chem. Soc. Jpn. 1965, 38, 517–522. [Google Scholar] [CrossRef] [PubMed]
  475. Hausmann, W.K.; Struck, A.H.; Martin, J.H.; Barritt, R.H.; Bohonos, N. Structure determination of fatty acids from the antibiotic Aspartocin. Antimicrob. Agents Chemother. 1963, 33, 352–359. [Google Scholar]
  476. Hausmann, W.K.; Borders, D.B.; Lancaster, J.E.; Heinemann, B. Alpha, beta-diaminobutyric acid obtained from aspartocin. J. Antibiot. 1969, 22, 207–210. [Google Scholar] [CrossRef] [Green Version]
  477. Bodanszky, M.; Chaturvedi, N.C.; Scozzie, J.A. The structure of fatty acids from the antibiotic amphomycin. J. Antibiot. 1969, 22, 399–408. [Google Scholar] [CrossRef]
  478. Hinuma, Y. Zaomycin, a new antibiotic from a Streptomyces sp. J. Antibiot. 1954, 7A, 134–136. [Google Scholar]
  479. Huber, F.M.; Pieper, R.L.; Tietz, A.J. The formation of daptomycin by supplying decanoic acid to Strptomyces roseosporus cultures producing the antibiotic complex A21978C. J. Biotechnol. 1988, 7, 283–292. [Google Scholar] [CrossRef]
  480. Martin, J.H.; Hausmann, W.K. Isolation and identification of D-α-pepecolic acid, a,b-methylaspartic acid and a,b-diaminobutyric acid from the polypeptide antibiotic aspartocin. J. Am. Chem. Soc. 1960, 82, 2079. [Google Scholar] [CrossRef]
  481. Naganawa, H.; Takita, T.; Maeda, K.; Umezawa, H. A novel fatty acid from laspartomycin. J. Antibiot. 1970, 23, 423–424. [Google Scholar] [CrossRef] [PubMed]
  482. Naganawa, H.; Hamada, M.; Maeda, K.; Okami, Y.; Takeushi, T. Laspartomycin, a new anti-staphylococcal peptide. J. Antibiot. 1968, 21, 55–62. [Google Scholar] [CrossRef] [PubMed]
  483. Shay, A.J.; Adam, J.; Martin, J.H.; Hausmann, W.K.; Shu, P.; Bohonos, N. Aspartocin. I. Production, isolation, and characteristics. Antibiot. Ann. 1959, 2, 194–198. [Google Scholar]
  484. Shoji, J.-I. Studies on tsushimycin. I. Isolation and characterization of an acidic acylpeptide containing a new fatty acid. J. Antibiot. 1968, 21, 439–443. [Google Scholar] [CrossRef]
  485. Wooda, T.M.; Martin, N.I. The calcium-dependent lipopeptide antibiotics: Structure, mechanism, & medicinal chemistry. Med. Chem. Commun. 2019, 10, 634–646. [Google Scholar]
  486. Engstrom, G.W.; De Lance, J.V.; Richard, J.L.; Baetz, A.L. Purification and characterization of roseotoxin b, a toxic cyclodepsipeptide from Trichothecium roseum. J. Agric. Food Chem. 1975, 23, 244–253. [Google Scholar] [CrossRef]
  487. Springer, J.P.; Cole, R.J.; Dorner, J.W.; Cox, R.H.; Richard, J.L.; Barnes, C.L.; Van der Helm, D. Structure and conformation of roseotoxin B. J. Am. Chem. Soc. 1984, 106, 2388–2392. [Google Scholar] [CrossRef]
  488. Silber, J.; Ohlendorf, B.; Labes, A.; Näther, C.; Imhoff, J.F. Calcaripeptides A–C, cyclodepsipeptides from a Calcarisporium strain. J. Nat. Prod. 2013, 76, 1461–1467. [Google Scholar] [CrossRef]
  489. Wang, X.; Zhang, A.; Gao, J.; Chen, W.; Wang, S.; Wu, X.; Shen, Y. Trichomide A, a natural cyclodepsipeptide, exerts immunosuppressive activity against activated T lymphocytes by upregulating SHP2 activation to overcome contact dermatitis. J. Invest. Dermatol. 2014, 134, 2737–2746. [Google Scholar] [CrossRef] [Green Version]
  490. Wang, N.; Dong, Y.; Yang, Y.; Xu, R.; Li, C.W.; Cui, C.B. Penicimutanin C, a new alkaloidal compound, isolated from a neomycin-resistant mutant 3-f-31of Penicillium purpurogenum G59. Chem. Biodivers. 2020, 17, e2000241. [Google Scholar] [CrossRef]
  491. Suzuki, K.; Yamaizumi, M.; Tateishi, S.; Monnai, Y.; Uyeda, M. Topostatin, a novel inhibitor of topoisomerases I and II produced by Thermomonospora alba strain No. 1520. III. Inhibitory properties. J. Antibiot. 1999, 52, 460–465. [Google Scholar] [CrossRef] [PubMed]
  492. Bills, G.F.; Platas, G.; Peláez, F.; Masurekar, P. Reclassification of a pneumocandin-producing anamorph, Glarea lozoyensis gen. et sp. nov., previously identified as Zalerion arboricola. Mycolog. Res. 1999, 103, 179–192. [Google Scholar] [CrossRef]
  493. Schwartz, R.E.; Sesin, D.F.; Joshua, H.; Wilson, K.E. Pneumocandins from Zalerion arboricola I. Discovery and isolation. J. Antibiot. 1992, 45, 1853–1866. [Google Scholar] [CrossRef] [PubMed]
  494. Schmatz, D.M.; Abruzzo, G.; Powles, M.A. Pneumocandins from Zalerion arboricola. IV. Biological evaluation of natural and semisynthetic pneumocandins for activity against Pneumocystis carinii and Candida species. J. Antibiot. 1992, 45, 1886–1891. [Google Scholar] [CrossRef] [PubMed]
  495. Pohanka, A.; Broberg, A.; Johansson, M.; Kenne, L.; Levenfors, J. Pseudotrienic acids A and B, two bioactive metabolites from Pseudomonas sp. MF381-IODS. J. Nat. Prod. 2005, 68, 1380–1385. [Google Scholar] [CrossRef]
  496. Pettit, R.K. Mixed fermentation for natural product drug discovery. Appl. Microbiol. Biotechnol. 2009, 83, 19–25. [Google Scholar] [CrossRef]
  497. Kunze, B.; Jansen, R.; Sasse, F.; Hofle, G. Chondramides A–D, new antifungal and cytostatic depsipeptides from Chondromyces crocatus (Myxobacteria). J. Antibiot. 1995, 48, 1262–1266. [Google Scholar] [CrossRef]
  498. Furumai, R.; Matsuyama, A.; Kobashi, N.; Lee, K.H.; Nishiyama, M.; Nakajima, H. FK228 (Depsipeptide) as a natural prodrug that inhibits class I. Histone deacetylases. Cancer Res. 2002, 62, 4916–4921. [Google Scholar]
  499. Ueda, H.; Nakajima, H.; Hori, Y.; Fujita, T.; Nishimura, M.; Goto, T.; Okuhara, M. FR901228, a novel antitumor bicyclic depsipeptide produced by Chromobacterium violaceum No. 968. I. Taxonomy, fermentation, isolation, physico-chemical and biological properties, and antitumor activity. J. Antibiot. 1994, 47, 301–310. [Google Scholar] [CrossRef]
  500. Kjaerulff, L.; Nielsen, A.; Mansson, M.; Gram, L.; Larsen, T.O. Identification of four new agr quorum sensing-interfering cyclodepsipeptides from a marine Photobacterium. Mar. Drugs 2013, 11, 5051–5062. [Google Scholar] [CrossRef] [Green Version]
  501. Adachi, K.; Kawabata, Y.; Kasai, H.; Katsuta, M.; Shizuri, Y. New Antibiotic. Jpn. Pat. Appl. JP 2007230911A, 13 September 2007. [Google Scholar]
  502. Plaza, A.; Garcia, R.; Bifulco, G.; Martinez, J.P. Aetheramides A and B, potent HIV-inhibitory depsipeptides from a Myxobacterium of the new genus “Aetherobacter”. Org. Lett. 2012, 14, 2854–2857. [Google Scholar] [CrossRef] [PubMed]
  503. Weissmana, K.J.; Müller, R. Myxobacterial secondary metabolites: Bioactivities and modes-of-action. Nat. Prod. Rep. 2010, 27, 1276–1295. [Google Scholar] [CrossRef] [PubMed]
  504. Leibold, T.; Sasse, F.; Reichenbach, H. Cyrmenins, novel antifungal peptides containing a nitrogen-linked β-methoxyacrylate pharmacophore: Isolation and structural elucidation. Eur. J. Org. Chem. 2004, 2, 431–435. [Google Scholar] [CrossRef]
  505. Sasse, F.; Leibold, T.; Kunze, B.; Höfle, G. Cyrmenins, new β-methoxyacrylate inhibitors of the electron transport production, isolation, physico-chemical and biological properties. J. Antibiot. 2003, 56, 827–831. [Google Scholar] [CrossRef]
  506. Oka, M.; Konishi, M. BU-2867T Peptide Antibiotics. U.S. Patent 4777160A, 18 September 1986. [Google Scholar]
  507. Oka, M.; Numata, K.; Konishi, M. Semi-Synthetic Peptide Antibiotics. U.S. Patent 4789731, 6 December 1988. [Google Scholar]
  508. Ray, L.; Yamanaka, K.; Moore, B.S. A peptidyl-transesterifying type I thioesterase in salinamide biosynthesis. Angew. Chem. Int. Ed. Engl. 2016, 55, 364–367. [Google Scholar] [CrossRef]
  509. Ohlendorf, B.; Kehraus, S.; König, G.M. Myxochromide B3, a new member of the myxochromide family of secondary metabolites. J. Nat. Prod. 2008, 71, 1708–1713. [Google Scholar] [CrossRef]
  510. PASS Online URL. Available online: http://www.way2drug.com/passonline/ (accessed on 24 February 2022).
Figure 1. Molecular structure of cryptophycin-1 and retrosynthetic division on the subunits. Shortly after the discovery of cryptophycin-1, biological screening assays showed a high cytotoxicity against human cervical carcinoma (KB) and human breast adenocarcinoma (MCF-7).
Figure 1. Molecular structure of cryptophycin-1 and retrosynthetic division on the subunits. Shortly after the discovery of cryptophycin-1, biological screening assays showed a high cytotoxicity against human cervical carcinoma (KB) and human breast adenocarcinoma (MCF-7).
Hydrobiology 01 00024 g001
Figure 2. The genus of cyanobacteria Nostoc: (a), Nostoc sp.; (b), Nostoc commune; (c), N. commune; (d), N. commune, which inhabit various environments, such as the bottom of both fresh and salt lakes, and form colonies consisting of filaments of moniliform cells in a gelatinous membrane. Cyanobacteria of the genus Nostoc produce saturated and unsaturated fatty acids, lipopeptides, depsipeptides, oligopeptides and toxins. All photos are taken from sites where permission was granted for non-commercial use.
Figure 2. The genus of cyanobacteria Nostoc: (a), Nostoc sp.; (b), Nostoc commune; (c), N. commune; (d), N. commune, which inhabit various environments, such as the bottom of both fresh and salt lakes, and form colonies consisting of filaments of moniliform cells in a gelatinous membrane. Cyanobacteria of the genus Nostoc produce saturated and unsaturated fatty acids, lipopeptides, depsipeptides, oligopeptides and toxins. All photos are taken from sites where permission was granted for non-commercial use.
Hydrobiology 01 00024 g002
Figure 3. Chemical diversity of FA incorporated into cryptophycins.
Figure 3. Chemical diversity of FA incorporated into cryptophycins.
Hydrobiology 01 00024 g003
Figure 4. 3D graph showing the predicted and calculated activity of cryptophycin−1 and its four subunits A in Figure 1 ((5S,6S,E)-5-hydroxy-6-((2R,3R)-3-phenyloxiran-2-yl)-hept-2-enoic acid); B in Figure 1 ((R)-2-amino-3-(3-chloro-4-methoxyphenyl)-propanoic acid); C in Figure 1 (3-aminopropanoic acid) and D in Figure 1 ((S)-2-hydroxy-4-methylpentanoic acid), with the highest degree of confidence being more than 74%. Designations A, B, C, and D are shown in Figure 1. The depsipeptide named cryptophycin−1 was produced by the cyanobacterium from the strain Nostoc.
Figure 4. 3D graph showing the predicted and calculated activity of cryptophycin−1 and its four subunits A in Figure 1 ((5S,6S,E)-5-hydroxy-6-((2R,3R)-3-phenyloxiran-2-yl)-hept-2-enoic acid); B in Figure 1 ((R)-2-amino-3-(3-chloro-4-methoxyphenyl)-propanoic acid); C in Figure 1 (3-aminopropanoic acid) and D in Figure 1 ((S)-2-hydroxy-4-methylpentanoic acid), with the highest degree of confidence being more than 74%. Designations A, B, C, and D are shown in Figure 1. The depsipeptide named cryptophycin−1 was produced by the cyanobacterium from the strain Nostoc.
Hydrobiology 01 00024 g004
Figure 5. The unbranched filamentous cyanobacterium, Lyngbya majuscula (ac), commonly referred to as “mermaids’ hair” or “fireweed,” is found in coastal tropical and subtropical marine and estuarine areas around the world. The cyanobacterium usually grows up to 30 m below the surface and attaches to rocks, sand, or algae (a,c), or can form large mats (b).
Figure 5. The unbranched filamentous cyanobacterium, Lyngbya majuscula (ac), commonly referred to as “mermaids’ hair” or “fireweed,” is found in coastal tropical and subtropical marine and estuarine areas around the world. The cyanobacterium usually grows up to 30 m below the surface and attaches to rocks, sand, or algae (a,c), or can form large mats (b).
Hydrobiology 01 00024 g005
Figure 6. Fatty acids from peptide molecules produced by cyanobacteria belonging to the strain Nostoc.
Figure 6. Fatty acids from peptide molecules produced by cyanobacteria belonging to the strain Nostoc.
Hydrobiology 01 00024 g006
Figure 7. 3D Graph showing the predicted and calculated anti−psoriatic activity of a 2,2-dimethyl-3-hydroxy-7-octynoic acid (15). This FA is a subunit in many linear and cyclic lipopeptides that are synthesized by cyanobacteria belonging to the genus Lyngbya.
Figure 7. 3D Graph showing the predicted and calculated anti−psoriatic activity of a 2,2-dimethyl-3-hydroxy-7-octynoic acid (15). This FA is a subunit in many linear and cyclic lipopeptides that are synthesized by cyanobacteria belonging to the genus Lyngbya.
Hydrobiology 01 00024 g007
Figure 8. FA derived from peptide molecules produced by cyanobacteria belonging to the strain Lyngbya.
Figure 8. FA derived from peptide molecules produced by cyanobacteria belonging to the strain Lyngbya.
Hydrobiology 01 00024 g008
Figure 9. 3D graph showing the predicted and calculated Iron antagonist activity of unusual FA (31) with the highest degree of confidence being more than 95%. This acid is part of a neuroactive cyclic depsipeptide called hoiamide. These cyclic depsipeptides are produced by two different filamentous cyanobacteria, Lyngbya majuscula and Phormidium gracila. It is a very rare case that a fatty acid exhibits the surprising property of being an iron antagonist with a high degree of certainty.
Figure 9. 3D graph showing the predicted and calculated Iron antagonist activity of unusual FA (31) with the highest degree of confidence being more than 95%. This acid is part of a neuroactive cyclic depsipeptide called hoiamide. These cyclic depsipeptides are produced by two different filamentous cyanobacteria, Lyngbya majuscula and Phormidium gracila. It is a very rare case that a fatty acid exhibits the surprising property of being an iron antagonist with a high degree of certainty.
Hydrobiology 01 00024 g009
Figure 10. FA derived from lipopeptides and the cyclic depsipeptides produced by Moorea and Lyngbya.
Figure 10. FA derived from lipopeptides and the cyclic depsipeptides produced by Moorea and Lyngbya.
Hydrobiology 01 00024 g010
Figure 11. 3D Graph showing the predicted and calculated activities with the dominance of hypolipemic and anti-psoriatic properties (these are two highs in the red zone), as seen in (S)-7,7-dichloro-3-hydroxy-2,2-dimethyloctanoic acid (34). This FA was found in the cytotoxic lipopeptide hectochlorin which was isolated from the cyanobacterium Lyngbya majuscula.
Figure 11. 3D Graph showing the predicted and calculated activities with the dominance of hypolipemic and anti-psoriatic properties (these are two highs in the red zone), as seen in (S)-7,7-dichloro-3-hydroxy-2,2-dimethyloctanoic acid (34). This FA was found in the cytotoxic lipopeptide hectochlorin which was isolated from the cyanobacterium Lyngbya majuscula.
Hydrobiology 01 00024 g011
Figure 12. 3D Graph showing the predicted and calculated activity with the dominance of antineoplastic properties (these are four highs in the red zone) of (4S,5S,6E,8E)-5-hydroxy-4,6,8,10,10-pentamethyl-3-methyleneundeca-6,8-dienoic acid (37), two cyclic lipopeptides, antillatoxins A (41) and B (42), and a synthetic analogue (43). The cyclic lipopeptides, ATx-A (41) and ATx-B (42) have been isolated from a marine cyanobacteria Lyngbya majuscula.
Figure 12. 3D Graph showing the predicted and calculated activity with the dominance of antineoplastic properties (these are four highs in the red zone) of (4S,5S,6E,8E)-5-hydroxy-4,6,8,10,10-pentamethyl-3-methyleneundeca-6,8-dienoic acid (37), two cyclic lipopeptides, antillatoxins A (41) and B (42), and a synthetic analogue (43). The cyclic lipopeptides, ATx-A (41) and ATx-B (42) have been isolated from a marine cyanobacteria Lyngbya majuscula.
Hydrobiology 01 00024 g012
Figure 13. Cyanobacteria: (a), Leptolyngbya sp.; (b), Oscillatoria sp.; (c), Tolypothrix sp.; (d), Symploca hydnoides, found in marine and freshwater environments, typically develop on lake, pond, or marine sediments. All these cyanobacteria produce various bioactive metabolites, including linear and cyclic peptides and lipopeptides containing unusual FA.
Figure 13. Cyanobacteria: (a), Leptolyngbya sp.; (b), Oscillatoria sp.; (c), Tolypothrix sp.; (d), Symploca hydnoides, found in marine and freshwater environments, typically develop on lake, pond, or marine sediments. All these cyanobacteria produce various bioactive metabolites, including linear and cyclic peptides and lipopeptides containing unusual FA.
Hydrobiology 01 00024 g013
Figure 14. Neo, aromatic, and other FA derived from peptides produced by cyanobacteria.
Figure 14. Neo, aromatic, and other FA derived from peptides produced by cyanobacteria.
Hydrobiology 01 00024 g014
Figure 15. 3D graph showing the predicted and calculated with dominance of the antidiabetic activity of 2-hydroxy-3-(4-hydroxyphenyl)-propanoic acid (51), (52), (53), and (2R,3S)-3-((2S,3E,5E,7R,8S)-2,8-dihydroxy-3,7-dimethyl-8-phenylocta-3,5-dien-1-yl)-2-methyloxirane-2-carboxylic acid (54). These FA were derived from lipopeptides produced by the cyanobacterium Nostoc sp. and myxobacterium Nannocystis sp.
Figure 15. 3D graph showing the predicted and calculated with dominance of the antidiabetic activity of 2-hydroxy-3-(4-hydroxyphenyl)-propanoic acid (51), (52), (53), and (2R,3S)-3-((2S,3E,5E,7R,8S)-2,8-dihydroxy-3,7-dimethyl-8-phenylocta-3,5-dien-1-yl)-2-methyloxirane-2-carboxylic acid (54). These FA were derived from lipopeptides produced by the cyanobacterium Nostoc sp. and myxobacterium Nannocystis sp.
Hydrobiology 01 00024 g015
Figure 16. FA derived from peptides produced by marine bacteria and cyanobacteria.
Figure 16. FA derived from peptides produced by marine bacteria and cyanobacteria.
Hydrobiology 01 00024 g016
Figure 17. 3D Graph showing the predicted and calculated with dominance of antiviral activity of (2E,4E)-octa-2,4-dienoic (66) acid. This FA was found in lipopeptides, which are produced by the marine bacteria Saccharomonospora sp. and Saccharomonospora coelicolor and can be used as a strong antiviral agent.
Figure 17. 3D Graph showing the predicted and calculated with dominance of antiviral activity of (2E,4E)-octa-2,4-dienoic (66) acid. This FA was found in lipopeptides, which are produced by the marine bacteria Saccharomonospora sp. and Saccharomonospora coelicolor and can be used as a strong antiviral agent.
Hydrobiology 01 00024 g017
Figure 18. 3D graph showing the predicted and calculated activity, with the dominance of lipid metabolism-regulator properties, of FA (62), (63), (68), and (70), with the highest degree of confidence being more than 93%. These FA derived from lipopeptides produced by the marine cyanobacterium Moorea produns, bacteria Pseudomonas sp. and Bacillus sp.
Figure 18. 3D graph showing the predicted and calculated activity, with the dominance of lipid metabolism-regulator properties, of FA (62), (63), (68), and (70), with the highest degree of confidence being more than 93%. These FA derived from lipopeptides produced by the marine cyanobacterium Moorea produns, bacteria Pseudomonas sp. and Bacillus sp.
Hydrobiology 01 00024 g018
Figure 19. Chloro-containing, sulfur-containing and other FA derived from lipopeptides.
Figure 19. Chloro-containing, sulfur-containing and other FA derived from lipopeptides.
Hydrobiology 01 00024 g019
Figure 20. 3D graph showing the predicted and calculated antineoplastic activity of 4,4-dichloro-3-methylbutanoic acid (78) with the highest degree of confidence being more than 96%. This halogenated acid was detected in depsipeptides of the Eastern Caribbean collection of Lyngbya majuscula.
Figure 20. 3D graph showing the predicted and calculated antineoplastic activity of 4,4-dichloro-3-methylbutanoic acid (78) with the highest degree of confidence being more than 96%. This halogenated acid was detected in depsipeptides of the Eastern Caribbean collection of Lyngbya majuscula.
Hydrobiology 01 00024 g020
Figure 21. Unsaturated and other FA derived from peptides.
Figure 21. Unsaturated and other FA derived from peptides.
Hydrobiology 01 00024 g021
Figure 22. 3D graph showing the predicted and calculated antineoplastic activity of (5S,6S,7R,8R,E)-5,7-dihydroxy-2,6,8-trimethyldec-2-enoic acid (98) as a fragment of lagunamides A and B. These cyclic depsipeptides were isolated from the marine cyanobacterium Lyngbya majuscula obtained from Pulau Hantu Besar (Singapore).
Figure 22. 3D graph showing the predicted and calculated antineoplastic activity of (5S,6S,7R,8R,E)-5,7-dihydroxy-2,6,8-trimethyldec-2-enoic acid (98) as a fragment of lagunamides A and B. These cyclic depsipeptides were isolated from the marine cyanobacterium Lyngbya majuscula obtained from Pulau Hantu Besar (Singapore).
Hydrobiology 01 00024 g022
Figure 23. Bioactive glycosidic FA derived from cyanobacterial peptides.
Figure 23. Bioactive glycosidic FA derived from cyanobacterial peptides.
Hydrobiology 01 00024 g023
Figure 24. Photos of different types of bacteria: (a), Actiniplanus sp.; (b), Pseudomonas aeruginosa; (c), Saccharomonospora viridis; (d), Streptomyces sp., which inhabit various environments and produce lipopeptides with rare and unusual FA.
Figure 24. Photos of different types of bacteria: (a), Actiniplanus sp.; (b), Pseudomonas aeruginosa; (c), Saccharomonospora viridis; (d), Streptomyces sp., which inhabit various environments and produce lipopeptides with rare and unusual FA.
Hydrobiology 01 00024 g024
Figure 25. 3D graph showing the predicted and calculated activity of an anti-infective agent of glycosidic FA (106). This acid is a fragment of the glycolipopeptide hassallidin B, which was isolated from the cyanobacterium Hassallia sp.
Figure 25. 3D graph showing the predicted and calculated activity of an anti-infective agent of glycosidic FA (106). This acid is a fragment of the glycolipopeptide hassallidin B, which was isolated from the cyanobacterium Hassallia sp.
Hydrobiology 01 00024 g025
Figure 26. Graphical display of the chemical structure of the green alga Derbesia marina lipopeptide and the free FA formed by hydrolysis of the amide bond.
Figure 26. Graphical display of the chemical structure of the green alga Derbesia marina lipopeptide and the free FA formed by hydrolysis of the amide bond.
Hydrobiology 01 00024 g026
Figure 27. FA incorporated into lipopeptides derived from lipid extracts of the green algae Bryopsis pennata, B. plumosa and the sacoglossan mollusc Elysia rufescens.
Figure 27. FA incorporated into lipopeptides derived from lipid extracts of the green algae Bryopsis pennata, B. plumosa and the sacoglossan mollusc Elysia rufescens.
Hydrobiology 01 00024 g027
Figure 28. The green algae Bryopsis plumosa (a) and B. pennata (b) are the staple food for the sea slug, Elysia rufescens (c,d). This mollusk is similar to nudibranch, but is not classified in this order of gastropods, but belongs instead to a closely related clade, Sacoglossa. These molluscs synthesize a class of cyclic depsipeptides called kahalalides.
Figure 28. The green algae Bryopsis plumosa (a) and B. pennata (b) are the staple food for the sea slug, Elysia rufescens (c,d). This mollusk is similar to nudibranch, but is not classified in this order of gastropods, but belongs instead to a closely related clade, Sacoglossa. These molluscs synthesize a class of cyclic depsipeptides called kahalalides.
Hydrobiology 01 00024 g028
Figure 29. 3D Graph showing the predicted and calculated cerebral anti-ischemic activity of FA (115). This acid is a fragment of cyclic depsipeptides, kahalalides P and Q.
Figure 29. 3D Graph showing the predicted and calculated cerebral anti-ischemic activity of FA (115). This acid is a fragment of cyclic depsipeptides, kahalalides P and Q.
Hydrobiology 01 00024 g029
Figure 30. Graphical display of the chemical structure of lipopeptide isolated from the marine sponge Theonella aff. mirabilis and the free FA formed by hydrolysis of the amide bond.
Figure 30. Graphical display of the chemical structure of lipopeptide isolated from the marine sponge Theonella aff. mirabilis and the free FA formed by hydrolysis of the amide bond.
Hydrobiology 01 00024 g030
Figure 31. Samples of marine sponges: (a), Theonella cylindrica; (b), T. swinhoei; (c), T. swinhoei; (d), T. swinhoei. It is known that sea sponges from the genus Theonella are home to many associated bacteria that occupy up to 40% of their body volume. Entotheonella sp. (Tectomicrobia) is a filamentous symbiont that produces almost all known biologically active compounds derived from the sponge Theonella swinhoei.
Figure 31. Samples of marine sponges: (a), Theonella cylindrica; (b), T. swinhoei; (c), T. swinhoei; (d), T. swinhoei. It is known that sea sponges from the genus Theonella are home to many associated bacteria that occupy up to 40% of their body volume. Entotheonella sp. (Tectomicrobia) is a filamentous symbiont that produces almost all known biologically active compounds derived from the sponge Theonella swinhoei.
Hydrobiology 01 00024 g031
Figure 32. Branched, saturated, neo-, and unsaturated FA isolated from sponge lipopeptides.
Figure 32. Branched, saturated, neo-, and unsaturated FA isolated from sponge lipopeptides.
Hydrobiology 01 00024 g032
Figure 33. The marine sponges belonging to the genus Haliclona contain more than fifty species of actinobacteria belonging to the genera Streptomyces, Nocardiopsis, Micromonospora and Verrucosispora. Members of this genus produce large amounts of bioactive metabolites such as lipids, steroids, FA, lipopeptides and amino acids.
Figure 33. The marine sponges belonging to the genus Haliclona contain more than fifty species of actinobacteria belonging to the genera Streptomyces, Nocardiopsis, Micromonospora and Verrucosispora. Members of this genus produce large amounts of bioactive metabolites such as lipids, steroids, FA, lipopeptides and amino acids.
Hydrobiology 01 00024 g033
Figure 34. Branched, unsaturated and glycosidic FA derived from sponge lipopeptides.
Figure 34. Branched, unsaturated and glycosidic FA derived from sponge lipopeptides.
Hydrobiology 01 00024 g034
Figure 35. 3D graph showing the predicted and calculated dermatologic activity of methyl-branched FA (127). This acid is incorporated into the cyclic depsipeptide halipeptin A, which is found in the marine sponge Haliclona sp.
Figure 35. 3D graph showing the predicted and calculated dermatologic activity of methyl-branched FA (127). This acid is incorporated into the cyclic depsipeptide halipeptin A, which is found in the marine sponge Haliclona sp.
Hydrobiology 01 00024 g035
Figure 36. 3D graph showing the predicted and calculated anti-infective and antineoplastic activities of glycosidic FA (145, 146 and 147). These acids are incorporated into cyclic peptides called aciculitins A–C and are produced by the lithistid sponge Aciculites orientalis.
Figure 36. 3D graph showing the predicted and calculated anti-infective and antineoplastic activities of glycosidic FA (145, 146 and 147). These acids are incorporated into cyclic peptides called aciculitins A–C and are produced by the lithistid sponge Aciculites orientalis.
Hydrobiology 01 00024 g036
Figure 37. Branched, saturated, and unsaturated FA derived from sponge lipopeptides.
Figure 37. Branched, saturated, and unsaturated FA derived from sponge lipopeptides.
Hydrobiology 01 00024 g037
Figure 38. 3D graph showing the predicted and calculated activity of a regulator lipid metabolism of FA (153). Acid 153 has a similar activity since the structures of both metabolites are similar. These acids are found in lipopeptides, bengamides P and Q, and are produced by marine sponges belonging to the Jaspidae family.
Figure 38. 3D graph showing the predicted and calculated activity of a regulator lipid metabolism of FA (153). Acid 153 has a similar activity since the structures of both metabolites are similar. These acids are found in lipopeptides, bengamides P and Q, and are produced by marine sponges belonging to the Jaspidae family.
Hydrobiology 01 00024 g038
Figure 39. Chlorinated FA derived from sponge lipopeptides.
Figure 39. Chlorinated FA derived from sponge lipopeptides.
Hydrobiology 01 00024 g039
Figure 40. 3D graph showing the predicted and calculated antineoplastic activity of chlorinated FA (170). The figure shows that a single peak (red zone) dominates, which corresponds to the strong antitumor activity of (S)−4,4−dichloro−3−methylbutanoic acid.
Figure 40. 3D graph showing the predicted and calculated antineoplastic activity of chlorinated FA (170). The figure shows that a single peak (red zone) dominates, which corresponds to the strong antitumor activity of (S)−4,4−dichloro−3−methylbutanoic acid.
Hydrobiology 01 00024 g040
Figure 41. Miscellaneous FA derived from sponge lipopeptides.
Figure 41. Miscellaneous FA derived from sponge lipopeptides.
Hydrobiology 01 00024 g041
Figure 42. 3D graph showing the predicted and calculated antineoplastic activity of tetrahydro−2H−pyran−containing FA (173). This acid is part of the onnamide B lipopeptide, which exhibits highly cytotoxic activity against the P388 cell line.
Figure 42. 3D graph showing the predicted and calculated antineoplastic activity of tetrahydro−2H−pyran−containing FA (173). This acid is part of the onnamide B lipopeptide, which exhibits highly cytotoxic activity against the P388 cell line.
Hydrobiology 01 00024 g042
Figure 43. FA derived from mollusc lipopeptides.
Figure 43. FA derived from mollusc lipopeptides.
Hydrobiology 01 00024 g043
Figure 44. Marine molluscs: (a), The pulmonate mollusc Onchidium sp.; (b), a cephalaspidean mollusc, Philinopsis speciosa; (c), Sea hare Dolabella auricularia. All these marine molluscs share the ability to synthesize lipopeptides, although it is possible that they can obtain related lipopeptides from algae.
Figure 44. Marine molluscs: (a), The pulmonate mollusc Onchidium sp.; (b), a cephalaspidean mollusc, Philinopsis speciosa; (c), Sea hare Dolabella auricularia. All these marine molluscs share the ability to synthesize lipopeptides, although it is possible that they can obtain related lipopeptides from algae.
Hydrobiology 01 00024 g044
Figure 45. 3D graph showing the predicted and calculated anti-psoriatic activity of FA (183). (S)−3−hydroxy−2,2−dimethyloct−7−enoic acid (183) was incorporated into the lipopeptide kulolide 2 from cephalaspidean molluscs, Philinopsis speciosa.
Figure 45. 3D graph showing the predicted and calculated anti-psoriatic activity of FA (183). (S)−3−hydroxy−2,2−dimethyloct−7−enoic acid (183) was incorporated into the lipopeptide kulolide 2 from cephalaspidean molluscs, Philinopsis speciosa.
Hydrobiology 01 00024 g045
Figure 46. FA derived from Tunicata lipopeptides.
Figure 46. FA derived from Tunicata lipopeptides.
Hydrobiology 01 00024 g046
Figure 47. 3D graph showing the predicted and calculated FA (208–210) activities as an anti-restenosis agent. Recurrence, called restenosis or vasoconstriction, is rare. Various drugs, mechanical devices such as stents, genetic treatments such as gene transfer or stem cell infusion, or combinations of the above are commonly used to treat restenosis. For FA, this property is apparently described for the first time.
Figure 47. 3D graph showing the predicted and calculated FA (208–210) activities as an anti-restenosis agent. Recurrence, called restenosis or vasoconstriction, is rare. Various drugs, mechanical devices such as stents, genetic treatments such as gene transfer or stem cell infusion, or combinations of the above are commonly used to treat restenosis. For FA, this property is apparently described for the first time.
Hydrobiology 01 00024 g047
Figure 48. Unusual FA derived from fungal lipopeptides.
Figure 48. Unusual FA derived from fungal lipopeptides.
Hydrobiology 01 00024 g048
Figure 49. 3D graph showing the predicted and calculated activity of an acute neurological disorder agent, FA (214). This acid is part of the cyclic lipopeptides of fusaristatin A and B, as well as the linear lipopeptide YM-170320.
Figure 49. 3D graph showing the predicted and calculated activity of an acute neurological disorder agent, FA (214). This acid is part of the cyclic lipopeptides of fusaristatin A and B, as well as the linear lipopeptide YM-170320.
Hydrobiology 01 00024 g049
Figure 50. Saturated, unsaturated, and phenolic FA derived from fungal lipopeptides.
Figure 50. Saturated, unsaturated, and phenolic FA derived from fungal lipopeptides.
Hydrobiology 01 00024 g050
Figure 51. 3D graph showing the predicted and calculated antiviral activity of unsaturated FA (234, 237 and 243) with over 90% confidence. FA (234 and 237) are found in the depsipeptide which is synthesized by Streptomyces hawaiiensis, and the acid (243) is present in the depsipeptide which is produced by Streptomyces graminofaciens.
Figure 51. 3D graph showing the predicted and calculated antiviral activity of unsaturated FA (234, 237 and 243) with over 90% confidence. FA (234 and 237) are found in the depsipeptide which is synthesized by Streptomyces hawaiiensis, and the acid (243) is present in the depsipeptide which is produced by Streptomyces graminofaciens.
Hydrobiology 01 00024 g051
Figure 52. Unusual FA derived from fungi and fungal endophyte lipopeptides.
Figure 52. Unusual FA derived from fungi and fungal endophyte lipopeptides.
Hydrobiology 01 00024 g052
Figure 53. 3D graph shows the predicted and calculated antihypertensive activity of a rare sulfur-containing FA (253). The antihypertensive property of this acid appears to be related to the presence of sulfur monoxide. It is known that many antihypertensive drugs, such as indapamide, chlorthalidone, metalozone, xipamide or clopamide, contain sulfur dioxide molecules.
Figure 53. 3D graph shows the predicted and calculated antihypertensive activity of a rare sulfur-containing FA (253). The antihypertensive property of this acid appears to be related to the presence of sulfur monoxide. It is known that many antihypertensive drugs, such as indapamide, chlorthalidone, metalozone, xipamide or clopamide, contain sulfur dioxide molecules.
Hydrobiology 01 00024 g053
Figure 54. Rare and unusual FA derived from fungal lipopeptides.
Figure 54. Rare and unusual FA derived from fungal lipopeptides.
Hydrobiology 01 00024 g054
Figure 55. 3D graph showing the predicted and calculated Myasthenia Gravis activity of epoxy FA (282, 283 and 284) at 78 to 96% confidence. Aspergillus japonicus TPR-64 synthesized acid (282), the actinomycete Kitasatospora cystarginea produces acid (283), and the fungus Streptomyces produces acid (284).
Figure 55. 3D graph showing the predicted and calculated Myasthenia Gravis activity of epoxy FA (282, 283 and 284) at 78 to 96% confidence. Aspergillus japonicus TPR-64 synthesized acid (282), the actinomycete Kitasatospora cystarginea produces acid (283), and the fungus Streptomyces produces acid (284).
Hydrobiology 01 00024 g055
Figure 56. Unique, rare, and unusual FA derived from fungal and bacterial lipopeptides.
Figure 56. Unique, rare, and unusual FA derived from fungal and bacterial lipopeptides.
Hydrobiology 01 00024 g056
Figure 57. 3D graph shows the predicted and calculated antineoplastic activity of FA (290, 293 and 299) with a confidence level of over 94%.
Figure 57. 3D graph shows the predicted and calculated antineoplastic activity of FA (290, 293 and 299) with a confidence level of over 94%.
Hydrobiology 01 00024 g057
Figure 58. 3D graph shows the predicted and calculated antibiotic glycopeptide activity of FA (297 and 300) with a confidence level of over 90%.
Figure 58. 3D graph shows the predicted and calculated antibiotic glycopeptide activity of FA (297 and 300) with a confidence level of over 90%.
Hydrobiology 01 00024 g058
Figure 59. Unusual and rare FA derived from fungal and bacterial lipopeptides.
Figure 59. Unusual and rare FA derived from fungal and bacterial lipopeptides.
Hydrobiology 01 00024 g059
Figure 60. 3D graph shows the predicted and calculated anti-infective activity of glycosidic FA (312) with a confidence level of over 94%. This acid produced by fungus Alternaria alternata f. sp. mali.
Figure 60. 3D graph shows the predicted and calculated anti-infective activity of glycosidic FA (312) with a confidence level of over 94%. This acid produced by fungus Alternaria alternata f. sp. mali.
Hydrobiology 01 00024 g060
Figure 61. 3D Graph shows the predicted and calculated activity of FA (315 and 316) as lipid metabolism regulators with a confidence level of over 96%. Both unsaturated acids have been incorporated into liposidomycines A and K which are synthesized by the fungus Streptomyces griseosporeus.
Figure 61. 3D Graph shows the predicted and calculated activity of FA (315 and 316) as lipid metabolism regulators with a confidence level of over 96%. Both unsaturated acids have been incorporated into liposidomycines A and K which are synthesized by the fungus Streptomyces griseosporeus.
Hydrobiology 01 00024 g061
Figure 62. Unusual FA derived from fungal and bacterial lipopeptides.
Figure 62. Unusual FA derived from fungal and bacterial lipopeptides.
Hydrobiology 01 00024 g062
Figure 63. 3D graph shows the predicted and calculated antiviral activity of FA (346, 347, 348 and 349) with a confidence level of over 95%. All dienoic acids were produced by proteobacteria Polyangium brachysporum.
Figure 63. 3D graph shows the predicted and calculated antiviral activity of FA (346, 347, 348 and 349) with a confidence level of over 95%. All dienoic acids were produced by proteobacteria Polyangium brachysporum.
Hydrobiology 01 00024 g063
Figure 64. Specimens of Streptomyces fungus of various strains inhabiting the various ecosystems. Many fungi belonging to this genus synthesizes a lot of biologically active metabolites, including linear and cyclic peptides and their FA. Pictures adapted by the author.
Figure 64. Specimens of Streptomyces fungus of various strains inhabiting the various ecosystems. Many fungi belonging to this genus synthesizes a lot of biologically active metabolites, including linear and cyclic peptides and their FA. Pictures adapted by the author.
Hydrobiology 01 00024 g064
Table 1. Biological activity cryptophycin-1 and its subunits A, B, C and D.
Table 1. Biological activity cryptophycin-1 and its subunits A, B, C and D.
Fragment NamePredicted Biological Activity, Pa *Reported Activity [46,47,48,49]
Cryptophycin-1Antifungal (0.845)
Antimitotic (0.784)
Antineoplastic (0.771)
Antineoplastic (solid tumors) (0.631)
Apoptosis agonist (0.625)
Antifungal
Anticancer
Apoptosis
Subunit A (1)Antineoplastic (0.856)
Antileukemic (0.783)
Antiviral (Arbovirus) (0.775)
Antifungal (0.773)
Cytoprotectant (0.675)
Apoptosis agonist (0.669)
Fibrinolytic (0.664)
Antimitotic (0.634)
Antithrombotic (0.624)
No published data
Subunit BPreneoplastic conditions treatment (0.836)
Antiviral (Arbovirus) (0.774)
Acute neurologic disorders treatment (0.759)
Antiviral (Picornavirus) (0.715)
No published data
Subunit CFibrinolytic (0.814)
Preneoplastic conditions treatment (0.803)
Antiviral (Arbovirus) (0.760)
Antimutagenic (0.737)
Anticonvulsant (0.676)
Antiviral (Picornavirus) (0.662)
No published data
Subunit DAnti-ischemic, cerebral (0.921)
Sclerosant (0.871)
Antihypertensive (0.756)
Anti-hypoxic (0.741)
Antiviral (Arbovirus) (0.740)
No published data
* Only activities with Pa > 0.5 are shown. The numbers in brackets show the level of biological activity. 100% activity level is 1.000.
Table 2. Predicted biological activity of subunit A (FA) incorporated into cryptophycins.
Table 2. Predicted biological activity of subunit A (FA) incorporated into cryptophycins.
No.Predicted Biological Activity, Pa *
2Antiviral (Arbovirus) (0.780); Antineoplastic (0.766); Antifungal (0.707)
Lipid metabolism regulator (0.705); Hypolipemic (0.694)
Preneoplastic conditions treatment (0.667); Apoptosis agonist (0.664)
3Antiviral (Arbovirus) (0.804); Antineoplastic (0.795); Antidiabetic (0.701)
Antifungal (0.695); Anti-inflammatory (0.680); Cytoprotectant (0.661)
Immunosuppressant (0.634); Anti-hypercholesterolemic (0.629)
Antithrombotic (0.611)
4Antiviral (Arbovirus) (0.833); Lipid metabolism regulator (0.827)
Anti-inflammatory (0.765); Hypolipemic (0.759); Cytoprotectant (0.729)
Anti-hypercholesterolemic (0.715); Antineoplastic (0.697)
5Antineoplastic (0.850); Antileukemic (0.781); Antifungal (0.727)
Anti-hypoxic (0.702); Antiviral (Arbovirus) (0.683); Cytoprotectant (0.638)
6Antineoplastic (0.764); Apoptosis agonist (0.762)
Antiviral (Arbovirus) (0.728); Antimitotic (0.664); Antifungal (0.560)
7Antiviral (Arbovirus) (0.730); Antifungal (0.658); Antineoplastic (0.611)
8Antineoplastic (0.776); Antifungal (0.694); Anti-helmintic (0.691)
Antidiabetic (0.660); Acute neurologic disorders treatment (0.625)
Antibacterial (0.624); Antiviral (Arbovirus) (0.621)
Antiviral (Picornavirus) (0.608)
9Antineoplastic (0.856); Antileukemic (0.783); Antiviral (Arbovirus) (0.775)
Antifungal (0.773); Anti-hypercholesterolemic (0.742); Apoptosis agonist (0.669)
* Only activities with Pa > 0.5 are shown. The numbers in brackets show the level of biological activity. 100% activity level is 1.000.
Table 3. Predicted biological activity of FA from cyanobacteria of the strain Nostoc.
Table 3. Predicted biological activity of FA from cyanobacteria of the strain Nostoc.
No.Predicted Biological Activity, Pa *
10Neuroprotector (0.806); Sclerosant (0.779); Anticonvulsant (0.734)
Acute neurologic disorders treatment (0.684); Anti-inflammatory (0.681)
Antineoplastic (0.631); Preneoplastic conditions treatment (0.628); Anti-neurogenic pain (0.610)
11Neuroprotector (0.827); Antineoplastic (0.708); Preneoplastic conditions treatment (0.672)
Anticonvulsant (0.672); Antiviral (Arbovirus) (0.647); Psychostimulant (0.643)
12Neuroprotector (0.806); Sclerosant (0.779); Acute neurologic disorders treatment (0.684)
Antineoplastic (0.631); Preneoplastic conditions treatment (0.628)
13Anticonvulsant (0.797); Hypolipemic (0.762); Acute neurologic disorders treatment (0.759)
Neuroprotector (0.739); Sclerosant (0.727); Antineoplastic (0.638)
14Sclerosant (0.767); Neuroprotector (0.748); Antineoplastic (0.666)
Acute neurologic disorders treatment (0.645); Antiviral (Arbovirus) (0.595)
15Anti-psoriatic (0.923); Antineoplastic (0.883); Neuroprotector (0.675)
Antiviral (Arbovirus) (0.635); Neurodegenerative diseases treatment (0.620)
Alzheimer’s disease treatment (0.591)
16Sclerosant (0.767); Antifungal (0.735); Antineoplastic (0.730)
17Antineoplastic (0.812); Anti-inflammatory (0.763); Apoptosis agonist (0.691)
18Sclerosant (0.767); Antineoplastic (0.666); Acute neurologic disorders treatment (0.645)
19Acute neurologic disorders treatment (0.795); Sclerosant (0.754)
Lipid metabolism regulator (0.749); Antiviral (Arbovirus) (0.704)
20Sclerosant (0.910); Antiviral (Arbovirus) (0.784); Acute neurologic disorders treatment (0.747)
Preneoplastic conditions treatment (0.714); Lipid metabolism regulator (0.667)
21Antineoplastic (0.758); Neuroprotector (0.752); Antiviral (Arbovirus) (0.636)
22Neuroprotector (0.756); Periodontitis treatment (0.744); Antineoplastic (0.680)
Preneoplastic conditions treatment (0.655); Psychostimulant (0.545)
23Antineoplastic (0.765); Neuroprotector (0.724); Lipid metabolism regulator (0.720)
Apoptosis agonist (0.625); Acute neurologic disorders treatment (0.566)
24Neuroprotector (0.800); Antineoplastic (0.738); Anticonvulsant (0.678)
25Anti-psoriatic (0.924); Lipid metabolism regulator (0.889); Antineoplastic (0.867)
* Only activities with Pa > 0.5 are shown.
Table 4. Predicted biological activity of FA from cyanobacteria of the strain Lyngbya.
Table 4. Predicted biological activity of FA from cyanobacteria of the strain Lyngbya.
No.Predicted Biological Activity, Pa *
26Antineoplastic (0.855); Lipid metabolism regulator (0.842)
Apoptosis agonist (0.803)
27Sclerosant (0.871); Acute neurologic disorders treatment (0.870)
Antineoplastic (0.714)
28Antineoplastic (0.906); Hypolipemic (0.902); Lipid metabolism regulator (0.879)
Apoptosis agonist (0.811); Acute neurologic disorders treatment (0.670)
Atherosclerosis treatment (0.662); Proliferative diseases treatment (0.562)
29Cholesterol antagonist (0.649); Antianginal (0.628)
Lipid metabolism regulator (0.532)
30Cholesterol antagonist (0.649); Antianginal (0.628)
Lipid metabolism regulator (0.532)
31Iron antagonist (0.952); Antineoplastic (0.736)
Microtubule formation stimulant (0.578)
32Neurodegenerative diseases treatment (0.688); Bone diseases treatment (0.528)
33Dermatologic (0.667); Anti-psoriatic (0.663)
* Only activities with Pa > 0.5 are shown.
Table 5. Predicted biological activity of FA from peptides of Lyngbya majuscula.
Table 5. Predicted biological activity of FA from peptides of Lyngbya majuscula.
No.Predicted Biological Activity, Pa *
34Hypolipemic (0.921); Anti-psoriatic (0.913); Antidiabetic (0.897); Antineoplastic (0.799); Anti-obesity (0.783); Antihypertriglyceridemic (0.766); Lipid metabolism regulator (0.633)
35Hypolipemic (0.910); Antidiabetic (0.902); Antihypertriglyceridemic (0.761)
36Sclerosant (0.741); Inflammatory Bowel disease treatment (0.607); Antibacterial (0.588)
37Antineoplastic (0.880); Hypolipemic (0.858); Lipid metabolism regulator (0.608)
38Antineoplastic (0.822); Hypolipemic (0.801); Lipid metabolism regulator (0.700)
39Hypolipemic (0.890); Lipid metabolism regulator (0.813); Anti-hypercholesterolemic (0.704)
Atherosclerosis treatment (0.657); Cholesterol synthesis inhibitor (0.542)
40Hypolipemic (0.858); Antineoplastic (0.810); Lipid metabolism regulator (0.608)
41Antineoplastic (0.897); Antifungal (0.787); Apoptosis agonist (0.696)
42Antineoplastic (0.869); Antifungal (0.728); Apoptosis agonist (0.650)
43Antineoplastic (0.895); Antifungal (0.802); Apoptosis agonist (0.734)
44Antineoplastic (0.854); Antifungal (0.801); Antibacterial (0.708); Apoptosis agonist (0.653)
AHypolipemic (0.910); Antidiabetic (0.902); Antihypertriglyceridemic (0.761)
BAnti-ischemic, cerebral (0.756); Neuroprotector (0.718); Antiviral (Arbovirus) (0.687); Genital warts treatment (0.648); Antineoplastic (liver cancer) (0.582); Antimetastatic (0.540)
CAntiviral (Arbovirus) (0.732); Neuroprotector (0.726); Antineoplastic (liver cancer) (0.633); Acute neurologic disorders treatment (0.568); Antimitotic (0.567)
DAntineoplastic (liver cancer) (0.923); Antineoplastic (0.685); Anti-ischemic, cerebral (0.664)
* Only activities with Pa > 0.5 are shown.
Table 6. Predicted biological activity of FA from peptides of cyanobacteria.
Table 6. Predicted biological activity of FA from peptides of cyanobacteria.
No.Predicted Biological Activity, Pa *
45Lipid metabolism regulator (0.928); Anti-hypercholesterolemic (0.738)
Preneoplastic conditions treatment (0.722); Antihypoxic (0.711)
Hypolipemic (0.676)
46Psychostimulant (0.768); Antiviral (0.766)
Preneoplastic conditions treatment (0.678)
47Psychostimulant (0.731); Antiviral (0.731)
Preneoplastic conditions treatment (0.718)
48Lipid metabolism regulator (0.885); Hypolipemic (0.772)
Anti-hypercholesterolemic (0.735)
49Lipid metabolism regulator (0.843); Preneoplastic conditions treatment (0.832)
Antimutagenic (0.832); Acute neurologic disorders treatment (0.691)
50Fibrinolytic (0.893); Preneoplastic conditions treatment (0.804)
51Antidiabetic (0.886); Inflammatory Bowel disease treatment (0.852)
52Antidiabetic (0.886); Inflammatory Bowel disease treatment (0.852)
53Antidiabetic (0.886); Inflammatory Bowel disease treatment (0.852)
54Antidiabetic (0.916); Antineoplastic (0.831); Immunosuppressant (0.681)
55Hypolipemic (0.790); Lipid metabolism regulator (0.738)
Atherosclerosis treatment (0.679)
56Lipid metabolism regulator (0.908); Hypolipemic (0.881)
Anti-hypercholesterolemic (0.765)
Acute neurologic disorders treatment (0.734); Atherosclerosis treatment (0.722)
57Antineoplastic (0.834); Lipid metabolism regulator (0.831)
Apoptosis agonist (0.818)
Acute neurologic disorders treatment (0.795); Hypolipemic (0.725); Atherosclerosis treatment (0.617)
58Antineoplastic (0.703); Lipid metabolism regulator (0.678)
Antiviral (Arbovirus) (0.643)
59Lipid metabolism regulator (0.886); Hypolipemic (0.823)
Anti-hypercholesterolemic (0.742)
60Anti-infective (0.877); Lipid metabolism regulator (0.866)
Antiviral (Arbovirus) (0.827)
* Only activities with Pa > 0.5 are shown.
Table 7. Predicted biological activity of FA derived from cyanobacterial peptides.
Table 7. Predicted biological activity of FA derived from cyanobacterial peptides.
No.Predicted Biological Activity, Pa *
61Hypolipemic (0.861); Antineoplastic (0.854); Lipid metabolism regulator (0.849)
62Lipid metabolism regulator (0.934); Hypolipemic (0.903); Sclerosant (0.869)
Acute neurologic disorders treatment (0.845); Anti-hypercholesterolemic (0.831)
Atherosclerosis treatment (0.705); Cholesterol synthesis inhibitor (0.519)
63Lipid metabolism regulator (0.959); Anti-hypercholesterolemic (0.900); Hypolipemic (0.895)
Acute neurologic disorders treatment (0.893); Atherosclerosis treatment (0.696)
Multiple sclerosis treatment (0.548); Antibacterial (0.537); Cholesterol synthesis inhibitor (0.526)
64Hypolipemic (0.878); Acute neurologic disorders treatment (0.780)
Lipid metabolism regulator (0.771); Atherosclerosis treatment (0.675)
65Lipid metabolism regulator (0.941); Hypolipemic (0.829); Anti-hypercholesterolemic (0.792)
Acute neurologic disorder treatment (0.697); Atherosclerosis treatment (0.647)
66Antiviral (Arbovirus) (0.947); Lipid metabolism regulator (0.884); Antimutagenic (0.793)
Antiviral (Picornavirus) (0.782); Hypolipemic (0.724); Anti-hypercholesterolemic (0.695)
67Hypolipemic (0.835); Lipid metabolism regulator (0.830); Antiviral (Arbovirus) (0.821)
Antiviral (Picornavirus) (0.761); Preneoplastic conditions treatment (0.722)
68Lipid metabolism regulator (0.953); Antiviral (Arbovirus) (0.885)
Hypolipemic (0.884); Anti-hypercholesterolemic (0.856); Leukopoiesis stimulant (0.826); Atherosclerosis treatment (0.692)
69Lipid metabolism regulator (0.934); Hypolipemic (0.903); Anti-hypercholesterolemic (0.831)
Atherosclerosis treatment (0.705); Cholesterol synthesis inhibitor (0.519)
70Lipid metabolism regulator (0.934); Hypolipemic (0.903); Anti-hypercholesterolemic (0.831)
Atherosclerosis treatment (0.705); Cholesterol synthesis inhibitor (0.519)
71Lipid metabolism regulator (0.934); Hypolipemic (0.903); Anti-hypercholesterolemic (0.831)
Atherosclerosis treatment (0.705); Cholesterol synthesis inhibitor (0.519)
72Vasoprotector (0.890); Lipid metabolism regulator (0.884); Antiviral (Arbovirus) (0.824)
73Antipsoriatic (0.957); Antineoplastic (0.886); Antiviral (Arbovirus) (0.796)
Lipid metabolism regulator (0.792); Alzheimer’s disease treatment (0.613)
74Lipid metabolism regulator (0.929); Antiviral (Arbovirus) (0.891); Antimutagenic (0.866)
Anti-hypercholesterolemic (0.786); Hypolipemic (0.721); Atherosclerosis treatment (0.629)
* Only activities with Pa > 0.5 are shown.
Table 8. Predicted biological activity of FA from cyanopeptides.
Table 8. Predicted biological activity of FA from cyanopeptides.
No.Predicted Biological Activity, Pa *
75Antifungal (0.791); Antineoplastic (0.744)
Lipid metabolism regulator (0.680)
76Antieczematic (0.869); Anesthetic general (0.722)
Neuroprotector (0.714)
77Antieczematic (0.900); Lipid metabolism regulator (0.859)
Antifungal (0.756)
78Antineoplastic (0.960); Preneoplastic conditions treatment (0.661)
Antiviral (Arbovirus) (0.618); Antiviral (Picornavirus) (0.520)
79Cystic fibrosis treatment (0.861); Antiviral (Arbovirus) (0.717); Anesthetic general (0.702)
80Cystic fibrosis treatment (0.850); Anesthetic general (0.733)
Antiviral (Arbovirus) (0.687)
81Anti-hypercholesterolemic (0.902); Apoptosis agonist (0.779); Antineoplastic (0.779)
82Lipid metabolism regulator (0.888); Angiogenesis stimulant (0.869); Expectorant (0.715)
83Acute neurologic disorders treatment (0.870); Anti-inflammatory (0.776)
84Acute neurologic disorders treatment (0.870); Anti-inflammatory (0.776)
85Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842)
Hypolipemic (0.830)
86Lipid metabolism regulator (0.848); Anti-hypercholesterolemic (0.770)
Hypolipemic (0.760)
87Lipid metabolism regulator (0.848); Anti-hypercholesterolemic (0.770)
88Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842)
Hypolipemic (0.830)
89Lipid metabolism regulator (0.848); Anti-hypercholesterolemic (0.770)
Hypolipemic (0.760)
* Only activities with Pa > 0.5 are shown.
Table 9. Predicted biological activity of FA from cyanobacterial peptides.
Table 9. Predicted biological activity of FA from cyanobacterial peptides.
No.Predicted Biological Activity, Pa *
90Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842) Hypolipemic (0.830)
Atherosclerosis treatment (0.635); Multiple sclerosis treatment (0.507)
91Lipid metabolism regulator (0.848); Anti-hypercholesterolemic (0.770); Hypolipemic (0.760)
Acute neurologic disorders treatment (0.696); Atherosclerosis treatment (0.551)
92Lipid metabolism regulator (0.848); Anti-hypercholesterolemic (0.770); Hypolipemic (0.760)
Acute neurologic disorders treatment (0.696); Atherosclerosis treatment (0.551)
93Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842); Hypolipemic (0.830)
Acute neurologic disorders treatment (0.690); Atherosclerosis treatment (0.635)
94Antineoplastic (0.859); Apoptosis agonist (0.765); Antimitotic (0.689)
95Antineoplastic (0.778); Apoptosis agonist (0.724); Antiviral (Arbovirus) (0.695)
Preneoplastic conditions treatment (0.551)
96Antineoplastic (0.800); Apoptosis agonist (0.794); Lipid metabolism regulator (0.780)
Antifungal (0.750); Preneoplastic conditions treatment (0.622); Spasmolytic (0.540)
97Antineoplastic (0.877); Apoptosis agonist (0.818); Hypolipemic (0.809); Antifungal (0.775)
Anti-inflammatory (0.748); Preneoplastic conditions treatment (0.540)
98Antineoplastic (0.881); Apoptosis agonist (0.814); Hypolipemic (0.810)
Antifungal (0.767); Anti-inflammatory (0.746); Lipid metabolism regulator (0.706)
99Antineoplastic (0.871); Hypolipemic (0.815); Antifungal (0.775); Apoptosis agonist (0.764)
Anti-inflammatory (0.729); Lipid metabolism regulator (0.643)
100Antineoplastic (0.812); Anti-inflammatory (0.763); Immunosuppressant (0.712)
Antifungal (0.695); Apoptosis agonist (0.691); Hypolipemic (0.667)
* Only activities with Pa > 0.5 are shown.
Table 10. Predicted biological activity of FA derived from cyanopeptides.
Table 10. Predicted biological activity of FA derived from cyanopeptides.
No.Predicted Biological Activity, Pa *
101Anti-hypercholesterolemic (0.936); Antifungal (0.848); Antibacterial (0.798)
Hypolipemic (0.723); Anti-infective (0.718); Atherosclerosis treatment (0.522)
102Anti-infective (0.936); Anti-hypercholesterolemic (0.928); Antifungal (0.853)
Antioxidant (0.848); Antineoplastic (0.833); Antidiabetic (0.807)
Antibacterial (0.774); Hypolipemic (0.774)
Acute neurologic disorders treatment (0.702)
Proliferative diseases treatment (0.690); Atherosclerosis treatment (0.603)
103Vasoprotector (0.970); Anti-infective (0.966); Hemostatic (0.950)
Neuroprotector (0.942)
Anti-hypercholesterolemic (0.915); Lipid metabolism regulator (0.903)
Acute neurologic disorders treatment (0.845); Hypolipemic (0.770)
Atherosclerosis treatment (0.624); DNA synthesis inhibitor (0.584)
Dementia treatment (0.582)
104Anti-infective (0.961); Vasoprotector (0.960); Neuroprotector (0.905)
Anti-hypercholesterolemic (0.875); Antithrombotic (0.850)
Hypolipemic (0.757); Atherosclerosis treatment (0.627); DNA synthesis inhibitor (0.590)
105Anti-infective (0.961); Vasoprotector (0.960); Neuroprotector (0.905); Sclerosant (0.891)
Anti-hypercholesterolemic (0.875); Antithrombotic (0.850); Lipid metabolism regulator (0.807)
Hypolipemic (0.757); Atherosclerosis treatment (0.627); DNA synthesis inhibitor (0.590)
106Anti-infective (0.966); Vasoprotector (0.953); Anti-hypercholesterolemic (0.890)
Antihypoxic (0.881); Lipid metabolism regulator (0.853); Antineoplastic (0.835)
Hypolipemic (0.751); Acute neurologic disorders treatment (0.684); DNA synthesis inhibitor (0.566)
107Vasodilator (0.868); Anti-infective (0.865); Antifungal (0.829); Antineoplastic (0.809)
Anti-hypercholesterolemic (0.609); Hypolipemic (0.566); Antimycobacterial (0.545)
* Only activities with Pa > 0.5 are shown.
Table 11. Predicted biological activity of FA from peptides of seaweeds and molluscs.
Table 11. Predicted biological activity of FA from peptides of seaweeds and molluscs.
No.Predicted Biological Activity, Pa *
108Hypolipemic (0.858); Lipid metabolism regulator (0.835); Anti-hypercholesterolemic (0.634)
Antifungal (0.648); Atherosclerosis treatment (0.603)
109Hypolipemic (0.858); Lipid metabolism regulator (0.835); Anti-hypercholesterolemic (0.634)
Antifungal (0.648); Atherosclerosis treatment (0.603)
110Preneoplastic conditions treatment (0.779); Hypolipemic (0.775); Anesthetic general (0.772)
Lipid metabolism regulator (0.768); Acute neurologic disorders treatment (0.694)
111Preneoplastic conditions treatment (0.805); Acute neurologic disorders treatment (0.723)
Antiviral (Arbovirus) (0.716); Anti-inflammatory (0.650); Antiviral (Picornavirus) (0.649)
112Lipid metabolism regulator (0.890); Hypolipemic (0.870); Anti-hypercholesterolemic (0.802)
Atherosclerosis treatment (0.692); Cholesterol synthesis inhibitor (0.511)
113Lipid metabolism regulator (0.895); Preneoplastic conditions treatment (0.778)
Anti-hypercholesterolemic (0.777); Hypolipemic (0.758); Atherosclerosis treatment (0.683)
114Mucositis treatment (0.886); Anesthetic general (0.852); Lipid metabolism regulator (0.842)
Autoimmune disorders treatment (0.798); Transplant rejection treatment (0.795)
115Anti-ischemic, cerebral (0.943); Acute neurologic disorders treatment (0.797)
Anticonvulsant (0.702); Anti-hypercholesterolemic (0.642); Antihypertensive (0.627)
116Lipid metabolism regulator (0.822); Vasodilator, peripheral (0.803); Vasoprotector (0.793)
Hypolipemic (0.757); Anti-hypercholesterolemic (0.677); Atherosclerosis treatment (0.647)
117Lipid metabolism regulator (0.890); Hypolipemic (0.870); Anti-hypercholesterolemic (0.802)
Atherosclerosis treatment (0.692); Cholesterol synthesis inhibitor (0.511)
* Only activities with Pa > 0.5 are shown.
Table 12. Predicted biological activity of FA from peptides of marine sponges.
Table 12. Predicted biological activity of FA from peptides of marine sponges.
No.Predicted Biological Activity, Pa *
118Antineoplastic (0.883); Lipid metabolism regulator (0.836); Anti-inflammatory (0.845)
Apoptosis agonist (0.847); Acute neurologic disorders treatment (0.795); Antifungal (0.793)
119Phobic disorders treatment (0.859); Psychostimulant (0.731); Antiviral (0.731)
Acute neurologic disorders treatment (0.586); Neuroprotector (0.574)
120Antiarthritic (0.805); Preneoplastic conditions treatment (0.730); Sclerosant (0.726)
Acute neurologic disorders treatment (0.696); Anti-inflammatory (0.641)
121Antineoplastic (0.813); Antiviral (Arbovirus) (0.748); Lipid metabolism regulator (0.693)
Cytoprotectant (0.668); Antiviral (Picornavirus) (0.585); Hypolipemic (0.575)
122Lipid metabolism regulator (0.880); Antineoplastic (0.863); Hypolipemic (0.816)
Anti-hypercholesterolemic (0.672); Atherosclerosis treatment (0.590)
123Lipid metabolism regulator (0.924); Antineoplastic (0.873); Hypolipemic (0.839)
Anti-hypercholesterolemic (0.642); Atherosclerosis treatment (0.592)
124Hypolipemic (0.911); Lipid metabolism regulator (0.829); Anti-inflammatory (0.765)
Anti-hypercholesterolemic (0.718); Acute neurologic disorders treatment (0.715)
125Lipid metabolism regulator (0.929); Hypolipemic (0.908)
Anti-hypercholesterolemic (0.825); Atherosclerosis treatment (0.680)
126Antineoplastic (0.788); Hypolipemic (0.754); Acute neurologic disorders treatment (0.687)
127Dermatologic (0.909); Anti-psoriatic (0.888); Anti-eczematic (0.856); Antifungal (0.605)
128Anti-psoriatic (0.862); Antineoplastic (0.846); Antifungal (0.625); Anti-eczematic (0.601)
129Hypolipemic (0.880); Antineoplastic (0.821); Antifungal (0.707); Antiviral (Arbovirus) (0.654)
130Hypolipemic (0.880); Antineoplastic (0.821); Antifungal (0.707); Antibacterial (0.555)
131Hypolipemic (0.880); Antineoplastic (0.821); Antifungal (0.707); Antibacterial (0.555)
132Antifungal (0.688); Antiprotozoal (Plasmodium) (0.570); Antibacterial (0.514)
133Antineoplastic (0.876); Antifungal (0.771); Lipid metabolism regulator (0.763); Hypolipemic (0.707)
* Only activities with Pa > 0.5 are shown.
Table 13. Predicted biological activity of FA derived from sponge peptides.
Table 13. Predicted biological activity of FA derived from sponge peptides.
No.Predicted Biological Activity, Pa *
134Sclerosant (0.834); Hypolipemic (0.825); Antineoplastic (0.779); Anti-inflammatory (0.731)
135Sclerosant (0.835); Antineoplastic (0.788); Hypolipemic (0.754); Anti-inflammatory (0.716)
136Sclerosant (0.853); Hypolipemic (0.824); Antineoplastic (0.775); Anti-inflammatory (0.734)
137Sclerosant (0.815); Hypolipemic (0.807); Antineoplastic (0.781); Anti-inflammatory (0.730)
138Sclerosant (0.834); Hypolipemic (0.825); Antineoplastic (0.779); Anti-inflammatory (0.731)
139Lipid metabolism regulator (0.903); Hypolipemic (0.848); Antineoplastic (0.805); Antifungal (0.782)
140Sclerosant (0.834); Hypolipemic (0.825); Antineoplastic (0.779); Anti-inflammatory (0.731)
141Restenosis treatment (0.827); Sclerosant (0.738); Neurodegenerative diseases treatment (0.722)
142Sclerosant (0.834); Hypolipemic (0.825); Antineoplastic (0.779); Anti-inflammatory (0.731)
143Antineoplastic (0.880); Antiviral (Arbovirus) (0.829); Apoptosis agonist (0.804)
Hypolipemic (0.794); Antiprotozoal (Coccidial) (0.684); Antiviral (Picornavirus) (0.599)
144Antineoplastic (0.885); Antiviral (Arbovirus) (0.814); Apoptosis agonist (0.800)
Hypolipemic (0.794); Antiprotozoal (Coccidial) (0.621); Antiviral (Picornavirus) (0.598)
145Anti-infective (0.934); Anti-hypercholesterolemic (0.916); Vasodilator (0.915)
Antineoplastic (0.911); Vasoprotector (0.864); Lipid metabolism regulator (0.856)
146Anti-infective (0.934); Anti-hypercholesterolemic (0.916); Vasodilator (0.915)
Antineoplastic (0.911); Vasoprotector (0.864); Lipid metabolism regulator (0.856)
147Anti-infective (0.934); Anti-hypercholesterolemic (0.916); Vasodilator (0.915)
Antineoplastic (0.911); Vasoprotector (0.864); Lipid metabolism regulator (0.856)
* Only activities with Pa > 0.5 are shown.
Table 14. Predicted biological activity of FA derived from peptides of marine sponges.
Table 14. Predicted biological activity of FA derived from peptides of marine sponges.
No.Predicted Biological Activity, Pa *
148Antineoplastic (0.752); Lipid metabolism regulator (0.715); Antiviral (Arbovirus) (0.642)
149Antineoplastic (0.752); Lipid metabolism regulator (0.715); Antiviral (Arbovirus) (0.642)
150Cell adhesion molecule inhibitor (0.883); Lipid metabolism regulator (0.801); Apoptosis agonist (0.752)
151Cell adhesion molecule inhibitor (0.866); Hypolipemic (0.816); Antineoplastic (0.780)
152Lipid metabolism regulator (0.931); Antineoplastic (0.826); Apoptosis agonist (0.634)
153Lipid metabolism regulator (0.931); Antineoplastic (0.826); Apoptosis agonist (0.634)
154Antineoplastic (0.813); Antiviral (Arbovirus) (0.748); Lipid metabolism regulator (0.693)
156Antineoplastic (0.813); Antiviral (Arbovirus) (0.748); Lipid metabolism regulator (0.693)
157Antineoplastic (0.779); Acute neurologic disorders treatment (0.681); Antiviral (Arbovirus) (0.680)
158Sclerosant (0.815); Antineoplastic (0.781); Acute neurologic disorders treatment (0.722)
159Sclerosant (0.834); Antineoplastic (0.799); Acute neurologic disorders treatment (0.725)
160Sclerosant (0.835); Antineoplastic (0.795); Acute neurologic disorders treatment (0.731)
161Sclerosant (0.815); Antineoplastic (0.781); Acute neurologic disorders treatment (0.722)
162Sclerosant (0.835); Antineoplastic (0.781); Acute neurologic disorders treatment (0.764)
163Sclerosant (0.815); Antineoplastic (0.781); Acute neurologic disorders treatment (0.722)
164Antineoplastic (0.881); Hypolipemic (0.797); Antifungal (0.793); Antimitotic (0.787)
* Only activities with Pa > 0.5 are shown.
Table 15. Predicted biological activity of FA from lipopeptides of Dysidea species.
Table 15. Predicted biological activity of FA from lipopeptides of Dysidea species.
No.Predicted Biological Activity, Pa *
165Antineoplastic (0.841); Preneoplastic conditions treatment (0.689); Antiprotozoal (0.586)
166Antineoplastic (0.841); Preneoplastic conditions treatment (0.689); Antiprotozoal (0.586)
167Antineoplastic (0.960); Anti-infective (0.613); Acute neurologic disorders treatment (0.572)
168Antineoplastic (0.774); Preneoplastic conditions treatment (0.690); Antiprotozoal (0.557)
169Antineoplastic (0.850); Antiviral (Arbovirus) (0.765); Acute neurologic disorders treatment (0.574)
170Antineoplastic (0.965); Anti-infective (0.628); Acute neurologic disorders treatment (0.589)
* Only activities with Pa > 0.5 are shown.
Table 16. Predicted biological activity of FA from sponge lipopeptides.
Table 16. Predicted biological activity of FA from sponge lipopeptides.
No.Predicted Biological Activity, Pa *
171Anti-hypercholesterolemic (0.873); Lipid metabolism regulator (0.713); Atherosclerosis treatment (0.559)
172Antineoplastic (0.962); Apoptosis agonist (0.955); Antiparasitic (0.703); Antiprotozoal (0.590)
173Antineoplastic (0.962); Apoptosis agonist (0.955); Antiparasitic (0.703); Antiprotozoal (0.590)
174Antineoplastic (0.962); Apoptosis agonist (0.955); Antiparasitic (0.703); Antiprotozoal (0.590)
175Antiseptic (0.945); Antiinfective (0.900); Preneoplastic conditions treatment (0.818)
176Acute neurologic disorders treatment (0.858); Antidiabetic (0.818); Antidiabetic (type 2) (0.645)
177Atherosclerosis treatment (0.857); Sweetener (0.635); Restenosis treatment (0.602)
178Antihypertensive (0.765); Antidiabetic (0.757); Antithrombotic (0.522)
179Antineoplastic (0.855); Transplant rejection treatment (0.591); Autoimmune disorders treatment (0.574)
180Antineoplastic (0.667); Angiogenesis stimulant (0.566); Antidiabetic (0.531)
* Only activities with Pa > 0.5 are shown.
Table 17. Predicted biological activity of FA from lipopeptides of molluscs.
Table 17. Predicted biological activity of FA from lipopeptides of molluscs.
No.Predicted Biological Activity, Pa *
181Antineoplastic (0.871); Apoptosis agonist (0.764); Anti-inflammatory (0.729)
182Anti-psoriatic (0.923); Anti-eczematic (0.721); Alzheimer’s disease treatment (0.591)
183Anti-psoriatic (0.924); Anti-eczematic (0.879); Antifungal (0.603)
184Anti-inflammatory (0.751); Antiviral (Arbovirus) (0.595); Antifungal (0.579)
185Anti-hypoxic (0.768); Antihypertensive (0.730); Antiviral (Picornavirus) (0.613)
186Anti-inflammatory (0.751); Antiviral (Arbovirus) (0.595); Antifungal (0.579)
187Lipid metabolism regulator (0.822); Hypolipemic (0.757); Anti-hypercholesterolemic (0.677)
188Lipid metabolism regulator (0.889); Hypolipemic (0.685); Anti-hypercholesterolemic (0.644)
189Antineoplastic (0.859); Apoptosis agonist (0.765); Atherosclerosis treatment (0.640)
190Antidiabetic (0.916); Leukopoiesis stimulant (0.822); Atherosclerosis treatment (0.568)
191Antineoplastic (0.834); Apoptosis agonist (0.818); Acute neurologic disorders treatment (0.795)
192Anesthetic general (0.759); Antiviral (Arbovirus) (0.737); Antitoxic (0.728); Antihypoxic (0.726)
193Antihypoxic (0.758); Antiviral (Arbovirus) (0.735); Antiviral (Picornavirus) (0.671)
194Antineoplastic (0.796); Apoptosis agonist (0.719); Antiparasitic (0.711); Antifungal (0.675)
195Acute neurologic disorders treatment (0.754); Antiviral (Arbovirus) (0.733); Antineurotic (0.555)
196Lipid metabolism regulator (0.809); Hypolipemic (0.758); Antihypertensive (0.741)
Anti-hypercholesterolemic (0.616); Atherosclerosis treatment (0.535)
197Hypolipemic (0.796); Antineoplastic (0.614); Preneoplastic conditions treatment (0.547)
198Hypolipemic (0.804); Antineoplastic (0.689); Preneoplastic conditions treatment (0.588)
* Only activities with Pa > 0.5 are shown.
Table 18. Predicted biological activity of FA from lipopeptides of tunicates.
Table 18. Predicted biological activity of FA from lipopeptides of tunicates.
No.Predicted Biological Activity, Pa *
199Neuroprotector (0.748); Immunosuppressant (0.710); Acute neurologic disorders treatment (0.645)
200Lipid metabolism regulator (0.901); Macrophage stimulant (0.829); Antineoplastic (0.777)
201Neuroprotector (0.748); Immunosuppressant (0.710); Acute neurologic disorders treatment (0.645)
202Lipid metabolism regulator (0.850); Antidiabetic symptomatic (0.740)
Atherosclerosis treatment (0.654); Hypolipemic (0.653); Antidiabetic (0.605)
203Lipid metabolism regulator (0.850); Antidiabetic symptomatic (0.740)
Atherosclerosis treatment (0.654); Hypolipemic (0.653); Antidiabetic (0.605)
204Lipid metabolism regulator (0.850); Antidiabetic symptomatic (0.740)
Atherosclerosis treatment (0.654); Hypolipemic (0.653); Antidiabetic (0.605)
205Lipid metabolism regulator (0.857); Apoptosis agonist (0.729); Immunosuppressant (0.689)
206Lipid metabolism regulator (0.879); Immunosuppressant (0.760); Apoptosis agonist (0.744)
207Hypolipemic (0.880); Lipid metabolism regulator (0.861); Anti-hypercholesterolemic (0.683)
Atherosclerosis treatment (0.634); Neurodegenerative diseases treatment (0.610)
208Restenosis treatment (0.827); Neurodegenerative diseases treatment (0.722); Antidiabetic (0.655)
209Restenosis treatment (0.827); Neurodegenerative diseases treatment (0.722); Antidiabetic (0.655)
210Restenosis treatment (0.827); Neurodegenerative diseases treatment (0.722); Antidiabetic (0.655)
211Mucositis treatment (0.803); Antimutagenic (0.727); Cytoprotectant (0.719)
* Only activities with Pa > 0.5 are shown.
Table 19. Predicted biological activity of FA from fungal lipopeptides.
Table 19. Predicted biological activity of FA from fungal lipopeptides.
No.Predicted Biological Activity, Pa *
212Hypolipemic (0.845); Antihypertensive (0.622); Lipid metabolism regulator (0.621)
213Lipid metabolism regulator (0.791); Hypolipemic (0.778); Antihypertensive (0.643)
214Acute neurologic disorders treatment (0.944); Lipid metabolism regulator (0.868)
215Natural killer cell stimulant (0.785); Antineoplastic (0.785); Leukopoiesis stimulant (0.733)
Acute neurologic disorders treatment (0.701); Immunosuppressant (0.697); Neuroprotector (0.664)
216Natural killer cell stimulant (0.785); Antineoplastic (0.785); Leukopoiesis stimulant (0.733)
Acute neurologic disorders treatment (0.701); Immunosuppressant (0.697); Neuroprotector (0.664)
217Hypolipemic (0.807); Acute neurologic disorders treatment (0.722); Immunosuppressant (0.710)
Leukopoiesis stimulant (0.687); Natural killer cell stimulant (0.685); Erythropoiesis stimulant (0.551)
218Natural killer cell stimulant (0.795); Leukopoiesis stimulant (0.784); Hypolipemic (0.765)
Immunosuppressant (0.702); Acute neurologic disorders treatment (0.696); Neuroprotector (0.684)
219Hypolipemic (0.851); Acute neurologic disorders treatment (0.844); Anticonvulsant (0.733)
Natural killer cell stimulant (0.715); Leukopoiesis stimulant (0.713)
220Hypolipemic (0.851); Acute neurologic disorders treatment (0.844); Anticonvulsant (0.733)
Natural killer cell stimulant (0.715); Leukopoiesis stimulant (0.713)
221Natural killer cell stimulant (0.795); Leukopoiesis stimulant (0.784); Hypolipemic (0.765)
Immunosuppressant (0.702); Acute neurologic disorders treatment (0.696); Neuroprotector (0.684)
222Hypolipemic (0.828); Acute neurologic disorders treatment (0.774); Natural killer cell stimulant (0.746)
Leukopoiesis stimulant (0.736); Immunosuppressant (0.726); Lipid metabolism regulator (0.685)
223Lipid metabolism regulator (0.881); Hypolipemic (0.851); Leukopoiesis stimulant (0.822)
Natural killer cell stimulant (0.795); Anti-hypercholesterolemic (0.728)
224Lipid metabolism regulator (0.881); Hypolipemic (0.851); Leukopoiesis stimulant (0.822)
Natural killer cell stimulant (0.795); Anti-hypercholesterolemic (0.728)
225Platelet antagonist (0.800); Acute neurologic disorders treatment (0.701); Fibrinolytic (0.700)
Natural killer cell stimulant (0.696); Erythropoiesis stimulant (0.619); Immunosuppressant (0.612)
226Leukopoiesis stimulant (0.780); Antineoplastic (0.773); Hypolipemic (0.748)
Natural killer cell stimulant (0.744); Lipid metabolism regulator (0.650)
227Lipid metabolism regulator (0.809); Leukopoiesis stimulant (0.776); Hypolipemic (0.758)
Antihypertensive (0.741); Natural killer cell stimulant (0.730); Antineoplastic (0.683)
228Antidiabetic (0.781); Leukopoiesis stimulant (0.725); Bone diseases treatment (0.679)
* Only activities with Pa > 0.5 are shown.
Table 20. Predicted biological activity of FA derived from fungal lipopeptides.
Table 20. Predicted biological activity of FA derived from fungal lipopeptides.
No.Predicted Biological Activity, Pa *
229Lipid metabolism regulator (0.861); Leukopoiesis stimulant (0.840); Hypolipemic (0.685)
230Fibrinolytic (0.760); Natural killer cell stimulant (0.746); Leukopoiesis stimulant (0.746)
231Antiviral (Arbovirus) (0.834); Antimutagenic (0.827); Mucositis treatment (0.812)
Preneoplastic conditions treatment (0.766); Lipid metabolism regulator (0.718)
232Antiviral (Arbovirus) (0.861); Lipid metabolism regulator (0.844); Antiviral (Picornavirus) (0.776)
233Antiviral (Arbovirus) (0.861); Lipid metabolism regulator (0.844); Antiviral (Picornavirus) (0.776)
234Antiviral (Arbovirus) (0.947); Lipid metabolism regulator (0.884); Antiviral (Picornavirus) (0.782)
235Antiviral (Arbovirus) (0.891); Lipid metabolism regulator (0.743); Antiviral (Picornavirus) (0.730)
236Apoptosis agonist (0.886); Antineoplastic (0.763); Chemoprotective (0.700)
237Antiviral (Arbovirus) (0.904); Lipid metabolism regulator (0.828); Antiviral (Picornavirus) (0.688)
238Antiviral (Arbovirus) (0.886); Lipid metabolism regulator (0.874); Antiviral (Picornavirus) (0.744)
239Lipid metabolism regulator (0.942); Anti-hypercholesterolemic (0.855); Hypolipemic (0.793)
240Antiviral (Arbovirus) (0.853); Antiviral (Picornavirus) (0.760); Apoptosis agonist (0.731)
241Antiviral (Arbovirus) (0.873); Lipid metabolism regulator (0.850); Antiviral (Picornavirus) (0.735)
242Anti-ischemic, cerebral (0.910); Leukopoiesis stimulant (0.708); Antidiabetic (0.698)
243Antiviral (Arbovirus) (0.944); Lipid metabolism regulator (0.882); Antiviral (Picornavirus) (0.766)
244Anti-inflammatory (0.785); Antineoplastic (0.782); Preneoplastic conditions treatment (0.661)
245Lipid metabolism regulator (0.846); Anti-hypercholesterolemic (0.827); Atherosclerosis treatment (0.538)
246Lipid metabolism regulator (0.943); Hypolipemic (0.860); Anti-hypercholesterolemic (0.844)
247Antineoplastic (0.844); Hypolipemic (0.782); Lipid metabolism regulator (0.699)
* Only activities with Pa > 0.5 are shown.
Table 21. Predicted biological activity of FA from lipopeptides of fungi.
Table 21. Predicted biological activity of FA from lipopeptides of fungi.
No.Predicted Biological Activity, Pa *
248Anti-ischemic, cerebral (0.919); Mucositis treatment (0.913); Lipid metabolism regulator (0.885)
Antithrombotic (0.745); Platelet antagonist (0.685); Antihypertensive (0.555)
249Anti-ischemic, cerebral (0.919); Mucositis treatment (0.913); Lipid metabolism regulator (0.885)
Antithrombotic (0.745); Platelet antagonist (0.685); Antihypertensive (0.555)
250Lipid metabolism regulator (0.923); Mucositis treatment (0.910); Antithrombotic (0.805)
Anti-ischemic, cerebral (0.801); Platelet antagonist (0.751); Fibrinolytic (0.684)
251Mucositis treatment (0.960); Lipid metabolism regulator (0.869); Platelet antagonist (0.836)
Antithrombotic (0.773); Anti-ischemic, cerebral (0.740); Fibrinolytic (0.717)
252Antineoplastic (0.844); Antiviral (Arbovirus) (0.790); Antiviral (Picornavirus) (0.659)
253Antihypertensive (0.962); Anti-Helicobacter pylori (0.839); Chemoprotective (0.718)
Antineoplastic (myeloid leukemia) (0.707); Apoptosis agonist (0.674); Antineoplastic (0.656)
254Anti-Helicobacter pylori (0.744); Antiviral (Arbovirus) (0.715); Antiviral (Picornavirus) (0.547)
255Hypolipemic (0.795); Antineoplastic (0.791); Antiviral (Arbovirus) (0.742)
Preneoplastic conditions treatment (0.740); Lipid metabolism regulator (0.739)
256Lipid metabolism regulator (0.902); Anti-hypercholesterolemic (0.871)
Hypolipemic (0.812); Atherosclerosis treatment (0.603)
257Antiviral (Arbovirus) (0.890); Antiviral (Picornavirus) (0.817); Mucositis treatment (0.743)
258Hypolipemic (0.845); Antihypertensive (0.622); Lipid metabolism regulator (0.621)
Atherosclerosis treatment (0.590); Antiprotozoal (Coccidial) (0.569); Antibacterial (0.562)
259Lipid metabolism regulator (0.874); Anti-hypercholesterolemic (0.785); Hypolipemic (0.707)
260Lipid metabolism regulator (0.942); Anti-hypercholesterolemic (0.855); Hypolipemic (0.793)
Atherosclerosis treatment (0.635); Multiple sclerosis treatment (0.521)
261Antiviral (Arbovirus) (0.927); Lipid metabolism regulator (0.887); Antiviral (Picornavirus) (0.669)
YM-170320Antifungal (0.867); Antibacterial (0.744); Antineoplastic (0.711); Chemoprotective (0.594)
262Acute neurologic disorders treatment (0.944); Lipid metabolism regulator (0.868)
Antineoplastic (0.814); Antifungal (0.829); Antibacterial (0.698)
* Only activities with Pa > 0.5 are shown.
Table 22. Predicted biological activity of FA derived from lipopeptides of fungi.
Table 22. Predicted biological activity of FA derived from lipopeptides of fungi.
No.Predicted Biological Activity, Pa *
263Antiviral (Arbovirus) (0.927); Lipid metabolism regulator (0.887); Antiviral (Picornavirus) (0.669)
264Antitoxic (0.733); Leukopoiesis stimulant (0.730); Preneoplastic conditions treatment (0.721)
265Antiviral (Arbovirus) (0.861); Lipid metabolism regulator (0.844); Antiviral (Picornavirus) (0.776)
266Antiviral (Arbovirus) (0.860); Antiviral (Picornavirus) (0.751); Antidiabetic (0.713)
267Antiviral (Arbovirus) (0.861); Lipid metabolism regulator (0.844); Antiviral (Picornavirus) (0.776)
268Antiviral (Arbovirus) (0.947); Lipid metabolism regulator (0.884); Antiviral (Picornavirus) (0.782)
269Antineoplastic (0.902); DNA synthesis inhibitor (0.675); Angiogenesis inhibitor (0.668)
270Antineoplastic (0.872); Apoptosis agonist (0.653); Antiarthritic (0.611)
271Hypolipemic (0.808); Anti-hypercholesterolemic (0.695); Lipid metabolism regulator (0.691)
272Lipid metabolism regulator (0.887); Atherosclerosis treatment (0.690); Hypolipemic (0.647)
273Lipid metabolism regulator (0.886); Hypolipemic (0.788); Anti-hypercholesterolemic (0.755)
274Lipid metabolism regulator (0.886); Hypolipemic (0.788); Anti-hypercholesterolemic (0.755)
275Apoptosis agonist (0.886); Antineoplastic (0.763); Chemoprotective (0.700)
276Apoptosis agonist (0.781); Antimutagenic (0.777); Preneoplastic conditions treatment (0.600)
277Natural killer cell stimulant (0.795); Leukopoiesis stimulant (0.784); Antineoplastic (0.778)
278Natural killer cell stimulant (0.795); Leukopoiesis stimulant (0.784); Antineoplastic (0.778)
279Leukopoiesis stimulant (0.802); Natural killer cell stimulant (0.769); Immunosuppressant (0.686)
280Lipid metabolism regulator (0.929); Hypolipemic (0.786); Lymphocytopoiesis inhibitor (0.726)
281Antidiabetic (0.763); Antihypoxic (0.736); Natural killer cell stimulant (0.570)
282Myasthenia Gravis treatment (0.962); Cell adhesion molecule inhibitor (0.863); Antidiabetic (0.757)
283Angiogenesis stimulant (0.915); Lipid metabolism regulator (0.913); Myasthenia Gravis treatment (0.779)
Cell adhesion molecule inhibitor (0.745); Antidiabetic (0.664)
284Angiogenesis stimulant (0.914); Myasthenia Gravis treatment (0.844)
Cell adhesion molecule inhibitor (0.788); Antidiabetic (0.700)
285Angiogenesis stimulant (0.915); Lipid metabolism regulator (0.913); Myasthenia Gravis treatment (0.779)
Cell adhesion molecule inhibitor (0.745); Antidiabetic (0.664)
286Antihypertensive (0.824); DNA intercalator (0.707); Antidiabetic (0.692)
287Leukopoiesis stimulant (0.763); Natural killer cell stimulant (0.721); Antidiabetic symptomatic (0.659)
288Myasthenia Gravis treatment (0.962); Cell adhesion molecule inhibitor (0.863); Antidiabetic (0.757)
* Only activities with Pa > 0.5 are shown.
Table 23. Predicted biological activity of FA of fungal lipopeptides.
Table 23. Predicted biological activity of FA of fungal lipopeptides.
No.Predicted Biological Activity, Pa *
289Antibiotic Glycopeptide-like (0.908); Antineoplastic (0.877); Apoptosis agonist (0.798)
290Antineoplastic (0.942); Apoptosis agonist (0.897); Antibiotic Glycopeptide-like (0.785)
291Antibiotic Glycopeptide-like (0.892); Antineoplastic (0.850); Antiprotozoal (Plasmodium) (0.834)
292Antineoplastic (0.890); Antibiotic Glycopeptide-like (0.883); Apoptosis agonist (0.802)
293Antineoplastic (0.942); Apoptosis agonist (0.897); Antibiotic Glycopeptide-like (0.785)
294Antineoplastic (0.913); Apoptosis agonist (0.882); Antibiotic Glycopeptide-like (0.771)
295Antineoplastic (0.938); Apoptosis agonist (0.881); Antibiotic Glycopeptide-like (0.769)
296Antibiotic Glycopeptide-like (0.813); Antineoplastic (0.800); Apoptosis agonist (0.684)
297Antibiotic Glycopeptide-like (0.908); Antineoplastic (0.877); Apoptosis agonist (0.798)
298Antibiotic Glycopeptide-like (0.879); Antineoplastic (0.857); Apoptosis agonist (0.803)
299Antineoplastic (0.942); Apoptosis agonist (0.897); Antibiotic Glycopeptide-like (0.785)
300Antibiotic Glycopeptide-like (0.906); Antineoplastic (0.865); Apoptosis agonist (0.763)
301Antineoplastic (0.913); Antibiotic Glycopeptide-like (0.782); Apoptosis agonist (0.779)
302Antidiabetic symptomatic (0.732); Leukopoiesis stimulant (0.671); Multiple sclerosis treatment (0.663)
303Natural killer cell stimulant (0.739); Antidiabetic symptomatic (0.733); Leukopoiesis stimulant (0.687)
304Antiviral (Arbovirus) (0.927); Lipid metabolism regulator (0.887); Antiviral (Picornavirus) (0.669)
305Lipid metabolism regulator (0.881); Hypolipemic (0.851); Anti-hypercholesterolemic (0.728)
306Lipid metabolism regulator (0.819); Natural killer cell stimulant (0.797); Hypolipemic (0.759)
307Lipid metabolism regulator (0.919); Anti-hypercholesterolemic (0.803); Hypolipemic (0.763)
308Anti-hypercholesterolemic (0.897); Lipid metabolism regulator (0.819); Antimutagenic (0.784)
309Anti-inflammatory (0.842); Analgesic (0.766); Leukopoiesis stimulant (0.584)
310Anti-inflammatory (0.814); Analgesic (0.746); Erythropoiesis stimulant (0.611)
* Only activities with Pa > 0.5 are shown.
Table 24. Predicted biological activity of FA derived from fungal peptides.
Table 24. Predicted biological activity of FA derived from fungal peptides.
No.Predicted Biological Activity, Pa *
311Acute neurologic disorders treatment (0.793); Natural killer cell stimulant (0.711)
312Anti-infective (0.945); Antitoxic (0.908); Natural killer cell stimulant (0.900)
313Lipid metabolism regulator (0.923); Apoptosis agonist (0.849); Antineoplastic (0.803)Acute neurologic disorders treatment (0.799); Preneoplastic conditions treatment (0.649)
314Lipid metabolism regulator (0.923); Apoptosis agonist (0.849); Antineoplastic (0.803)Acute neurologic disorders treatment (0.799); Preneoplastic conditions treatment (0.649)
315Lipid metabolism regulator (0.960); Hypolipemic (0.898); Anti-hypercholesterolemic (0.886)
316Lipid metabolism regulator (0.960); Hypolipemic (0.898); Anti-hypercholesterolemic (0.886)
317Lipid metabolism regulator (0.848); Anti-hypercholesterolemic (0.770); Hypolipemic (0.760)
318Lipid metabolism regulator (0.848); Anti-hypercholesterolemic (0.770); Hypolipemic (0.760)
319Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842); Hypolipemic (0.830)
320Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842); Hypolipemic (0.830)
321Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842); Hypolipemic (0.830)
322Lipid metabolism regulator (0.930); Anti-hypercholesterolemic (0.842); Hypolipemic (0.830)
323Lipid metabolism regulator (0.846); Anti-hypercholesterolemic (0.827)
324Lipid metabolism regulator (0.846); Anti-hypercholesterolemic (0.827)
325Anti-ischemic, cerebral (0.907); Cell adhesion molecule inhibitor (0.876); Antidiabetic (0.702)
326Antineoplastic (0.857); Apoptosis agonist (0.746); Preneoplastic conditions treatment (0.517)
327Anti-ischemic, cerebral (0.835); Acute neurologic disorders treatment (0.783); Hypolipemic (0.749)
328Anti-ischemic, cerebral (0.845); Leukopoiesis stimulant (0.783); Antitoxic (0.675)
329Hypolipemic (0.879); Lipid metabolism regulator (0.825); Anti-hypercholesterolemic (0.769)
330Acute neurologic disorders treatment (0.947); Antineoplastic (0.816); Apoptosis agonist (0.771)
331Preneoplastic conditions treatment (0.770); Acute neurologic disorders treatment (0.757)
332Lipid metabolism regulator (0.937); Acute neurologic disorders treatment (0.832)
333Lipid metabolism regulator (0.937); Hypolipemic (0.866)Acute neurologic disorders treatment (0.832); Atherosclerosis treatment (0.653)
* Only activities with Pa > 0.5 are shown.
Table 25. Predicted biological activity of FA from fungal peptides.
Table 25. Predicted biological activity of FA from fungal peptides.
No.Predicted Biological Activity, Pa *
334Natural killer cell stimulant (0.795); Leukopoiesis stimulant (0.784); Antineurotic (0.700)
335Hypolipemic (0.828); Acute neurologic disorders treatment (0.774); Leukopoiesis stimulant (0.736)
336Hypolipemic (0.880); Antineoplastic (0.858); Lipid metabolism regulator (0.732)
337Lipid metabolism regulator (0.859); Acute neurologic disorders treatment (0.811)
Hypolipemic (0.800); Mucositis treatment (0.756); Antidiabetic symptomatic (0.696)
338Lipid metabolism regulator (0.891); Hypolipemic (0.861); Anti-hypercholesterolemic (0.784)
339Lipid metabolism regulator (0.888); Acute neurologic disorders treatment (0.761)
340Hypolipemic (0.816); Lipid metabolism regulator (0.793); Mucositis treatment (0.779)
341Antifungal (0.836); Antibacterial (0.653); Antiparasitic (0.614)
342Antifungal (0.836); Antibacterial (0.653); Antiparasitic (0.614)
343Lipid metabolism regulator (0.952); Antineoplastic (liver cancer) (0.909)
Anti-hypercholesterolemic (0.815); Antineoplastic (0.777); Atherosclerosis treatment (0.649)
344Lipid metabolism regulator (0.923); Apoptosis agonist (0.849); Antineoplastic (0.803)
Hypolipemic (0.712); Atherosclerosis treatment (0.651); Preneoplastic conditions treatment (0.649)
345Lipid metabolism regulator (0.942); Antiviral (Arbovirus) (0.903); Hypolipemic (0.741)
Acute neurologic disorders treatment (0.728); Antiviral (Picornavirus) (0.676)
346Antiviral (Arbovirus) (0.952); Lipid metabolism regulator (0.903); Antiviral (Picornavirus) (0.790)
347Antiviral (Arbovirus) (0.952); Lipid metabolism regulator (0.903); Antiviral (Picornavirus) (0.790)
348Antiviral (Arbovirus) (0.952); Lipid metabolism regulator (0.903); Antiviral (Picornavirus) (0.790)
349Antiviral (Arbovirus) (0.952); Lipid metabolism regulator (0.903); Antiviral (Picornavirus) (0.790)
350Platelet antagonist (0.800); Anticoagulant (0.702); Fibrinolytic (0.700)
351Apoptosis agonist (0.879); Antineoplastic (0.878); Preneoplastic conditions treatment (0.618)
352Antiviral (Arbovirus) (0.861); Lipid metabolism regulator (0.844); Antiviral (Picornavirus) (0.776)
353Antiviral (Arbovirus) (0.917); Antiviral (Picornavirus) (0.781); Antimutagenic (0.699)
354Antiviral (Arbovirus) (0.861); Lipid metabolism regulator (0.844); Antiviral (Picornavirus) (0.776)
355Antiviral (Arbovirus) (0.917); Antiviral (Picornavirus) (0.781); Antimutagenic (0.699)
* Only activities with Pa > 0.5 are shown.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dembitsky, V.M. Hydrobiological Aspects of Fatty Acids: Unique, Rare, and Unusual Fatty Acids Incorporated into Linear and Cyclic Lipopeptides and Their Biological Activity. Hydrobiology 2022, 1, 331-432. https://doi.org/10.3390/hydrobiology1030024

AMA Style

Dembitsky VM. Hydrobiological Aspects of Fatty Acids: Unique, Rare, and Unusual Fatty Acids Incorporated into Linear and Cyclic Lipopeptides and Their Biological Activity. Hydrobiology. 2022; 1(3):331-432. https://doi.org/10.3390/hydrobiology1030024

Chicago/Turabian Style

Dembitsky, Valery M. 2022. "Hydrobiological Aspects of Fatty Acids: Unique, Rare, and Unusual Fatty Acids Incorporated into Linear and Cyclic Lipopeptides and Their Biological Activity" Hydrobiology 1, no. 3: 331-432. https://doi.org/10.3390/hydrobiology1030024

Article Metrics

Back to TopTop