Next Article in Journal
Utilization of Fishery Processing By-Product Squid Pens for α-Glucosidase Inhibitors Production by Paenibacillus sp.
Next Article in Special Issue
Characterization and Trypanocidal Activity of a Novel Pyranaphthoquinone
Previous Article in Journal
Current Status and Future Prospects of Marine Natural Products (MNPs) as Antimicrobials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Marine Pharmacology in 2012–2013: Marine Compounds with Antibacterial, Antidiabetic, Antifungal, Anti-Inflammatory, Antiprotozoal, Antituberculosis, and Antiviral Activities; Affecting the Immune and Nervous Systems, and Other Miscellaneous Mechanisms of Action †

by
Alejandro M. S. Mayer
1,*,
Abimael D. Rodríguez
2,
Orazio Taglialatela-Scafati
3 and
Nobuhiro Fusetani
4
1
Department of Pharmacology, Chicago College of Osteopathic Medicine, Midwestern University, 555 31st Street, Downers Grove, IL 60515, USA
2
Molecular Sciences Research Center, University of Puerto Rico, 1390 Ponce de León Avenue, San Juan, PR 00926, USA
3
Department of Pharmacy, University of Naples “Federico II”, Via D. Montesano 49, 80131 Napoli, Italy
4
Fisheries and Oceans Hakodate, Hakodate 041-8611, Japan
*
Author to whom correspondence should be addressed.
This review is dedicated to the memory of the late Professor Ernesto Fattorusso on the occasion of what would have been his 80th birthday, and the late Professor Robert S. Jacobs on the occasion of what would have been his 84th birthday.
Mar. Drugs 2017, 15(9), 273; https://doi.org/10.3390/md15090273
Submission received: 20 July 2017 / Revised: 17 August 2017 / Accepted: 21 August 2017 / Published: 29 August 2017

Abstract

:
The peer-reviewed marine pharmacology literature from 2012 to 2013 was systematically reviewed, consistent with the 1998–2011 reviews of this series. Marine pharmacology research from 2012 to 2013, conducted by scientists from 42 countries in addition to the United States, reported findings on the preclinical pharmacology of 257 marine compounds. The preclinical pharmacology of compounds isolated from marine organisms revealed antibacterial, antifungal, antiprotozoal, antituberculosis, antiviral and anthelmitic pharmacological activities for 113 marine natural products. In addition, 75 marine compounds were reported to have antidiabetic and anti-inflammatory activities and affect the immune and nervous system. Finally, 69 marine compounds were shown to display miscellaneous mechanisms of action which could contribute to novel pharmacological classes. Thus, in 2012–2013, the preclinical marine natural product pharmacology pipeline provided novel pharmacology and lead compounds to the clinical marine pharmaceutical pipeline, and contributed significantly to potentially novel therapeutic approaches to several global disease categories.

1. Introduction

The aim of the present review is to consolidate preclinical marine pharmacology in 2012–2013, with a format similar to the previous 8 reviews of this series, which cover the period 1998–2011 [1,2,3,4,5,6,7,8]. The peer-reviewed articles were retrieved from searches of several databases, including MarinLit, PubMed, Chemical Abstracts®, ISI Web of Knowledge and Google Scholar. The review only includes bioactivity and/or pharmacology of structurally characterized marine chemicals, which we have classified using a modification of Schmitz’s chemical classification [9] into six major chemical classes; namely, polyketides, terpenes, peptides, alkaloids, shikimates, and sugars. The preclinical antibacterial, antifungal, antiprotozoal, antituberculosis, antiviral and anthelmintic pharmacology of marine chemicals is reported in Table 1, with the structures shown in Figure 1. Marine compounds that affect the immune and nervous systems, as well as those with antidiabetic and anti-inflammatory effects, are exhibited in Table 2, with their structures presented in Figure 2. Finally, marine compounds that affected a variety of cellular and molecular targets are noted in Table 3, and their structures presented in Figure 3.
A number of publications during 2012–2013 reported extracts or structurally uncharacterized marine compounds, with novel and interesting preclinical and/or clinical pharmacology: in vitro antimalarial activity in crude extracts from Fiji marine organisms using a semi-automated RNA fluorescence-based high-content live cell-imaging assay [10]; the first report of in vitro liver stage antiplasmodial activity and dual stage inhibitory potential of British seaweeds [11]; anti-hepatitis C virus activity affecting the viral helicase NS3 and replication, in crude extracts from the marine feather star Alloeocomatella polycladia [12]; anti-herpes simplex virus HSV-1 and HSV-2 activity in a purified sulfoglycolipid fraction from the Brazilian marine alga Osmundaria obtusiloba [13]; in vivo anti-inflammatory activity of a heterofucan from the Brazilian seaweed Dictyota menstrualis that inhibited leukocyte migration to sites of tissue injury by binding to the cell membrane [14]; in vivo antinociceptive and anti-inflammatory activity in a crude methanolic extract of the red alga Bryothamnion triquetrum [15]; in vivo anti-inflammatory activity in a sulfate polysaccharide fraction from the red alga Gracilaria caudata resulting in significant inhibition of neutrophil migration and cytokine release [16]; in vitro anti-inflammatory effect of a hexane-soluble fraction of the brown alga Laminaria japonica that inhibited nitric oxide, prostaglandin E2, interleukin (IL)-1β and IL-6 release from lipopolysaccharide-stimulated macrophages via inactivation of nuclear factor-κB transcription factor [17]; in vivo anti-inflammatory of a polysaccharide-rich fraction from the marine red alga Lithothamnion muelleri that reduced organ injury and lethality, as well as pro-inflammatory cytokines and chemokines, associated with graft-versus-host disease in mice [18]; in vivo clinical effectiveness in an osteoarthritis trial by PCSO-524TM, a nonpolar lipid extract from the New Zealand marine green lipped mussel Perna canaliculus, which may offer “potential alternative complementary therapy with no side effects for osteoarthritis patients” [19]; enhanced antioxidant activity of chitosan nanoparticles as compared to chitosan on hydrogen peroxide-induced stress injury in mouse macrophages in vitro [20]; induction of concentration-dependent vasoconstrictive activity on isolated rat aorta by a tentacle extract from the jellyfish Cyanea capillata [21]; significant antioxidant effect of a sulfated-polysaccharide fraction of the marine red alga Gracilaria birdiae which prevented naproxen-induced gastrointestinal damage in rats by reversing glutathione depletion [22]; in vitro antioxidant properties of a polysaccharide from the brown seaweed Sargassum graminifolium (Turn.) that was also observed to inhibit calcium oxalate crystallization, a constituent of urinary kidney stones [23]; antioxidant activity in organic extracts from 30 species of Hawaiian marine algae, with the carotenoid fucoxanthin identified as the major bioactive antioxidant compound in the brown alga T. ornata [24]; screening of antioxidant activity in 18 cyanobacteria and 23 microalgae cell extracts identified Scenedesmus obliquus strain M2-1, which protected against DNA oxidative damage induced by copper (II)-ascorbic acid [25]; anxiolytic-like effect of a salmon phospholipopeptidic complex composed of polyunsaturated fatty acids and bioactive peptides associated with strong free radical scavenging properties [26]; antinociceptive activity in extracts of the skin of the Brazilian planehead filefish Stephanolepis hispidus with partial activation of opioid receptors in the nervous system [27]; strong in vitro acetylcholinesterase inhibition, an enzyme targeted by drugs used to treat Alzheimer’s disease, myasthenia gravis and glaucoma, by an extract from the polar marine sponge Latrunculia sp. [28]; central nervous system activity of a phlorotannin-rich extract from the edible brown seaweed Ecklonia cava targeting gamma-aminobutyric acid type A benzodiazepine receptors [29]; and novel protease inhibitors from Norwegian spring spawning herring determined by screening of marine extracts with assays combining fluorescence resonance energy transfer activity and surface plasmon resonance spectroscopy-based binding [30].

2. Marine Compounds with Antibacterial, Antifungal, Antiprotozoal, Antituberculosis, Antiviral and Anthelmintic Activities

Table 1 presents 2012–2013 preclinical pharmacological research on the antibacterial, antifungal, antiprotozoal, antituberculosis, antiviral and anthelmintic activities of marine natural products (1113) shown in Figure 1.

2.1. Antibacterial Activity

During 2012–2013, 31 studies reported antibacterial marine natural products (150) isolated from bacteria, fungi, tunicates, sponges, and algae, a global effort that may contribute to the search for novel leads for developing newer drugs to treat drug-resistant bacterial infections.
As shown in Table 1 and Figure 1, three papers reported molecular mechanism of action studies with marine antibacterial compounds. Jang and colleagues reported a potent antianthrax antibiotic, anthracimycin (1), derived from a marine actinomycete with significant activity against Bacillus anthracis, by a mechanism that “…remains to be fully defined…” but that appears to involve DNA/RNA synthesis inhibition [31]. Keffer and colleagues extended the mechanism of action of bis-diarylbutene macrocycle chrysophaentins (2,3), isolated from the chrysophyte alga Chrysophaeum taylori, by determining that they competitively inhibited the biochemical activity of the Gram-positive and Gram-negative cell division protein FtsZ by binding to its GTP-binding site [32]. Sakoulas and colleagues reported the antibacterial activity of merochlorin A (4), a meroterpenoid isolated from a marine-derived actinomycete strain CNH189, which demonstrated activity against Gram-positive bacteria including Clostridium difficile, but not against Gram-negative bacteria, by a mechanism that appeared to involve “…global inhibition of DNA, RNA, protein, and cell wall synthesis…” [33].
As shown in Table 1 and Figure 1, 46 marine chemicals (550), some of them novel, were reported to exhibit antibacterial activity with MICs < 10 μg/mL or 10 μM against several bacterial strains, although the mechanism of action for these compounds remained undetermined: a novel aflatoxin B2b (5), isolated from the fungus Aspergillus flavus; 092008, isolated from the root of the mangrove H. tiliaceus from Hainan, China [34]; a new alkaloid ageloxime B (6), isolated from the South China Sea marine sponge Agelas mauritiana [35]; several known yet bioactive compounds namely altersolanol C (7), macrosporin (8) and alterporriol C (9) isolated from a soft-coral derived from South China Sea fungus Alternaria sp. [36]; a novel antimycin A analogue, antimycin B2 (10), derived from the actinomycete Streptomyces lusitanus, isolated from the mangrove Avicennia mariana in Fujian, China [37]; a new bisabolane-type sesquiterpenoid (−)-sydonol (11) from a South China Sea sponge-derived fungus Aspergillus sp. [38]; three new pyrimidine diterpenes designated axistatins 1 (12), 2 (13) and 3 (14), isolated from the marine sponge Agelas axifera collected in the Republic of Palau [39]; two new diterpene-benzoate compounds bromophycoic acid A (15) and E (16) from a Fijian red alga Callophycus sp. [40]; new butenolide cadiolides C–F (1720) from a Korean tunicate Pseudodistoma antinboja [41]; novel tris-aromatic furanones cadiolides G-I (2123) from the Korean dark red ascidian Synoicum sp. [42]; xanthones citreamicins θ A and B (24,25), isolated from the Red Sea marine Streptomyces caelestis [43]; two new aromatic polyketides, communols A and F (26,27), isolated from the marine Penicillium commune 518, associated with the gorgonian Muricella abnormalis [44]; two dolabellane diterpenes (28,29), isolated from the Greek brown alga Dilophus spiralis [45]; a novel enhygrolide A (30), isolated from the obligate marine myxobacterium Enhygromyxa salina from a mud sample from Prerow, Germany [46]; a new β-carboline alkaloid eudistomin Y11 (31), isolated from a purple-colored ascidian Synoicum sp. [47]; a new capoamycin-type antibiotic fradimycin B (32), isolated from the marine Streptomyces fradiae strain PTZ0025 [48]; three novel cyclic bis-1,3 dialkylpyridiniums (3335) from a Korean sponge Halyclona sp. [49]; a novel bisindole alkaloid hyrtimomine D (36), isolated from an Okinawan marine sponge Hyrtios sp. [50]; a new bromotyrosine-derived metabolite, ianthellisformisamine A (37), reported from the Australian marine sponge Suberea ianthelliformis [51]; a new thiazolyl peptide kocurin (38) from the marine-derived bacterium Kocuria palustris [52]; the known alkaloid lamellarin O (39), isolated from a southern Australian sponge Ianthella sp. [53]; three new halogenated sesquiterpenes (4042), isolated from the Chinese marine red alga Laurencia okamurai [54]; a new spirotetronate antibiotic, lobophorin H (43) from a South China Sea-Streptomyces sp. 12A35 [55]; a new cyclopeptide marthiapeptide A (44), isolated from the South China Sea-derived bacterium Marinactinospora thermotolerans [56]; two known napyradiomycin A1 (45) and napyradiomycin B3 (46) from a Chinese marine-derived Streptomyces sp. strain SCSIO [57,58]; two new hydroanthraquinone analogues 4a-epi-9α-methoxydihydrodeoxybostrycin (47) and 10-deoxy-bostrycin (48), isolated from a South China Sea marine-derived fungus Nigrospora sp., isolated from an unidentified sea anemone [59]; a novel cyclic peptide ohmyungsamycin A (49) from a Korean Streptomyces sp. strain SNJ042 [60]; and a novel benzofuran penicifuran A (50), obtained from a South China Sea sponge-derived fungus Penicillium sp. strain MWZ14-4 [61].
Furthermore, during 2012–2013, several other marine natural products, some of them novel, reported MICs or IC50s ranging from 10 to 50 μg/mL, or 10–50 μM, respectively, and thus, because of their lower antibacterial potency, were excluded from Table 1 and Figure 1: guaiazulene-derived terpenoids from a Chinese gorgonian Anthogorgia sp. (MIC = 12.7–18 μg/mL) [110]; novel fulvynes antimicrobial polyoxygenated acetylenes from the Mediterranean sponge Haliclona fulva (IC50 = 12–60 μM) [111]; bioactive polyhydroxylated halicrasterols (MIC = 4–32 μg/mL) from the Chinese marine sponge Haliclona crassiloba [112]; hunanamycin A, an antibiotic (MIC = 12.4 μM), isolated from the Bahamanian marine-derived Bacillus hunanensis [113]; three new dimeric bromopyrrole alkaloids, nagelamides X–Z (MIC = 8–32 μg/mL) from an Okinawan marine sponge Agelas sp. [69]; a new anthraquinone-citrin derivative (MIC = 16 μg/mL), isolated from the sea fan-derived fungus Penicillium citrinum PSU-F51 [114]; and a new chlorinated benzophenone derivative, (±)-pestalachloride C (MIC = 5–20 μM) from a South China Sea soft coral-derived fungus Pestalotiopsis sp. [115]. Finally, during 2012–2013, the novel marine lipopeptides, peptidolipins B–F (MIC = 64 μg/mL), were isolated from an ascidian-derived Gram positive Nocardia sp. bacterium [116].

2.2. Antifungal Activity

Eleven studies during 2012–2013 reported on the antifungal activity of several novel marine natural products (6,36,5160), isolated from marine fungi, sponges, sea cucumbers and algae, a slight decrease from our last review [7], and previous reviews of this series.
As shown in Table 1 and Figure 1, two reports described antifungal marine chemicals with novel mechanisms of action. Rubiolo and colleagues investigated the guanidine antifungal alkaloid crambescidin-816 (51), previously isolated from the Mediterranean sponge Crambe crambe [62]. Detailed cell cycle studies in the yeast Saccharomyces cerevisiae demonstrated that this compound induced G2/M cell cycle arrest followed by apoptosis and mitochondrial disfunction, suggesting that although cytotoxic crambescidin-816 “….could serve as a lead compound to fight fungal infections”. Yibmantasiri and colleagues investigated the molecular basis for the fungicidal action of the triterpene glycoside neothyonidioside (52) isolated from the sea cucumber Australostichopus mollis [63], demonstrating that neothyonidioside binds directly to fungal ergosterol affecting membrane curvature and fusion capability essential for membrane recycling and lysosomal degradation.
Furthermore, as shown in Table 1 and Figure 1, several marine natural products showed significant antifungal activity (i.e., MICs that were either less than 10 μg/mL, 10 μM, or 10 μg/disk), although no mechanism of action studies were reported in the published articles: a novel alkaloid ageloxime B (6), isolated from the South China Sea sponge Agelas mauritiana [35]; a novel tetramic acid glycoside, aurantoside K (53), isolated from a Fijian marine sponge Melophlus sp. [64]; a new prenylated para-xylene caulerprenylol A (54), isolated from the green alga Caulerpa racemosa collected in the Zhanjiang coastline, China [65]; a new 4-hydroxy-2-pyridone alkaloid didymellamide A (55), isolated from the Japanese marine-derived fungus S. cucurbitacearum [66]; a new polyketide hippolachnin A (56), reported from the South China Sea sponge Hippospongia lachne [67]; novel triterpene glycosides holotoxin F and G (57,58), isolated from the sea cucumber Apostichopus japonicus Selenka, “a traditional tonic with high economic value” in China [68]; a novel bisindole alkaloid hyrtimomine D and E (36,59), isolated from an Okinawan marine sponge Hyrtios sp. [50]; a novel dimeric alkaloid nagelamide Z (60), isolated from a Japanese sponge Agelas sp. [69]; and a new linear polyketide woodylide A (61), isolated from the South China Sea sponge Plakortis simplex [70]. Ongoing mechanism of action studies with these potent marine compounds will be required to characterize their molecular pharmacology.
Finally, several novel structurally-characterized marine molecules demonstrated MICs or IC50s greater than 10 μg/mL, 10 μM, or 10 μg/disk, and therefore, because of the reported weaker antifungal activity, were excluded from Table 1 and Figure 1: three triterpene glycosides, cucumariosides A1, A6 and A15 (MIC = 20 μg/mL), isolated from the Pacific Sea cucumber Eupentacta fraudatrix [117]; a tetranorditerpenoid derivative isolated from Aspergillus wentii EN-48 (MIC = 16 μg/mL), a fungus isolated from an unidentified marine brown algae [118]; and bromophenol-aconitic acid adduct, symphyocladin G, isolated from the marine red alga Symphyocladia latiuscula (MIC = 10 μg/mL) [119]. These novel marine compounds may contribute to ongoing research for clinically useful antifungal agents.

2.3. Antiprotozoal and Antituberculosis Activity

As shown in Table 1, during 2012–2013 twenty five studies contributed to novel findings on antiprotozoal (antimalarial, antileishmanial and antitrypanosomal) and antituberculosis pharmacology of structurally characterized marine natural products (6292), a decrease from previous 1998–2011 reviews [1,2,3,4,5,6,7,8].
Malaria, which is caused by protozoa of the genus Plasmodium (P. falciparum, P. ovale, P. vivax and P. malariae), affects millions of people worldwide. Contributing to the global search for novel antimalarial drugs, and as presented in Table 1, seventeen novel marine molecules (6278), isolated from bacteria, ascidians, fungi, sponges, and tunicates, were shown during 2012–2013 to possess antimalarial activity, although mechanism of action studies were not reported for these compounds.
As shown in Table 1 and Figure 1, potent (IC50 < 2 μM) to moderate (IC50 > 2–10 μM) antimalarial activity was reported for several marine natural products (6278), isolated from ascidians, sponges, bacteria and fungi. Mani and colleagues reported antiplasmodial activity in the bromotyrosine derivative araplysillin I (62) from the South Pacific Solomon Islands sponge Suberea ianthelliformis [71]. Lam and colleagues extended the pharmacology of the New Zealand ascidian dioxothiazino-quinoline-quinone metabolite ascidiathiazone A (63) by demonstrating it to be a moderate growth inhibitor of chloroquine and a pyrimethamine resistant P. falciparum K1 strain, and noting that changing the quinolone-based structure to incorporate benzofuran or benzothiophene moieties yielded particularly potent antimalarials [72]. Farokhi and colleagues characterized new glycosphingolipids axidjiferoside A–C (6466) from the Senegal marine sponge Axinyssa djiferi with potent antimalarial activity against chloroquine-resistant FcB1/Colombia P. falciparum strain [73]. Beau and colleagues reported that epigenetic tailoring of the marine fungus Leucostoma persoonii enhanced production of the known polyketide cytosporone E (67), which inhibited P. falciparum with significant selectivity [74]. Calcul and colleagues reported a massive screening of Chinese mangrove endophytic fungi and discovered several new compounds, including a novel dimeric tetrahydroxanthone polyketide dicerandrol D (68), which was potent against “a robust and validated” drug-sensitive P. falciparum strain 3D7 [75]. Ilias and colleagues reported a novel pentacyclic ingamine alkaloid dihydroingenamine D (69), isolated from a sponge Petrosid Ng5 sp.5, which showed strong antiplasmodial activity against P. falciparum D6 and W2 strains [76]. Mudianta and colleagues reported that the novel alkaloid 19-hydroxypsammaplysin E (70) from the Indonesian marine sponge Aplysinella strongylata had notable antimalarial activity against the P. falciparum chloroquine-sensitive 3D7 strain [77]. Sirirak and colleagues reported a new trisoxazole macrolide kabiramide L (71) from the Thai marine sponge Pachatrissa nux that had moderate activity against a P. falciparum K1 multidrug-resistant strain [78]. Bharate and colleagues extended the pharmacology of the known meridianin C and G alkaloids (72,73), originally isolated from the marine tunicate Aplidium meridianum, by reporting that they inhibited both chloroquine-resistant D6 and sensitive W2 clones of P. falciparum [79]. Liew and colleagues identified orthidine F (74), a metabolite from the New Zealand ascidian Aplidium orthium of low toxicity and a moderate growth inhibitor of P. falciparum K1 strain dual drug-resistant strain [80]. Lin and colleagues isolated a new polyketide endoperoxide plakortide U (75) from the Fijian sponge Plakinastrella mamillaris with potent antimalarial activity against chloroquine-resistant P. falciparum FcM29 strain [81]. Davis and colleagues isolated several novel thiazine alkaloids from the Australian marine sponge Plakortis lita, one of which thiaplakortone A (76), showed potent activity against the human malaria parasite Plasmodium falciparum strains 3D7 and Dd2 with low cytotoxicity [82]. Davis and colleagues reported a novel bispyrroloiminoquinone alkaloid tsitikammamine C (77) from an Australian sponge Zyzzya sp. that displayed potent activity against P. falciparum chloroquine-sensitive 3D7 and -resistant dd2 strains [83]. Supong and colleagues reported a novel C-glycosylated benz[a]anthraquinone derivative, urdamycinone E (78) isolated from a marine Streptomyces sp. BCC45596 that potently inhibited P. falciparum K1 strain [84].
As shown in Table 1 and Figure 1, nine marine compounds (7986) isolated from bacteria, ascidians, sponges, soft corals and algae were reported to possess bioactivity towards so-called neglected protozoal diseases, namely leishmaniasis, caused by the genus Leishmania (L.), amebiasis, trichomoniasis, and both African sleeping sickness (caused by Trypanosoma (T.) brucei rhodesiense and T. brucei gambiense) and American sleeping sickness or Chagas disease (caused by T. cruzi).
As shown in Table 1, three reports described four antitrypanosomal marine chemicals (7982) as well as their mechanisms of action. Sanchez and colleagues examined the mode of action of almiramides (79,80), originally isolated from the cyanobacterium Lyngbya majuscula, and demonstrated for the first time that these compounds inhibited T. brucei by disrupting the parasite’s glycosomal function by targeting two membrane proteins, and were thus considered “encouraging candidates for further lead development” [85]. Abdelmohsen and colleagues reported that the dibenzodiazepine alkaloid diazepinomicin (81) isolated from a strain of Micromonospora sp. RV115 associated with the Croatian marine sponge Aplysina aerophoba showed activity against T. brucei trypmastigote forms and inhibited the parasite protease rhodesain [86]. Desoti and colleagues extended the pharmacology of (−)-elatol (82), a sesquiterpene isolated from the Brazilian red alga Laurencia dendroidea shown to affect trypomastigotes of T. cruzi, demonstrating that it induced initial depolarization of the parasite’s mitochondrial membrane, followed by an increase in superoxide generation, as well as loss of cell membrane and DNA integrity [87].
As shown in Table 1 and Figure 1, five marine natural products (63,8386) were characterized to exhibit antileishmanial and antiprotozoal activity, although the mechanism of action remained undetermined. Lam and colleagues reported that the known dioxothiazino-quinoline-quinone metabolite ascidiathiazone A (63), isolated from a New Zealand ascidian, moderately inhibited the growth of T. brucei rhodesiense, but was ineffective against T. cruzi and L. donovani [72]. Balunas and colleagues isolated the polyketide coibacin A (83) from a Panamanian marine cyanobacterium Oscillatoria sp., and observed potent activity against L. donovani axenic amastigotes [88]. Ishigami and colleagues isolated a new xenicane diterpenoid cristaxenicin A (84) from the deep-sea gorgonian Acanthoprimnoa cristata, which showed potent activity against L. amazonensis and T. congolense [89]. Chianese and colleagues completed structure-activity relationship studies with several natural and semisynthetic manadoperoxide B analogues (85,86), isolated from the Indonesian sponge Plakortis sfr. lita, and determined that both were highly active towards the parasite T. brucei rhodesiense, highlighting the 1,2-dioxane ring to be a key pharmacophore [90].
Because of the surge in drug-resistant strains of the intracellular pathogen Mycobacterium tuberculosis (Mtb), there is a global need for the development of novel drugs with novel mechanisms of action. As shown in Table 1 and Figure 1, seven novel marine natural products (78,8792), isolated from bacteria, sponges and fungi, contributed to the ongoing global search for novel antituberculosis agents. Although these marine natural products were characterized to exhibit antituberculosis activity, unfortunately the mechanism of action of these compounds remained undetermined.
Huang and colleagues reported a novel sesterterpenoid asperterpenoid A (87) from a mangrove endophytic fungus Aspergillus sp. that demonstrated strong inhibitory activity against M. tuberculosis protein tyrosine phosphatase B, an enzyme that is “…considered a promissory target for pulmonary tuberculosis cure” [91]. Song and colleagues isolated a new dimeric diketopiperazine, brevianamide S (88), from Aspergillus versicolor collected in the Bohai Sea, China, which demonstrated selective antibacterial activity against Bacille Calmette-Guérin (BCG), “suggestive of a new mechanism of action that could inform the development of next generation antitubercular drugs…if translated to M. tuberculosis…” [92]. Chen and colleagues reported a new spirotetronate, lobophorin G (89), from a marine-derived Streptomyces sp. MS100061 which exhibited strong anti-M. bovis BCG activity, providing relevant pharmacological information as this screen is thought to “serve as a useful screening surrogate for M. tuberculosis” [93]. Yamano and colleagues discovered a new cyclic depsipeptide neamphamide B (90) in a Japanese marine sponge Neamphius sp., which showed activity against M. bovis BCG in “both actively growing and dormant states” [94]. Avilés and colleagues isolated two new tricyclic diterpenes (91,92) from the Bahamian marine sponge Svenzea flava that displayed moderate antimycobacterial activity against M. tuberculosis H37Rv, the data suggesting that “the isoneoamphilectane backbone” may be “responsible for the observed activity” [95]. In addition to the antimalarial activity described earlier, Supong and colleagues reported that the novel C-glycosylated benz[a]anthraquinone derivative, urdamycinone E (78), inhibited M. tuberculosis strain H37Rv [84].

2.4. Antiviral Activity

As shown in Table 1 and Figure 1, thirteen reports were published during 2012–2013 on the antiviral pharmacology of marine natural products (93102) against hepatitis C, human immunodeficiency virus type-1 (HIV-1), influenza virus, human rhinovirus (HRV) and herpes simplex virus (HSV).
As shown in Table 1, only six reports described antiviral marine chemicals and their mechanisms of action. Da Rosa Guimarães and colleagues extended the pharmacology of the known steroids halistanol sulfate (93) and halistanol sulfate C (94), isolated from the Brazilian marine sponge Petromica citrina, by demonstrating that the compounds inhibited attachment and penetration of the “early events of HSV-1 infection” [96]. Ellithey and colleagues investigated several known metabolites (9597) from the Red Sea soft coral Litophyton arboreum and demonstrated selective inhibition of the HIV-1 protease by a mechanism that “confirms the contribution of the hydrophobicity of inhibitors of HIV protease” [97]. Salam and colleagues reported a novel pharmacological activity for the sesterterpene manoalide (98), which was observed to affect the hepatitis C virus NS3 helicase by inhibiting RNA binding and ATPase activity [98]. Park and colleagues reported that two polybromocatechol compounds (99,100), isolated from the red alga Neorhodomela aculeate, inhibited infection and cytopathic effects on a HeLa cell line by HRV2 and HRV3, causal agents of viral respiratory infections and common colds [99]. Ma and colleagues determined that the novel phenylspirodrimane stachybotrin D (101), isolated from the fungus Stachybotrys chartarum MXH-X73 derived from the Chinese marine sponge Xestospongia testudinaria, inhibited HIV-1 replication of wild-type and five non-nucleoside reverse transcriptase inhibitor (NNRTI)-resistant HIV-1 strains by inhibiting the reverse transcriptase, and thus “provides a new class of chemotype for the search of NNRT inhibitors” [100]. Jiao and colleagues reported that streptoseolactone (102), derived from the actinomycete Streptomyces seoulensis strain isolated from the shrimp Penasus orientalis, inhibited neuraminidase by a noncompetitive mechanism, a finding “of value in terms of drug discovery for the treatment of influenza” [101].
As shown in Table 1 and Figure 1, several marine natural products (103111) were characterized to exhibit antiviral activity, although the mechanism of action of these compounds remained undetermined. He and colleagues isolated a novel cyclic tetrapeptide asperterrestide A (103) from the marine-derived fungus Aspergillus terreus SCSGAF0162, which inhibited influenza virus strains H1N1 and H3N2 [102].
Two contributions by Peng and colleagues reported two novel indole alkaloids (104,105), produced by the mangrove-derived fungus Cladosporium sp. PJX-41, that inhibited influenza A virus H1N1 [103], and a new pyronepolyene C-glucoside iso-D8646-2-6 (108), from a sponge-associated fungus Epicoccum sp. JJY40, that also inhibited the influenza virus H1N1 [106]. Hawas and colleagues isolated the novel isorhodoptilometrin-1-methyl ether (106) from the Red Sea marine fungus Aspergillus versicolor, which exhibited hepatitis virus C NS3/4A protease activity [104]. Zhang and colleagues isolated a novel polyketide massarilactone H (107) from the marine-derived fungus Phoma herbarum which displayed moderate neuraminidase inhibitory activity [105]. Ahmed and colleagues purified a novel polyhydroxylated sterol (109) and a new ceramide (110) from the Red Sea soft coral Sinularia candidula, which inhibited the H5N1 avian influenza viral strain [107]. Plouguerné and colleagues characterized the antiviral activity of a sulfoquinovosyldiacylglycerol (111) from the Brazilian brown seaweed Sargassum vulgare, demonstrating that it inhibited both HSV-1 and HSV-2 more potently than acyclovir, a clinically used antiherpetic agent [108].

2.5. Anthelmintic Activity

As shown in Table 1, only one report was published during 2012–2013 on the anthelmintic pharmacology of marine natural products. Melek and colleagues isolated triterpene glycosides echinosides A and B (112,113) from the sea cucumbers Actinopyga echinites and Holothuria polii that displayed “potential in vitro schisotomicidal activity against worms of Schistosoma mansoni”, suggesting that these compounds may be “promising lead compounds for the development of new schistosomicidal agents” [109].

3. Marine Compounds with Antidiabetic and Anti-Inflammatory Activity, and Affecting the Immune and Nervous System

Table 2 presents the 2012–2013 preclinical pharmacology of marine chemicals (114188), which demonstrated either antidiabetic or anti-inflammatory activity, as well as those affecting the immune or nervous system; their structures are depicted in Figure 2.

3.1. Antidiabetic Activity

Lee and colleagues reported the pharmacology of octaphlorethol A (114), a novel phenolic compound isolated from the marine brown alga Ishige foliacea, by showing that octaphlorethol A enhanced glucose uptake in L6 rat myoblast cells by increasing glucose transporter 4 translocation to the plasma membrane and protein kinase B and AMP-activated protein kinase activity [120].

3.2. Anti-Inflammatory Activity

As shown in Table 2 and Figure 2, there was a remarkable increase in marine anti-inflammatory pharmacology research during 2012–2013. The molecular mechanism of action of marine natural products (115134) was investigated in both in vitro and in vivo preclinical pharmacological studies which were completed using a variety of in vitro models including bone marrow-derived macrophages, human U937 monocytic cells, murine RAW 264.7 macrophages, human epidermoid carcinoma A431 cell line, human polymorphonuclear leukocytes, rat brain microglia, and mouse peritoneal macrophages.
Chae and colleagues evaluated the anti-inflammatory properties of apo-9′-fucoxanthinone (115), isolated from the marine edible brown alga Sargassum muticum [121] in unmethylated CpG DNA-stimulated bone marrow-derived macrophages and dendritic cells. Inhibition of interleukin-12 p40, interleukin-6 (IL-6) and tumor necrosis factor-α (TNF-α) production, as well as concomitant attenuation of the mitogen-activated protein kinase pathways, was observed, leading the authors to conclude that apo-9′-fucoxanthinone may have “potential therapeutic use…for inflammatory disease”. In a detailed mechanistic study, Speranza and colleagues investigated the antioxidant marine carotenoid astaxanthin (116), showing that it inhibited hydrogen peroxide-stimulated production of pro-inflammatory cytokines IL-1, IL-6 and TNF-α in a human U937 monocytic cell line by selectively restoring physiological levels and function of the tyrosine phosphatase SHP-1, thus proposing that astaxanthin might become a novel agent for the therapy of inflammatory diseases [122]. Johnson and colleagues identified the alkaloids bengamide A and B (117,118) as potent inhibitors of NFκB and LPS-induced expression of cytokines IL-6, TNF-α and chemokine monocyte chemoattractant protein-1 (MCP-1) release from murine RAW 264.7 macrophages, concluding that these compounds may “serve as therapeutic leads for immune disorders involving inflammation” [123]. Song and colleagues determined that bis-N-norgliovictin (119) derived from a marine fungus S3-1-c inhibited TNF-α, IL-6, interferon-β, and MCP-1 production by LPS-stimulated RAW 264.7 macrophages and affecting Toll-like receptor 4 (TLR-4) signal transduction pathways, as well as LPS-induced septic shock in mice, thus suggesting bis-N-norgliovictin might result in a useful therapeutic candidate for “sepsis and other inflammatory diseases” [124]. Investigations by Yang and colleagues with phlorotannin 6,6’-bieckol (120), isolated from the marine brown alga Ecklonia cava, showed that the compound inhibited expression and release of nitric oxide, prostaglandin E2, TNF-α and IL-6 in LPS-stimulated macrophages, with concomitant inhibition of NFκB activation, suggesting that compound 120 is potentially useful for the treatment of inflammatory diseases [125]. Balunas and colleagues determined that the polyketide coibacin B (121), isolated from the Panamanian marine cyanobacterium, cf. Oscillatoria sp. possessed not only antileishmanial activity, but also significant anti-inflammatory activity, as it significantly decreased LPS-induced nitric oxide, TNF-α and IL-6 release from RAW 264.7 macrophages [88]. Hsu and colleages reported that the soft coral S. flexibilis-derived 11-epi-sinulariolide acetate (122) inhibited cyclooxygenase-2 and interleukin-8 expression in human epidermoid carcinoma A431 cells in vitro by inhibition of Ca2+ signaling, suggesting that it might become a lead compound to target “store-operated calcium signaling-dependent inflammatory diseases” [126]. Choi and colleagues demonstrated that the novel honaucin A (123) from the Hawaiian cyanobacterium Leptolyngbya crossbyana, which inhibited LPS-induced nitric oxide production, and TNF-α, IL-1β, IL-6 and iNOS gene transcription in RAW 264.7 macrophages, had functional groups “critical for anti-inflammatory... activity” [127]. Rat brain microglia, a macrophage type involved in neuroinflammation and neurodegeneration [180] was used by Mayer and colleagues to investigate several known diterpene isocyanide amphilectane metabolites (124,125) from the Caribbean marine sponge Hymeniacidon sp., which potently inhibited thromboxane B2 generation from LPS activated rat neonatal microglia in vitro, with concomitant low lactate dehydrogenase release and minimal mitochondrial dehydrogenase inhibition. The authors concluded that the potency of these compounds warranted “further investigation…as lead compounds to modulate…activated microglia in neuroinflammatory disorders” [128]. Ahmed and colleagues extended the pharmacology of largazole (126), originally isolated from a marine cyanobacterium Symploca sp., by reporting that largazole inhibited class I histone deacetylase 6 in vitro in human rheumatoid arthritis. Furthermore, largazole-enhanced expression of intercellular adhesion molecule-1 and vascular cell adhesion molecule-1 was observed to be mediated by activation of the p38 and Akt signal transduction pathways in synovial fibroblasts [129]. Lee and colleagues reported that the sesquiterpenoid lemnalol (127), isolated from the Japanese soft coral Lemnalia tenuis, attenuated monosodium urate-induced gouty rat arthritis, by a mechanism that involved inhibition of inducible nitric oxide synthase and cyclooxygenase-2, thus becoming a potential new candidate for “development of a new treatment for gout” [130]. Kim and colleagues reported that the diketopiperazine-type indole alkaloid neoechinulin A (128), isolated from an Antarctic marine fungus Eurotium sp. SF-5989, inhibited LPS-stimulated RAW264.7 macrophages expression, release of nitric oxide and prostaglandin E2, with concomitant inhibition of NFκB activation, and reduced inhibitor NFκB and p38 mitogen-activated protein kinase (MAPK) phosphorylation [131]. In a detailed study, Lee and colleagues investigated penstyrylpyrone (129), isolated from a marine-derived fungus Penicillium sp. JF-55, and determined that the inhibition of LPS-treated murine peritoneal macrophage production of NO, PGE2, TNF-α, IL-1β, was correlated with suppression of iκB-α and NF-κB and concomitant expression of heme oxygenase-1 [132]. Vilasi and colleagues extended the molecular pharmacology of the novel cyclic octapepetide perthamide C (130), isolated from the marine sponge Theonella swinhoei, by investigating its effect on the proteome of murine macrophages J774.A1 using two-dimensional proteomics, and determining differential effect on several cytosolic and ER-associated proteins, mainly involved in cellular folding processes, thus “shed(ding) more light on the…mechanisms of action” of this natural product [133]. Reina and colleagues reported that R-prostaglandins (131,132) isolated from the Caribbean Colombian soft coral Plexaura homomalla inhibited 12-O-tetradecanoylphorbol-13-acetate-induced mouse ear inflammation in vivo and decreased human polymorphonuclear leukocytes degranulation, as well as myeloperoxidase and elastase levels in vitro, thus concluding that prostaglandins from “…P. homomalla are promising molecules with an interesting anti-inflammatory activity profile” [134]. Huang and colleagues extended the pharmacology of the known compound sinularin (133), demonstrating that it modulates nociceptive responses and spinal neuroinflammation by a mechanism that may involve inhibition of leukocyte iNOS and cyclooxygenase-2 (COX-2) and the upregulation of the anti-inflammatory cytokine transforming growth factor-β [135]. Marino and colleagues reported the molecular pharmacology of the novel polyhydroxylated steroid swinhosterol B (134) isolated from the Solomon Islands marine sponge T. swinhoei [136]. Swinhosterol B was shown to be a highly specific agonist for the human pregnane-X-receptor (PXR), and in transgenic PXR murine monocytes, it attenuated pro-inflammatory cytokine production in vitro, thus supporting “the exploitation of this compound in rodent model(s) of liver inflammation and cholestasis”.
As shown in Table 2, and in contrast to the 20 marine compounds (115134) with described anti-inflammatory mechanisms of action, for marine compounds (135157), only anti-inflammatory activity, namely IC50, was reported, but the molecular mechanism of action remained undetermined: A. polyacanthus steroids (135,136) [137]; barettin (137) [138]; briarenolide F (138) [139]; diketopiperazine (139) [140]; 6-epi-cladieunicellin F (140) [141]; crassarosteroside A (141) [142]; cystodione A (142) [143]; densanins A and B (143,144) [144]; dissesterol (145) [145]; echinohalimane A (146) [146]; eunicidiol (147) [147]; flexibilisolide C (148) [148]; flexibilisquinone (149) [149]; lobocrassin F (150) [150]; perthamide J (151) [151]; pseudoalteromone A (152) [152]; sarcocrassocolide M (153) [153]; sclerosteroids K and M (154,155) [154]; seco-briarellinone (156) [155]; and sinularioside (157) [156].

3.3. Marine Compounds with Activity on the Immune System

In 2012–2013 preclinical pharmacology of marine compounds that affected the immune system showed a decline as previously reported in this series.
Lin and colleagues reported that the cembrane-type diterpenoid lobocrassin B (158), isolated from the marine soft coral Lobophytum crissum, demonstrated immunomodulatory effects on bone marrow-derived dendritic cells (DC), a cell type known to be an important link between the innate and adaptive immune response [157]. Lobocrassin B was shown to attenuate DC maturation and activation with concomitant inhibition of toll-like receptor-stimulated translocation of NF-κB and TNF-α production, data that suggested that lobocrassin B might have “therapeutic applications in certain immune disfunctions”. Chen and colleagues reported that a novel mycophenolic acid derivative, penicacid B (159), isolated from a South China sea fungus Penicillium sp. SOF07, inhibited splenocyte lymphocyte proliferation by a mechanism that involved inhibition of inosine 5′-monophosphate dehydrogenase, an essential rate-limiting enzyme in purine metabolic pathway and an “important drug target for immunosuppressive” activity [158].

3.4. Marine Compounds Affecting the Nervous System

In 2012–2013, the preclinical marine nervous system pharmacology with compounds (160188), which is consolidated in Table 2 and Figure 2, was focused on sodium and potassium channels, nicotinic acetylcholine receptors, as well as, analgesia, antinociception, and neuroprotection.
Four marine compounds (160163) were shown to bind to sodium (Na+) and potassium (K+) channels. Jensen and colleagues determined the effect of cyclisation on the stability of the sea anemone peptide APETx2 (160). Cyclization with either a six-, seven- or eight-residue linker appeared to be a “promising strategy” to increase protease resistance of APETx2, but it decreased its potency against non-voltage gated, pH-sensitive Na+ channel ASIC3 (IC50 = 61 nM). Furthermore, truncation at either N- and C-terminus significantly affected APETx2 binding to ASIC3, demonstrating their critical role in this process [159]. Li and colleagues reported the discovery of a cysteine-crosslinked peptide asteropsin A (161), isolated from a Korean marine sponge Asteropus sp., that affected neuronal Ca2+ influx by a mechanism that involved murine cerebrocortical neurons agonist-induced Na+ channel activation, and may thus represent “…a valuable contribution to the cysteine knot peptide-based drug development as a model scaffold” [160]. Orts and colleagues published the biochemical and electrophysiological characterization of two novel sea anemone type 1 potassium toxins, namely Bcs Tx1 (162) and Bcs Tx2 (163) isolated from the Atlantic sea anemone Bunodosoma caissarum, and demonstrated by electrophysiological screening of 12 subtypes of voltage-gated Kv K+ channels, that BcsTx1 showed highest affinity for rKv1.2 (IC50 = 0.03 ± 0.006 nM) while Bcs Tx2 potently inhibited rKv1.6 (IC50 = 7.76 ± 1.90 nM) [161].
Four studies extended the pharmacology of conopeptides (164167). Favreau and colleagues reported that a novel μ-conopeptide CnIIIIC (164) isolated from the venom of the marine snail C. consors strongly decreased mouse hemidiaphragm contraction by a mechanism that involved potently blocking muscle Nav1.4 (IC50 = 1.3 nM) and rat brain Nav1.2 (IC50 < 1 μM) voltage-gated Na+ channels in a “virtually irreversible” manner, which will probably result in potential development of 164 “…as a myorelaxing drug candidate” [162]. Vetter and colleagues reported the isolation and characterization of a novel hydrophobic 32-residue μO-conotoxin MfVIA (165), isolated from the venom of marine snail C. magnificus, and by using a variety of electrophysiological techniques demonstrated that it preferentially inhibited Nav1.8 (IC50 = 96 nM) and Nav1.4 (IC50 < 81 nM) voltage-gated Na+ channels, leading the authors to propose it as a “drug lead for development of improved analgesic molecules… to improve pain management” [163]. Franco and colleagues isolated an α4/7-conotoxin RegIIA (166) from the venom of the marine cone snail C. regius, and demonstrated that it potently inhibited α3β4 neuronal nicotinic acetylcholine receptors (IC50 = 33 nM) by a mechanism that will require continuous investigation to determine “the precise binding mode of this peptide” [164]. Bernáldez and colleagues described the isolation and biochemical characterization of the first Conus regularis conotoxin designated RsXXIVA (167) with an eight-cysteine framework, which “diverges from other known conotoxins” and that inhibited Cav2.2 channels (IC50 = 2.8 μM) in rat superior cervical ganglion neurons, and also displayed both analgesic and anti-nociceptive activity in the hot-plate and formalin murine in vivo assays, which may contribute to the “design of analgesic peptides” [165].
Two studies reported marine compounds (168,169) that contributed to nociceptive pharmacology. Figuereido and colleagues extended the pharmacology of convolutamydine A (168), isolated from the Floridian marine bryozoan Amantia convoluta, demonstrating that it caused peripheral anti-nociceptive and anti-inflammatory effects in several acute pain models, an effect probably mediated by the cholinergic, opioid and nitric oxide systems and “comparable to morphine’s effects” [166]. Andreev and colleagues contributed an extensive in vitro and in vivo pharmacological study of two polypeptides APHC1 and PAHC3 (169), isolated from the sea anemone Heteractis crispa, shown to have significant anti-nociceptive and analgesic activity in a number of in vivo murine models with associated hypothermia. Furthermore, the two compounds were proposed as a new class of vanilloid 1 receptors modulators based on detailed in vitro biochemical studies [167].
Neuroprotective activity of marine compounds (170,171) was reported in two studies. Feng and colleagues observed that the novel octopamine derivative ianthellamide A (170), isolated from the Australian marine sponge Ianthella quadrangulate, increased endogenous kynurenic acid in rat brain, as well as selectively inhibited the kynurenine 3-hydroxylase in vitro, thus revealing that modulation of the kynurenine pathway of tryptophan metabolism by this compound suggested “potential as a neuroprotective agent” [168]. Burgy and colleagues completed an extensive pharmacological study on the selectivity, co-crystal structures and neuroprotective properties of the leucettines, analogues of the marine sponge alkaloid leucettamine B (171), originally isolated from the calcareous sponge Leucetta microraphis. An optimized product, leucettine L41, with multi-target selectivity that resulted in neuroprotective effects was proposed for “further optimization as potential therapeutics against neurodegenerative diseases such as Alzheimer’s disease” [169].
As shown in Table 2, additional marine compounds (172174) were shown to modulate other molecular targets, i.e., TRPV1 and cannabinoid receptors, and the acetylcholinesterase enzyme. Guzii and colleagues reported that a novel guanidine-containing compound pulchranin A (172), isolated from the marine sponge Monanchora pulchra inhibited TRPV1 receptor, an ionic channel involved in the regulation of pain and body temperature. Pulchranin A, “the first marine non-peptide inhibitor of TRPV1 channels”, led to a decrease of Ca2+ response in a CHO cell line expressing the rat TRPV1 channel by a mechanism the authors propose may result from “direct action on the channel pore” [170]. Montaser and colleagues reported a new fatty acid amide, serinolamide B (173), isolated from the Guam cyanobacterium Lyngbya majuscula that bound with higher selectivity to cannabinoid receptor CB2 and inhibited forskolin-stimulated cAMP accumulation in Chinese hamster ovary cells expressing the CB1 and CB2 receptors, a finding that “introduces a new structural lead to the cannabimimetic” field of research [171]. Huang and colleagues reported the isolation of a new α-pyrone meroterpene arigsugacin I (174), isolated from an endophytic fungus Penicillium sp. Sk5GW1L [172] that was observed to potently inhibit acetylcholinesterase, thus contributing to the “best-established treatment target for the design of anti-Alzheimer’s drugs”.
In contrast to the 15 marine compounds (160174) affecting the nervous system with investigated mechanisms of action discussed above, for marine compounds 175188, only an IC50 was reported and consolidated in Table 2, but their respective molecular mechanisms of action remained undetermined: asperterpenol A (175) [173]; cymatherelactone (176) [174]; dictyodendrin H (177) [175]; geranylphenazinediol (178) [176]; halomadurones C and D (179,180) [177]; lamellarin O (39) [53]; ircinianin lactams A (181,182) [178]; and polar steroids (183188) [179].
Finally, marine bioprospecting resulting from deep sequencing of transcriptomes of marine organisms may ultimately enhance the search for new nervous system drug candidates, as demonstrated by a study of the adult polyp transcriptomes of two cold-water sea anemone species that revealed 15 new neurotoxin peptide candidates [181].

4. Marine Compounds with Miscellaneous Mechanisms of Action

Table 3 presents 2012–2013 preclinical pharmacological research of 69 marine compounds (189257) with miscellaneous mechanisms of action; their structures are shown in Figure 3. Because comprehensive pharmacological characterization data for these compounds were unavailable, it was not possible to assign these compounds to a particular drug class.
Table 3 presents a pharmacological activity, an IC50, and a molecular mechanism of action for 36 marine natural products as reported in the peer-reviewed literature: astaxanthin (189) [182]; biselyngbyaside (190) [183]; Callyspongia sp. bisacetylenic alcohol (191) [184]; conicasterol E (192) [185]; 6”-debromohamacanthin A (193) [186]; dieckol (194) [187]; fructigenine A (195) [188]; geoditin A (196) [189]; gorgosterol (197) [190]; gracilioether B (198) [191]; gracilioether K (199) [192]; herdmanine K (200) [193]; hyrtioreticulin A (201) [194]; new Kunitz-type protease inhibitor InHVJ (202) [195]; jaspamide (203) [196]; latonduine A (204) [197]; leucettine L41 (205) [169]; manzamine A (206) [198]; nahuoic acid A (207) [199]; namalide (208) [200]; ningalins C and D (209,210) [201]; octaphlorethol A (114) [120]; petrosaspongiolide M (211) [202]; petrosiol A (212) [203]; phidianidine A (213) [204]; Poly-APS (214) [205]; Pseudoceratina sp. dibromotyrosine (215) [206]; pseudopterosin A (216) [207]; sargachromanol G (217) [208]; S. graminifolium polysaccharide (218) [209]; S. patens phloroglucinol (219) [210]; S. xiamenensis benzopyran (220) [211]; theonellasterol (221) [212]; toluquinol (222) [213]; and U. lactuca fatty acid (223) [214].
Also described in Table 3 is the pharmacological activity of 34 additional compounds. Albeit an IC50 for enzyme or receptor inhibition is provided, no mechanism of action studies were reported at the time of publication: alotaketal C (224) [215]; aspergentisyl A (225) [216]; A. terreus butyrolactone (226) [217]; caulerpine (227) [218]; conicasterol F (228) [219]; D. avara sesquiterpene (229) [220]; D. gigantea sterols (230,231) [221]; dysidavarone A (232) [222]; galvaquinone B (233) [223]; halicloic acids A and B (234,235) [224]; isochromophilone XI (236) [225]; leucettamols A and B (237,238) [226]; manadosterol A (239) [227]; marilines A1 and A2 (240,241) [228]; methyl sarcotroate B (242) [229]; P. citrinum sorbicillinoid (243) [230]; phosphoiodyn A (244) [231]; purpuroines A and D (245,246) [232]; santacruzamate A (247) [233]; sarcophytonolide N (248) [234]; sargassumol (249) [235]; sesquibastadin 1 (250) [236]; S. glaucum cembranoids (251253) [237]; symplocin A (254) [238]; tsitsikammamine A derivative (255) [239]; V. lanosa bromophenol (256) [240]; and X. testudinaria fatty acid (257) [241].

5. Reviews on Marine Pharmacology

In 2012–2013, several reviews were published covering general and/or specific areas of marine preclinical pharmacology: (a) marine pharmacology and marine pharmaceuticals: new marine natural products and relevant biological activities published in 2010 and 2011 [243,244]; natural products drug discovery as a continuing source of novel drug leads [245]; guiding principles for natural product drug discovery [246]; challenges and triumphs to genomic-based natural product discovery and pharmacology [247]; future of marine natural products drug discovery [248]; bioactive marine natural products from Antarctic and Arctic organisms [249]; biological activities of terpenes from the soft coral genus Sarcophyton [250]; pharmacologically active marine peptides from fish and shellfish [251]; preclinical pharmacology of marine diterpene glycosides [252]; bioactivity of fucoidan, a complex algal sulfated polysaccharide [253]; therapeutic application of marine fucanomics and galactanomics in drug development [254]; marine pharmacology of cosmopolitan brown alga Cystoseira genus secondary metabolites [255]; pharmacological activity of sulfated polysaccharides from marine algae [256]; biological activities and functions of halogenated organic molecules of red algae Rhodomelaceae [257]; pharmacological potential of marine cyanobacterial secondary metabolites [258]; pharmaceutical agents from filamentous marine cyanobacteria [259]; chemistry and preclinical pharmacology of sponge glycosides [260]; sea cucumbers as drug candidates [261]; bioactives from microalgal dinoflagellates [262]; the global marine pharmaceutical pipeline in 2017: U.S. Food and Drug Administration-approved compounds and those in Phase I, II and III of clinical development http://marinepharmacology.midwestern.edu/clinPipeline.htm; (b) antimicrobial marine pharmacology: antimicrobial non-ribosomal peptides from abundant α-, γ- and δ-marine Proteobacteria classes [263]; marine bacteria as potential sources for compounds to overcome methicillin-resistant Staphylococcus aureus [264]; marine coral alkaloids and antibacterial activities [265]; marine fish and invertebrates as sources of antimicrobial peptides [266]; marine actinomycetes as an emerging resource for drug development [267]; chemistry and biological activity of marine Bacillus sp. secondary metabolites [268]; marine compounds with therapeutic potential in Gram-negative sepsis [269]; antimicrobial properties of tunichromes [270]; drug discovery from marine microbes [271]; (c) antiviral marine pharmacology: marine natural products with anti-HIV activities in the last decade [272]; fucoidans as potential inhibitors of human immunodeficiency virus type 1 (HIV-1) [273]; discovery of potent broad spectrum antivirals derived from marine Actinobacteria [274]; algal lectins for prevention of HIV transmission [275]; (d) antiprotozoal, antimalarial, antituberculosis and antifungal marine pharmacology: trypanocidal activity of marine natural products [276]; natural sesquiterpenes as lead compounds for the design of trypanocidal drugs [277]; antifungal compounds from marine fungi [278]; (e) immuno- and anti-inflammatory marine pharmacology: immunoregulatory properties of bryostatin [279]; bioactive marine peptides as potential anti-inflammatory therapeutics [280]; anti-inflammatory soft coral marine natural products from Taiwan [281]; marine natural products with potential for the therapeutics of inflammatory diseases [282]; antioxidant properties of crude extracts and compounds from brown marine algae [283]; (f) cardiovascular and antidiabetic marine pharmacology: oxidation of marine omega-3 supplements and human health [284]; marine peptides for prevention of metabolic syndrome [285]; antidiabetic effect of marine brown algae-derived phlorotannins [286]; marine bioactive peptides as potential antioxidants [287]; cardioprotective peptides from marine sources [288]; antioxidant and antidiabetic pharmacology of fucoxantin [289]; marine-derived bioactive peptides as new anticoagulants [290]; (g) nervous system marine pharmacology: marine neurotoxins, structures, molecular targets and pharmacology [291]; the phosphatase inhibitor okadaic acid as a tool to identify phosphoepitopes relevant to neurodegeneration [292]; marine toxins and drug discovery targeting nicotinic acetylcholine receptors [293]; marine-derived marine secondary metabolites and neuroprotection [294]; cone snail polyketides active in neurological assays [295]; and (h) miscellaneous molecular targets and uses: small-molecule inhibitors of clinically validated protein and lipid kinases of marine origin [296]; natural products as kinase inhibitors [297]; marine natural products with protein tyrosine phosphatase 1B activity [298]; current development strategies for marine conotoxins and their mimetics as therapeutic leads [299]; therapeutic potential of novel conotoxins reported in 2007–2011 [300]; computational studies of marine toxins targeting ion channels [301]; marine invertebrates as sources of skeletal proteins for bone regeneration [302]; marine algal compounds in cosmeceuticals [303]; and marine sponge steroids as nuclear receptor ligands [304].

6. Conclusions

The purpose of the current marine pharmacology review was to continue the marine preclinical pharmacology pipeline review series that was initiated in 1998 [1,2,3,4,5,6,7,8] by consolidating preclinical marine pharmacological research published during 2012–2013 in the global literature. The large number of peer-reviewed publications we have reviewed demonstrates that the global research effort involved chemists and pharmacologists from 43 countries, namely, Argentina, Australia, Austria, Belgium, Brazil, Canada, Chile, China, Colombia, Egypt, Fiji, France, French Polynesia, Germany, Greece, India, Indonesia, Ireland, Israel, Italy, Japan, Malaysia, Mexico, Morocco, the Netherlands, New Zealand, Norway, Pakistan, Panama, Papua New Guinea, Russian Federation, Saudi Arabia, Slovenia, South Africa, South Korea, Spain, Sri Lanka, Switzerland, Taiwan, Thailand, United Kingdom, Vietnam, and the United States. Thus, during 2012–2013 the marine preclinical pharmaceutical pipeline continued to provide novel pharmacological lead compounds that enriched the marine clinical pharmaceutical pipeline. Currently, the clinical pharmaceutical pipeline consists of 6 pharmaceuticals approved by the U.S. Food and Drug Administration, and 29 compounds in Phase I, II and III of clinical pharmaceutical development, as shown at a dedicated website: http://marinepharmacology.midwestern.edu/clinPipeline.htm.

Acknowledgments

We thank the contributions of Hillary Kerns, Michelle Nguyen, and Patrycja Kalwajtys from the Chicago College of Pharmacy for database and literature retrieval. We also thank the secretarial assistance of Victoria Sears, Laura Phelps and Mary Hall from the Pharmacology Department, CCOM for careful review of this manuscript. We gratefully acknowledge financial support from Midwestern University to AMSM; and NIH-SC1 Award (Grant 1SC1GM086271-01A1) of the University of Puerto Rico to ADR, and Italian MIUR (Grant 20154JRJPP) to OTS. The content of this review is solely the responsibility of the authors and does not necessarily represent the official views of the NIH. Article retrieval by library staff members, and students from the Chicago College of Pharmacy, Midwestern University, is gratefully acknowledged. The authors are especially grateful to Mary Hall for her careful review of the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mayer, A.M.S.; Lehmann, V.K.B. Marine pharmacology in 1998: Marine compounds with antibacterial, anticoagulant, antifungal, anti-inflammatory, anthelmintic, antiplatelet, antiprotozoal, and antiviral activities;with actions on the cardiovascular, endocrine, immune, and nervous systems; and other miscellaneous mechanisms of action. Pharmacologist 2000, 42, 62–69. [Google Scholar]
  2. Mayer, A.M.S.; Hamann, M.T. Marine pharmacology in 1999: Compounds with antibacterial, anticoagulant, antifungal, anti-inflammatory, anthelmintic, anti-inflammatory, antiplatelet, antiprotozoal and antiviral activities; affecting the cardiovascular, endocrine, immune, and nervous systems; and other miscellaneous mechanisms of action. Comp. Biochem. Physiol. C Pharmacol. Toxicol. Endocrinol. 2002, 132, 315–339. [Google Scholar]
  3. Mayer, A.M.S.; Hamann, M.T. Marine pharmacology in 2000: Marine compounds with antibacterial, anticoagulant, antifungal, anti-inflammatory, antimalarial, antiplatelet, antituberculosis, and antiviral activities; affecting the cardiovascular, immune, and nervous systems and other miscellaneous mechanisms of action. Mar. Biotechnol. 2004, 6, 37–52. [Google Scholar] [PubMed]
  4. Mayer, A.M.S.; Hamann, M.T. Marine pharmacology in 2001–2002: Marine compounds with anthelmintic, antibacterial, anticoagulant, antidiabetic, antifungal, anti-inflammatory, antimalarial, antiplatelet, antiprotozoal, antituberculosis, and antiviral activities; affecting the cardiovascular, immune and nervous systems and other miscellaneous mechanisms of action. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2005, 140, 265–286. [Google Scholar] [PubMed]
  5. Mayer, A.M.S.; Rodriguez, A.D.; Berlinck, R.G.; Hamann, M.T. Marine pharmacology in 2003-4: Marine compounds with anthelmintic antibacterial, anticoagulant, antifungal, anti-inflammatory, antimalarial, antiplatelet, antiprotozoal, antituberculosis, and antiviral activities; affecting the cardiovascular, immune and nervous systems, and other miscellaneous mechanisms of action. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2007, 145, 553–581. [Google Scholar] [PubMed]
  6. Mayer, A.M.S.; Rodriguez, A.D.; Berlinck, R.G.; Hamann, M.T. Marine pharmacology in 2005-6: Marine compounds with anthelmintic, antibacterial, anticoagulant, antifungal, anti-inflammatory, antimalarial, antiprotozoal, antituberculosis, and antiviral activities; affecting the cardiovascular, immune and nervous systems, and other miscellaneous mechanisms of action. Biochim. Biophys. Acta 2009, 1790, 283–308. [Google Scholar] [PubMed]
  7. Mayer, A.M.S.; Rodriguez, A.D.; Berlinck, R.G.; Fusetani, N. Marine pharmacology in 2007-8: Marine compounds with antibacterial, anticoagulant, antifungal, anti-inflammatory, antimalarial, antiprotozoal, antituberculosis, and antiviral activities; affecting the immune and nervous system, and other miscellaneous mechanisms of action. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2011, 153, 191–222. [Google Scholar] [PubMed]
  8. Mayer, A.M.S.; Rodriguez, A.D.; Taglialatela-Scafati, O.; Fusetani, N. Marine Pharmacology in 2009–2011: Marine Compounds with Antibacterial, Antidiabetic, Antifungal, Anti-Inflammatory, Antiprotozoal, Antituberculosis, and Antiviral Activities; Affecting the Immune and Nervous Systems, and other Miscellaneous Mechanisms of Action. Mar. Drugs 2013, 11, 2510–2573. [Google Scholar] [PubMed]
  9. Schmitz, F.J.; Bowden, B.F.; Toth, S.I. Antitumor and Cytotoxic Compounds from Marine Organisms. In Marine Biotechnology, Pharmaceutical and Bioactive Natural Products; Attaway, D.H., Zaborsky, O.R., Eds.; Plenum Press: New York, NY, USA; London, UK, 1993; pp. 197–308. [Google Scholar]
  10. Cervantes, S.; Stout, E.P.; Prudhomme, J.; Engel, S.; Bruton, M.; Cervantes, M.; Carter, D.; Tae-Chang, Y.; Hay, M.E.; Aalbersberg, W.; et al. High content live cell imaging for the discovery of new antimalarial marine natural products. BMC Infect. Dis. 2012, 12, 1. [Google Scholar] [CrossRef] [PubMed]
  11. Spavieri, J.; Allmendinger, A.; Kaiser, M.; Itoe, M.A.; Blunden, G.; Mota, M.M.; Tasdemir, D. Assessment of dual life stage antiplasmodial activity of british seaweeds. Mar. Drugs 2013, 11, 4019–4034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Yamashita, A.; Salam, K.A.; Furuta, A.; Matsuda, Y.; Fujita, O.; Tani, H.; Fujita, Y.; Fujimoto, Y.; Ikeda, M.; Kato, N.; et al. Inhibition of hepatitis C virus replication and viral helicase by ethyl acetate extract of the marine feather star Alloeocomatella polycladia. Mar. Drugs 2012, 10, 744–761. [Google Scholar] [CrossRef] [PubMed]
  13. De Souza, L.M.; Sassaki, G.L.; Romanos, M.T.; Barreto-Bergter, E. Structural characterization and anti-HSV-1 and HSV-2 activity of glycolipids from the marine algae Osmundaria obtusiloba isolated from Southeastern Brazilian coast. Mar. Drugs 2012, 10, 918–931. [Google Scholar] [CrossRef] [PubMed]
  14. Albuquerque, I.R.; Cordeiro, S.L.; Gomes, D.L.; Dreyfuss, J.L.; Filgueira, L.G.; Leite, E.L.; Nader, H.B.; Rocha, H.A. Evaluation of Anti-Nociceptive and Anti-Inflammatory Activities of a Heterofucan from Dictyota menstrualis. Mar. Drugs 2013, 11, 2722–2740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Cavalcante-Silva, L.H.; da Matta, C.B.; de Araujo, M.V.; Barbosa-Filho, J.M.; de Lira, D.P.; de Oliveira Santos, B.V.; de Miranda, G.E.; Alexandre-Moreira, M.S. Antinociceptive and anti-inflammatory activities of crude methanolic extract of red alga Bryothamnion triquetrum. Mar. Drugs 2012, 10, 1977–1992. [Google Scholar] [CrossRef] [PubMed]
  16. Chaves, L.S.; Nicolau, L.A.; Silva, R.O.; Barros, F.C.; Freitas, A.L.; Aragao, K.S.; Ribeiro, R.A.; Souza, M.H.; Barbosa, A.L.; Medeiros, J.V. Antiinflammatory and antinociceptive effects in mice of a sulfated polysaccharide fraction extracted from the marine red algae Gracilaria caudata. Immunopharmacol. Immunotoxicol. 2013, 35, 93–100. [Google Scholar] [CrossRef] [PubMed]
  17. Lee, J.Y.; Lee, M.S.; Choi, H.J.; Choi, J.W.; Shin, T.; Woo, H.C.; Kim, J.I.; Kim, H.R. Hexane fraction from Laminaria japonica exerts anti-inflammatory effects on lipopolysaccharide-stimulated RAW 264.7 macrophages via inhibiting NF-kappa B pathway. Eur. J. Nutr. 2013, 52, 409–421. [Google Scholar] [CrossRef] [PubMed]
  18. Rezende, B.M.; Bernardes, P.T.; Resende, C.B.; Arantes, R.M.; Souza, D.G.; Braga, F.C.; Castor, M.G.; Teixeira, M.M.; Pinho, V. Lithothamnion muelleri controls inflammatory responses, target organ injury and lethality associated with graft-versus-host disease in mice. Mar. Drugs 2013, 11, 2595–2615. [Google Scholar] [CrossRef] [PubMed]
  19. Zawadzki, M.; Janosch, C.; Szechinski, J. Perna canaliculus lipid complex PCSO-524 demonstrated pain relief for osteoarthritis patients benchmarked against fish oil, a randomized trial, without placebo control. Mar. Drugs 2013, 11, 1920–1935. [Google Scholar] [CrossRef] [PubMed]
  20. Wen, Z.S.; Liu, L.J.; Qu, Y.L.; Ouyang, X.K.; Yang, L.Y.; Xu, Z.R. Chitosan nanoparticles attenuate hydrogen peroxide-induced stress injury in mouse macrophage RAW264.7 cells. Mar. Drugs 2013, 11, 3582–3600. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, B.; Zhang, B.; Wang, Q.; Zhang, Z.; Nie, F.; Liu, G.; Zheng, J.; Xiao, L.; Zhang, L. Pharmacological studies of tentacle extract from the jellyfish Cyanea capillata in isolated rat aorta. Mar. Drugs 2013, 11, 3335–3349. [Google Scholar] [CrossRef] [PubMed]
  22. Silva, R.O.; Santana, A.P.; Carvalho, N.S.; Bezerra, T.S.; Oliveira, C.B.; Damasceno, S.R.; Chaves, L.S.; Freitas, A.L.; Soares, P.M.; Souza, M.H.; et al. A sulfated-polysaccharide fraction from seaweed Gracilaria birdiae prevents naproxen-induced gastrointestinal damage in rats. Mar. Drugs 2012, 10, 2618–2633. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, C.Y.; Wu, W.H.; Wang, J.; Lan, M.B. Antioxidant properties of polysaccharide from the brown seaweed Sargassum graminifolium (Turn.), and its effects on calcium oxalate crystallization. Mar. Drugs 2012, 10, 119–130. [Google Scholar] [CrossRef] [PubMed]
  24. Kelman, D.; Posner, E.K.; McDermid, K.J.; Tabandera, N.K.; Wright, P.R.; Wright, A.D. Antioxidant activity of Hawaiian marine algae. Mar. Drugs 2012, 10, 403–416. [Google Scholar] [CrossRef] [PubMed]
  25. Guedes, A.C.; Giao, M.S.; Seabra, R.; Ferreira, A.C.; Tamagnini, P.; Moradas-Ferreira, P.; Malcata, F.X. Evaluation of the antioxidant activity of cell extracts from microalgae. Mar. Drugs 2013, 11, 1256–1270. [Google Scholar] [CrossRef] [PubMed]
  26. Belhaj, N.; Desor, F.; Gleizes, C.; Denis, F.M.; Arab-Tehrany, E.; Soulimani, R.; Linder, M. Anxiolytic-like effect of a salmon phospholipopeptidic complex composed of polyunsaturated fatty acids and bioactive peptides. Mar. Drugs 2013, 11, 4294–4317. [Google Scholar] [CrossRef] [PubMed]
  27. Carvalho, V.; Fernandes, L.; Conde, T.; Zamith, H.; Silva, R.; Surrage, A.; Frutuoso, V.; Castro-Faria-Neto, H.; Amendoeira, F. Antinociceptive activity of Stephanolepis hispidus skin aqueous extract depends partly on opioid system activation. Mar. Drugs 2013, 11, 1221–1234. [Google Scholar] [CrossRef] [PubMed]
  28. Turk, T.; Ambrozic, A.J.; Batista, U.; Strugar, G.; Kosmina, R.; Civovic, S.; Janussen, D.; Kauferstein, S.; Mebs, D.; Sepcic, K. Biological activities of ethanolic extracts from deep-sea Antarctic marine sponges. Mar. Drugs 2013, 11, 1126–1139. [Google Scholar] [CrossRef] [PubMed]
  29. Cho, S.; Han, D.; Kim, S.B.; Yoon, M.; Yang, H.; Jin, Y.H.; Jo, J.; Yong, H.; Lee, S.H.; Jeon, Y.J.; et al. Depressive Effects on the Central Nervous System and Underlying Mechanism of the Enzymatic Extract and Its Phlorotannin-Rich Fraction from Ecklonia cava Edible Brown Seaweed. Biosci. Biotechnol. Biochem. 2012, 76, 163–168. [Google Scholar] [CrossRef] [PubMed]
  30. Christopeit, T.; Overbo, K.; Danielson, U.H.; Nilsen, I.W. Efficient screening of marine extracts for protease inhibitors by combining FRET based activity assays and surface plasmon resonance spectroscopy based binding assays. Mar. Drugs 2013, 11, 4279–4293. [Google Scholar] [CrossRef] [PubMed]
  31. Jang, K.H.; Nam, S.J.; Locke, J.B.; Kauffman, C.A.; Beatty, D.S.; Paul, L.A.; Fenical, W. Anthracimycin, a potent anthrax antibiotic from a marine-derived actinomycete. Angew. Chem. Int. Ed. Engl. 2013, 52, 7822–7824. [Google Scholar] [CrossRef] [PubMed]
  32. Keffer, J.L.; Huecas, S.; Hammill, J.T.; Wipf, P.; Andreu, J.M.; Bewley, C.A. Chrysophaentins are competitive inhibitors of FtsZ and inhibit Z-ring formation in live bacteria. Bioorg. Med. Chem. 2013, 21, 5673–5678. [Google Scholar] [CrossRef] [PubMed]
  33. Sakoulas, G.; Nam, S.J.; Loesgen, S.; Fenical, W.; Jensen, P.R.; Nizet, V.; Hensler, M. Novel Bacterial Metabolite Merochlorin A Demonstrates in vitro Activity against Multi-Drug Resistant Methicillin-Resistant Staphylococcus aureus. PLoS ONE 2012, 7, e29439. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, H.; Lu, Z.; Qu, H.J.; Liu, P.; Miao, C.; Zhu, T.; Li, J.; Hong, K.; Zhu, W. Antimicrobial aflatoxins from the marine-derived fungus Aspergillus flavus 092008. Arch. Pharm. Res. 2012, 35, 1387–1392. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, F.; Hamann, M.T.; Zou, Y.; Zhang, M.Y.; Gong, X.B.; Xiao, J.R.; Chen, W.S.; Lin, H.W. Antimicrobial metabolites from the Paracel Islands sponge Agelas mauritiana. J. Nat. Prod. 2012, 75, 774–778. [Google Scholar] [CrossRef] [PubMed]
  36. Zheng, C.J.; Shao, C.L.; Guo, Z.Y.; Chen, J.F.; Deng, D.S.; Yang, K.L.; Chen, Y.Y.; Fu, X.M.; She, Z.G.; Lin, Y.C.; et al. Bioactive hydroanthraquinones and anthraquinone dimers from a soft coral-derived Alternaria sp. fungus. J. Nat. Prod. 2012, 75, 189–197. [Google Scholar] [CrossRef] [PubMed]
  37. Han, Z.; Xu, Y.; McConnell, O.; Liu, L.; Li, Y.; Qi, S.; Huang, X.; Qian, P. Two antimycin A analogues from marine-derived actinomycete Streptomyces lusitanus. Mar. Drugs 2012, 10, 668–676. [Google Scholar] [CrossRef] [PubMed]
  38. Li, D.; Xu, Y.; Shao, C.L.; Yang, R.Y.; Zheng, C.J.; Chen, Y.Y.; Fu, X.M.; Qian, P.Y.; She, Z.G.; de Voogd, N.J.; et al. Antibacterial bisabolane-type sesquiterpenoids from the sponge-derived fungus Aspergillus sp. Mar. Drugs 2012, 10, 234–241. [Google Scholar] [CrossRef] [PubMed]
  39. Pettit, G.R.; Tang, Y.; Zhang, Q.; Bourne, G.T.; Arm, C.A.; Leet, J.E.; Knight, J.C.; Pettit, R.K.; Chapuis, J.C.; Doubek, D.L.; et al. Isolation and structures of axistatins 1–3 from the Republic of Palau marine sponge Agelas axifera Hentschel. J. Nat. Prod. 2013, 76, 420–424. [Google Scholar] [CrossRef] [PubMed]
  40. Teasdale, M.E.; Shearer, T.L.; Engel, S.; Alexander, T.S.; Fairchild, C.R.; Prudhomme, J.; Torres, M.; Le, R.K.; Aalbersberg, W.; Hay, M.E.; et al. Bromophycoic acids: Bioactive natural products from a Fijian red alga Callophycus sp. J. Org. Chem. 2012, 77, 8000–8006. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, W.; Kim, H.; Nam, S.J.; Rho, B.J.; Kang, H. Antibacterial butenolides from the Korean tunicate Pseudodistoma antinboja. J. Nat. Prod. 2012, 75, 2049–2054. [Google Scholar] [CrossRef] [PubMed]
  42. Won, T.H.; Jeon, J.E.; Kim, S.H.; Lee, S.H.; Rho, B.J.; Oh, D.C.; Oh, K.B.; Shin, J. Brominated aromatic furanones and related esters from the ascidian Synoicum sp. J. Nat. Prod. 2012, 75, 2055–2061. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, L.L.; Xu, Y.; Han, Z.; Li, Y.X.; Lu, L.; Lai, P.Y.; Zhong, J.L.; Guo, X.R.; Zhang, X.X.; Qian, P.Y. Four new antibacterial xanthones from the marine-derived actinomycetes Streptomyces caelestis. Mar. Drugs 2012, 10, 2571–2583. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, J.; Liu, P.; Wang, Y.; Wang, H.; Li, J.; Zhuang, Y.; Zhu, W. Antimicrobial Aromatic Polyketides from Gorgonian-Associated Fungus, Penicillium commune 518. Chin. J. Chem. 2012, 30, 1236–1242. [Google Scholar] [CrossRef]
  45. Ioannou, E.; Quesada, A.; Rahman, M.M.; Gibbons, S.; Vagias, C.; Roussis, V. Structures and Antibacterial Activities of Minor Dolabellanes from the Brown Alga Dilophus spiralis. Eur. J. Org. Chem. 2012, 2012, 5177–5186. [Google Scholar] [CrossRef]
  46. Felder, S.; Kehraus, S.; Neu, E.; Bierbaum, G.; Schaberle, T.F.; Konig, G.M. Salimyxins and enhygrolides: Antibiotic, sponge-related metabolites from the obligate marine myxobacterium Enhygromyxa salina. Chembiochem 2013, 14, 1363–1371. [Google Scholar] [CrossRef] [PubMed]
  47. Won, T.H.; Jeon, J.E.; Lee, S.H.; Rho, B.J.; Oh, K.B.; Shin, J. Beta-carboline alkaloids derived from the ascidian Synoicum sp. Bioorg. Med. Chem. 2012, 20, 4082–4087. [Google Scholar] [CrossRef] [PubMed]
  48. Xin, W.; Ye, X.; Yu, S.; Lian, X.Y.; Zhang, Z. New capoamycin-type antibiotics and polyene acids from marine Streptomyces fradiae PTZ0025. Mar. Drugs 2012, 10, 2388–2402. [Google Scholar] [CrossRef] [PubMed]
  49. Lee, Y.; Jang, K.H.; Jeon, J.E.; Yang, W.Y.; Sim, C.J.; Oh, K.B.; Shin, J. Cyclic Bis-1,3-dialkylpyridiniums from the sponge Haliclona sp. Mar. Drugs 2012, 10, 2126–2137. [Google Scholar] [CrossRef] [PubMed]
  50. Tanaka, N.; Momose, R.; Takahashi, Y.; Kubota, T.; Takahashi-Nakaguchi, A.; Gonoi, T.; Fromont, J.; Kobayashi, J. Hyrtimomines D and E, bisindole alkaloids from a marine sponge Hyrtios sp. Tetrahedron Lett. 2013, 54, 4038–4040. [Google Scholar] [CrossRef]
  51. Xu, M.; Davis, R.A.; Feng, Y.; Sykes, M.L.; Shelper, T.; Avery, V.M.; Camp, D.; Quinn, R.J. Ianthelliformisamines A–C, antibacterial bromotyrosine-derived metabolites from the marine sponge Suberea ianthelliformis. J. Nat. Prod. 2012, 75, 1001–1005. [Google Scholar] [CrossRef] [PubMed]
  52. Martin, J.; da S Sousa, T.; Crespo, G.; Palomo, S.; Gonzalez, I.; Tormo, J.R.; de la Cruz, M.; Anderson, M.; Hill, R.T.; Vicente, F.; et al. Kocurin, the true structure of PM181104, an anti-methicillin-resistant Staphylococcus aureus (MRSA) thiazolyl peptide from the marine-derived bacterium Kocuria palustris. Mar. Drugs 2013, 11, 387–398. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, H.; Conte, M.M.; Huang, X.C.; Khalil, Z.; Capon, R.J. A search for BACE inhibitors reveals new biosynthetically related pyrrolidones, furanones and pyrroles from a southern Australian marine sponge, Ianthella sp. Org. Biomol. Chem. 2012, 10, 2656–2663. [Google Scholar] [CrossRef] [PubMed]
  54. Li, X.D.; Miao, F.P.; Li, K.; Ji, N.Y. Sesquiterpenes and acetogenins from the marine red alga Laurencia okamurai. Fitoterapia 2012, 83, 518–522. [Google Scholar] [CrossRef] [PubMed]
  55. Pan, H.Q.; Zhang, S.Y.; Wang, N.; Li, Z.L.; Hua, H.M.; Hu, J.C.; Wang, S.J. New spirotetronate antibiotics, lobophorins H and I, from a South China Sea-derived Streptomyces sp. 12A35. Mar. Drugs 2013, 11, 3891–3901. [Google Scholar] [CrossRef] [PubMed]
  56. Zhou, X.; Huang, H.; Chen, Y.; Tan, J.; Song, Y.; Zou, J.; Tian, X.; Hua, Y.; Ju, J. Marthiapeptide A, an anti-infective and cytotoxic polythiazole cyclopeptide from a 60 L scale fermentation of the deep sea-derived Marinactinospora thermotolerans SCSIO 00652. J. Nat. Prod. 2012, 75, 2251–2255. [Google Scholar] [CrossRef] [PubMed]
  57. Wu, Z.; Li, S.; Li, J.; Chen, Y.; Saurav, K.; Zhang, Q.; Zhang, H.; Zhang, W.; Zhang, W.; Zhang, S.; et al. Antibacterial and cytotoxic new napyradiomycins from the marine-derived Streptomyces sp. SCSIO 10428. Mar. Drugs 2013, 11, 2113–2125. [Google Scholar] [CrossRef] [PubMed]
  58. Cheng, Y.B.; Jensen, P.R.; Fenical, W. Cytotoxic and Antimicrobial Napyradiomycins from Two Marine-Derived Streptomyces Strains. Eur. J. Org. Chem. 2013, 3751–3757. [Google Scholar] [CrossRef] [PubMed]
  59. Yang, K.L.; Wei, M.Y.; Shao, C.L.; Fu, X.M.; Guo, Z.Y.; Xu, R.F.; Zheng, C.J.; She, Z.G.; Lin, Y.C.; Wang, C.Y. Antibacterial anthraquinone derivatives from a sea anemone-derived fungus Nigrospora sp. J. Nat. Prod. 2012, 75, 935–941. [Google Scholar] [CrossRef] [PubMed]
  60. Um, S.; Choi, T.J.; Kim, H.; Kim, B.Y.; Kim, S.H.; Lee, S.K.; Oh, K.B.; Shin, J.; Oh, D.C. Ohmyungsamycins A and B: Cytotoxic and Antimicrobial Cyclic Peptides Produced by Streptomyces sp. from a Volcanic Island. J. Org. Chem. 2013, 78, 12321–12329. [Google Scholar] [CrossRef] [PubMed]
  61. Qi, J.; Shao, C.L.; Li, Z.Y.; Gan, L.S.; Fu, X.M.; Bian, W.T.; Zhao, H.Y.; Wang, C.Y. Isocoumarin derivatives and benzofurans from a sponge-derived Penicillium sp. fungus. J. Nat. Prod. 2013, 76, 571–579. [Google Scholar] [CrossRef] [PubMed]
  62. Rubiolo, J.A.; Ternon, E.; Lopez-Alonso, H.; Thomas, O.P.; Vega, F.V.; Vieytes, M.R.; Botana, L.M. Crambescidin-816 acts as a fungicidal with more potency than crambescidin-800 and -830, inducing cell cycle arrest, increased cell size and apoptosis in Saccharomyces cerevisiae. Mar. Drugs 2013, 11, 4419–4434. [Google Scholar] [CrossRef] [PubMed]
  63. Yibmantasiri, P.; Leahy, D.C.; Busby, B.P.; Angermayr, S.A.; Sorgo, A.G.; Boeger, K.; Heathcott, R.; Barber, J.M.; Moraes, G.; Matthews, J.H.; et al. Molecular basis for fungicidal action of neothyonidioside, a triterpene glycoside from the sea cucumber, Australostichopus mollis. Mol. Biosyst. 2012, 8, 902–912. [Google Scholar] [CrossRef] [PubMed]
  64. Kumar, R.; Subramani, R.; Feussner, K.D.; Aalbersberg, W. Aurantoside K, a new antifungal tetramic acid glycoside from a Fijian marine sponge of the genus Melophlus. Mar. Drugs 2012, 10, 200–208. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, A.H.; Liu, D.Q.; Liang, T.J.; Yu, X.Q.; Feng, M.T.; Yao, L.G.; Fang, Y.; Wang, B.; Feng, L.H.; Zhang, M.X.; et al. Caulerprenylols A and B, two rare antifungal prenylated para-xylenes from the green alga Caulerpa racemosa. Bioorg. Med. Chem. Lett. 2013, 23, 2491–2494. [Google Scholar] [CrossRef] [PubMed]
  66. Haga, A.; Tamoto, H.; Ishino, M.; Kimura, E.; Sugita, T.; Kinoshita, K.; Takahashi, K.; Shiro, M.; Koyama, K. Pyridone alkaloids from a marine-derived fungus, Stagonosporopsis cucurbitacearum, and their activities against azole-resistant Candida albicans. J. Nat. Prod. 2013, 76, 750–754. [Google Scholar] [CrossRef] [PubMed]
  67. Piao, S.J.; Song, Y.L.; Jiao, W.H.; Yang, F.; Liu, X.F.; Chen, W.S.; Han, B.N.; Lin, H.W. Hippolachnin A, a New Antifungal Polyketide from the South China Sea Sponge Hippospongia lachne. Org. Lett. 2013, 15, 3526–3529. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, Z.; Zhang, H.; Yuan, W.; Gong, W.; Tang, H.; Liu, B.; Krohn, K.; Li, L.; Yi, Y.; Zhang, W. Antifungal nortriterpene and triterpene glycosides from the sea cucumber Apostichopus japonicus Selenka. Food Chem. 2013, 132, 295–300. [Google Scholar] [CrossRef] [PubMed]
  69. Tanaka, N.; Kusama, T.; Takahashi-Nakaguchi, A.; Gonoi, T.; Fromont, J.; Kobayashi, J. Nagelamides X–Z, dimeric bromopyrrole alkaloids from a marine sponge Agelas sp. Org. Lett. 2013, 15, 3262–3265. [Google Scholar] [CrossRef] [PubMed]
  70. Yu, H.B.; Liu, X.F.; Xu, Y.; Gan, J.H.; Jiao, W.H.; Shen, Y.; Lin, H.W. Woodylides A–C, new cytotoxic linear polyketides from the South China Sea sponge Plakortis simplex. Mar. Drugs 2012, 10, 1027–1036. [Google Scholar] [CrossRef] [PubMed]
  71. Mani, L.; Jullian, V.; Mourkazel, B.; Valentin, A.; Dubois, J.; Cresteil, T.; Folcher, E.; Hooper, J.N.; Erpenbeck, D.; Aalbersberg, W.; et al. New antiplasmodial bromotyrosine derivatives from Suberea ianthelliformis Lendenfeld, 1888. Chem. Biodivers. 2012, 9, 1436–1451. [Google Scholar] [CrossRef] [PubMed]
  72. Lam, C.F.; Pearce, A.N.; Tan, S.H.; Kaiser, M.; Copp, B.R. Discovery and evaluation of thiazinoquinones as anti-protozoal agents. Mar. Drugs 2013, 11, 3472–3499. [Google Scholar] [CrossRef] [PubMed]
  73. Farokhi, F.; Grellier, P.; Clement, M.; Roussakis, C.; Loiseau, P.M.; Genin-Seward, E.; Kornprobst, J.M.; Barnathan, G.; Wielgosz-Collin, G. Antimalarial activity of axidjiferosides, new beta-galactosylceramides from the African sponge Axinyssa djiferi. Mar. Drugs 2013, 11, 1304–1315. [Google Scholar] [CrossRef] [PubMed]
  74. Beau, J.; Mahid, N.; Burda, W.N.; Harrington, L.; Shaw, L.N.; Mutka, T.; Kyle, D.E.; Barisic, B.; van, O.A.; Baker, B.J. Epigenetic tailoring for the production of anti-infective cytosporones from the marine fungus Leucostoma persoonii. Mar. Drugs 2012, 10, 762–774. [Google Scholar] [CrossRef] [PubMed]
  75. Calcul, L.; Waterman, C.; Ma, W.S.; Lebar, M.D.; Harter, C.; Mutka, T.; Morton, L.; Maignan, P.; van, O.A.; Kyle, D.E.; et al. Screening mangrove endophytic fungi for antimalarial natural products. Mar. Drugs 2013, 11, 5036–5050. [Google Scholar] [CrossRef] [PubMed]
  76. Ilias, M.; Ibrahim, M.A.; Khan, S.I.; Jacob, M.R.; Tekwani, B.L.; Walker, L.A.; Samoylenko, V. Pentacyclic ingamine alkaloids, a new antiplasmodial pharmacophore from the marine sponge Petrosid Ng5 Sp5. Planta Med. 2012, 78, 1690–1697. [Google Scholar] [CrossRef] [PubMed]
  77. Mudianta, I.W.; Skinner-Adams, T.; Andrews, K.T.; Davis, R.A.; Hadi, T.A.; Hayes, P.Y.; Garson, M.J. Psammaplysin derivatives from the Balinese marine sponge Aplysinella strongylata. J. Nat. Prod. 2012, 75, 2132–2143. [Google Scholar] [CrossRef] [PubMed]
  78. Sirirak, T.; Brecker, L.; Plubrukarn, A. Kabiramide L, a new antiplasmodial trisoxazole macrolide from the sponge Pachastrissa nux. Nat. Prod. Res. 2013, 27, 1213–1219. [Google Scholar] [CrossRef] [PubMed]
  79. Bharate, S.B.; Rammohan, R.Y.; Khan, S.I.; Tekwani, B.L.; Jacob, M.R.; Khan, I.A.; Vishwakarma, R.A. Meridianin G and its analogs as antimalarial agents. Med. Chem. Commun. 2013, 4, 1042–1048. [Google Scholar] [CrossRef]
  80. Liew, L.P.; Kaiser, M.; Copp, B.R. Discovery and preliminary structure-activity relationship analysis of 1,14-sperminediphenylacetamides as potent and selective antimalarial lead compounds. Bioorg. Med. Chem. Lett. 2013, 23, 452–454. [Google Scholar] [CrossRef] [PubMed]
  81. Festa, C.; De Marino, S.; D’Auria, M.V.; Taglialatela-Scafati, O.; Deharo, E.; Petek, S.; Zampella, A. New antimalarial polyketide endoperoxides from the marine sponge Plakinastrella mamillaris collected at Fiji Islands. Tetrahedron 2013, 69, 3706–3713. [Google Scholar] [CrossRef]
  82. Davis, R.A.; Duffy, S.; Fletcher, S.; Avery, V.M.; Quinn, R.J. Thiaplakortones A-D: Antimalarial thiazine alkaloids from the Australian marine sponge Plakortis lita. J. Org. Chem. 2013, 78, 9608–9613. [Google Scholar] [CrossRef] [PubMed]
  83. Davis, R.A.; Buchanan, M.S.; Duffy, S.; Avery, V.M.; Charman, S.A.; Charman, W.N.; White, K.L.; Shackleford, D.M.; Edstein, M.D.; Andrews, K.T.; et al. Antimalarial activity of pyrroloiminoquinones from the Australian marine sponge Zyzzya sp. J. Med. Chem. 2012, 55, 5851–5858. [Google Scholar] [CrossRef] [PubMed]
  84. Supong, K.; Thawai, C.; Suwanborirux, K.; Choowong, W.; Supothina, S.; Pittayakhajonwut, P. Antimalarial and antitubercular C-glycosylated benz[a]anthraquinones from the marine-derived Streptomyces sp. BCC45596. Phytochem. Lett. 2012, 5, 651–656. [Google Scholar] [CrossRef]
  85. Sanchez, L.M.; Knudsen, G.M.; Helbig, C.; De, M.G.; Mascuch, S.M.; Mackey, Z.B.; Gerwick, L.; Clayton, C.; McKerrow, J.H.; Linington, R.G. Examination of the mode of action of the almiramide family of natural products against the kinetoplastid parasite Trypanosoma brucei. J. Nat. Prod. 2013, 76, 630–641. [Google Scholar] [CrossRef] [PubMed]
  86. Abdelmohsen, U.R.; Szesny, M.; Othman, E.M.; Schirmeister, T.; Grond, S.; Stopper, H.; Hentschel, U. Antioxidant and anti-protease activities of diazepinomicin from the sponge-associated Micromonospora strain RV115. Mar. Drugs 2012, 10, 2208–2221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Desoti, V.C.; Lazarin-Bidoia, D.; Sudatti, D.B.; Pereira, R.C.; Alonso, A.; Ueda-Nakamura, T.; Dias Filho, B.P.; Nakamura, C.V.; Silva, S.O. Trypanocidal action of (−)-elatol involves an oxidative stress triggered by mitochondria dysfunction. Mar. Drugs 2012, 10, 1631–1646. [Google Scholar] [CrossRef] [PubMed]
  88. Balunas, M.J.; Grosso, M.F.; Villa, F.A.; Engene, N.; McPhail, K.L.; Tidgewell, K.; Pineda, L.M.; Gerwick, L.; Spadafora, C.; Kyle, D.E.; et al. Coibacins A–D, antileishmanial marine cyanobacterial polyketides with intriguing biosynthetic origins. Org. Lett. 2012, 14, 3878–3881. [Google Scholar] [CrossRef] [PubMed]
  89. Ishigami, S.T.; Goto, Y.; Inoue, N.; Kawazu, S.; Matsumoto, Y.; Imahara, Y.; Tarumi, M.; Nakai, H.; Fusetani, N.; Nakao, Y. Cristaxenicin A, an antiprotozoal xenicane diterpenoid from the deep sea gorgonian Acanthoprimnoa cristata. J. Org. Chem. 2012, 77, 10962–10966. [Google Scholar] [CrossRef] [PubMed]
  90. Chianese, G.; Scala, F.; Calcinai, B.; Cerrano, C.; Dien, H.A.; Kaiser, M.; Tasdemir, D.; Taglialatela-Scafati, O. Natural and semisynthetic analogues of manadoperoxide B reveal new structural requirements for trypanocidal activity. Mar. Drugs 2013, 11, 3297–3308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Huang, X.; Huang, H.; Li, H.; Sun, X.; Huang, H.; Lu, Y.; Lin, Y.; Long, Y.; She, Z. Asperterpenoid A, a new sesterterpenoid as an inhibitor of Mycobacterium tuberculosis protein tyrosine phosphatase B from the culture of Aspergillus sp. 16–5c. Org. Lett. 2013, 15, 721–723. [Google Scholar] [CrossRef] [PubMed]
  92. Song, F.; Liu, X.; Guo, H.; Ren, B.; Chen, C.; Piggott, A.M.; Yu, K.; Gao, H.; Wang, Q.; Liu, M.; et al. Brevianamides with antitubercular potential from a marine-derived isolate of Aspergillus versicolor. Org. Lett. 2012, 14, 4770–4773. [Google Scholar] [CrossRef] [PubMed]
  93. Chen, C.; Wang, J.; Guo, H.; Hou, W.; Yang, N.; Ren, B.; Liu, M.; Dai, H.; Liu, X.; Song, F.; et al. Three antimycobacterial metabolites identified from a marine-derived Streptomyces sp. MS100061. Appl. Microbiol. Biotechnol. 2013, 97, 3885–3892. [Google Scholar] [CrossRef] [PubMed]
  94. Yamano, Y.; Arai, M.; Kobayashi, M. Neamphamide B, new cyclic depsipeptide, as an anti-dormant mycobacterial substance from a Japanese marine sponge of Neamphius sp. Bioorg. Med. Chem. Lett. 2012, 22, 4877–4881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Aviles, E.; Rodriguez, A.D.; Vicente, J. Two rare-class tricyclic diterpenes with antitubercular activity from the Caribbean sponge Svenzea flava. Application of vibrational circular dichroism spectroscopy for determining absolute configuration. J. Org. Chem. 2013, 78, 11294–11301. [Google Scholar] [CrossRef] [PubMed]
  96. Da Rosa Guimaraes, T.; Quiroz, C.G.; Borges, C.R.; de Oliveira, S.Q.; de Almeida, M.T.; Bianco, E.M.; Moritz, M.I.; Carraro, J.L.; Palermo, J.A.; Cabrera, G.; et al. Anti HSV-1 activity of halistanol sulfate and halistanol sulfate C isolated from Brazilian marine sponge Petromica citrina (Demospongiae). Mar. Drugs 2013, 11, 4176–4192. [Google Scholar] [CrossRef] [PubMed]
  97. Ellithey, M.S.; Lall, N.; Hussein, A.A.; Meyer, D. Cytotoxic, cytostatic and HIV-1 PR inhibitory activities of the soft coral Litophyton arboreum. Mar. Drugs 2013, 11, 4917–4936. [Google Scholar] [CrossRef] [PubMed]
  98. Salam, K.A.; Furuta, A.; Noda, N.; Tsuneda, S.; Sekiguchi, Y.; Yamashita, A.; Moriishi, K.; Nakakoshi, M.; Tsubuki, M.; Tani, H.; et al. Inhibition of hepatitis C virus NS3 helicase by manoalide. J. Nat. Prod. 2012, 75, 650–654. [Google Scholar] [CrossRef] [PubMed]
  99. Park, S.H.; Song, J.H.; Kim, T.; Shin, W.S.; Park, G.M.; Lee, S.; Kim, Y.J.; Choi, P.; Kim, H.; Kim, H.S.; et al. Anti-human rhinoviral activity of polybromocatechol compounds isolated from the rhodophyta, Neorhodomela aculeata. Mar. Drugs 2012, 10, 2222–2233. [Google Scholar] [CrossRef] [PubMed]
  100. Ma, X.; Li, L.; Zhu, T.; Ba, M.; Li, G.; Gu, Q.; Guo, Y.; Li, D. Phenylspirodrimanes with anti-HIV activity from the sponge-derived fungus Stachybotrys chartarum MXH-X73. J. Nat. Prod. 2013, 76, 2298–2306. [Google Scholar] [CrossRef] [PubMed]
  101. Jiao, R.H.; Xu, H.; Cui, J.T.; Ge, H.M.; Tan, R.X. Neuraminidase Inhibitors from marine-derived actinomycete Streptomyces seoulensis. J. Appl. Microbiol. 2013, 114, 1046–1053. [Google Scholar] [CrossRef] [PubMed]
  102. He, F.; Bao, J.; Zhang, X.Y.; Tu, Z.C.; Shi, Y.M.; Qi, S.H. Asperterrestide A, a cytotoxic cyclic tetrapeptide from the marine-derived fungus Aspergillus terreus SCSGAF0162. J. Nat. Prod. 2013, 76, 1182–1186. [Google Scholar] [CrossRef] [PubMed]
  103. Peng, J.; Lin, T.; Wang, W.; Xin, Z.; Zhu, T.; Gu, Q.; Li, D. Antiviral alkaloids produced by the mangrove-derived fungus Cladosporium sp. PJX-41. J. Nat. Prod. 2013, 76, 1133–1140. [Google Scholar] [CrossRef] [PubMed]
  104. Hawas, U.W.; El-Beih, A.A.; El-Halawany, A.M. Bioactive anthraquinones from endophytic fungus Aspergillus versicolor isolated from red sea algae. Arch. Pharm. Res. 2012, 35, 1749–1756. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, G.F.; Han, W.B.; Cui, J.T.; Ng, S.W.; Guo, Z.K.; Tan, R.X.; Ge, H.M. Neuraminidase inhibitory polyketides from the marine-derived fungus Phoma herbarum. Planta Med. 2012, 78, 76–78. [Google Scholar] [CrossRef] [PubMed]
  106. Peng, J.; Jiao, J.; Li, J.; Wang, W.; Gu, Q.; Zhu, T.; Li, D. Pyronepolyene C-glucosides with NF-kappaB inhibitory and anti-influenza A viral (H1N1) activities from the sponge-associated fungus Epicoccum sp. JJY40. Bioorg. Med. Chem. Lett. 2012, 22, 3188–3190. [Google Scholar] [CrossRef] [PubMed]
  107. Ahmed, S.; Ibrahim, A.; Satar Arafa, A. Anti-H5N1 virus metabolites from the Red Sea soft coral, Sinularia candidula. Tetrahedron Lett. 2013, 54, 2377–2381. [Google Scholar] [CrossRef]
  108. Plouguerne, E.; de Souza, L.M.; Sassaki, G.L.; Cavalcanti, J.F.; Villela Romanos, M.T.; da Gama, B.A.; Pereira, R.C.; Barreto-Bergter, E. Antiviral Sulfoquinovosyldiacylglycerols (SQDGs) from the Brazilian brown seaweed Sargassum vulgare. Mar. Drugs 2013, 11, 4628–4640. [Google Scholar] [CrossRef] [PubMed]
  109. Melek, F.R.; Tadros, M.M.; Yousif, F.; Selim, M.A.; Hassan, M.H. Screening of marine extracts for schistosomicidal activity in vitro. Isolation of the triterpene glycosides echinosides A and B with potential activity from the Sea Cucumbers Actinopyga echinites and Holothuria polii. Pharm. Biol. 2012, 50, 490–496. [Google Scholar] [CrossRef] [PubMed]
  110. Chen, D.; Yu, S.; Van, O.L.; Proksch, P.; Lin, W. Anthogorgienes A–O, new guaiazulene-derived terpenoids from a Chinese gorgonian Anthogorgia species, and their antifouling and antibiotic activities. J. Agric. Food Chem. 2012, 60, 112–123. [Google Scholar] [CrossRef] [PubMed]
  111. Nuzzo, G.; Ciavatta, M.L.; Villani, G.; Manzo, E.; Zanfardino, A.; Varcamonti, M.; Gavagnin, M. Fulvynes, antimicrobial polyoxygenated acetylenes from the Mediterranean sponge Haliclona fulva. Tetrahedron 2012, 68, 754–760. [Google Scholar] [CrossRef]
  112. Cheng, Z.B.; Xiao, H.; Fan, C.Q.; Lu, Y.N.; Zhang, G.; Yin, S. Bioactive polyhydroxylated sterols from the marine sponge Haliclona crassiloba. Steroids 2013, 78, 1353–1358. [Google Scholar] [CrossRef] [PubMed]
  113. Hu, Y.; Wang, K.; MacMillan, J.B. Hunanamycin A, an antibiotic from a marine-derived Bacillus hunanensis. Org. Lett. 2013, 15, 390–393. [Google Scholar] [CrossRef] [PubMed]
  114. Khamthong, N.; Rukachaisirikul, V.; Phongpaichit, S.; Preedanon, S.; Sakayaroj, J. Bioactive polyketides from the sea fan-derived fungus Penicillium citrinum PSU-F51. Tetrahedron 2012, 68, 8245–8250. [Google Scholar] [CrossRef]
  115. Wei, M.Y.; Li, D.; Shao, C.L.; Deng, D.S.; Wang, C.Y. (+/−)-Pestalachloride D, an antibacterial racemate of chlorinated benzophenone derivative from a soft coral-derived fungus Pestalotiopsis sp. Mar. Drugs 2013, 11, 1050–1060. [Google Scholar] [CrossRef] [PubMed]
  116. Wyche, T.P.; Hou, Y.; Vazquez-Rivera, E.; Braun, D.; Bugni, T.S. Peptidolipins B–F, antibacterial lipopeptides from an ascidian-derived Nocardia sp. J. Nat. Prod. 2012, 75, 735–740. [Google Scholar] [CrossRef] [PubMed]
  117. Silchenko, A.S.; Kalinovsky, A.I.; Avilov, S.A.; Andryjaschenko, P.V.; Dmitrenok, P.S.; Martyyas, E.A.; Kalinin, V.I. Triterpene glycosides from the sea cucumber Eupentacta fraudatrix. Structure and biological action of cucumariosides I1, I3, I4, three new minor disulfated pentaosides. Nat. Prod. Commun. 2013, 8, 1053–1058. [Google Scholar] [PubMed]
  118. Sun, H.F.; Li, X.M.; Meng, L.; Cui, C.M.; Gao, S.S.; Li, C.S.; Huang, C.G.; Wang, B.G. Asperolides A–C, tetranorlabdane diterpenoids from the marine alga-derived endophytic fungus Aspergillus wentii EN-48. J. Nat. Prod. 2012, 75, 148–152. [Google Scholar] [CrossRef] [PubMed]
  119. Xu, X.; Yin, L.; Gao, L.; Gao, J.; Chen, J.; Li, J.; Song, F. Two new bromophenols with radical scavenging activity from marine red alga Symphyocladia latiuscula. Mar. Drugs 2013, 11, 842–847. [Google Scholar] [CrossRef] [PubMed]
  120. Lee, S.H.; Kang, S.M.; Ko, S.C.; Lee, D.H.; Jeon, Y.J. Octaphlorethol A, a novel phenolic compound isolated from a brown alga, Ishige foliacea, increases glucose transporter 4-mediated glucose uptake in skeletal muscle cells. Biochem. Biophys. Res. Commun. 2012, 420, 576–581. [Google Scholar] [CrossRef] [PubMed]
  121. Chae, D.; Manzoor, Z.; Kim, S.C.; Kim, S.; Oh, T.H.; Yoo, E.S.; Kang, H.K.; Hyun, J.W.; Lee, N.H.; Ko, M.H.; et al. Apo-9′-fucoxanthinone, isolated from Sargassum muticum, inhibits CpG-induced inflammatory response by attenuating the mitogen-activated protein kinase pathway. Mar. Drugs 2013, 11, 3272–3287. [Google Scholar] [CrossRef] [PubMed]
  122. Speranza, L.; Pesce, M.; Patruno, A.; Franceschelli, S.; de Lutiis, M.A.; Grilli, A.; Felaco, M. Astaxanthin treatment reduced oxidative induced pro-inflammatory cytokines secretion in U937: SHP-1 as a novel biological target. Mar. Drugs 2012, 10, 890–899. [Google Scholar] [CrossRef] [PubMed]
  123. Johnson, T.A.; Sohn, J.; Vaske, Y.M.; White, K.N.; Cohen, T.L.; Vervoort, H.C.; Tenney, K.; Valeriote, F.A.; Bjeldanes, L.F.; Crews, P. Myxobacteria versus sponge-derived alkaloids: The bengamide family identified as potent immune modulating agents by scrutiny of LC-MS/ELSD libraries. Bioorg. Med. Chem. 2012, 20, 4348–4355. [Google Scholar] [CrossRef] [PubMed]
  124. Song, Y.; Dou, H.; Gong, W.; Liu, X.; Yu, Z.; Li, E.; Tan, R.; Hou, Y. Bis-N-norgliovictin, a small-molecule compound from marine fungus, inhibits LPS-induced inflammation in macrophages and improves survival in sepsis. Eur. J. Pharmacol. 2013, 705, 49–60. [Google Scholar] [CrossRef] [PubMed]
  125. Yang, Y.I.; Shin, H.C.; Kim, S.H.; Park, W.Y.; Lee, K.T.; Choi, J.H. 6,6′-Bieckol, isolated from marine alga Ecklonia cava, suppressed LPS-induced nitric oxide and PGE(2) production and inflammatory cytokine expression in macrophages: The inhibition of NFkappaB. Int. Immunopharmacol. 2012, 12, 510–517. [Google Scholar] [CrossRef] [PubMed]
  126. Hsu, W.L.; Chiu, S.J.; Tsai, Y.T.; Chang, C.M.; Wang, J.Y.; Wang, E.T.; Hou, M.F.; Huang, C.Y.; Sheu, J.H.; Chang, W.C. A soft coral natural product, 11-episinulariolide acetate, inhibits gene expression of cyclooxygenase-2 and interleukin-8 through attenuation of calcium signaling. Molecules 2013, 18, 7023–7034. [Google Scholar] [CrossRef] [PubMed]
  127. Choi, H.; Mascuch, S.J.; Villa, F.A.; Byrum, T.; Teasdale, M.E.; Smith, J.E.; Preskitt, L.B.; Rowley, D.C.; Gerwick, L.; Gerwick, W.H. Honaucins A–C, potent inhibitors of inflammation and bacterial quorum sensing: Synthetic derivatives and structure-activity relationships. Chem. Biol. 2012, 19, 589–598. [Google Scholar] [CrossRef] [PubMed]
  128. Mayer, A.M.; Aviles, E.; Rodriguez, A.D. Marine sponge Hymeniacidon sp. amphilectane metabolites potently inhibit rat brain microglia thromboxane B2 generation. Bioorg. Med. Chem. 2012, 20, 279–282. [Google Scholar] [CrossRef] [PubMed]
  129. Ahmed, S.; Riegsecker, S.; Beamer, M.; Rahman, A.; Bellini, J.V.; Bhansali, P.; Tillekeratne, L.M. Largazole, a class I histone deacetylase inhibitor, enhances TNF-alpha-induced ICAM-1 and VCAM-1 expression in rheumatoid arthritis synovial fibroblasts. Toxicol. Appl. Pharmacol. 2013, 270, 87–96. [Google Scholar] [CrossRef] [PubMed]
  130. Lee, H.P.; Huang, S.Y.; Lin, Y.Y.; Wang, H.M.; Jean, Y.H.; Wu, S.F.; Duh, C.Y.; Wen, Z.H. Soft coral-derived lemnalol alleviates monosodium urate-induced gouty arthritis in rats by inhibiting leukocyte infiltration and iNOS, COX-2 and c-Fos protein expression. Mar. Drugs 2013, 11, 99–113. [Google Scholar] [CrossRef] [PubMed]
  131. Kim, K.S.; Cui, X.; Lee, D.S.; Sohn, J.H.; Yim, J.H.; Kim, Y.C.; Oh, H. Anti-inflammatory effect of neoechinulin a from the marine fungus Eurotium sp. SF-5989 through the suppression of NF-small ka, CyrillicB and p38 MAPK Pathways in lipopolysaccharide-stimulated RAW264.7 macrophages. Molecules 2013, 18, 13245–13259. [Google Scholar] [CrossRef] [PubMed]
  132. Lee, D.S.; Jang, J.H.; Ko, W.; Kim, K.S.; Sohn, J.H.; Kang, M.S.; Ahn, J.S.; Kim, Y.C.; Oh, H. PTP1B inhibitory and anti-inflammatory effects of secondary metabolites isolated from the marine-derived fungus Penicillium sp. JF-55. Mar. Drugs 2013, 11, 1409–1426. [Google Scholar] [CrossRef] [PubMed]
  133. Vilasi, A.; Monti, M.C.; Tosco, A.; De Marino, S.; Margarucci, L.; Riccio, R.; Casapullo, A. Differential in gel electrophoresis (DIGE) comparative proteomic analysis of macrophages cell cultures in response to perthamide C treatment. Mar. Drugs 2013, 11, 1288–1299. [Google Scholar] [CrossRef] [PubMed]
  134. Reina, E.; Ramos, F.A.; Castellanos, L.; Aragon, M.; Ospina, L.F. Anti-inflammatory R-prostaglandins from Caribbean Colombian soft coral Plexaura homomalla. J. Pharm. Pharmacol. 2013, 65, 1643–1652. [Google Scholar] [CrossRef] [PubMed]
  135. Huang, S.Y.; Chen, N.F.; Chen, W.F.; Hung, H.C.; Lee, H.P.; Lin, Y.Y.; Wang, H.M.; Sung, P.J.; Sheu, J.H.; Wen, Z.H. Sinularin from indigenous soft coral attenuates nociceptive responses and spinal neuroinflammation in carrageenan-induced inflammatory rat model. Mar. Drugs 2012, 10, 1899–1919. [Google Scholar] [CrossRef] [PubMed]
  136. De, M.S.; Ummarino, R.; D’Auria, M.V.; Chini, M.G.; Bifulco, G.; D’Amore, C.; Renga, B.; Mencarelli, A.; Petek, S.; Fiorucci, S.; et al. 4-Methylenesterols from Theonella swinhoei sponge are natural pregnane-X-receptor agonists and farnesoid-X-receptor antagonists that modulate innate immunity. Steroids 2012, 77, 485–495. [Google Scholar]
  137. Thao, N.P.; Cuong, N.X.; Luyen, B.T.; Quang, T.H.; Hanh, T.T.; Kim, S.; Koh, Y.S.; Nam, N.H.; Van, K.P.; Van, M.C.; et al. Anti-inflammatory components of the starfish Astropecten polyacanthus. Mar. Drugs 2013, 11, 2917–2926. [Google Scholar] [CrossRef] [PubMed]
  138. Lind, K.F.; Hansen, E.; Osterud, B.; Eilertsen, K.E.; Bayer, A.; Engqvist, M.; Leszczak, K.; Jorgensen, T.O.; Andersen, J.H. Antioxidant and anti-inflammatory activities of barettin. Mar. Drugs 2013, 11, 2655–2666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Hong, P.H.; Su, Y.D.; Su, J.H.; Chen, Y.H.; Hwang, T.L.; Weng, C.F.; Lee, C.H.; Wen, Z.H.; Sheu, J.H.; Lin, N.C.; et al. Briarenolides F and G, new briarane diterpenoids from a Briareum sp. octocoral. Mar. Drugs 2012, 10, 1156–1168. [Google Scholar] [CrossRef] [PubMed]
  140. Chen, J.H.; Lan, X.P.; Liu, Y.; Jia, A.Q. The effects of diketopiperazines from Callyspongia sp. on release of cytokines and chemokines in cultured J774A.1 macrophages. Bioorg. Med. Chem. Lett. 2012, 22, 3177–3180. [Google Scholar] [CrossRef] [PubMed]
  141. Chen, Y.H.; Hwang, T.L.; Su, Y.D.; Chang, Y.C.; Chen, Y.H.; Hong, P.H.; Hu, L.C.; Yen, W.H.; Hsu, H.Y.; Huang, S.J.; et al. New 6-hydroxyeunicellins from a soft coral Cladiella sp. Chem. Pharm. Bull. 2012, 60, 160–163. [Google Scholar] [CrossRef] [PubMed]
  142. Chao, C.H.; Chou, K.J.; Huang, C.Y.; Wen, Z.H.; Hsu, C.H.; Wu, Y.C.; Dai, C.F.; Sheu, J.H. Steroids from the soft coral Sinularia crassa. Mar. Drugs 2012, 10, 439–450. [Google Scholar] [CrossRef] [PubMed]
  143. De Los, R.C.; Zbakh, H.; Motilva, V.; Zubia, E. Antioxidant and anti-inflammatory meroterpenoids from the brown alga Cystoseira usneoides. J. Nat. Prod. 2013, 76, 621–629. [Google Scholar] [CrossRef] [PubMed]
  144. Hwang, B.S.; Oh, J.S.; Jeong, E.J.; Sim, C.J.; Rho, J.R. Densanins A and B, new macrocyclic pyrrole alkaloids isolated from the marine sponge Haliclona densaspicula. Org. Lett. 2012, 14, 6154–6157. [Google Scholar] [CrossRef] [PubMed]
  145. Thao, N.P.; Nam, N.H.; Cuong, N.X.; Tai, H.; Quang, T.H.; Ngan, N.T.; Luyen, B.T.T.; Yang, S.Y.; Choi, C.H.; Kim, S.; et al. Steroidal Constituents from the Soft Coral Sinularia dissecta and Their Inhibitory Effects on Lipopolysaccharide-Stimulated Production of Pro-inflammatory Cytokines in Bone Marrow-Derived Dendritic Cells. Bull. Korean Chem. Soc. 2013, 34, 949–952. [Google Scholar]
  146. Chung, H.M.; Hu, L.C.; Yen, W.H.; Su, J.H.; Lu, M.C.; Hwang, T.L.; Wang, W.H.; Sung, P.J. Echinohalimane A, a bioactive halimane-type diterpenoid from a Formosan gorgonian Echinomuricea sp. (Plexauridae). Mar. Drugs 2012, 10, 2246–2253. [Google Scholar] [CrossRef] [PubMed]
  147. Marchbank, D.H.; Berrue, F.; Kerr, R.G. Eunicidiol, an anti-inflammatory dilophol diterpene from Eunicea fusca. J. Nat. Prod. 2012, 75, 1289–1293. [Google Scholar] [CrossRef] [PubMed]
  148. Shih, H.J.; Tseng, Y.J.; Huang, C.Y.; Wen, Z.H.; Dai, C.F.; Sheu, J.H. Cytotoxic and anti-inflammatory diterpenoids from the Dongsha Atoll soft coral Sinularia flexibilis. Tetrahedron 2013, 68, 244–249. [Google Scholar] [CrossRef]
  149. Lin, Y.F.; Kuo, C.Y.; Wen, Z.H.; Lin, Y.Y.; Wang, W.H.; Su, J.H.; Sheu, J.H.; Sung, P.J. Flexibilisquinone, a new anti-inflammatory quinone from the cultured soft coral Sinularia flexibilis. Molecules 2013, 18, 8160–8167. [Google Scholar] [CrossRef] [PubMed]
  150. Lee, C.H.; Kao, C.Y.; Kao, S.Y.; Chang, C.H.; Su, J.H.; Hwang, T.L.; Kuo, Y.H.; Wen, Z.H.; Sung, P.J. Terpenoids from the octocorals Menella sp. (Plexauridae) and Lobophytum crassum (Alcyonacea). Mar. Drugs 2012, 10, 427–438. [Google Scholar] [CrossRef] [PubMed]
  151. Festa, C.; De Marino, S.; D’Auria, M.V.; Monti, M.C.; Bucci, M.; Vellecco, V.; Debitus, C.; Zampella, A. Anti-inflammatory cyclopeptides from the marine sponge Theonella swinhoei. Tetrahedron 2012, 68, 2851–2857. [Google Scholar] [CrossRef]
  152. Chen, Y.H.; Lu, M.C.; Chang, Y.C.; Hwang, T.L.; Wang, W.H.; Weng, C.F.; Kuo, J.; Sung, P.J. Pseudoalteromone A: A novel bioactive ubiquinone from a marine bacterium Pseudoalteromonas sp. CGH2XX (Pseudoalteromonadaceae). Tetrahedron Lett. 2012, 53, 1675–1677. [Google Scholar] [CrossRef]
  153. Lin, W.Y.; Lu, Y.; Chen, B.W.; Huang, C.Y.; Su, J.H.; Wen, Z.H.; Dai, C.F.; Kuo, Y.H.; Sheu, J.H. Sarcocrassocolides M–O, bioactive cembranoids from the Dongsha Atoll soft coral Sarcophyton crassocaule. Mar. Drugs 2012, 10, 617–626. [Google Scholar] [CrossRef] [PubMed]
  154. Fang, H.Y.; Hsu, C.H.; Chao, C.H.; Wen, Z.H.; Wu, Y.C.; Dai, C.F.; Sheu, J.H. Cytotoxic and anti-inflammatory metabolites from the soft coral Scleronephthya gracillimum. Mar. Drugs 2013, 11, 1853–1865. [Google Scholar] [CrossRef] [PubMed]
  155. Gomez-Reyes, J.F.; Salazar, A.; Guzman, H.M.; Gonzalez, Y.; Fernandez, P.L.; Ariza-Castolo, A.; Gutierrez, M. Seco-Briarellinone and briarellin S, two new eunicellin-based diterpenoids from the Panamanian octocoral Briareum asbestinum. Mar. Drugs 2012, 10, 2608–2617. [Google Scholar] [CrossRef] [PubMed]
  156. Putra, M.Y.; Ianaro, A.; Panza, E.; Bavestrello, G.; Cerrano, C.; Fattorusso, E.; Taglialatela-Scafati, O. Sinularioside, a triacetylated glycolipid from the Indonesian soft coral Sinularia sp., is an inhibitor of NO release. Bioorg. Med. Chem. Lett. 2012, 22, 2723–2725. [Google Scholar] [CrossRef] [PubMed]
  157. Lin, C.Y.; Lu, M.C.; Su, J.H.; Chu, C.L.; Shiuan, D.; Weng, C.F.; Sung, P.J.; Huang, K.J. Immunomodulatory effect of marine cembrane-type diterpenoids on dendritic cells. Mar. Drugs 2013, 11, 1336–1350. [Google Scholar] [CrossRef] [PubMed]
  158. Chen, Z.; Zheng, Z.; Huang, H.; Song, Y.; Zhang, X.; Ma, J.; Wang, B.; Zhang, C.; Ju, J. Penicacids A–C, three new mycophenolic acid derivatives and immunosuppressive activities from the marine-derived fungus Penicillium sp. SOF07. Bioorg. Med. Chem. Lett. 2012, 22, 3332–3335. [Google Scholar] [CrossRef] [PubMed]
  159. Jensen, J.E.; Mobli, M.; Brust, A.; Alewood, P.F.; King, G.F.; Rash, L.D. Cyclisation increases the stability of the sea anemone peptide APETx2 but decreases its activity at acid-sensing ion channel 3. Mar. Drugs 2012, 10, 1511–1527. [Google Scholar] [CrossRef] [PubMed]
  160. Li, H.; Bowling, J.J.; Fronczek, F.R.; Hong, J.; Jabba, S.V.; Murray, T.F.; Ha, N.C.; Hamann, M.T.; Jung, J.H. Asteropsin A: An unusual cystine-crosslinked peptide from porifera enhances neuronal Ca2+ influx. Biochim. Biophys. Acta 2013, 1830, 2591–2599. [Google Scholar] [CrossRef] [PubMed]
  161. Orts, D.J.; Peigneur, S.; Madio, B.; Cassoli, J.S.; Montandon, G.G.; Pimenta, A.M.; Bicudo, J.E.; Freitas, J.C.; Zaharenko, A.J.; Tytgat, J. Biochemical and electrophysiological characterization of two sea anemone type 1 potassium toxins from a geographically distant population of Bunodosoma caissarum. Mar. Drugs 2013, 11, 655–679. [Google Scholar] [CrossRef] [PubMed]
  162. Favreau, P.; Benoit, E.; Hocking, H.G.; Carlier, L.; D’ hoedt, D.; Leipold, E.; Markgraf, R.; Schlumberger, S.; Cordova, M.A.; Gaertner, H.; et al. A novel micro-conopeptide, CnIIIC, exerts potent and preferential inhibition of NaV1.2/1.4 channels and blocks neuronal nicotinic acetylcholine receptors. Br. J. Pharmacol. 2012, 166, 1654–1668. [Google Scholar] [CrossRef] [PubMed]
  163. Vetter, I.; Dekan, Z.; Knapp, O.; Adams, D.J.; Alewood, P.F.; Lewis, R.J. Isolation, characterization and total regioselective synthesis of the novel muO-conotoxin MfVIA from Conus magnificus that targets voltage-gated sodium channels. Biochem. Pharmacol. 2012, 84, 540–548. [Google Scholar] [CrossRef] [PubMed]
  164. Franco, A.; Kompella, S.N.; Akondi, K.B.; Melaun, C.; Daly, N.L.; Luetje, C.W.; Alewood, P.F.; Craik, D.J.; Adams, D.J.; Mari, F. RegIIA: An α4/7-conotoxin from the venom of Conus regius that potently blocks α3β4 nAChRs. Biochem. Pharmacol. 2012, 83, 419–426. [Google Scholar] [CrossRef] [PubMed]
  165. Bernaldez, J.; Roman-Gonzalez, S.A.; Martinez, O.; Jimenez, S.; Vivas, O.; Arenas, I.; Corzo, G.; Arreguin, R.; Garcia, D.E.; Possani, L.D.; et al. A Conus regularis conotoxin with a novel eight-cysteine framework inhibits CaV2.2 channels and displays an anti-nociceptive activity. Mar. Drugs 2013, 11, 1188–1202. [Google Scholar] [CrossRef] [PubMed]
  166. Figueiredo, G.S.; Zardo, R.S.; Silva, B.V.; Violante, F.A.; Pinto, A.C.; Fernandes, P.D. Convolutamydine A and synthetic analogues have antinociceptive properties in mice. Pharmacol. Biochem. Behav. 2013, 103, 431–439. [Google Scholar] [CrossRef] [PubMed]
  167. Andreev, Y.A.; Kozlov, S.A.; Korolkova, Y.V.; Dyachenko, I.A.; Bondarenko, D.A.; Skobtsov, D.I.; Murashev, A.N.; Kotova, P.D.; Rogachevskaja, O.A.; Kabanova, N.V.; et al. Polypeptide modulators of TRPV1 produce analgesia without hyperthermia. Mar. Drugs 2013, 11, 5100–5115. [Google Scholar] [CrossRef] [PubMed]
  168. Feng, Y.; Bowden, B.F.; Kapoor, V. Ianthellamide A, a selective kynurenine-3-hydroxylase inhibitor from the Australian marine sponge Ianthella quadrangulata. Bioorg. Med. Chem. Lett. 2012, 22, 3398–3401. [Google Scholar] [CrossRef] [PubMed]
  169. Burgy, G.; Tahtouh, T.; Durieu, E.; Foll-Josselin, B.; Limanton, E.; Meijer, L.; Carreaux, F.; Bazureau, J.P. Chemical synthesis and biological validation of immobilized protein kinase inhibitory Leucettines. Eur. J. Med. Chem. 2013, 62, 728–737. [Google Scholar] [CrossRef] [PubMed]
  170. Guzii, A.G.; Makarieva, T.N.; Korolkova, Y.V.; Andreev, Y.A.; Mosharova, I.V.; Tabakmaher, K.M.; Denisenko, V.A.; Dmitrenok, P.S.; Ogurtsova, E.L.; Antonov, A.S.; et al. Pulchranin A, isolated from the Far-Eastern marine sponge, Monanchora pulchra: The first marine non-peptide inhibitor of TRPV-1 channels. Tetrahedron Lett. 2013, 54, 1247–1250. [Google Scholar] [CrossRef]
  171. Montaser, R.; Paul, V.J.; Luesch, H. Marine cyanobacterial fatty acid amides acting on cannabinoid receptors. Chembiochem 2012, 13, 2676–2681. [Google Scholar] [CrossRef] [PubMed]
  172. Huang, X.; Sun, X.; Ding, B.; Lin, M.; Liu, L.; Huang, H.; She, Z. A new anti-acetylcholinesterase alpha-pyrone meroterpene, arigsugacin I, from mangrove endophytic fungus Penicillium sp. sk5GW1L of Kandelia candel. Planta Med. 2013, 79, 1572–1575. [Google Scholar] [PubMed]
  173. Xiao, Z.; Huang, H.; Shao, C.; Xia, X.; Ma, L.; Huang, X.; Lu, Y.; Lin, Y.; Long, Y.; She, Z. Asperterpenols A and B, new sesterterpenoids isolated from a mangrove endophytic fungus Aspergillus sp. 085242. Org. Lett. 2013, 15, 2522–2525. [Google Scholar] [CrossRef] [PubMed]
  174. Choi, H.; Proteau, P.J.; Byrum, T.; Pereira, A.R.; Gerwick, W.H. Cymatherelactone and cymatherols A–C, polycyclic oxylipins from the marine brown alga Cymathere triplicata. Phytochemistry 2012, 73, 134–141. [Google Scholar] [CrossRef] [PubMed]
  175. Zhang, H.; Conte, M.M.; Khalil, Z.; Huang, X.C.; Capon, R.J. New dictyodendrins as BACE inhibitors from a southern Australian marine sponge, Ianthella sp. RSC Adv. 2012, 2, 4209–4214. [Google Scholar] [CrossRef]
  176. Ohlendorf, B.; Schulz, D.; Erhard, A.; Nagel, K.; Imhoff, J.F. Geranylphenazinediol, an acetylcholinesterase inhibitor produced by a Streptomyces species. J. Nat. Prod. 2012, 75, 1400–1404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Wyche, T.P.; Standiford, M.; Hou, Y.; Braun, D.; Johnson, D.A.; Johnson, J.A.; Bugni, T.S. Activation of the nuclear factor e2-related factor 2 pathway by novel natural products halomadurones a–d and a synthetic analogue. Mar. Drugs 2013, 11, 5089–5099. [Google Scholar] [CrossRef] [PubMed]
  178. Balansa, W.; Islam, R.; Fontaine, F.; Piggott, A.M.; Zhang, H.; Xiao, X.; Webb, T.I.; Gilbert, D.F.; Lynch, J.W.; Capon, R.J. Sesterterpene glycinyl-lactams: A new class of glycine receptor modulator from Australian marine sponges of the genus Psammocinia. Org. Biomol. Chem. 2013, 11, 4695–4701. [Google Scholar] [CrossRef] [PubMed]
  179. Palyanova, N.V.; Pankova, T.M.; Starostina, M.V.; Kicha, A.A.; Ivanchina, N.V.; Stonik, V.A. Neuritogenic and neuroprotective effects of polar steroids from the Far East starfishes Patiria pectinifera and Distolasterias nipon. Mar. Drugs 2013, 11, 1440–1455. [Google Scholar] [CrossRef] [PubMed]
  180. Mayer, A.M.S.; Murphy, J.; Macadam, D.; Osterbauer, C.; Baseer, I.; Hall, M.L.; Feher, D.; Williams, P. Classical and Alternative Activation of Cyanobacterium Oscillatoria sp. Lipopolysaccharide-Treated Rat Microglia in vitro. Toxicol. Sci. 2016, 149, 484–495. [Google Scholar] [CrossRef] [PubMed]
  181. Urbarova, I.; Karlsen, B.O.; Okkenhaug, S.; Seternes, O.M.; Johansen, S.D.; Emblem, A. Digital marine bioprospecting: Mining new neurotoxin drug candidates from the transcriptomes of cold-water sea anemones. Mar. Drugs 2012, 10, 2265–2279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Dona, G.; Kozuh, I.; Brunati, A.M.; Andrisani, A.; Ambrosini, G.; Bonanni, G.; Ragazzi, E.; Armanini, D.; Clari, G.; Bordin, L. Effect of astaxanthin on human sperm capacitation. Mar. Drugs 2013, 11, 1909–1919. [Google Scholar] [CrossRef] [PubMed]
  183. Yonezawa, T.; Mase, N.; Sasaki, H.; Teruya, T.; Hasegawa, S.; Cha, B.Y.; Yagasaki, K.; Suenaga, K.; Nagai, K.; Woo, J.T. Biselyngbyaside, isolated from marine cyanobacteria, inhibits osteoclastogenesis and induces apoptosis in mature osteoclasts. J. Cell. Biochem. 2012, 113, 440–448. [Google Scholar] [CrossRef] [PubMed]
  184. Shirouzu, T.; Watari, K.; Ono, M.; Koizumi, K.; Saiki, I.; Tanaka, C.; van Soest, R.W.; Miyamoto, T. Structure, synthesis, and biological activity of a C-20 bisacetylenic alcohol from a marine sponge Callyspongia sp. J. Nat. Prod. 2013, 76, 1337–1342. [Google Scholar] [CrossRef] [PubMed]
  185. Sepe, V.; Ummarino, R.; D’Auria, M.V.; Chini, M.G.; Bifulco, G.; Renga, B.; D’Amore, C.; Debitus, C.; Fiorucci, S.; Zampella, A. Conicasterol E, a small heterodimer partner sparing farnesoid X receptor modulator endowed with a pregnane X receptor agonistic activity, from the marine sponge Theonella swinhoei. J. Med. Chem. 2012, 55, 84–93. [Google Scholar] [CrossRef] [PubMed]
  186. Kim, G.D.; Cheong, O.J.; Bae, S.Y.; Shin, J.; Lee, S.K. 6′′-Debromohamacanthin A, a bis (indole) alkaloid, inhibits angiogenesis by targeting the VEGFR2-mediated PI3K/AKT/mTOR signaling pathways. Mar. Drugs 2013, 11, 1087–1103. [Google Scholar] [CrossRef] [PubMed]
  187. Kang, S.M.; Heo, S.J.; Kim, K.N.; Lee, S.H.; Yang, H.M.; Kim, A.D.; Jeon, Y.J. Molecular docking studies of a phlorotannin, dieckol isolated from Ecklonia cava with tyrosinase inhibitory activity. Bioorg. Med. Chem. 2012, 20, 311–316. [Google Scholar] [CrossRef] [PubMed]
  188. Sohn, J.H.; Lee, Y.R.; Lee, D.S.; Kim, Y.C.; Oh, H. PTP1B inhibitory secondary metabolites from marine-derived fungal strains Penicillium spp. and Eurotium sp. J. Microbiol. Biotechnol. 2013, 23, 1206–1211. [Google Scholar] [CrossRef] [PubMed]
  189. Cheung, F.W.; Guo, J.; Ling, Y.H.; Che, C.T.; Liu, W.K. Anti-melanogenic property of geoditin A in murine B16 melanoma cells. Mar. Drugs 2012, 10, 465–476. [Google Scholar] [CrossRef] [PubMed]
  190. Putra, M.Y.; Bavestrello, G.; Cerrano, C.; Renga, B.; D’Amore, C.; Fiorucci, S.; Fattorusso, E.; Taglialatela-Scafati, O. Polyhydroxylated sterols from the Indonesian soft coral Sinularia sp. and their effect on farnesoid X-activated receptor. Steroids 2012, 77, 433–440. [Google Scholar] [CrossRef] [PubMed]
  191. Festa, C.; Lauro, G.; De Marino, S.; D’Auria, M.V.; Monti, M.C.; Casapullo, A.; D’Amore, C.; Renga, B.; Mencarelli, A.; Petek, S.; et al. Plakilactones from the marine sponge Plakinastrella mamillaris. Discovery of a new class of marine ligands of peroxisome proliferator-activated receptor gamma. J. Med. Chem. 2012, 55, 8303–8317. [Google Scholar] [CrossRef] [PubMed]
  192. Festa, C.; D’Amore, C.; Renga, B.; Lauro, G.; De Marino, S.; D’Auria, M.V.; Bifulco, G.; Zampella, A.; Fiorucci, S. Oxygenated polyketides from Plakinastrella mamillaris as a new chemotype of PXR agonists. Mar. Drugs 2013, 11, 2314–2327. [Google Scholar] [CrossRef] [PubMed]
  193. Li, J.L.; Xiao, B.; Park, M.; Yoo, E.S.; Shin, S.; Hong, J.; Chung, H.Y.; Kim, H.S.; Jung, J.H. PPAR-gamma agonistic metabolites from the ascidian Herdmania momus. J. Nat. Prod. 2012, 75, 2082–2087. [Google Scholar] [CrossRef] [PubMed]
  194. Yamanokuchi, R.; Imada, K.; Miyazaki, M.; Kato, H.; Watanabe, T.; Fujimuro, M.; Saeki, Y.; Yoshinaga, S.; Terasawa, H.; Iwasaki, N.; et al. Hyrtioreticulins A–E, indole alkaloids inhibiting the ubiquitin-activating enzyme, from the marine sponge Hyrtios reticulatus. Bioorg. Med. Chem. 2012, 20, 4437–4442. [Google Scholar] [CrossRef] [PubMed]
  195. Gladkikh, I.; Monastyrnaya, M.; Leychenko, E.; Zelepuga, E.; Chausova, V.; Isaeva, M.; Anastyuk, S.; Andreev, Y.; Peigneur, S.; Tytgat, J.; et al. Atypical reactive center Kunitz-type inhibitor from the sea anemone Heteractis crispa. Mar. Drugs 2012, 10, 1545–1565. [Google Scholar] [CrossRef] [PubMed]
  196. Schweikart, K.; Guo, L.; Shuler, Z.; Abrams, R.; Chiao, E.T.; Kolaja, K.L.; Davis, M. The effects of jaspamide on human cardiomyocyte function and cardiac ion channel activity. Toxicol. In Vitro 2013, 27, 745–751. [Google Scholar] [CrossRef] [PubMed]
  197. Carlile, G.W.; Keyzers, R.A.; Teske, K.A.; Robert, R.; Williams, D.E.; Linington, R.G.; Gray, C.A.; Centko, R.M.; Yan, L.; Anjos, S.M.; et al. Correction of F508del-CFTR trafficking by the sponge alkaloid latonduine is modulated by interaction with PARP. Chem. Biol. 2012, 19, 1288–1299. [Google Scholar] [CrossRef] [PubMed]
  198. Eguchi, K.; Fujiwara, Y.; Hayashida, A.; Horlad, H.; Kato, H.; Rotinsulu, H.; Losung, F.; Mangindaan, R.E.; de Voogd, N.J.; Takeya, M.; et al. Manzamine A, a marine-derived alkaloid, inhibits accumulation of cholesterol ester in macrophages and suppresses hyperlipidemia and atherosclerosis in vivo. Bioorg. Med. Chem. 2013, 21, 3831–3838. [Google Scholar] [CrossRef] [PubMed]
  199. Williams, D.E.; Dalisay, D.S.; Li, F.; Amphlett, J.; Maneerat, W.; Chavez, M.A.; Wang, Y.A.; Matainaho, T.; Yu, W.; Brown, P.J.; et al. Nahuoic acid A produced by a Streptomyces sp. isolated from a marine sediment is a selective SAM-competitive inhibitor of the histone methyltransferase SETD8. Org. Lett. 2013, 15, 414–417. [Google Scholar] [CrossRef] [PubMed]
  200. Cheruku, P.; Plaza, A.; Lauro, G.; Keffer, J.; Lloyd, J.R.; Bifulco, G.; Bewley, C.A. Discovery and synthesis of namalide reveals a new anabaenopeptin scaffold and peptidase inhibitor. J. Med. Chem. 2012, 55, 735–742. [Google Scholar] [CrossRef] [PubMed]
  201. Plisson, F.; Conte, M.; Khalil, Z.; Huang, X.C.; Piggott, A.M.; Capon, R.J. Kinase inhibitor scaffolds against neurodegenerative diseases from a Southern Australian ascidian, Didemnum sp. ChemMedChem 2012, 7, 983–990. [Google Scholar] [CrossRef] [PubMed]
  202. Margarucci, L.; Tosco, A.; De Simone, R.; Riccio, R.; Monti, M.C.; Casapullo, A. Modulation of proteasome machinery by natural and synthetic analogues of the marine bioactive compound petrosaspongiolide M. Chembiochem 2012, 13, 982–986. [Google Scholar] [CrossRef] [PubMed]
  203. Choi, B.K.; Cha, B.Y.; Yagyu, T.; Woo, J.T.; Ojika, M. Sponge-derived acetylenic alcohols, petrosiols, inhibit proliferation and migration of platelet-derived growth factor (PDGF)-induced vascular smooth muscle cells. Bioorg. Med. Chem. 2013, 21, 1804–1810. [Google Scholar] [CrossRef] [PubMed]
  204. Vitale, R.M.; Gatti, M.; Carbone, M.; Barbieri, F.; Felicita, V.; Gavagnin, M.; Florio, T.; Amodeo, P. Minimalist hybrid ligand/receptor-based pharmacophore model for CXCR4 applied to a small-library of marine natural products led to the identification of phidianidine a as a new CXCR4 ligand exhibiting antagonist activity. ACS Chem. Biol. 2013, 8, 2762–2770. [Google Scholar] [CrossRef] [PubMed]
  205. Lunder, M.; Drevensek, G.; Hawlina, S.; Sepcic, K.; Ziberna, L. Cardiovascular effects induced by polymeric 3-alkylpyridinium salts from the marine sponge Reniera sarai. Toxicon 2012, 60, 1041–1048. [Google Scholar] [CrossRef] [PubMed]
  206. Su, J.H.; Chen, Y.C.; El-Shazly, M.; Du, Y.C.; Su, C.W.; Tsao, C.W.; Liu, L.L.; Chou, Y.; Chang, W.B.; Su, Y.D.; et al. Towards the small and the beautiful: A small dibromotyrosine derivative from Pseudoceratina sp. sponge exhibits potent apoptotic effect through targeting IKK/NFkappaB signaling pathway. Mar. Drugs 2013, 11, 3168–3185. [Google Scholar] [CrossRef] [PubMed]
  207. Day, D.R.; Jabaiah, S.; Jacobs, R.S.; Little, R.D. Cyclodextrin formulation of the marine natural product pseudopterosin A uncovers optimal pharmacodynamics in proliferation studies of human umbilical vein endothelial cells. Mar. Drugs 2013, 11, 3258–3271. [Google Scholar] [CrossRef] [PubMed]
  208. Yoon, W.J.; Kim, K.N.; Heo, S.J.; Han, S.C.; Kim, J.; Ko, Y.J.; Kang, H.K.; Yoo, E.S. Sargachromanol G inhibits osteoclastogenesis by suppressing the activation NF-kappaB and MAPKs in RANKL-induced RAW 264.7 cells. Biochem. Biophys. Res. Commun. 2013, 434, 892–897. [Google Scholar] [CrossRef] [PubMed]
  209. Zhang, C.Y.; Kong, T.; Wu, W.H.; Lan, M.B. The protection of polysaccharide from the Brown Seaweed Sargassum graminifolium against ethylene glycol-induced mitochondrial damage. Mar. Drugs 2013, 11, 870–880. [Google Scholar] [CrossRef] [PubMed]
  210. Kawamura-Konishi, Y.; Watanabe, N.; Saito, M.; Nakajima, N.; Sakaki, T.; Katayama, T.; Enomoto, T. Isolation of a new phlorotannin, a potent inhibitor of carbohydrate-hydrolyzing enzymes, from the brown alga Sargassum patens. J. Agric. Food Chem. 2012, 60, 5565–5570. [Google Scholar] [CrossRef] [PubMed]
  211. Xu, M.J.; Liu, X.J.; Zhao, Y.L.; Liu, D.; Xu, Z.H.; Lang, X.M.; Ao, P.; Lin, W.H.; Yang, S.L.; Zhang, Z.G.; et al. Identification and characterization of an anti-fibrotic benzopyran compound isolated from mangrove-derived Streptomyces xiamenensis. Mar. Drugs 2012, 10, 639–654. [Google Scholar] [CrossRef] [PubMed]
  212. Sepe, V.; Ummarino, R.; D’Auria, M.V.; Taglialatela-Scafati, O.; Marino, S.D.; D’Amore, C.; Renga, B.; Chini, M.G.; Bifulco, G.; Nakao, Y.; et al. Preliminary structure-activity relationship on theonellasterol, a new chemotype of FXR antagonist, from the marine sponge Theonella swinhoei. Mar. Drugs 2012, 10, 2448–2466. [Google Scholar] [CrossRef] [PubMed]
  213. Garcia-Caballero, M.; Mari-Beffa, M.; Canedo, L.; Medina, M.A.; Quesada, A.R. Toluquinol, a marine fungus metabolite, is a new angiosuppresor that interferes with the Akt pathway. Biochem. Pharmacol. 2013, 85, 1727–1740. [Google Scholar] [CrossRef] [PubMed]
  214. Wang, R.; Paul, V.J.; Luesch, H. Seaweed extracts and unsaturated fatty acid constituents from the green alga Ulva lactuca as activators of the cytoprotective Nrf2-ARE pathway. Free Radic. Biol. Med. 2013, 57, 141–153. [Google Scholar] [CrossRef] [PubMed]
  215. Daoust, J.; Chen, M.; Wang, M.; Williams, D.E.; Garcia Chavez, M.A.; Wang, Y.A.; Merchant, C.E.; Fontana, A.; Kieffer, T.J.; Andersen, R.J. Sesterterpenoids isolated from a northeastern Pacific Phorbas sp. J. Org. Chem. 2013, 78, 8267–8273. [Google Scholar] [CrossRef] [PubMed]
  216. Sun, S.W.; Ji, C.Z.; Gu, Q.Q.; Li, D.H.; Zhu, T.J. Three new polyketides from marine-derived fungus Aspergillus glaucus HB1-19. J. Asian Nat. Prod. Res. 2013, 15, 956–961. [Google Scholar] [CrossRef] [PubMed]
  217. Haroon, M.H.; Premaratne, S.R.; Choudhry, M.I.; Dharmaratne, H.R. A new beta-glucuronidase inhibiting butyrolactone from the marine endophytic fungus Aspergillus terreus. Nat. Prod. Res. 2013, 27, 1060–1066. [Google Scholar] [CrossRef] [PubMed]
  218. Cavalcante-Silva, L.H.; de Carvalho Correia, A.C.; Barbosa-Filho, J.M.; da Silva, B.A.; de Oliveira Santos, B.V.; de Lira, D.P.; Sousa, J.C.; de Miranda, G.E.; de Andrade, C.F.; Alexandre-Moreira, M.S. Spasmolytic effect of caulerpine involves blockade of Ca2+ influx on guinea pig ileum. Mar. Drugs 2013, 11, 1553–1564. [Google Scholar] [CrossRef] [PubMed]
  219. Chini, M.G.; Jones, C.R.; Zampella, A.; D’Auria, M.V.; Renga, B.; Fiorucci, S.; Butts, C.P.; Bifulco, G. Quantitative NMR-derived interproton distances combined with quantum mechanical calculations of 13C chemical shifts in the stereochemical determination of conicasterol F, a nuclear receptor ligand from Theonella swinhoei. J. Org. Chem. 2012, 77, 1489–1496. [Google Scholar] [CrossRef] [PubMed]
  220. Hamed, A.N.; Watjen, W.; Schmitz, R.; Chovolou, Y.; Edrada-Ebel, R.; Youssef, D.T.; Kamel, M.S.; Proksch, P. A new bioactive sesquiterpenoid quinone from the Mediterranean Sea marine sponge Dysidea avara. Nat. Prod. Commun. 2013, 8, 289–292. [Google Scholar] [PubMed]
  221. Shin, K.; Chin, J.; Hahn, D.; Lee, J.; Hwang, H.; Won, D.H.; Ham, J.; Choi, H.; Kang, E.; Kim, H.; et al. Sterols from a soft coral, Dendronephthya gigantea as farnesoid X-activated receptor antagonists. Steroids 2012, 77, 355–359. [Google Scholar] [CrossRef] [PubMed]
  222. Jiao, W.H.; Huang, X.J.; Yang, J.S.; Yang, F.; Piao, S.J.; Gao, H.; Li, J.; Ye, W.C.; Yao, X.S.; Chen, W.S.; et al. Dysidavarones A–D, new sesquiterpene quinones from the marine sponge Dysidea avara. Org. Lett. 2012, 14, 202–205. [Google Scholar] [CrossRef] [PubMed]
  223. Hu, Y.; Martinez, E.D.; MacMillan, J.B. Anthraquinones from a marine-derived Streptomyces spinoverrucosus. J. Nat. Prod. 2012, 75, 1759–1764. [Google Scholar] [CrossRef] [PubMed]
  224. Williams, D.E.; Steino, A.; de Voogd, N.J.; Mauk, A.G.; Andersen, R.J. Halicloic acids A and B isolated from the marine sponge Haliclona sp. collected in the Philippines inhibit indoleamine 2,3-dioxygenase. J. Nat. Prod. 2012, 75, 1451–1458. [Google Scholar] [CrossRef] [PubMed]
  225. Jansen, N.; Ohlendorf, B.; Erhard, A.; Bruhn, T.; Bringmann, G.; Imhoff, J.F. Helicusin E, isochromophilone X and isochromophilone XI: New chloroazaphilones produced by the fungus Bartalinia robillardoides strain LF550. Mar. Drugs 2013, 11, 800–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Chianese, G.; Fattorusso, E.; Putra, M.Y.; Calcinai, B.; Bavestrello, G.; Moriello, A.S.; De Petrocellis, L.; Di Marzo, V.; Taglialatela-Scafati, O. Leucettamols, bifunctionalized marine sphingoids, act as modulators of TRPA1 and TRPM8 channels. Mar. Drugs 2012, 10, 2435–2447. [Google Scholar] [CrossRef] [PubMed]
  227. Ushiyama, S.; Umaoka, H.; Kato, H.; Suwa, Y.; Morioka, H.; Rotinsulu, H.; Losung, F.; Mangindaan, R.E.; de Voogd, N.J.; Yokosawa, H.; et al. Manadosterols A and B, sulfonated sterol dimers inhibiting the Ubc13-Uev1A interaction, isolated from the marine sponge Lissodendryx fibrosa. J. Nat. Prod. 2012, 75, 1495–1499. [Google Scholar] [CrossRef] [PubMed]
  228. Almeida, C.; Hemberger, Y.; Schmitt, S.M.; Bouhired, S.; Natesan, L.; Kehraus, S.; Dimas, K.; Gutschow, M.; Bringmann, G.; Konig, G.M. Marilines A–C: Novel phthalimidines from the sponge-derived fungus Stachylidium sp. Chemistry 2012, 18, 8827–8834. [Google Scholar] [CrossRef] [PubMed]
  229. Liang, L.F.; Kurtan, T.; Mandi, A.; Yao, L.G.; Li, J.; Zhang, W.; Guo, Y.W. Unprecedented diterpenoids as a PTP1B inhibitor from the Hainan soft coral Sarcophyton trocheliophorum Marenzeller. Org. Lett. 2013, 15, 274–277. [Google Scholar] [CrossRef] [PubMed]
  230. Kawahara, T.; Takagi, M.; Shin-Ya, K. JBIR-124: A novel antioxidative agent from a marine sponge-derived fungus Penicillium citrinum SpI080624G1f01. J. Antibiot. 2012, 65, 45–47. [Google Scholar] [CrossRef] [PubMed]
  231. Kim, H.; Chin, J.; Choi, H.; Baek, K.; Lee, T.G.; Park, S.E.; Wang, W.; Hahn, D.; Yang, I.; Lee, J.; et al. Phosphoiodyns A and B, unique phosphorus-containing iodinated polyacetylenes from a Korean sponge Placospongia sp. Org. Lett. 2013, 15, 100–103. [Google Scholar] [CrossRef] [PubMed]
  232. Shen, S.; Liu, D.; Wei, C.; Proksch, P.; Lin, W. Purpuroines A–J, halogenated alkaloids from the sponge Iotrochota purpurea with antibiotic activity and regulation of tyrosine kinases. Bioorg. Med. Chem. 2012, 20, 6924–6928. [Google Scholar] [CrossRef] [PubMed]
  233. Pavlik, C.M.; Wong, C.Y.; Ononye, S.; Lopez, D.D.; Engene, N.; McPhail, K.L.; Gerwick, W.H.; Balunas, M.J. Santacruzamate A, a potent and selective histone deacetylase inhibitor from the Panamanian marine cyanobacterium cf. Symploca sp. J. Nat. Prod. 2013, 76, 2026–2033. [Google Scholar] [CrossRef] [PubMed]
  234. Liang, L.F.; Gao, L.X.; Li, J.; Taglialatela-Scafati, O.; Guo, Y.W. Cembrane diterpenoids from the soft coral Sarcophyton trocheliophorum Marenzeller as a new class of PTP1B inhibitors. Bioorg. Med. Chem. 2013, 21, 5076–5080. [Google Scholar] [CrossRef] [PubMed]
  235. Kim, C.; Lee, I.K.; Cho, G.Y.; Oh, K.H.; Lim, Y.W.; Yun, B.S. Sargassumol, a novel antioxidant from the brown alga Sargassum micracanthum. J. Antibiot. 2012, 65, 87–89. [Google Scholar] [CrossRef] [PubMed]
  236. Niemann, H.; Lin, W.; Muller, W.E.; Kubbutat, M.; Lai, D.; Proksch, P. Trimeric hemibastadin congener from the marine sponge Ianthella basta. J. Nat. Prod. 2013, 76, 121–125. [Google Scholar] [CrossRef] [PubMed]
  237. Hegazy, M.E.; Gamal Eldeen, A.M.; Shahat, A.A.; Abdel-Latif, F.F.; Mohamed, T.A.; Whittlesey, B.R.; Pare, P.W. Bioactive hydroperoxyl cembranoids from the Red Sea soft coral Sarcophyton glaucum. Mar. Drugs 2012, 10, 209–222. [Google Scholar] [CrossRef] [PubMed]
  238. Molinski, T.F.; Reynolds, K.A.; Morinaka, B.I. Symplocin A, a linear peptide from the Bahamian cyanobacterium Symploca sp. Configurational analysis of N,N-dimethylamino acids by chiral-phase HPLC of naphthacyl esters. J. Nat. Prod. 2012, 75, 425–431. [Google Scholar] [CrossRef] [PubMed]
  239. Dolusic, E.; Larrieu, P.; Meinguet, C.; Colette, D.; Rives, A.; Blanc, S.; Ferain, T.; Pilotte, L.; Stroobant, V.; Wouters, J.; et al. Indoleamine 2,3-dioxygenase inhibitory activity of derivatives of marine alkaloid tsitsikammamine A. Bioorg. Med. Chem. Lett. 2013, 23, 47–54. [Google Scholar] [CrossRef] [PubMed]
  240. Olsen, E.K.; Hansen, E.; Isaksson, J.; Andersen, J.H. Cellular Antioxidant Effect of Four Bromophenols from the Red Algae, Vertebrata lanosa. Mar. Drugs 2013, 11, 2769–2784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Matsunaga, S.; Akiyama, T.; Takada, K.; Oikawa, T.; Matsuura, N.; Ise, Y.; Okada, S.; Matsunaga, S. Stimulators of adipogenesis from the marine sponge Xestospongia testudinaria. Tetrahedron 2013, 69, 6560–6564. [Google Scholar]
  242. Dong, L.Y.; Jin, J.; Lu, G.; Kang, X.L. Astaxanthin attenuates the apoptosis of retinal ganglion cells in db/db mice by inhibition of oxidative stress. Mar. Drugs 2013, 11, 960–974. [Google Scholar] [CrossRef] [PubMed]
  243. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2012, 29, 144–222. [Google Scholar] [CrossRef] [PubMed]
  244. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2013, 30, 237–323. [Google Scholar] [CrossRef] [PubMed]
  245. Cragg, G.M.; Newman, D.J. Natural products: A continuing source of novel drug leads. Biochim. Biophys. Acta 2013, 1830, 3670–3695. [Google Scholar] [CrossRef] [PubMed]
  246. Camp, D.; Davis, R.A.; Evans-Illidge, E.A.; Quinn, R.J. Guiding principles for natural product drug discovery. Future Med. Chem. 2012, 4, 1067–1084. [Google Scholar] [CrossRef] [PubMed]
  247. Jensen, P.R.; Chavarria, K.L.; Fenical, W.; Moore, B.S.; Ziemert, N. Challenges and triumphs to genomics-based natural product discovery. J. Ind. Microbiol Biotechnol. 2013, 41, 203–209. [Google Scholar] [CrossRef] [PubMed]
  248. Gerwick, W.H.; Moore, B.S. Lessons from the past and charting the future of marine natural products drug discovery and chemical biology. Chem. Biol. 2012, 19, 85–98. [Google Scholar] [CrossRef] [PubMed]
  249. Liu, J.T.; Lu, X.L.; Liu, X.Y.; Gao, Y.; Hu, B.; Jiao, B.H.; Zheng, H. Bioactive natural products from the antarctic and arctic organisms. Mini Rev. Med. Chem. 2013, 13, 617–626. [Google Scholar] [CrossRef] [PubMed]
  250. Liang, L.F.; Guo, Y.W. Terpenes from the soft corals of the genus Sarcophyton: Chemistry and biological activities. Chem. Biodivers. 2013, 10, 2161–2196. [Google Scholar] [CrossRef] [PubMed]
  251. Harnedy, P.A.; FitzGerald, R.J. Bioactive peptides from marine processing waste and shellfish: A review. J. Funct. Foods 2013, 4, 6–24. [Google Scholar] [CrossRef]
  252. Berrue, F.; McCulloch, M.W.; Kerr, R.G. Marine diterpene glycosides. Bioorg. Med. Chem. 2011, 19, 6702–6719. [Google Scholar] [CrossRef] [PubMed]
  253. Morya, V.K.; Kim, J.; Kim, E.K. Algal fucoidan: Structural and size-dependent bioactivities and their perspectives. Appl. Microbiol. Biotechnol. 2012, 93, 71–82. [Google Scholar] [CrossRef] [PubMed]
  254. Pomin, V.H. Fucanomics and galactanomics: Marine distribution, medicinal impact, conceptions, and challenges. Mar. Drugs 2012, 10, 793–811. [Google Scholar] [CrossRef] [PubMed]
  255. Gouveia, V.; Seca, A.M.; Barreto, M.C.; Pinto, D.C. Di- and sesquiterpenoids from Cystoseira genus: Structure, intra-molecular transformations and biological activity. Mini Rev. Med. Chem. 2013, 13, 1150–1159. [Google Scholar] [CrossRef] [PubMed]
  256. Ngo, D.H.; Kim, S.K. Sulfated polysaccharides as bioactive agents from marine algae. Int. J. Biol. Macromol. 2013, 62, 70–75. [Google Scholar] [CrossRef] [PubMed]
  257. Wang, B.G.; Gloer, J.B.; Ji, N.Y.; Zhao, J.C. Halogenated organic molecules of Rhodomelaceae origin: Chemistry and biology. Chem. Rev. 2013, 113, 3632–3685. [Google Scholar] [CrossRef] [PubMed]
  258. Nagarajan, M.; Maruthanayagam, V.; Sundararaman, M. A review of pharmacological and toxicological potentials of marine cyanobacterial metabolites. J. Appl. Toxicol. 2012, 32, 153–185. [Google Scholar] [CrossRef] [PubMed]
  259. Tan, L.T. Pharmaceutical agents from filamentous marine cyanobacteria. Drug Discov. Today 2013, 18, 863–871. [Google Scholar] [CrossRef] [PubMed]
  260. Kalinin, V.I.; Ivanchina, N.V.; Krasokhin, V.B.; Makarieva, T.N.; Stonik, V.A. Glycosides from marine sponges (Porifera, Demospongiae): Structures, taxonomical distribution, biological activities and biological roles. Mar. Drugs 2012, 10, 1671–1710. [Google Scholar] [CrossRef] [PubMed]
  261. Kiew, P.L.; Don, M.M. Jewel of the seabed: Sea cucumbers as nutritional and drug candidates. Int. J. Food Sci. Nutr. 2011, 63, 616–636. [Google Scholar] [CrossRef] [PubMed]
  262. Gallardo-Rodriguez, J.; Sanchez-Miron, A.; Garcia-Camacho, F.; Lopez-Rosales, L.; Chisti, Y.; Molina-Grima, E. Bioactives from microalgal dinoflagellates. Biotechnol. Adv. 2012, 30, 1673–1684. [Google Scholar] [CrossRef] [PubMed]
  263. Desriac, F.; Jegou, C.; Balnois, E.; Brillet, B.; Le, C.P.; Fleury, Y. Antimicrobial peptides from marine proteobacteria. Mar. Drugs 2013, 11, 3632–3660. [Google Scholar] [CrossRef] [PubMed]
  264. Eom, S.H.; Kim, Y.M.; Kim, S.K. Marine bacteria: Potential sources for compounds to overcome antibiotic resistance. Appl. Microbiol. Biotechnol. 2013, 97, 4763–4773. [Google Scholar] [CrossRef] [PubMed]
  265. Gao, C.; Yi, X.; Huang, R.; Yan, F.; He, B.; Chen, B. Alkaloids from corals. Chem. Biodivers. 2013, 10, 1435–1447. [Google Scholar] [CrossRef] [PubMed]
  266. Hoang, V.L.; Kim, S.K. Antimicrobial peptides from marine sources. Curr. Protein Pept. Sci. 2013, 14, 205–211. [Google Scholar] [CrossRef] [PubMed]
  267. Zotchev, S.B. Marine actinomycetes as an emerging resource for the drug development pipelines. J. Biotechnol. 2012, 158, 168–175. [Google Scholar] [CrossRef] [PubMed]
  268. Mondol, M.A.; Shin, H.J.; Islam, M.T. Diversity of secondary metabolites from marine Bacillus species: Chemistry and biological activity. Mar. Drugs 2013, 11, 2846–2872. [Google Scholar] [CrossRef] [PubMed]
  269. Solov’eva, T.; Davydova, V.; Krasikova, I.; Yermak, I. Marine compounds with therapeutic potential in gram-negative sepsis. Mar. Drugs 2013, 11, 2216–2229. [Google Scholar] [CrossRef] [PubMed]
  270. Sugumaran, M.; Robinson, W.E. Structure, biosynthesis and possible function of tunichromes and related compounds. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2012, 163, 1–25. [Google Scholar] [CrossRef] [PubMed]
  271. Gerwick, W.H.; Fenner, A.M. Drug discovery from marine microbes. Microb. Ecol. 2013, 65, 800–806. [Google Scholar] [CrossRef] [PubMed]
  272. Zhou, X.; Liu, J.; Yang, B.; Lin, X.; Yang, X.W.; Liu, Y. Marine natural products with anti-HIV activities in the last decade. Curr. Med. Chem. 2013, 20, 953–973. [Google Scholar] [PubMed]
  273. Prokofjeva, M.M.; Imbs, T.I.; Shevchenko, N.M.; Spirin, P.V.; Horn, S.; Fehse, B.; Zvyagintseva, T.N.; Prassolov, V.S. Fucoidans as potential inhibitors of HIV-1. Mar. Drugs 2013, 11, 3000–3014. [Google Scholar] [CrossRef] [PubMed]
  274. Raveh, A.; Delekta, P.C.; Dobry, C.J.; Peng, W.; Schultz, P.J.; Blakely, P.K.; Tai, A.W.; Matainaho, T.; Irani, D.N.; Sherman, D.H.; et al. Discovery of potent broad spectrum antivirals derived from marine actinobacteria. PLoS ONE 2013, 8, e82318. [Google Scholar] [CrossRef] [PubMed]
  275. Huskens, D.; Schols, D. Algal lectins as potential HIV microbicide candidates. Mar. Drugs 2012, 10, 1476–1497. [Google Scholar] [CrossRef] [PubMed]
  276. Jones, A.J.; Grkovic, T.; Sykes, M.L.; Avery, V.M. Trypanocidal activity of marine natural products. Mar. Drugs 2013, 11, 4058–4082. [Google Scholar] [CrossRef] [PubMed]
  277. Saeidnia, S.; Gohari, A.R.; Haddadi, A. Biogenic trypanocidal sesquiterpenes: Lead compounds to design future trypanocidal drugs—A mini review. J. Fac. Pharm. Tehran Univ. Med. Sci. 2013, 21, 35. [Google Scholar] [CrossRef] [PubMed]
  278. Bladt, T.T.; Frisvad, J.C.; Knudsen, P.B.; Larsen, T.O. Anticancer and antifungal compounds from Aspergillus, Penicillium and other filamentous fungi. Molecules 2013, 18, 11338–11376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Ruan, B.F.; Zhu, H.L. The chemistry and biology of the bryostatins: Potential PKC inhibitors in clinical development. Curr. Med. Chem. 2012, 19, 2652–2664. [Google Scholar] [CrossRef] [PubMed]
  280. Kim, J.A.; Kim, S.K. Bioactive peptides from marine sources as potential anti-inflammatory therapeutics. Curr. Protein Pept. Sci. 2013, 14, 177–182. [Google Scholar] [CrossRef] [PubMed]
  281. Wei, W.C.; Sung, P.J.; Duh, C.Y.; Chen, B.W.; Sheu, J.H.; Yang, N.S. Anti-inflammatory activities of natural products isolated from soft corals of Taiwan between 2008 and 2012. Mar. Drugs 2013, 11, 4083–4126. [Google Scholar] [CrossRef] [PubMed]
  282. D’Orazio, N.; Gammone, M.A.; Gemello, E.; De, G.M.; Cusenza, S.; Riccioni, G. Marine bioactives: Pharmacological properties and potential applications against inflammatory diseases. Mar. Drugs 2012, 10, 812–833. [Google Scholar] [CrossRef] [PubMed]
  283. Balboa, E.M.; Conde, E.; Moure, A.; Falque, E.; Dominguez, H. In vitro antioxidant properties of crude extracts and compounds from brown algae. Food Chem. 2013, 138, 1764–1785. [Google Scholar] [CrossRef] [PubMed]
  284. Albert, B.B.; Cameron-Smith, D.; Hofman, P.L.; Cutfield, W.S. Oxidation of marine omega-3 supplements and human health. Biomed. Res. Int. 2013, 2013, 464921. [Google Scholar] [CrossRef] [PubMed]
  285. Ko, S.C.; Jeon, Y.J. Marine peptides for preventing metabolic syndrome. Curr. Protein Pept. Sci. 2013, 14, 183–188. [Google Scholar] [CrossRef] [PubMed]
  286. Lee, S.H.; Jeon, Y.J. Anti-diabetic effects of brown algae derived phlorotannins, marine polyphenols through diverse mechanisms. Fitoterapia 2013, 86, 129–136. [Google Scholar] [CrossRef] [PubMed]
  287. Ngo, D.H.; Kim, S.K. Marine bioactive peptides as potential antioxidants. Curr. Protein Pept. Sci. 2013, 14, 189–198. [Google Scholar] [CrossRef] [PubMed]
  288. Harnedy, P.A.; FitzGerald, R.J. Cardioprotective peptides from marine sources. Curr. Protein Pept. Sci. 2013, 14, 162–172. [Google Scholar] [CrossRef] [PubMed]
  289. D’Orazio, N.; Gemello, E.; Gammone, M.A.; De Girolamo, M.; Ficoneri, C.; Riccioni, G. Fucoxantin: A treasure from the sea. Mar. Drugs 2012, 10, 604–616. [Google Scholar]
  290. Nasri, R.; Nasri, M. Marine-derived bioactive peptides as new anticoagulant agents: A review. Curr. Protein Pept. Sci. 2013, 14, 199–204. [Google Scholar] [CrossRef] [PubMed]
  291. Cusick, K.D.; Sayler, G.S. An overview on the marine neurotoxin, saxitoxin: Genetics, molecular targets, methods of detection and ecological functions. Mar. Drugs 2013, 11, 991–1018. [Google Scholar] [CrossRef] [PubMed]
  292. Medina, M.; Avila, J.; Villanueva, N. Use of okadaic acid to identify relevant phosphoepitopes in pathology: A focus on neurodegeneration. Mar. Drugs 2013, 11, 1656–1668. [Google Scholar] [CrossRef] [PubMed]
  293. Molgo, J.; Araoz, R.; Benoit, E.; Iorga, B.I. Physical and virtual screening methods for marine toxins and drug discovery targeting nicotinic acetylcholine receptors. Expert Opin. Drug Discov. 2013, 8, 1203–1223. [Google Scholar] [CrossRef] [PubMed]
  294. Pangestuti, R.; Kim, K. Marine-derived bioactive materials for neuroprotection. Food Sci. Biotechnol. 2013, 22, 1175–1186. [Google Scholar] [CrossRef]
  295. Lin, Z.; Torres, J.P.; Ammon, M.A.; Marett, L.; Teichert, R.W.; Reilly, C.A.; Kwan, J.C.; Hughen, R.W.; Flores, M.; Tianero, M.D.; et al. A bacterial source for mollusk pyrone polyketides. Chem. Biol. 2013, 20, 73–81. [Google Scholar] [CrossRef] [PubMed]
  296. Bharate, S.B.; Sawant, S.D.; Singh, P.P.; Vishwakarma, R.A. Kinase inhibitors of marine origin. Chem. Rev. 2013, 113, 6761–6815. [Google Scholar] [CrossRef] [PubMed]
  297. Liu, J.; Hu, Y.; Waller, D.L.; Wang, J.; Liu, Q. Natural products as kinase inhibitors. Nat. Prod. Rep. 2012, 29, 392–403. [Google Scholar] [CrossRef] [PubMed]
  298. Jiang, C.S.; Liang, L.F.; Guo, Y.W. Natural products possessing protein tyrosine phosphatase 1B (PTP1B) inhibitory activity found in the last decades. Acta Pharmacol. Sin. 2012, 33, 1217–1245. [Google Scholar] [CrossRef] [PubMed]
  299. Brady, R.M.; Baell, J.B.; Norton, R.S. Strategies for the development of conotoxins as new therapeutic leads. Mar. Drugs 2013, 11, 2293–2313. [Google Scholar] [CrossRef] [PubMed]
  300. Essack, M.; Bajic, V.B.; Archer, J.A. Conotoxins that confer therapeutic possibilities. Mar. Drugs 2012, 10, 1244–1265. [Google Scholar] [CrossRef] [PubMed]
  301. Rashid, M.H.; Mahdavi, S.; Kuyucak, S. Computational studies of marine toxins targeting ion channels. Mar. Drugs 2013, 11, 848–869. [Google Scholar] [CrossRef] [PubMed]
  302. Green, D.W.; Padula, M.P.; Santos, J.; Chou, J.; Milthorpe, B.; Ben-Nissan, B. A therapeutic potential for marine skeletal proteins in bone regeneration. Mar. Drugs 2013, 11, 1203–1220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Thomas, N.V.; Kim, S.K. Beneficial effects of marine algal compounds in cosmeceuticals. Mar. Drugs 2013, 11, 146–164. [Google Scholar] [CrossRef] [PubMed]
  304. Fiorucci, S.; Distrutti, E.; Bifulco, G.; D’Auria, M.V.; Zampella, A. Marine sponge steroids as nuclear receptor ligands. Trends Pharmacol. Sci. 2012, 33, 591–601. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Marine pharmacology in 2012–2013: marine compounds with antibacterial, antifungal, antiprotozoal, antituberculosis, and antiviral activities.
Figure 1. Marine pharmacology in 2012–2013: marine compounds with antibacterial, antifungal, antiprotozoal, antituberculosis, and antiviral activities.
Marinedrugs 15 00273 g001aMarinedrugs 15 00273 g001bMarinedrugs 15 00273 g001cMarinedrugs 15 00273 g001dMarinedrugs 15 00273 g001eMarinedrugs 15 00273 g001fMarinedrugs 15 00273 g001gMarinedrugs 15 00273 g001hMarinedrugs 15 00273 g001i
Figure 2. Marine pharmacology in 2012–2013: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Figure 2. Marine pharmacology in 2012–2013: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Marinedrugs 15 00273 g002aMarinedrugs 15 00273 g002bMarinedrugs 15 00273 g002cMarinedrugs 15 00273 g002dMarinedrugs 15 00273 g002eMarinedrugs 15 00273 g002f
Figure 3. Marine pharmacology in 2012–2013: marine compounds with miscellaneous mechanisms of action.
Figure 3. Marine pharmacology in 2012–2013: marine compounds with miscellaneous mechanisms of action.
Marinedrugs 15 00273 g003aMarinedrugs 15 00273 g003bMarinedrugs 15 00273 g003cMarinedrugs 15 00273 g003dMarinedrugs 15 00273 g003eMarinedrugs 15 00273 g003fMarinedrugs 15 00273 g003g
Table 1. Marine pharmacology in 2012–2013: marine compounds with antibacterial, antifungal, antituberculosis, antiprotozoal, antiviral and anthelmintic activities.
Table 1. Marine pharmacology in 2012–2013: marine compounds with antibacterial, antifungal, antituberculosis, antiprotozoal, antiviral and anthelmintic activities.
Drug ClassCompound/Organism aChemistryPharmacologic ActivityIC50 bMMOA bCountry cReferences
Antibacterialanthracimycin (1)/bacteriumPolyketide dB. anthracis & S. aureus inhibition0.03–0.06 μg/mL +DNA/RNA inhibitionUSA[31]
Antibacterialchrysophaentins (2,3)/algaShikimate hGram-negative & -positive bacterial inhibition27–84 μM +Competitive inhibition of FtsZ GTP-binding siteESP, USA[32]
Antibacterialmerochlorin A (4)/bacteriumTerpenoid eC. dificile & S. aureus strains inhibition0.3–2 μg/mL +DNA, RNA, protein & cell wall synthesis inhibitionUSA[33]
Antibacterialaflatoxin B2b (5)/fungusPolyketide dB. subtilis & E. aerogenes inhibition1.7, 1.1 μM +UndeterminedCHN[34]
Antibacterialageloxime B (6)/spongeAlkaloid/terpenoid eS. aureus inhibition7.2–9.2 μg/mL *UndeterminedCHN, USA[35]
AntibacterialAlternaria sp. anthraquinones (79)/fungusPolyketide dE. coli & V. parahemolyticus inhibition0.62–5 μM +UndeterminedCHN[36]
Antibacterialantimycin B2 (10)/bacteriumShikimate/Polyketide dL. hongkongensis inhibition8 μg/mL +UndeterminedCHN[37]
AntibacterialAspergillus sp. (−)sydonol (11)/fungusTerpenoid eS. albus & M. tetragenus inhibition1.2–5 μg/mL +UndeterminedCHN, NLD[38]
Antibacterialaxistatins 1–3 (1214)/spongeAlkaloid/terpenoid eC. neoformans & S. aureus inhibition1–4 μg/mL +UndeterminedAUS, USA[39]
Antibacterialbromophycoic acid A & E (15,16)/algaTerpenoid eS. aureus & E. faecilis inhibition1.6 μg/mL +UndeterminedFJI, USA[40]
Antibacterialcadeolides C–F (1720)/tunicateShikimate hS. aureus inhibition0.13–3 μg/mL +UndeterminedS. KOR[41]
Antibacterialcadiolides E–I (2123)/ascidianShikimate hS. aureus & B. subtilis inhibition0.8–12 μg/mL +UndeterminedS. KOR[42]
Antibacterialcitreamicin θ A & B (24,25)/bacteriumPolyketide dS. aureus inhibition0.25–1 μg/mL *UndeterminedCHN, SAU[43]
Antibacterialcommunol A & F (26,27)/fungusPolyketide dE. coli inhibition4.1, 6.4 μg/mL +UndeterminedCHN[44]
AntibacterialD. spiralis dolabellanes (28,29)/algaTerpenoid eS. aureus inhibition2–8 μg/mL +UndeterminedGRC, ESP, UK[45]
Antibacterialenhygrolide A (30)/bacteriumShikimate hA. cristallopoietes inhibition4 μg/mL +UndeterminedDEU[46]
Antibacterialeudistomin Y11 (31)/ascidianAlkaloid fB. subtilis & S. typhimurium inhibition3.12 μg/mL +UndeterminedS. KOR[47]
Antibacterialfradimycin B (32)/bacteriumPolyketide dS. aureus inhibition2.0 μg/mL +UndeterminedCHN[48]
AntibacterialHaliclona diAPS (3335)/spongeAlkaloid fM. luteus inhibition3.1 μg/mL +UndeterminedS. KOR[49]
Antibacterialhyrtimomine D (36)/spongeAlkaloid fS. aureus inhibition4 μg/mL +UndeterminedJPN[50]
Antibacterialianthelliformisamine A (37)/spongeAlkaloid fP. aeruginosa inhibition6.8 μMUndeterminedAUS[51]
Antibacterialkocurin (38)/bacteriumPeptide fMR S. aureus inhibition0.25 μg/mL +UndeterminedESP, USA[52]
Antibacteriallamellarin O (39)/spongeAlkaloid fB. subtilis inhibition2.5 μMUndeterminedAUS[53]
AntibacterialLaurencia sesquiterpenes (4042)/algaTerpenoid eE. coli & S. aureus inhibition5–7 μg/disk ++UndeterminedCHN, USA[54]
Antibacteriallobophorin H (43)/bacteriumTerpenoid glycosideB. subtilis inhibition1.57 μg/mL +UndeterminedCHN[55]
Antibacterialmarthiapeptide A (44)/bacteriumPeptide fM. luteus & B. thuringiensis inhibition2.0 μg/mL *UndeterminedCHN[56]
Antibacterialnapyradiomycin A1 & B3 (45,46)/bacteriumTerpenoid/polyketide dS. aureus inhibition0.5–2 μg/mL +UndeterminedCHN[57,58]
AntibacterialNigrospora sp. anthraquinones (47,48)/fungusPolyketide dE. coli & S. aureus inhibition0.6–0.7 μM +UndeterminedCHN[59]
Antibacterialohmyungsamycin A (49)/bacteriumPeptide fB. subtilis inhibition4.28 μM +UndeterminedS. KOR[60]
Antibacterialpenicifuran A (50)/fungusShikimate hS. albus inhibition3.1 μM +UndeterminedCHN[61]
Antifungalcrambescidin-816 (51)/spongeAlkaloid fS. cerevisiae growth inhibition1 μM +G2/M cell cycle arrest and apoptosisESP, FRA[62]
Antifungalneothyonidioside (52)/sea cucumberTerpenoid glycosideS. cerevisiae inhibition1 μM +Binding to plasma membrane sterolsNZL[63]
Antifungalageloxime B (6)/spongeAlkaloid/terpenoidC. neoformans inhibition4.9 μg/mL *UndeterminedCHN, USA[35]
Antifungalaurantoside K (53)/spongePolyketide/alkaloid glycosideC. albicans inhibition1.95 μg/mL +UndeterminedFJI[64]
Antifungalcaulerprenylol B (54)/algaTerpenoid eC. glabrata & C. neoformans inhibition4.0 μg/mL +UndeterminedCHN[65]
Antifungaldidymellamide A (55)/fungusAlkaloid fC. albicans inhibition3.1 μg/mL +UndeterminedJPN[66]
Antifungalhippolachnin A (56)/spongePolyketide dT. rubrum, M. gypseum & C. neoformans inhibition0.41 μM +UndeterminedCHN[67]
Antifungalholotoxins F & G (57,58)/sea cucumberTerpenoid glycosideC. albicans, Microsporum & Cryptococcus inhibition1.4–5.8 μM +UndeterminedCHN, DEU[68]
Antifungalhyrtimomine D & E (36,59)/spongeAlkaloid fC. albicans & C. neoformans inhibition4–16 μg/mL +UndeterminedJPN[50]
Antifungalnagelamide Z (60)/spongeAlkaloid fC. albicans inhibition0.25 μg/mL *UndeterminedJPN[69]
Antifungalwoodylide A (61)/spongePolyketide dC. neoformans inhibition3.7 μg/mL *UndeterminedCHN[70]
Antiprotozoalaraplysillin I (62)/spongeAlkaloid fP. falciparum FcB1 & 3D7 strain inhibition4.5 μMUndeterminedAUS, DEU, FJI, FRA[71]
Antiprotozoalascidiathiazone A (63)/ascidianAlkaloid fP. falciparum K1 strain inhibition3.3 μMUndeterminedNZL, CHE[72]
Antiprotozoalaxidjiferosides A–C (6466)/spongeGlycosphingolipidP. falciparum FcB1strain inhibition0.53 μMUndeterminedFRA[73]
Antiprotozoalcytosporone E (67)/fungusPolyketide dP. falciparum inhibition13 μM **UndeterminedUSA[74]
Antiprotozoaldicerandrol D (68)/fungusPolyketide dP. falciparum 3D7 strain inhibition0.6 μMUndeterminedCHN, TWN, USA[75]
Antiprotozoaldihydroingenamine D (69)/spongeAlkaloid fP. falciparum D6 & W2 strain inhibition57–72 ng/mLUndeterminedUSA[76]
Antiprotozoal19-hydroxypsammaplysin E (70)/spongeAlkaloid fP. falciparum 3D7strain inhibition6.4 μMUndeterminedAUS, IDN[77]
Antiprotozoalkabiramide L (71)/spongePolyketide dP. falciparum K1 strain inhibition2.6 μMUndeterminedTHAI, AUT[78]
Antiprotozoalmeridianin C & G (72,73)/tunicateAlkaloid fP. falciparum D6 & W2 strain inhibition4.4–14.4 μMUndeterminedIND[79]
Antiprotozoalorthidine F (74)/ascidianAlkaloid fP. falciparum K1 strain inhibition0.90 μMUndeterminedCHE, NZL[80]
Antiprotozoalplakortide U (75)/spongePolyketide dP. falciparum FcM29 strain inhibition0.8 μMUndeterminedFRA, ITA[81]
Antiprotozoalthiaplakortone A (76)/spongeAlkaloid fP. falciparum 3D7 & Dd2 strain inhibition0.006–0.051 μMUndeterminedAUS[82]
Antiprotozoaltsitikammamine C (77)/spongeAlkaloid fP. falciparum 3D7 & Dd2 strain inhibition13 & 18 nMUndeterminedAUS[83]
Antiprotozoalurdamycinone E (78)/bacteriumPolyketide dP. falciparum K1 strain inhibition0.05 μg/mLUndeterminedTHAI[84]
Antiprotozoalalmiramide (79,80)/bacteriumPeptide fT. brucei inhibition0.4–3.5 μMGlycosome function inhibitionUSA[85]
Antiprotozoaldiazepinomicin (81)/bacteriumAlkaloid/terpenoidT. brucei inhibition13.5 μMRhodesain inhibitionEGY, DEU[86]
Antiprotozoal(−)-elatol (82)/algaTerpenoid eT. cruzi inhibition1.5–3 μM *Mitochondrial disfunctionBRA[87]
Antiprotozoalascidiathiazone A (63)/ascidianAlkaloid fT. b. rhodesiense inhibition3.1 μMUndeterminedNZL, CHE[72]
Antiprotozoalcoibacin A (83)/bacteriumPolyketide dL. donovani inhibition2.4 μMUndeterminedUSA, PAN[88]
Antiprotozoalcristaxenicin A (84)/gorgonianTerpenoid eT. congolense & L. amazonensis inhibition0.25 & 0.088 μMUndeterminedJPN[89]
Antiprotozoalmanadoperoxide B analogues (85,86)/spongePolyketide dT. b. rhodesiense inhibition3–11 ng/mLUndeterminedITA, IDN, CHE, IRL[90]
Antituberculosisasperterpenoid A (87)/fungusTerpenoid eM. tuberculosis PTP inhibition2.2 μMUndeterminedCHN[91]
Antituberculosisbrevianamide S (88)/fungusAlkaloid fBCG inhibition6.25 μg/mL +UndeterminedAUS, CHN[92]
Antituberculosislobophorin G (89)/bacteriumTerpenoid e glycosideBCG inhibition1.56 μg/mL +UndeterminedCHN[93]
Antituberculosisneamphamide B (90)/spongePeptide fM. bovis inhibition1.56 μg/mL +UndeterminedJPN[94]
AntituberculosisS. flava diterpenes (91,92)/spongeTerpenoid eM. tuberculosis H37Rv inhibition15, 32 μg/mL +UndeterminedUSA[95]
Antituberculosisurdamycinone E (78)/bacteriumPolyketide dM. tuberculosis H37Ra inhibition3.13 μg/mL +UndeterminedTHAI[84]
Antiviralhalistanol sulfates (93,94)/spongeTerpenoid fHuman Herpes simplex virus-1 inhibition0.5–12.2 μg/mLAttachment & penetration inhibitionARG, BRA[96]
AntiviralL. arboreum metabolites (9597)/soft coralTerpenoid/sphingolipidHIV-1 protease inhibition4.8–7.2 μM *Molecular docking & HIV-1 protease receptorZAF[97]
Antiviralmanoalide (98)/spongeTerpenoid eHepatitis C virus inhibition15–70 μMRNA helicase and ATPase inhibitionJPN[98]
AntiviralN. aculeata metabolites (99,100)/algaPolyketide dHuman rhinoviruses 2 & 3 inhibition2.5–7.1 μg/mLCytopathic effect inhibitionS. KOR[99]
Antiviralstachybotrin D (101)/fungusAlkaloid/terpenoidHIV-1 replication inhibition8.4 μMReverse transcriptase inhibitionCHN[100]
Antiviralstreptoseolactone (102)/bacteriumTerpenoid fNeuraminidase inhibition3.9 μMNoncompetitive inhibitionCHN[101]
Antiviralasperterrestide A(103)/fungusPeptide fH3N2 influenza virus inhibition8.1 μMUndeterminedCHN[102]
AntiviralCladosporium sp. alkaloids (104,105)/fungusAlkaloid fH1N1 influenza virus inhibition82–85 μMUndeterminedCHN[103]
Antiviralisorhodoptilometrin-1-methyl ether (106)/fungusPolyketide dHepatitis C NS3/4A protease inhibition>1 ng/mL *UndeterminedEGY[104]
Antiviralmassarilactone H (107)/fungusPolyketide dInfluenza virus neuraminidase inhibition8.2 μMUndeterminedCHN, MYS[105]
Antiviralpyronepolyene C-glucoside (108)/fungusPolyketide dH1N1 influenza virus inhibition91.5 μMUndeterminedCHN[106]
AntiviralS. candidula sterol (109,110)/soft coralTerpenoid/sphingolipidH5N1 avian influenza virus inhibition1 ng/mL *UndeterminedEGY[107]
AntiviralS. vulgare glycolipid (111)/algaGlycolipidHuman herpes simplex virus-1 & 2 inhibition<50 μg/mLUndeterminedBRA[108]
Anthelminticechinosides A & B (112,113)/sea cucumberTerpenoid glycosideS. mansoni worm lethality0.19, 0.27 μg/mL +++UndeterminedEGY[109]
(a) Organism: Kingdom Animalia: ascidian (Phylum Chordata), gorgonian, coral (Phylum Cnidaria), sea cucumber (Phylum Echinodermata), sponge (Phylum Porifera); Kingdom Monera: bacterium (Phylum Cyanobacteria); Kingdom Fungi: fungus; Kingdom Plantae: alga; (b) IC50: concentration of a compound required for 50% inhibition in vitro, *: estimated IC50, **: IC90, +: MIC: minimum inhibitory concentration, ++: MID: minimum inhibitory concentration per disk; +++: LC50: concentration of a compound required for 50% lethality; MMOA: molecular mechanism of action; (c) Country: ARG: Argentina; AUS: Australia; AUT: Austria; BRA: Brazil; CHE: Switzerland; CHN: China; DEU: Germany; EGY: Egypt; ESP: Spain; FJI: Fiji; FRA: France; GRC: Greece; IDN: Indonesia; IND: India; IRL: Ireland; ITA: Italy; JPN: Japan; MYS: Malaysia; NLD: The Netherlands; NZL: New Zealand; PAN: Panama; SAU: Saudi Arabia; S. KOR: South Korea; THAI: Thailand; TWN: Taiwan; UK: United Kingdom; ZAF: S. Africa; Chemistry: (d) Polyketide; (e) Terpene; (f) Nitrogen-containing compound; (g) Polysaccharide, (h) Shikimate; Abbreviations: BCG: Bacille Calmette-Guérin; diAPS: dialkylpyridinium; MR: methicillin-resistant.
Table 2. Marine pharmacology in 2012–2013: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Table 2. Marine pharmacology in 2012–2013: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Drug ClassCompound/Organism aChemistryPharmacological ActivityIC50 bMMOA cCountry dReferences
Antidiabeticoctaphlorethol A (114)/algaPolyketide eIncreased glucose uptake in rat myoblast cells50 μM *Glucose transporter 4 translocationS. KOR[120]
Anti-inflammatoryapo-9′-fucoxanthinone (115)/algaTerpenoid fMacrophage TNF-α, IL-6 & 12 expression inhibition5–14 μMMAPK pathway inhibitionS. KOR[121]
Anti-inflammatoryastaxanthin (116)/algaTerpenoid fMacrophage cytokine inhibition10 μM *SHP-1 restorationITA[122]
Anti-inflammatorybengamide A & B (117,118)/spongeAlkaloid gMacrophage TNF-α & IL-6 inhibition0.5 μM *IĸBα phosphorylation inhibitionUSA[123]
Anti-inflammatorybis-N-norgliovictin (119)/fungusAlkaloid gMacrophage TNF-α, IL1-6, MCP-1 release inhibition in vitro0.5 μg/mL *Inflammatory gene inhibitionCHN[124]
Anti-inflammatory6,6′-bieckol (120)/algaPolyketide eMacrophage TNF-α & IL-6 expression inhibition25 μM *Inhibition of NFκBS. KOR, USA[125]
Anti-inflammatorycoibacin B (121)/bacteriumPolyketide eMacrophage NO inhibition5 μMiNOS, TNF-α, IL-1, IL-6 transcription inhibitionUSA, PAN[88]
Anti-inflammatory11-epi-sinulariolide acetate (122)/soft coralTerpenoid fMacrophage COX-2 & IL-8 expression inhibition10 μMCa2+ signaling inhibitionTWN[126]
Anti-inflammatoryhonaucin A (123)/bacteriumPolyketide eMacrophage NO inhibition4 μMiNOS, TNF-α, IL-1, IL-6 transcription inhibitionUSA, PAN[127]
Anti-inflammatoryHymeniacidon sp. amphilectanes (124,125)/spongeTerpenoid fBrain microglia TXB2 inhibition0.2 μMSOX independent & COX dependentUSA[128]
Anti-inflammatorylargazole (126)/bacteriumPeptide gModulation of human RA synovial fibroblasts in vitro5 μM *Enhanced HDAC6 & ICAM-1USA[129]
Anti-inflammatorylemnalol (127)/soft coralTerpenoid fIn vivo arthritis inhibition30 mg/kg*iNOS, COX-2 and c-Fos expression inhibitionTWN[130]
Anti-inflammatoryneoechinulin A (128)/fungusAlkaloidgMacrophage PGE2 and NO expression inhibition25–50 μM *Inhibition of NFκB & MAPKS. KOR; CHN[131]
Anti-inflammatorypenstyrylpyrone (129)/fungusShikimate/polyketideMacrophage NO, PGE2, IL1β inhibition9.3–13.5 μMPTP1B inhibitionS. KOR[132]
Anti-inflammatoryperthamide C (130)/spongePeptide gCarrageenan-induced paw edema inhibitionNDInduction of proteome changesITA[133]
Anti-inflammatoryR-prostaglandins (131,132)/soft coralPolyketide eTopical inflammation inhibitionNDPMN elastase inhibitionCOL[134]
Anti-inflammatorysinularin (133)/soft coralTerpenoid fCarrageenan-induced spinal neuroinflammation inhibition0.1 μM *iNOS & COX-2 inhibitionTWN[135]
Anti-inflammatoryswinhosterol B (134)/spongeTerpenoid fLymphocyte release of IL-1010 μM *Pregnane-X-receptor agonistITA, FRA[136]
Anti-inflammatoryA. polyacanthus steroids (135,136)/starfishTerpenoid fBone marrow-derived dendritic cells IL-6 and TNF-α inhibition1.8–7.0 μMUndeterminedS. KOR, VNM[137]
Anti-inflammatorybarettin (137)/spongeAlkaloid gMacrophage anti-inflammatory IL-10 release in vitro50 μg/mLUndeterminedNOR[138]
Anti-inflammatorybriarenolide F (138)/octocoralTerpenoid fNeutrophil superoxide inhibition3.82 μg/mLUndeterminedTWN[139]
Anti-inflammatoryCallyspongia sp. diketopiperazine (139)/spongePeptide gMacrophage IL1β release inhibition in vitro5 μg/mL *UndeterminedCHN[140]
Anti-inflammatory6-epi-cladieunicellin F (140)/octocoralTerpenoid fNeutrophil superoxide and elastase inhibition10 μM *UndeterminedTWN[141]
Anti-inflammatorycrassarosteroside A (141)/soft coralTerpenoid glycoside fMacrophage iNOS protein inhibition10 μM *UndeterminedTWN[142]
Anti-inflammatorycystodione A (142)/algaTerpenoid fRadical-scavenging and macrophage TNF-α inhibition in vitro8–22 μM *UndeterminedESP, MAR[143]
Anti-inflammatorydensanins A & B (143,144)/spongeAlkaloid gMacrophage NO release inhibition1–2.1 μMUndeterminedS. KOR[144]
Anti-inflammatorydissesterol (145)/soft coralTerpenoid fBone marrow dendritic cells IL-12 release inhibition4 μMUndeterminedS. KOR, VNM[145]
Anti-inflammatoryechinohalimane A (146)/gorgonianTerpenoid fNeutrophil elastase inhibition0.38 μg/mLUndeterminedTWN[146]
Anti-inflammatoryeunicidiol (147)/gorgonianTerpenoid fPMA-induced mouse ear edema inhibition100 μg/earUndeterminedCAN[147]
Anti-inflammatoryflexibilisolide C (148)/soft coralTerpenoid fMacrophage COX-2 & iNOS expression inhibition10 μM *UndeterminedTWN[148]
Anti-inflammatoryflexibilisquinone (149)/soft coralTerpenoid fMacrophage COX-2 & iNOS expression inhibition10–20 μM *UndeterminedTWN[149]
Anti-inflammatorylobocrassin F (150)/soft coralTerpenoid fNeutrophil elastase release inhibition6.3 μM *UndeterminedTWN[150]
Anti-inflammatoryperthamide J (151)/spongePeptide gCarrageenan-induced paw edema reduction0.3 mg/kg *UndeterminedITA, FRA[151]
Anti-inflammatorypseudoalteromone A (152)/bacteriumTerpenoid fNeutrophil elastase inhibition10 μg/mL *UndeterminedTWN[152]
Anti-inflammatorysarcocrassocolide M (153)/soft coralTerpenoid fMacrophage COX-2 & iNOS expression inhibition10 μM *UndeterminedTWN[153]
Anti-inflammatorysclerosteroids K & M (154,155)/soft coralTerpenoid fMacrophage COX-2 & iNOS expression inhibition10 μM *UndeterminedTWN[154]
Anti-inflammatoryseco-briarellinone (156)/octocoralTerpenoid fMacrophage NO release inhibition4.7 μMUndeterminedPAN[155]
Anti-inflammatorysinularioside (157)/soft coralGlycolipidMacrophage NO release inhibition30 μM *UndeterminedITA[156]
Immune systemlobocrassin B (158)/soft coralTerpenoid fDendritic cell activation inhibition39 μM *NF-κB translocation and TNF-α release inhibitionTWN[157]
Immune systempenicacid B(159)/fungusPolyketide eT lymphocyte proliferation inhibition0.23–20 μMIMPDH inhibitionCHN[158]
Nervous systemAPETx2 peptide (160)/sea anemonePeptide gASIC3 inhibition61 nMN- and C- termini truncation decrease inhibitionAUS[159]
Nervous systemasteropsin A (161)/spongePeptide gEnhancement of neuronal Ca2+ influx14 nMNo binding with VGSC site 2S. KOR, USA[160]
Nervous systemBcsTx peptides (162,163)/sea anemonePeptide grKv1.1 inhibition0.02–80 nMPotassium influx inhibitionBRA, BEL[161]
Nervous systemC. consors peptide (164)/cone snailPeptide gMuscle relaxation induction0.15 μMNav1.4 & Nav1.2 channel inhibitionBEL, FRA, CHE, CHL, DEU, NLD,[162]
Nervous systemC. magnificus conotoxin MfVIA(165)/cone snailPeptide gNeuronal Na+ current inhibition95 nMNav1.8 and Nav1.4 channel inhibitionAUS[163]
Nervous systemC. regius conotoxin RegIIA (166)/cone snailPeptide gACH-current inhibition33 nMΑ2β2 ACH receptorAUS, DEU, USA[164]
Nervous systemC. regularis peptide (167)/cone snailPeptide gAntinociceptive activity0.85 mg/kg *Cav2.2 channel inhibitionMEX[165]
Nervous systemconvolutamydine A (168)/bryozoaAlkaloid gAntinociceptive activity1 mg/kgCholinergic, opioid and nitric oxideBRA[166]
Nervous systemH. crispa polypeptides (169)/sea anemonePeptide gAntinociceptive and analgesic activity in vivo0.01–0.1 mg/kg *Inhibition of TRPV1 vanilloid 1 receptorRUS[167]
Nervous systemianthellamide A (170)/spongeAlkaloid gIncreased kynurenic acid in vivo200 mg/kg *Kynurenine 3- hydroxylase inhibitionAUS[168]
Nervous systemleucettamine B (171)/spongeAlkaloid gReduction of neurodegeneration in brain slices by analog leucettine L410.6–4.1 μMDual tyrosine phosphorylation kinase inhibitionFRA, UK, USA[169]
Nervous systempulchranin A (172)/spongeAlkaloid gTRPV1 receptor inhibition41.2 μMCa2+ response inhibitionRUS, S. KOR[170]
Nervous systemserinolamide B (173)/bacteriumAlkaloid gCB1 & CB2 binding**cAMP accumulation inhibitionUSA[171]
Nervous systemarigsugacin I (174)/fungusTerpenoid facetylcholinesterase inhibition0.64 μMUndeterminedCHN[172]
Nervous systemasperterpenol A (175)/fungusTerpenoid facetylcholinesterase inhibition2.3 μMUndeterminedCHN[173]
Nervous systemcymatherelactone (176)/algaPolyketide evoltage-gated sodium channel inhibition16 μMUndeterminedUSA[174]
Nervous systemdictyodendrin H (177)/spongeAlkaloid gBACE inhibition1 μMUndeterminedAUS[175]
Nervous systemgeranylphenazinediol (178)/bacteriumAlkaloid gacetylcholinesterase inhibition2.62 μMUndeterminedDEU[176]
Nervous systemhalomadurones C & D (179,180)/bacteriaTerpenoid eNrf2-ARE activation3.7 μM *UndeterminedUSA[177]
Nervous systemlamellarin O (39)/spongeAlkaloid gBACE inhibition<10 μMUndeterminedAUS[53]
Nervous systemPsammocinia sp. ircinianin lactams (181,182)/spongeTerpenoid fA3 GlyR potentiation8.5 μMUndeterminedAUS, DEU[178]
Nervous systemstarfish polar steroids (183188)/starfishTerpenoid fNeuritogenic and neuroprotective1–100 nMUndeterminedRUS[179]
(a) Organism: Kingdom Animalia: coral and sea anemone (Phylum Cnidaria); starfish (Phylum Echinodermata); cone snail (Phylum Mollusca); sponge (Phylum Porifera); Kingdom Fungi: fungus; Kingdom Plantae: alga; Kingdom Monera: bacterium; (b) IC50: concentration of a compound required for 50% inhibition, *: apparent IC50, **: Ki 16.4 and 2 μM, respectively; (c) MMOA: molecular mechanism of action; (d) Country: AUS: Australia; BEL: Belgium; BRA: Brazil; CHE: Switzerland; CHL: Chile; CHN: China; COL: Colombia; DEU: Germany; ESP: Spain; FRA: France; ITA: Italy; MAR: Morocco; MEX: Mexico; NLD: Netherlands; NOR: Norway; PAN: Panama; RUS: Russian Federation; S. KOR: South Korea; TWN: Taiwan; UK: United Kingdom; VNM: Vietnam; Chemistry: (e) Polyketide; (f) Terpene; (g) Nitrogen-containing compound; (h) polysaccharide. Abbreviations: ASIC3: pH-sensitive sodium ion channel 3; BACE: protease β-secretase; COX: cyclooxygenase; GlyR: glycine-gated chloride channel receptor; HDAC6: class II, histone deacetylase 6; ICAM: intercellular adhesion molecule-1; iNOS: inducible nitric oxide synthase; IMPDH: inosine 5′-monophosphate dehydrogenase; MAPK: mitogen-activated protein kinase pathway; NO: nitric oxide; Nrf2-ARE: nuclear transcription factor E2-related factor antioxidant response element; PTP1B: tyrosine protein phosphatase 1B; rKv1.1: voltage-gated potassium channel Kv subfamily; SHP1: SHP-1 protein tyrosine phosphatase; SOX: superoxide; TRPV1: transient receptor potential cationic channel of subfamily V.
Table 3. Marine pharmacology in 2012–2013: marine compounds with miscellaneous mechanisms of action.
Table 3. Marine pharmacology in 2012–2013: marine compounds with miscellaneous mechanisms of action.
Compound/Organism aChemistryPharmacological ActivityIC50 bMMOA cCountry dReferences
astaxanthin (189)/algaTerpenoid fHuman sperm capacitation2 μM *Increased tyrosine phosphorylationITA[182]
astaxanthin (189)/algaTerpenoid fApoptosis reduction in retinal ganglion cells2 μMH2O2 inhibitionCHN[242]
biselyngbyaside (190)/bacteriumPolyketide eOsteoclast apoptosis induction30 nM *c-Fos and NFATc1 inhibitionJPN[183]
Callyspongia sp. bisacetylenic alcohol (191)/spongePolyketide eLymphatic endothelial cell proliferation inhibition0.11 μMCell cycle arrestJPN, NLD[184]
conicasterol E (192)/spongeTerpenoid fBile acid detoxification10 μM *Farnesoid and pregnane receptor activity modulationITA, PYF[185]
6′′-debromohamacanthin A (193)/spongeAlkaloid gAngiogenesis inhibition14.8 μMPI3K/AKT/mTOR signaling inhibitionCAN, S. KOR[186]
dieckol (194)/algaPolyketide eInhibition of melanin synthesis>120 µM *Cellular tyrosinase inhibitionS. KOR[187]
fructigenine A (195)/fungusAlkaloid gPTP1B inhibition10.7 μMNoncompetitive inhibitionS. KOR[188]
geoditin A (196)/spongeTerpenoid fMelanogenesis inhibition1 μg/mLcAMP-dependent signaling inhibitionCHN, USA[189]
gorgosterol (197)/soft coralTerpenoid fFXR transactivation antagonism10 μMInhibition of OSTα & BSEP genesITA[190]
gracilioether B (198)/spongePolyketide ePPARγ binding5 μM *Cys285 covalent bindingFRA, ITA[191]
gracilioether K (199)/spongePolyketide ePXR agonistic activity10 μM *Binding to LBD by molecular dockingITA[192]
herdmanine K (200)/ascidianAlkaloid gPPAR-γ agonist1 μg/mL *mRNAexpression of target genesS. KOR[193]
hyrtioreticulin A (201)/spongeAlkaloid gUbiquitin-activating enzyme inhibition2.4 μMPutative ubiquitin-adenylate intermediate inhibitionIDN, JPN, NLD[194]
InhVJ protease inhibitor (202)/sea anemonePeptide gTrypsin and α-chymotrysin inhibition**Glu45 involved in InhVJ-trypsin complexBEL, RUS[195]
jaspamide (203)/spongePeptide gDecreased cardiomyocyte activity and function1–19 μM *Kv1.5 channel inhibitionUSA[196]
latonduine A (204)/spongeAlkaloid gF508del-CTFR correction1 μM *PARP-3 inhibitionCAN[197]
leucettine L41 (205)/spongeAlkaloid gDYR and CL tyrosine kinase inhibition21–77 nMPrimary and secondary targets identifiedFRA[169]
manzamine A (206)/spongeAlkaloid gCholesterol esters inhibition4.1 μMACAT inhibitionJPN[198]
nahuoic acid A (207)/bacteriumPolyketide eSETDH inhibition6.5 μMCompetitive inhibitionPNG, CAN[199]
namalide (208)/spongePeptide gCarbopeptidase A inhibition0.25 μMd-Lys presence required for activityITA, USA[200]
ningalins C & D (209,210)/ascidianAlkaloid gCK1δ and GSK3β inhibition0.2 μMBinding to ATP binding siteAUS[201]
octaphlorethol A (114)/algaPolyketide eGlucose tansporter 4 increase10 μM *AKT and AMPK activationS. KOR[120]
petrosaspongiolide M (211)/spongeTerpenoid fProteasome inhibition0.085–1.05 μMPro-apoptotic bax inductionITA[202]
petrosiol A (212)/spongePolyketide ePDGF-induced DNA synthesis inhibition0.73 μMPDGF receptor-β signaling inhibitionJPN[203]
phidianidine A (213)/molluscAlkaloid gCXCR4 ligand antagonist<50 μMCXCL12-dependent DNA synthesis inhibitionITA[204]
Poly-APS (214)/spongePolyketide eThoracic aorta contraction inhibition in vitro<10 μM *Concentration-dependent LDH releaseSVN[205]
Pseudoceratina sp. Dibromotyrosine (215)/spongeAlkaloid gApoptosis induction5 μg/mLMitochondrial disfunctionEGY, TWN[206]
pseudopterosin A (216)/soft coralTerpenoid fIncreased HUVEC proliferation13 nMEnhancement potency by HPβCDUSA[207]
sargachromanol G (217)/algaTerpenoid fOsteoclastogenesis inhibition20 Μm *NF-ĸB phosphorylation of MAPK kinases inhibitionS. KOR[208]
S. graminifolium polysaccharide (218)/algaPolysaccharide hImproved mitochondrial disfunction and oxidative stress25 mg/kg ***Increased activity of antioxidant enzymesCHN[209]
S. patens phloroglucinol (219)/algaPolyketide eα-amylase inhibition3.2 μg/mLCompetitive α-amylase inhibitorJPN[210]
S. xiamenensis benzopyran (220)/bacteriumMixed biogenesisFibrosis inhibition30 μg/mL *Anti-proliferation, anti-contractile and anti-adhesion activityCHN[211]
theonellasterol (221)/spongeTerpenoid fFarnesoid receptor transactivation inhibition50 μM *SAR showed OH at C-4 and oxidation at C-3 requiredITA, JPN[212]
toluquinol (222)/fungusShikimateAngiogenesis inhibition in vitro and in vivo2.5 μM *Cell cycle arrest inductionESP[213]
U. lactuca fatty acid (223)/algaPolyketide eARE activator10 μg/mL *Nrf2 transcription factor activationUSA[214]
alotaketal C (224)/spongeTerpenoid fcAMP signaling activation6.5 μMUndeterminedCAN[215]
aspergentisyl A(225)/fungusPolyketide eDPPH radical-scavenging9.3 μMUndeterminedCHN[216]
A. terreus butyrolactone (226)/fungusShikimateβ-glucuronidase inhibition6.2 μMUndeterminedLKA, PAK, USA[217]
caulerpine (227)/algaAlkaloid gSpasmolytic effect on guinea pig ileum0.05–0.13 μMUndeterminedBRA[218]
conicasterol F (228)/spongeTerpenoid fFXR antagonism10 μM *UndeterminedGBR, ITA[219]
D. avara sesquiterpene (229)/spongeTerpenoid fFAK, IGF1 & ERBB2 kinase inhibition1 μg/mL *UndeterminedDEU, GBR, EGY, SAU[220]
D. gigantea sterols (230,231)/soft coralTerpenoid fFarnesoid receptor transactivation inhibition14–15 μMUndeterminedS. KOR[221]
dysidavarone A (232)/spongeTerpenoid fPTP1B inhibition9.98 μMUndeterminedCHN[222]
galvaquinone B (233)/bacteriumPolyketide eEpigenetic activity1.0 μMUndeterminedUSA[223]
halicloic acids A & B (234,235)/spongeTerpenoid fIDO1 inhibition10 & 11 μMUndeterminedCAN, NLD[224]
isochromophilone XI (236)/fungusPolyketide ePD4 inhibition8.3 μMUndeterminedDEU[225]
leucettamols A & B (237,238)/spongeTerpenoid fTRPA1 and TRPM8 channel inhibition4.7–6.4 μMUndeterminedITA[226]
manadosterol A (239)/spongeTerpenoid fUbiquitin E2 enzyme UBc13-Uev1A complex inhibition90 nMUndeterminedIDN. JPN, NLD[227]
marilines A1 & A2 (240,241)/fungusMixed biogenesisHLE inhibition0.86 μMUndeterminedDEU, GRC, PAN[228]
methyl sarcotroate B (242)/soft coralTerpenoid fPTP1B inhibition6.97 μMUndeterminedCHN[229]
P. citrinum sorbicillinoid (243)/fungusPolyketide eAntioxidant30 μMUndeterminedJPN[230]
phosphoiodyn A (244)/spongePolyketide ehPPARδ inhibition23.7 nMUndeterminedAUS, S. KOR[231]
purpuroines A & D (245,246)/spongeAlkaloid gLCK kinase inhibition0.94, 2.35 μg/mLUndeterminedDEU, CHN[232]
santacruzamate A (247)/bacteriumAlkaloid gHDAC2 inhibition0.110 nMUndeterminedPAN, USA[233]
sarcophytonolide N (248)/soft coralTerpenoid fPTP1B inhibition5.9 μMUndeterminedCHN, ITA[234]
sargassumol (249)/algaPolyketide eAntioxidant47 μMUndeterminedS. KOR[235]
sesquibastadin 1 (250)/spongeAlkaloid gProtein kinases inhibition0.1–6.5 μMUndeterminedCHN, DEU[236]
S. glaucum cembranoids (251253)/soft coralTerpenoid fCytochrome P450 1A inhibition12.7–3.7 nM *UndeterminedEGY, SAU, USA[237]
symplocin A (254)/bacteriumPeptide gCathepsin E inhibition0.3 nMUndeterminedUSA[238]
tsitsikammamine A derivative (255)/spongeAlkaloid gIDO1 inhibition0.9 μMUndeterminedBEL, FRA[239]
V. lanosa bromophenol (256)/algaTerpenoid fBiochemical & cellular antioxidant activity30 μg/mLUndeterminedNOR[240]
X. testudinaria fatty acid (257)/spongePolyketide eAdipogenesis stimulation2 μMUndeterminedJPN[241]
(a) Organism: Kingdom Animalia: soft corals and sea anemone (Phylum Cnidaria), starfish (Phylum Echinodermata), mollusk (Phylum Mollusca); sponge (Phylum Porifera); Kingdom Plantae: alga; Kingdom Monera: bacterium; (b) IC50: concentration of a compound required for 50% inhibition in vitro; *: estimated IC50; **: Ki 7.4 × 10−8 M, and 9.9 × 10−7 M, respectively; ***: in vivo study; (c) MMOA: molecular mechanism of action; (d) Country: AUS: Australia; BEL: Belgium; BRA: Brazil; CAN: Canada; CHN: China; DEU: Germany; EGY: Egypt; FRA: France; ESP: Spain; GBR: United Kingdom; GRC: Greece; IDN: Indonesia; ITA: Italy; JPN: Japan; LKA: Sri Lanka; NLD: The Netherlands; NOR: Norway; PAN: Panama; PAK: Pakistan; PNG: Papua New Guinea; PYF: French Polynesia; RUS: Russian Federation; SAU: Saudi Arabia; S. KOR: South Korea; SVN: Slovenia; TWN: Taiwan; Chemistry: (e) Polyketide; (f) Terpene; (g) Nitrogen-containing compound; (h) polysaccharide; Abbreviations: ACAT: acyl-CoA:cholesterol acyl-transferase; Akt: protein kinase B; AMPK: AMP-activated protein kinase; ARE: antioxidant-response element; ASIC3: pH-sensitive sodium ion channel 3; CFTR: cystic fibrosis transmembrane conductance regulator; CXCR4: chemokine receptor; CKL: cdc2-like kinase; DYRK: dual-specificity, tyrosine phosphorylation regulated kinase; ERBB2: erb-b2 receptor tyrosine kinase; FAK: focal adhesion kinase; FXR: farnesoid-X-receptor; HDAC: histone deacetylase; HLE: human leukocyte elastase; HUVEC: human umbilical vein endothelial cells; HPβCD: hydroxypropyl-β-cyclodrextrin; IDO1: indoleamine 2, 3 dioxygenase; Kv1.5: Potassium voltage-gated ion channel; LBD: ligand binding domain; LCK: lymphocyte-specific protein tyrosine kinase; IGF1-R: insulin-like growth factor 1 receptor; PDGF: platelet-derived growth factor; PI3K: phosphoinositide 3-kinase; Poly-APS: polymeric 3-alkylpyridinium salts; PARP: poly(ADP-ribose) polymerase; PD4: phosphodiesterase 4; PPARγ: peroxisome proliferator-activated receptor γ; PTP1B: protein tyrosine phosphatase 1B; PXR: pregnane-X-receptor; SETDH: protein methyltransferase SETD8; TRPA1: ankyrin channel; TRPM8: melastin channel.

Share and Cite

MDPI and ACS Style

Mayer, A.M.S.; Rodríguez, A.D.; Taglialatela-Scafati, O.; Fusetani, N. Marine Pharmacology in 2012–2013: Marine Compounds with Antibacterial, Antidiabetic, Antifungal, Anti-Inflammatory, Antiprotozoal, Antituberculosis, and Antiviral Activities; Affecting the Immune and Nervous Systems, and Other Miscellaneous Mechanisms of Action. Mar. Drugs 2017, 15, 273. https://doi.org/10.3390/md15090273

AMA Style

Mayer AMS, Rodríguez AD, Taglialatela-Scafati O, Fusetani N. Marine Pharmacology in 2012–2013: Marine Compounds with Antibacterial, Antidiabetic, Antifungal, Anti-Inflammatory, Antiprotozoal, Antituberculosis, and Antiviral Activities; Affecting the Immune and Nervous Systems, and Other Miscellaneous Mechanisms of Action. Marine Drugs. 2017; 15(9):273. https://doi.org/10.3390/md15090273

Chicago/Turabian Style

Mayer, Alejandro M. S., Abimael D. Rodríguez, Orazio Taglialatela-Scafati, and Nobuhiro Fusetani. 2017. "Marine Pharmacology in 2012–2013: Marine Compounds with Antibacterial, Antidiabetic, Antifungal, Anti-Inflammatory, Antiprotozoal, Antituberculosis, and Antiviral Activities; Affecting the Immune and Nervous Systems, and Other Miscellaneous Mechanisms of Action" Marine Drugs 15, no. 9: 273. https://doi.org/10.3390/md15090273

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

Article Metrics

Back to TopTop