Next Article in Journal
Design and Preclinical Evaluation of Chitosan/Kaolin Nanocomposites with Enhanced Hemostatic Efficiency
Previous Article in Journal
Efficacy of Posidonia oceanica Extract against Inflammatory Pain: In Vivo Studies in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Marine Pharmacology in 2016–2017: 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,*,
Aimee J. Guerrero
1,
Abimael D. Rodríguez
2,
Orazio Taglialatela-Scafati
3,
Fumiaki Nakamura
4 and
Nobuhiro Fusetani
5
1
Department of Pharmacology, College of Graduate Studies, 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, I-80131 Napoli, Italy
4
Department of Chemistry and Biochemistry, Graduate School of Advanced Science and Engineering, Waseda University, 3-4-1 Okubo, Shinjuku-ku, Tokyo 169-8555, Japan
5
Fisheries and Oceans Hakodate, Hakodate 041-8611, Japan
*
Author to whom correspondence should be addressed.
Mar. Drugs 2021, 19(2), 49; https://doi.org/10.3390/md19020049
Submission received: 8 December 2020 / Revised: 13 January 2021 / Accepted: 15 January 2021 / Published: 21 January 2021

Abstract

:
The review of the 2016–2017 marine pharmacology literature was prepared in a manner similar as the 10 prior reviews of this series. Preclinical marine pharmacology research during 2016–2017 assessed 313 marine compounds with novel pharmacology reported by a growing number of investigators from 54 countries. The peer-reviewed literature reported antibacterial, antifungal, antiprotozoal, antituberculosis, and antiviral activities for 123 marine natural products, 111 marine compounds with antidiabetic and anti-inflammatory activities as well as affecting the immune and nervous system, while in contrast 79 marine compounds displayed miscellaneous mechanisms of action which upon further investigation may contribute to several pharmacological classes. Therefore, in 2016–2017, the preclinical marine natural product pharmacology pipeline generated both novel pharmacology as well as potentially new lead compounds for the growing clinical marine pharmaceutical pipeline, and thus sustained with its contributions the global research for novel and effective therapeutic strategies for multiple disease categories.

1. Introduction

The present review aims to consolidate the 2016–2017 preclinical marine pharmacology literature, with a format similar to our previous 10 reviews of this series which cover the period 1998–2015 [1,2,3,4,5,6,7,8,9,10]. All peer-reviewed articles were retrieved from the following databases: MarinLit, PubMed, Chemical Abstracts®, ISI Web of Knowledge, and Google Scholar. As in our previous reviews we have decided to limit the review to the bioactivity and/or pharmacology of structurally characterized marine chemicals, which we have classified using a modification of Schmitz’s chemical classification [11] into six major chemical classes, namely, polyketides, terpenes, peptides, alkaloids, shikimates, and sugars. The preclinical antibacterial, antifungal, antiprotozoal, antituberculosis, and antiviral pharmacology of marine chemicals is reported in Table 1, with the structures, shown in Figure 1. Marine compounds that showed immune and nervous systems activities, as well as antidiabetic and anti-inflammatory effects are exhibited in Table 2, with their respective structures consolidated in Figure 2. Finally, marine compounds affecting a variety of cellular and molecular targets are noted in Table 3, and their structures presented in Figure 3.
Several publications during 2016–2017 reported on several marine extracts or structurally uncharacterized compounds, with potentially novel preclinical and/or clinical pharmacology: In vitro antibacterial and antibiotic potentiating activities of tropical Mauritian marine sponge extracts [12]; a sterol-rich fraction with significant synergistic anti-inflammatory activity isolated from the South Korean soft coral Dendronephthya gigantea [13]; two metabolites with anti-methicillin-resistant Staphylococcus aureus activity from an Indian marine sponge Clathria procera associated actinomycete S. pharmamarensis ICN40 with antimicrobial potential “after structure identification and clinical studies” are completed [14]; 49 of 251 bacterial isolates from sediments from several Red Sea harbor and lagoon environments with the potential to produce secondary metabolites with “antimicrobial activity” [15]; antibacterial activity in extracts from the marine bacterium Salinispora arenicola from the Gulf of California, Mexico [16]; activation by cyanobacterium Oscillatoria sp. lipopolysaccharide of rat microglia and murine B cells with concomitant Toll-like receptor 4 signaling in vitro [17,18]; and lipophilic fractions from the Icelandic marine sponge Halichondria sitiens that decrease dendritic cells’ pro-inflammatory cytokine release [19].
Table 1. Marine pharmacology in 2016–2017: marine compounds with antibacterial, antifungal, antituberculosis, antiprotozoal and antiviral activities.
Table 1. Marine pharmacology in 2016–2017: marine compounds with antibacterial, antifungal, antituberculosis, antiprotozoal and antiviral activities.
Drug ClassCompound/
Organism a
ChemistryPharmacologic ActivityIC50 b1MMOA b2Country cReferences
Antibacterialageloline A (1)/bacteriumAlkaloid fC. trachomatis inhibition9.5 μMAntioxidant activityDEU[20]
Antibacterialblasticidin S analog (2)/spongeAlkaloid fS. aureus inhibition6.2 μg/mL +norA multidrug trasporter inactivationCAN, USA[21]
Antibacterialdibromohemibastadin-1(3)/spongePeptide fP. auruginosa biofilm disruption10 μM *Quorum sensing activityDEU, FRA, GBR[22]
Antibacterialecteinamycin (4)/bacteriumPolyketide dC. difficile inhibition0.059 μM +K+ transport dysregulationJPN, USA[23]
Antibacterial 5-episinuleptolide (5)/soft coralTerpenoid eA. baumannii biofilm formation inhibition20 μM *PNAG gene expression inhibitionTWN[24]
AntibacterialE. esculentus peptides (6,7)/sea urchinPeptide fGram-positive and negative inhibition0.1–3.1 μM +Heavy chains bioactiveNOR, SWE[25]
Antibacterialgranaticin and granatomycin D (8,9)/bacteriumPolyketide dB. subtilis and MR S. aureus inhibition1.6, 6.2 μg/mL +Co-culture enhanced MICUSA[26]
Antibacterialkeyicin (10)/bacteriumPolyketide dB. subtilis and MR S. aureus inhibition 2.5–9.9 μM +Fatty acid metabolism modulationUSA[27]
Antibacterial microcionamides C and D (11, 12)/spongePeptide fS. aureus inhibition6.2 μM +Depolarize cytoplasmic membranesDEU, IDN, IRN[28]
Antibacterialmyticalin A5 (13)/musselPeptide fGram-positive and negative inhibition2–8 μM +RNA synthesis inhibitionDEU, ITA[29]
Antibacterialplakofuranolactone
(14)/sponge
Polyketide d Quorum quenching inhibition0.1 μM *Specificity to QS systemsITA[30]
Antibacterial psammaplin A (15)/spongePeptide fV. vulnificus in vivo growth inhibition50 μg/mouse **Associated in vitro and in vivo pathology suppressedS. KOR[31]
Antibacterialabyssomicin 2 (16)/bacteriumPolyketide dB. thuringiensis and M. luteus inhibition3.6, 7.2 μg/mL +UndeterminedCHN[32]
Antibacterialactinomycins D, V, and X2 (1719)/bacteriumPeptide fMR S. aureus, B. subtilis and E. coli inhibition0.08–0.61 μMUndeterminedCHN, EGY, SAU[33,34]
Antibacterialaneurinifactin (20)/bacteriumLipopeptide fK. pneumoniae and S. aureus inhibition4, 8 μg/mL +UndeterminedIND[35]
Antibacterialaspewentins D and H (21, 22)/fungusTerpenoid eP. aeruginosa and M. luteus inhibition4 μg/mL +UndeterminedCHN[36]
AntibacterialB. subtilis furanoterpenoid (23)/bacteriumTerpenoid eV. vulnificus and parahaemolyticus inhibition3.12 μg/mL +UndeterminedIND[37]
Antibacterialbacillisporin A (24)/fungusPolyketide dB. subtilis inhibition0.12 μM +UndeterminedBGD, CHN[38]
Antibacterial bacilotetrin A (25)/bacteriumPeptide fMR S. aureus inhibition8 μg/mL +UndeterminedS. KOR[39]
Antibacterialbranimycin B (26)/bacteriumPolyketide dM. luteus and C. urealyticum inhibition1, 8 μg/mL +UndeterminedESP[40]
Antibacterialbrocazine G (27)/fungusAlkaloid fS. aureus inhibition0.25 μg/mL +UndeterminedCHN, DEU, HUN[41]
Antibacterialcadiolides K and M (28, 29)/ascidianPolyketide dMR S. aureus inhibition1–2 μg/mL +UndeterminedS. KOR[42]
Antibacterialcahuitamycin D (30)/bacteriumPeptide fA. baumannii biofilm inhibition8.4 μMUndeterminedCRI, USA[43]
Antibacterialchalcomycin (31)/bacteriumPolyketide dS. aureus inhibition4 μg/mL +UndeterminedCHN[44,45]
Antibacterial chermesins A and B (32, 33)/fungusTerpenoid eM. luteus inhibition8 μg/mL +UndeterminedCHN[46]
Antibacterialchloro-preussomerins A and B (34, 35)/fungusPolyketide dS. aureus inhibition3.2, 6.2 μg/mL +UndeterminedCHN[47]
Antibacterialcollismycin C (36)/bacteriumAlkaloid fMR S. aureus biofilm inhibition10 μg/mL *UndeterminedS. KOR[48]
Antibacterialengyodontochones A and B (37, 38)/fungusPolyketide dMR S. aureus inhibition0.17, 0.24 μMUndeterminedCHN, DEU[49]
Antibacterialhydroanthraquinones (3943/fungusPolyketide dS. aureus inhibition2–8 μg/mL *UndeterminedCHN, DEU[50]
Antibacteriallangcoquinone C (44)/spongeTerpenoid eB. subtilis inhibition6.2 μM +UndeterminedJPN, MMR, VNM[51]
Antibacterialluffariellolide (45)/spongeTerpenoid eS. enterica inhibition4 μg/mL +UndeterminedS. KOR[52]
Antibacterialmanzamine alkaloids (4648)/spongeAlkaloid fGram-positive and negative inhibition2–8 ng/mL +UndeterminedIDN, S. KOR[53]
Antibacterialnapyradiomycin A1 (49)/bacteriumTerpenoid eMR S. aureus inhibition0.5–1 μg/mL +UndeterminedESP[54]
Antibacterialoxysporizoline (50)/fungusAlkaloid fMR S. aureus inhibition6.25 μg/mL +UndeterminedS. KOR[55]
AntibacterialP. citrinum 1-(2,6-dihydroxyphenyl)butan-1-one (51)/fungusPolyketide dS. aureus inhibition6.95 μM +UndeterminedCHN[56]
Antibacterialpenicillstressols (52, 53)/fungusPolyketide dMR S. aureus inhibition0.5 μg/mL +UndeterminedCHN[57]
Antibacterialpestalone (54)/fungusPolyketide dMR S. aureus inhibition6.25 μM +UndeterminedCHN[58]
Antibacterialpestalotionol (55)/fungusPolyketide dB. subtilis and S. aureus inhibition2, 8 μg/mL +UndeterminedCHN, TWN[59]
Antibacterialphomaethers A and C (56, 57)/fungusPolyketide dE. coli and S. aureus inhibition0.15–1.25 μM +UndeterminedCHN[60]
Antibacterial4-methyl-3”-prenylcandidusin A (58)/fungusPolyketide dMR S. aureus and V. vulnificus inhibition3.8, 7.8 μg/mL +UndeterminedCHN[61]
AntibacterialPseudomonas sp. rhamnolipid (59)/bacteriumLipid eB. cenocepacia and S. aureus inhibition1.6–3.1 μg/mL +UndeterminedGBR, ITA[62]
Antibacterialsmenospongine (60)/spongeTerpenoid eB. cereus and S. aureus inhibition3.1 μM +UndeterminedCHN, USA[63]
AntibacterialS. cheonanensis phthalate (61)/bacteriumPolyketide dP. vulgaris inhibition4 μg/mL +UndeterminedIND[64]
Antibacterialsporalactam B (62)/bacteriumPolyketide dMR S. aureus and E. coli inhibition0.4–1.8 μM **UndeterminedCAN, PHL[65]
Antibacterialtetrocarcin A (63)/fungusPolyketide dB. subtilis inhibition0.03–0.125 μg/mL +Undetermined CHN[66,67]
Antibacterialtricepyridinium (64)/spongeAlkaloid fB. subtilis and S. aureus inhibition0.78–1.56 μg/mL +UndeterminedJPN[68]
Antibacterialtrocheliane (65)/soft coralTerpenoid eA. baumannii and S. aureus inhibition 4–4.2 μM +UndeterminedEGY, IDN, SAU[69]
Antibacterialtulongicin (66)/spongeAlkaloid fS. aureus inhibition1.2 μg/mL +UndeterminedITA, NZL, USA[70]
Antibacterialvineomycin A1 (67)/bacteriumPolyketide dS. aureus inhibition4 μg/mL+UndeterminedCHN[71]
Antifungalamphidinol 3 (68)/dinoflagellatePolyketide dPore formation requires cholesterol or ergosterol2.0 μM *Toroidal pore 2.6-4.0 nMJPN, PHL[72,73]
Antifungalavarol (69)/spongeTerpenoid eC. albicans inhibition6–8 µg/mL +UndeterminedSRB[74]
Antifungaldihydromaltophilin (70)/bacteriumPolyketide dC. albicans inhibition3 μMUndeterminedAUS, MEX[75]
Antifungalhippolide j (71a, 71b)/spongeTerpenoid eC. albicans inhibition0.1 µg/mL +UndeterminedCHN, GBR[76]
Antifungalilicicolin H (72)/fungusPolyketide dC. albicans inhibition<0.25 µg/mL +UndeterminedDEU, DNK, ESP[77]
Antifungal iturin F1 and F2 (73, 74)/bacteriumPeptide fA. flavus and P. griseofulvum inhibition3.1 µg/mL +UndeterminedJPN, S. KOR[78]
AntifungalP. meleagrinum macrolides (75, 76)/fungusPolyketide dC. albicans inhibition1–2 μg/mL **UndeterminedJPN[79]
Antifungalplakinic acid M (77)/spongePolyketide dC. gattii inhibition2.4 μM **UndeterminedUSA[80]
Antifungalpoecillastroside D (78)/spongeTerpenoid eA. fumigatus inhibition6 μg/mL **UndeterminedESP, FRA, IRL, OMN, SWE[81]
Antifungalrocheicoside A (79)/ bacteriumAlkaloid fC. albicans inhibition4 μg/mL +UndeterminedTUR[82]
Antimalarialdiacarperoxide A (80)/spongeTerpenoid eP. falciparum D6 and W2 strain inhibition1.9–2.0 μMUndeterminedCHN, USA[83]
Antimalarial dudawalamide A and D (81, 82)/cyanobacteriumPeptide fP. falciparum W2 strain inhibition3.5 μMUndeterminedJOR, PAN, USA[84]
Antimalarialeudistidine A (83)/soft coralAlkaloid fP. falciparum D6 and W2 strain inhibition1.1–1.4 μMUndeterminedCAN, USA[85]
Antimalarialnaseseazine C (84)/bacteriumAlkaloid fP. falciparum 3D7 inhibition3.5 μMUndeterminedAUS[86]
AntimalarialP. opacum β-carboline (85)/ascidianAlkaloid fP. falciparum FcB1inhibition3.8 μMUndeterminedFRA, NZL[87]
Antimalarialptilomycalin F
(86)/sponge
Alkaloid fP. falciparum 3D7 strain inhibition0.23 μMUndeterminedBEL, FRA, CHE, NLD[88]
Antimalarialpustulosaisonitrile-1 (88)/nudibranchTerpenoid eP. falciparum 3D7 strain inhibition1.08 μMUndeterminedAUS, USA[89]
AntileishmanialA. Niger fatty acids (89, 90)/spongeLipid eL. infantum inhibition0.17, 0.34 mg/mLTopIB inhibitionESP, USA[90]
Antileishmanialdudawalamide D (82)/cyanobacteriumPeptide fL. donovani inhibition2.6 μMUndeterminedJOR, PAN, USA[84]
AntileishmanialGorgonia sp. sterol (91) spongeTerpenoid eL. infantum inhibition>10 μM *UndeterminedESP, PAN[91]
Antileishmanialircinin-1 and 2 (92, 93)/spongeTerpenoid eL. donovani inhibition28–31 μMUndeterminedCHE, DEU, ITA, TUR[92]
Antitrypanosomaljanadolide (94)/cyanobacteriumPeptide fT. b. brucei inhibition0.047 μMUndeterminedJPN[93]
Antitrypanosomalmalformin A1 (95)/fungusPeptide fT. congolense inhibition0.015 µg/mL UndeterminedJPN, PHL[94]
Antitrypanosomalrhodozepinone (96)/bacteriumAlkaloid fT. b. brucei inhibition16.3 µg/mLUndeterminedDEU, EGY[95]
Antituberculosismelophlin A (97)/spongeAlkaloid fM. smegmatis inhibition0.8 μg/mL +BCG1083 & BCG1321c proteins targetedIDN, JPN[96]
Antituberculosismethoxypuupehenol (98)/spongeTerpenoid eDormant M. tuberculosis inhibition0.5 μg/mL +Bactericidal activityUSA[97]
Antituberculosisgliotoxin (99)/fungusAlkaloid fM. tuberculosis inhibition0.03 μM +UndeterminedCHN[98]
Antituberculosisproximicin B (100)/bacteriumPeptide fM. bovis Pasteur 1173P2 inhibition6.25 μg/mL +UndeterminedAUS, CHN, EGY, NGA[99]
Antituberculosissmenothiazole A (101)/spongePeptide fM. tuberculosis H37Rv inhibition4.1 μg/mL +UndeterminedPOL, USA[100]
Antituberculosissporalactam B (62)/bacteriumPolyketide dM. tuberculosis inhibition0.06 μM **UndeterminedCAN, PHL[65]
Antituberculosistalaramide A (102)/fungusAlkaloid fMycobacterial PknG inhibition 55 μMUndeterminedCHN[101]
Antituberculosisviomellein (103)/fungusPolyketide dDormant M. bovis BCG inhibition1.56 μg/mL +UndeterminedIDN, JPN[102]
Antiviralhymenialdisine (104)/spongeAlkaloid fHIV-1 inhibition>3.1 μM * Reverse transcriptase inhibitionDEU, SAU, USA[103]
Antiviralmetachromin A (105)/spongeTerpenoid eHBV inhibition0.8 μMViral promoter inhibitionJPN, NLD[104]
Antiviralperidinin (106)/coralTerpenoid eHTLV-1 infected T cell inhibition0.7-5.4 μMNF-κB inhibitionJPN[105]
Antiviralspiromastilactone D (107)/fungusPolyketide dH1N1 influenza A virus inhibition6.0 μMHA-sialic acid receptor binding inhibitionCHN, USA[106]
Antiviralxiamycin D (108)/bacteriumTerpenoid ePEDV virus inhibition0.93 μMVirion structural proteins inhibitionS. KOR[107]
Antiviralzoanthone A (109)/sea anemoneTerpenoid eDENV-2 virus inhibition19.6 μMRNA pocket tunnel bindingTWN[108]
AntiviralA. polycladia aromatic sulfate (110)/crinoidPolyketide dHCV NS3 helicase inhibition5 μMUndeterminedJPN[109]
Antiviral alotaketal C (111)/spongeTerpenoid eHIV expression activation1 μM *UndeterminedCAN[110]
Antiviralaspergillipeptide D (112)/fungusPeptide fHSV-1 inhibition9.5 μMUndeterminedCHN[111]
Antiviralaspergilols H and I (113, 114)/fungusPolyketide dHSV-1 inhibition4.7, 6.2 μMUndeterminedCHN[112]
Antiviralasteltoxin E (115)/fungusPolyketide dH1N1 and H3N2 influenza virus inhibition3.5, 6.2 μMUndeterminedCHN[113]
AntiviralS. verruca cyclopentenone (116)/soft coralPolyketide dHIV infection inhibition5.8 μMUndeterminedCHN, USA[114]
Antiviraleutypellazine E (117)/fungusAlkaloid fHIV-1 inhibition3.2 μMUndeterminedCHN, DEU[115]
Antiviralω-hydroxyemodin (118)/fungusPolyketide dHCV NS3 protease inhibition10.7 μMUndeterminedEGY, SAU[116]
Antiviralmalformin C (119)/fungusPeptide fHIV infection inhibition1.4 μMUndeterminedCHN[117]
Antiviralmanzamine A (120)/sponge Alkaloid fHSV-1 inhibition 1 μM +UndeterminedUSA[118]
Antiviralperidinin (106)/zoanthidTerpenoid eAnti-dengue virus 2 inhibition4.5 μMUndeterminedTWN[119]
Antiviralstachybonoid A (121)/fungusTerpenoid eDengue virus prM protein expression inhibition25 μMUndeterminedCHN[120]
Antiviralsubergorgols T and U (122, 123)/soft coralTerpenoid eH1N1 influenza A virus inhibition35–37 μMUndeterminedCHN, NLD[121]
a Organism, Kingdom Animalia: ascidian (Phylum Chordata); gorgonian, coral, crinoids, sea anemone, zoanthid (Phylum Cnidaria); sea urchin (Phylum Echinodermata), nudibranch (Phylum Mollusca), sponge (Phylum Porifera); Kingdom Monera: bacterium, cyanobacterium (Phylum Cyanobacteria); Kingdom Fungi: fungus; Kingdom Protista: dinoflagellates; b1 IC50: concentration of a compound required for 50% inhibition in vitro, *: estimated IC50, **: in vivo study; + MIC: minimum inhibitory concentration, b2 MMOA: molecular mechanism of action; c Country: AUS: Australia; BEL: Belgium; BGD: Bangladesh; CAN: Canada; CHE: Switzerland; CHN: China; CRI: Costa Rica; DEU: Germany; DNK: Denmark; EGY: Egypt; ESP: Spain; FRA: France; GBR: United Kingdom; HUN: Hungary; IDN: Indonesia; IND: India; IRL: Ireland; IRN: Iran; ITA: Italy; JOR: Jordan; JPN: Japan; MEX: Mexico; MMR: Myanmar; NGA: Nigeria; NLD: The Netherlands; NOR: Norway; NZL: New Zealand; OMN: Oman; PAN: Panama; PHL: Philippines; POL: Poland; SAU: Saudi Arabia; S. KOR: South Korea; SAU: Saudi Arabia; SRB: Serbia; SWE: Sweden; TWN: Taiwan; TUR: Turkey; VNM: Vietnam; Chemistry: d Polyketide; e Terpene; f Nitrogen-containing compound; Abbreviations: AHL: acylated homoserine lactones; DENV-2: dengue virus type 2; HA: hemagglutinin; HBV: hepatitis B virus; HCV: hepatitis C virus; HSV: herpes simplex virus; MR: methicillin-resistant; PEDV: porcine epidemic diarrhea virus; PknG: mycobacterial protein kinase G; PNAG: polysaccharide poly-β-(1,6)-N-acetylglucosamine; TopIB: topoisomerase IB.

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

Table 1 presents 2016–2017 preclinical pharmacological research on the antibacterial, antifungal, antiprotozoal, antituberculosis, and antiviral activities of marine natural products (1123) shown in Figure 1.
Figure 1. Marine pharmacology in 2016–2017: marine compounds with antibacterial, antifungal, antiprotozoal, antituberculosis and antiviral activities.
Figure 1. Marine pharmacology in 2016–2017: marine compounds with antibacterial, antifungal, antiprotozoal, antituberculosis and antiviral activities.
Marinedrugs 19 00049 g001aMarinedrugs 19 00049 g001bMarinedrugs 19 00049 g001cMarinedrugs 19 00049 g001dMarinedrugs 19 00049 g001eMarinedrugs 19 00049 g001fMarinedrugs 19 00049 g001gMarinedrugs 19 00049 g001hMarinedrugs 19 00049 g001iMarinedrugs 19 00049 g001j

2.1. Antibacterial Activity

During 2016–2017, 51 studies reported on antibacterial bioactivity in marine natural products (167) isolated from ascidians, bacteria, fungi, mussels, sea urchins, soft corals, and sponges, global research that contributed new preclinical pharmacology that may contribute to the ongoing search of novel therapeutics for multi-drug resistant bacterial infections.
As shown in Table 1, and Figure 1, eleven publications reported on the mode of action of marine-derived antibacterial compounds. Cheng and colleagues reported a new chlorinated quinolone, ageloline A (1) isolated from a Streptomyces sp. derived from the Mediterranean sponge Agelas oroides that inhibited growth of Chlamydia trachomatis inclusions resulting from a mechanism that “might be related to its antioxidant potential” [20]. Davison and colleagues discovered a new analog of the peptidyl nucleoside antibiotic blasticidin S (2) produced by the actinomycete Streptomyces griseochromogenes that demonstrated increased potency against both Gram-positive and Gram-negative bacteria, with the NorA multidrug transporter being a key factor involved in membrane permeability that “facilitates cellular entry of peptidyl nucleosides” [21]. Le Norcy and colleagues described the activity of dibromohemibastadin-1 (3), derived from natural bastadins discovered in the marine sponge Ianthella basta, and which prevented and disrupted Gram negative bacterial biofilms without toxicity by a mechanism that involved “regulation in the quorum sensing” which is a process considered important for “biofilm formation and organization” [22]. Wyche and colleagues demonstrated in a series of in vitro and in vivo studies that the polyether antibiotic ecteinamycin (4) isolated from a marine-derived bacterium Actinomadura sp., showed significant activity against the toxigenic strains of Clostridium difficile and proposed that the mechanism of action leading to detoxification and cell death likely involved “potassium transport dysregulation” [23]. Tseng and colleagues identified 5-episinuleptolide (5) isolated from the soft coral Sinularia leptoclados as an inhibitor of biofilm-associated Gram-negative bacterium Acinetobacter baumannii infections, shown to be of high incidence in immunocompromised individuals, by a mechanism that correlated with a decreased production of the extracellular polysaccharide poly-β-(1,6)-N-acetylglucosamine [24]. Solstand and colleagues described two new antimicrobial peptides eecentrocin 1 and eestrongylocin 2 (6, 7) from the haemocytes of the edible sea urchin Echinus esculentus with potent antimicrobial activity against both Gram-positive and Gram-negative bacteria, concluding that “a genomic approach to discover homologues in other echinoderms for the discovery of novel antimicrobial peptides could be a beneficial venture” [25]. Sung and colleagues observed the increased production of the antibiotics granaticin and granatomycin D (8, 9) from marine-derived Streptomyces PTY08712 isolated from the tunicate Styela canopus when co-cultured with several human bacterial pathogens, concluding that “utilization of co-culture experiments…may enhance metabolite production and further our understanding of … microbial interactions” [26]. Adnani and colleagues reported the isolation of new antibiotic anthracycline keyicin (10) from a co-culture of marine bacteria Rhodococcus sp. and Micromonospora sp. derived from the ascidian Ecteinascidia turbinata and the sponge Chondrilla nucula, which was found to be selective for Gram-positive bacteria by a unique mechanism that “does not involve nucleic acid damage” [27]. Mokhlesi and colleagues discovered two new peptides microcionamides C and D (11, 12) from the marine sponge Clathria basilana which inhibited Gram-positive bacterial growth with a mechanism that suggested “dissipation of the bacterial membrane potential”, which is an electrical gradient across the bacterial cytoplasmatic membrane that is required for both ATP generation and active transport processes [28]. Leoni and colleagues reported that the novel peptide myticalin A5 (13) isolated from the mussel Mytilus galloprovincialis had antimicrobial activity against both Gram-positive and Gram-negative bacteria by a mode of action that appeared to involve inhibition of RNA synthesis but “remains to be elucidated” [29]. Costantino and colleagues isolated a new γ-lactone plakofuranolactone (14) from the extract of the marine sponge Plakortis cf. lita that inhibited the bacterial LasI/R quorum sensing system, a “mechanism of cell-cell communication and gene regulation in bacteria” [30]. Lee and colleagues studied the activity of psammaplin A (15), a marine natural product isolated from sponges, on Gram-negative bacterium Vibrio vulnificus infections noting that it suppressed cytotoxicity in vitro and prolonged both the survival and pathology of V. vulnificus-infected mice [31].
As shown in Table 1 and Figure 1, 51 marine natural products (16–67), some of them novel, were reported to exhibit antibacterial activity with MICs < 10 μg/mL or 10 μM against several Gram-positive and Gram-negative bacterial strains, although their respective mechanisms of action remained undetermined: A polycyclic macrolactone abyssomicin 2 (16) isolated from a South China Sea Streptomyces koyangensis CSIO 5802 [32]; three bioactive actinomycins (17–19) isolated from a marine-derived Streptomyces sp. [33,34]; a new lipopeptide aneurinifactin (20) produced by the Indian marine Aneurinibacillus aneurinilyticus SBP-11 [35]; two new 20-nor-isopimarane diterpenoids aspewenins D and H (21, 22), isolated from a South China Sea marine fungus Aspergillus wentii SD-310 [36]; a polyketide furanoterpenoid (23) isolated from the Indian brown seaweed-associated heterotrophic bacterium Bacillus subtilis MTCC 10403 [37]; a chromone polyketide derivative bacillisporin A (24) isolated from a Chinese mangrove-derived marine fungus Penicillium aculeatum (No.9EB) [38]; a new cyclic-lipotetrapeptide bacilotetrin A (25) isolated from a Korean marine sediment-derived bacterium Bacillus subtilis strain 109GGC020 [39]; a new polyketide branimycin B (26) isolated from a Cantabrian Sea abyssal (3000 m) actinobacterium Pseudonocardia carboxydivorans M-227 [40]; a novel disulfide diketopiperazine alkaloid brocazine G (27) was characterized from a mangrove-derived endophytic fungus Penicillium brocae MA-231 [41]; two new polyphenyl butenolides cadiolides K and M (28, 29) identified from the Korean marine tunicate Pseudodistoma antinboja [42]; a novel peptide cahuitamycin D (30) isolated from the Costa Rican marine-sediment derived bacterium Streptomyces gandocaensis [43]; a known macrolide chalcomycin (31) isolated from a Chinese marine-sediment derived bacterium Streptomyces sp. HK-2006-1 [44,45]; two novel spiromeroterpenoids chermesins A and B (32, 33) isolated from the Chinese marine red alga Pterocladiella tenuis-derived endophytic fungus Penicillium chermesinum EN-480 [46]; two novel chloro-preussomerins A and B (34, 35) isolated from the Chinese mangrove endophytic fungus Lasiodiplodia theobromae ZJ-HG1 [47]; a bipyridine collismycin C (36) isolated from the Micronesian marine red alga-derived bacterium Streptomyces sp. MC025 [48]; two new polyketides engyodontochones A and B (37, 38) isolated from the Croatian marine sponge Cacospinga scalaris-derived fungus Engyodontium album LF069 [49]; five new polyhydroxylated hydroanthraquinone derivates (39–43) isolated from the Chinese marine red alga Laurencia okamurai-derived fungus Talaromyces islandicus EN-501 [50]; a novel sesquiterpene hydroxyquinone langcoquinone C (44) isolated from a Vietnamese marine sponge Spongia sp. [51]; the sesterterpene luffariellolide (45) isolated from the Philippine marine sponge Suberea sp. [52]; three new manzamine alkaloids (46–48) isolated from an Indonesian marine sponge Acanthostrongylophora sp. [53]; a naphtoquinone terpenoid napyradiomycin A1 (49) isolated from the Sao Tome and Principe marine ascidian-derived actinomycete Streptomyces sp. strain CA-271078 [54]; a new polycyclic quinazoline alkaloid oxysporizoline (50) isolated form a Korean marine mudflat-derived fungus Fusarium oxysporum [55]; a 1-(2,6-dihydroxyphenyl)butan-1-one (51) obtained from the South China Sea mangrove-derived endophytic fungus Penicillium citrinum HL-5126 [56]; two novel polyketides penicillstressol and isopenicillstressol (52, 53) isolated from the Chinese marine sediment-derived fungus Penicillium sp. BB1122 [57]; a dechlorinated benzophenone polyketide pestalone (54) isolated from the South China Sea soft coral-derived fungus Pestalotiopsis sp. [58]; a known phenol pestaltionol (55) isolated from a Taiwanese marine hydrothermal vent sediment-derived Penicillium sp. Y-5-2 [59]; two novel diphenyl ether derivatives phomaethers A and C (56, 57) isolated from a South China Sea gorgonian-derived fungus Phoma sp. [60]; a polyketide 4-methyl-3”-prenylcandidusin A (58) isolated from the Malaysian marine coral Galaxea fascicularis-derived fungus Aspergillus tritici SP2-8-1 [61]; a mono-rhamnolipid (59) isolated from Ross Sea (Antarctica) sediments-derived Pseudomonas sp. BTN1 [62]; the sesquiterpene smenospongine (60) isolated from the South China Sea “purple-colored encrusting” marine sponge Dysidea sp. [63]; a methyl butyl propyl phthalate (61) isolated from Indian mangrove sediments-derived Streptomyces cheonanensis VUK-A [64]; a new macrolide sporalactam B (62) isolated from Northeastern Pacific Canadian marine sediment-derived Micromonospora sp. RJA4480 [65]; a glycosidic spirotetronate polyketide tetrocarcin A (63) isolated from a Chinese marine sediment-derived Micromonospora sp. 5-297 [66]; a novel pyridinium polyketide tricepyridinium (64) was isolated from Escherichia coli EPI300 clone pDC113 transfected with metagenomic DNA prepared from the Japanese marine sponge Discodermia calyx [68]; a new tetracyclic biscembrane trocheliane (65) isolated from the Red Sea soft coral Sarcophyton trocheliophorum [69]; a new indole alkaloid tulongicin (66) isolated from a Paluan deep-water marine sponge Topsentia sp. [70]; and a polyketide vineomycin A1 (67) isolated from the culture broth of a Taiwan Strait marine sediment-derived actinomycete Streptomyces sp. A6H [71].
Furthermore, during 2016–2017, several other marine natural products, some of them novel, reported antimicrobial activity in 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: an alkaloid haliclonadiamine from the Okinawan sponge Halichondria sp. (MIC = 10 µg/disk) [122]; a sesquiterpene alismol from a Red Sea soft coral Lobophytum sp. (MIC = 15 μg/mL) [123]; a new unicellane diterpene eunicellol A from the Arctic soft coral Gersemia fruticosa (MIC90 = 28–48 μg/mL) [124]; a new cembrane diterpene16-hydroxycembra-1,3,7,11-tetraene (MIC = 25 μg/mL) from the Malaysian soft coral Sarcophyton sp. [125]; a new bisindole alkaloid hyrtinadine D (MIC = 16 μg/mL) isolated from the Okinawan marine sponge Hyrtios sp. [126]; a briarane diterpenoid dichotelide O (MIC = 2 mg/mL) isolated from the South China Sea gorgonian Dichotella gemmacea [127]; two new sesquiterpene aminoquinones langcoquinones A and B isolated from the Vietnamese sponge Spongia sp. (MIC = 12.5 μM) [128]; a new lobane diterpenoid prenyl-α-elemenone (MIC = 20 μg/mL) from a Malaysian soft coral Sinularia sp. [129]; a new diterpene scaffold darwinolide (MIC = 33.2 μM) from the Antarctic sponge Dendrilla membranosa [130]; echinochrome A and spinochrome C (MIC = 22.5 μM) from sea urchins D. savignyi, T. gratilla, E.mathaei and T. pileolus from Madagascar [131], the meroterpenoid verruculide B2 (MIC = 32 μg/mL) from the fermentation broth of Penicillium sp. SCS-KFD09 isolated from the Chinese marine worm Sipunculus nudus [132]; a bioactive sterol (MIC = 20 μg/mL) from a Red Sea soft coral S. terspilli [133]; a new norditerpene citrovirin (MIC = 12.4 μg/mL) isolated from a marine algicolous funcus Trichoderma citrinoviride [134]; two new bisabolene sesquiterpenes asperchondols A and B (MIC = 25 μM) isolated from the marine sponge Chondrilla nucula [135]; long-chain peptaibols peptides (MIC = 25 μg/mL) by a French marine blue mussel-derived strain of Trichoderma longibrachiatum [136]; a polyketide antibiotic haliangicin (MIC > 32 μg/mL) and its analogues isolated from the marine myxobacterium Haliangium ochraceum SMP-2 [137]; two curvalin macrolides (MIC = 20 μg/mL) isolated from the broth of a cultured marine actinomycete Pseudonocardia sp. HS7 [138]; a new halimane-type diterpenoid micromonohalimane B (MIC = 40 μg/mL) isolated from a Micromonospora sp. cultivated from the marine ascidian S. brakenhielmi [139]; a azaphilonidal derivative penicilazaphilone C (MIC = 15.6 μg/mL) isolated from the marine fungus strain Penicillium sclerotiorum M-22 [140]; a new 2′-acetoxy-7-chlorocitreorosein anthraquinone (MIC = 10 μM) from a South China Sea endophytic marine fungus Penicillium citrinum HL-5126 [141]; a new antibacterial macrolide borrelidin C (MIC = 6 μM) from a saltern-derived holophilic Nocardiopsis sp. [142]; a new antichlamydial dimeric indole derivative (IC50 = 46.6–96.4 μM) isolated from the South China Sea sponge-derived actinomycete Rubrobacter radiotolerans [143]; the anthraquinone emodin (MIC = 32 μg/mL) isolated from a culture of the endophytic fungus Eurotium chevalieri KUFA 0006 [144]; new thiodiketopiperazines eutypellazines P-R (MIC = 16–32 μM) isolated form a deep see fungus Eutypella sp. [145]; a new lipophilic cyclic hexapeptide thermoactinoamide A (MIC = 35 μM) isolated from the Icelandic thermophilic bacterium Thermoactinomyces vulgaris [146] ; and a new neo-actinomycin A (MIC = 16 μg/mL) isolated from a Chinese marine-derived Streptomyces sp. IMB094 [147].

2.2. Antifungal Activity

Eleven studies during 2016–2017 reported on the antifungal activity of several marine natural products (68–79) isolated from marine bacteria, dinoflagellates, fungi and sponges, a slight increase from our last review [10], and previous reviews of this marine pharmacology series.
As shown in Table 1 and Figure 1, two reports investigated an antifungal marine compound with a novel mechanism of action. Espiritu investigated the polydroxy polyene antifungal amphidinol 3 (68) isolated from cultures of the marine dinoflagellate Amphidinium klebsii [72], observing that membrane integrity loss by direct interaction with membrane lipids showed “an absolute dependence on the presence of sterols”. Furthermore, Iwamoto and colleagues, using atomic force microscopy, determined that amphidinol 3 formed “different types of sterol-aided polymorphic channels in a concentration dependent manner” [73].
As shown in Table 1 and Figure 1, ten marine natural products (69–79) showed antifungal activity with MICs that were either less than 10 μg/mL, 10 μM, or 10 μg/disk but no mechanism of action studies were reported in the papers: a sesquiterpene avarol (69) isolated from the Mediterranean sponge Dysidea avara [74]; a polycyclic tetramic macrolactam dihydromaltophilin (70) isolated from a marine Conus miles-derived Streptomyces sp. CMB-CS038 [75]; a pair of enantiomeric sesterterpenoids (±) hippolide J (71a, 71b) isolated from the South China Sea sponge Hippospongia lachne [76]; a hybrid polyketide ilicicolin H (72) isolated from the Danish marine-derived fungus Stilbella fimetaria [77] ; new cyclic lipopetides iturin F1 and iturin F2 (73, 74) from the Korean saltern-derived marine Bacillus sp. KCB14S006 [78]; two macrolide polyketides PF1163A (75) and -B (76) isolated from a Japanese unidentified marine alga-derived fungus Penicillium meleagrinum var. viridiflavum [79]; a terpene plakinic acid M (77) isolated from the Bahamian sponge association Plakortis halichondrioides-Xestospongia deweerdtae [80]; a new steroidal saponin poecillastroside D (78) isolated from a Mediterranean deep-sea sponge Poecillastra compressa [81]; a nucleoside analog rocheicoside A (79) isolated from a Turkish marine sediment-derived actinomycete Streptomyces rochei 06CM016 [82]. Further characterization of the antifungal pharmacology of these marine-derived natural compounds will require mechanism of action studies.
In addition, several novel structurally-characterized marine molecules with antifungal MICs or IC50s greater than 10 μg/mL, 10 μM, or 10 μg/disk, have been excluded from Table 1 and Figure 1 because of their weaker bioactivity: a aminobisabolene (IC90 = 17–72 μM) isolated from the tropical Micronesian sponge Halichondria sp. [148]; several new bromopyrrole alkaloids (apparent IC50 = 20 μM) isolated from the South China sea sponge Agelas sp. [149]; and a novel polyketide kalkipyrone B (IC50 = 13.4 μM) isolated from American Samoan cyanobacterium Leptolyngbya sp. [150].

2.3. Antiprotozoal and Antituberculosis Activity

As shown in Table 1, and as reported in the 1998–2015 marine pharmacology reviews of this series [1,2,3,4,5,6,7,8,9,10], 22 studies contributed novel findings in 2016–2017 on the antiprotozoal (antimalarial, antileishmanial and antitrypanosomal) and antituberculosis pharmacology of structurally characterized marine natural products (80103).
Malaria, a global disease caused by protozoan genus Plasmodium (P. falciparum, P. ovale, P. vivax and P. malariae), currently affects over 2 billion people worldwide. Further contributing to the global antimalarial drugs research, 8 marine molecules (8088) isolated from bacteria, molluscs, sponges, soft corals, and ascidians were shown to possess antimalarial activity during 2016–2017. 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 (8088) although the mechanism of action for these compounds remained undertermined at the time of publication: a norterpene cyclic peroxide diacarperoxide A (80) isolated from the South China Sea sponge Diacarnus megaspinorhabdosa shown to be a potent inhibitor of both P. falciparum W2 and D6 clones [83]; cyclic depsipeptides dudawalamides A and D (81, 82) from the Papua New Guinean cyanobacterium Moorea producens with moderate antimalarial activity against chloroquine-resistant P. falciparum strain W2 [84]; the alkaloid eudistidine A (83) isolated from Paluan marine ascidian Eudistoma sp. that potently inhibited chlorquine-sensitive P. falciparum strain D6 and chloroquine-resistant strain W2 [85]; a novel dimeric diketopiperazine naseseazine C (84) isolated from an Australian marine-sediment derived Streptomyces sp. that exhibited moderate inhibitory activity against chloroquine-sensitive P. falciparum strain 3D7 [86]; a new N-hydroxylated 1,2,3,4-tetrahydro-β-carboline alkaloid analogue of the known 7-bromohomotrypargine (85) isolated from the New Zealand ascidian Pseudodistoma opacum with moderate activity against P. falciparum chloroquine-resistant strain FcB1 [87]; a new guanidine alkaloid ptilomycalin F (86) and known fromiamycalin (87) from the Madagascar marine sponge Monanchora unguiculata which demonstrated potent activity against P. falciparum strain 3D7 [88], and a new isocyanoditerpene pustulosaisonitrile-1 (88) from the Australian nudibranch Phyllidiella pustulosa which showed potent activity against P. falciparum strain 3D7 [89].
As shown in Table 1 and Figure 1, 9 marine compounds (83, 8996) isolated from bacteria, fungi, sponges were reported to possess bioactivity towards the so-called neglected protozoal diseases: Leishmaniasis, caused by the genus Leishmania (L.), amebiasis, trichomoniasis, as well as African sleeping sickness (caused by Trypanosoma (T.) brucei rhodesiense and T. brucei gambiense) and American sleeping sickness or Chagas disease (caused by T. cruzi).
Only one report described two antitrypanosomal marine chemicals as well as their mechanisms of action. Carballeira and colleagues examined the mode of action of novel very long-chain α-methoxylated fatty acids (89, 90), isolated from the Caribbean sponge Asteropus niger, and demonstrated that they were toxic towards L. infantum amastigotes and free living promastigotes by inhibition of Leishmania topoisomerase 1B enzyme considered “an important therapeutic target against L. infantum” [90]. In addition, seven marine natural products (82, 9196) exhibited antileishmanial and antiprotozoal activity, although their mechanisms of action remained undetermined: The cyclic depsipeptide dudawalamide D (82) from the Papua New Guinean cyanobacterium Moorea producens with potent against L. donovani. [84]; a new oxysterol (91), isolated from a Panamanian octocoral Gorgonia sp. that moderately reduced the multiplication of L. infantum promastigotes thus suggesting “antileishmanial efficacy against intracellular amastigotes” [91]; two linear furanosesterterpenoids ircinin-1 and -2 (92, 93) from the Turkish sponge Ircinia oros with moderate activity against L. donovani, suggesting the compounds’ “bifuran terminus… positively influences the in vitro antiprotozoal activity” [92]; a new cyclic polyketide-peptide hybrid janadolide (94) from the Japanese marine cyanobaterium Okeania sp. which demonstrated very potent activity against T. brucei brucei, thus revealing potential for development as “new antitrypanosomal drugs” [93]; the cyclic pentapeptide malformin A1 (95) isolated from the Philippine marine seagrass-derived fungus Aspergillus tubingensis IFM 63452 highly active towards the parasite T. congolense and recommended as ”an antiprotozoal agent” [94]; a novel azepino-diindole alkaloid rhodozepinone (96) isolated from a Red Sea marine sponge-derived bacterium Rhodococcus sp. UA13 s with moderate activity against T. brucei brucei TC221 and perhaps a “promising future contribution to drug discovery” [95].
The emergence of drug-resistant Mycobacterium tuberculosis has continued to stimulate an ongoing global search for novel therapeutic leads with novel mechanisms of action, and, as shown in Table 1 and Figure 1, during 2016–2017, 8 novel marine natural products (62, 97103), isolated from bacteria, sponges and fungi, generated promising pharmacological activity and thus contributed to the search for novel antituberculosis agents. Arai and colleagues identified the alkaloid melophlin A (97) isolated from the Indonesian marine sponge Melophlus sp. that demonstrated strong inhibitory activity against dormant M. smegmatis by targeting the “BCG1083 protein of putative exopolyphosphatase and the BCG1321c protein of diadenosine 5′,5”-P1,P4-tetraphosphate phosphorylase” [96]. Rodrigues Felix and colleagues isolated the polyketide 15-α-methoxypuupehenol (98) from the marine sponge Petrosia sp. that demonstrated potent antibacterial activity against dormant M. tuberculosis, highlighting a mode of action in which bacterial killing “is observed only for dormant but not metabolically active bacteria” [97].
As shown in Table 1 and Figure 1, additional six marine natural products (99103) exhibited antituberculosis activity, although their mechanisms of action remained undetermined: the alkaloid gliotoxin (99) derived from a deep-sea fungus Aspergillus sp. SCSIO Ind09F01 that strongly inhibited “at very low µM level” M. tuberculosis in vitro [98]; the polyketide proximicin B (100) isolated from the South China Sea sediments-derived Verrucosispora sp. MS100047 which demonstrated “a good anti-BCG activity” [99]; the hybrid peptide/polyketide smenothiazole A (101) isolated from a sponge consortium of Puerto Rican marine sponge Plakortis symbiotica-Xestospongia deweerdtae and identified as a “new lead compound with high activity” against M. tuberculosis H37Rv [100]; a new macrolide sporalactam B (62) isolated from Northeastern Pacific Canadian marine sediment-derived Micromonospora sp. RJA4480 reported to demonstrate “selective and potent inhibition of M. tuberculosis” [65]; an “unusual” alkaloid talaramide A (102) isolated from the mangrove endophytic fungus Talaromyces sp. HZ-YX1 that inhibited a mycobacterial protein kinase C required for localization of mycobacterial in macrophage [101], and a naphthoquinone dimer viomellein (103) produced by the Indonesian sponge-derived Aspergillus sp. that showed potent activity against dormant M. bovis BCG [102].

2.4. Antiviral Activity

As shown in Table 1 and Figure 1, 18 reports were published during 2016–2017 on the antiviral pharmacology of marine natural products (104123) against dengue virus, human immunodeficiency virus type-1 (HIV-1), human T-cell leukemia virus type 1 (HTLV-1), human herpes simplex virus (HSV), influenza virus, hepatitis C virus, and porcine epidemic diarrhea virus.
As shown in Table 1, 6 reports described antiviral marine chemicals and their mechanisms of action. O’Rourke and colleagues communicated that the alkaloid hymenialdisine (104), isolated from the Red Sea sponge Stylissa carteri inhibited HIV infection and while the retroviral reverse transcriptase was not inhibited, the investigators concluded that it could “serve as starting scaffold(s) for further investigation” [103]. Yamashita and colleagues discovered that the merosesquiterpene metachromin A (105) isolated from the marine sponge Dactylospongia metachromia significantly inhibited the production of hepatitis B virus (HBV) by affecting the activities of the viral core promoter and reducing the hepatic nuclear factor α protein, a mechanism that may contribute to “ameliorating HBV-related disorders in the liver” [104]. Ishikawa and colleagues reported that the carotenoid peridin (106) isolated from the Japanese coral Isis hippuris inhibited the proliferation and survival of HTLV-1-infected T-cell lines by a mechanism that involved “suppression of NF-κB and Akt signaling” suggesting the compound was a “promising drug for HTLV-1-associated diseases” [105]. Niu and colleagues determined that the phenolic lactone spiromastilactone D (107) isolated from a South Atlantic deep-sea (2869 m) sediment-derived fungus Spiromastix sp. MCCC 3A00308 demonstrated “broad anti-influenza spectrum” by a mechanism that targeted viral attachment and entry by affecting “hemagglutinin protein-sialic acid receptor interaction” and viral genome replication by “targeting the viral RNP complex” [106]. Kim and colleagues investigated the indolosesquiterpenoid xiamycin D (108) isolated from the Korean saltern-derived halophilic actinomycete Streptomyces sp. strain HK18 and observed it displayed potent inhibition of porcine epidemic diarrhea virus by inhibiting genes encoding several essential structural proteins required for PEDV replication, thus demonstrating novel and “promising skeletons against PEDV-related viruses” [107]. Cheng and colleagues isolated a new ecdysone terpenoid zoanthone A (109) from a Taiwanese sea anemone Zoanthus spp. that demonstrated good activity against dengue virus 2 by a mechanism that inhibited viral replication by blocking the C-terminal RNA-dependent RNA polymerase domain of NS5, the “largest and the most conserved (non-structural) protein” of the virus [108].
An additional 15 marine natural products (106, 110123), listed in Table 1 and shown in Figure 1, demonstrated antiviral activity, but the mechanism of action of these compounds remained undetermined at the time of publication: A new aromatic terpenoid (110) isolated form the Japanese marine crinoid Alloeocomatella polycladia which showed moderate activity against the hepatitis C virus (HCV) NS3 helicase [109]; a known sesterterpenoid alotaketal C (111) isolated from the Canadian marine sponge Phorbas sp. shown to activate latent HIV-1 provirus expression [110]; a new cyclic pentapeptide aspergillipeptide D (112) isolated from the South China Sea gorgonian Melitodes squamata-derived fungus Aspergillus sp. SCSIO 41501 with moderate activity against herpes virus simplex type 1 (HSV-1) [111]; new anthraquinones aspergilols H and I (113, 114), isolated from a South China Sea deep sea sediment (2326 m)-derived fungus, Aspergillus versicolor SCSIO 41502, with moderate activity against HSV-1 [112]; a new polyketide asteltoxin E (115) isolated from a Chinese marine sponge Callyspongia sp.-derived fungus Aspergillus sp. SCSIO XWS02F40 with moderate activity against influenza virus subtype H1N1 and H3N2 [113]; a novel cyclopentenone derivative (116) was isolated from the South China Sea soft coral Sinularia verruca which was “moderately protective” against the cytopathic activity of in vitro HIV-1 infection [114]; a new thiodiketopiperazine-type alkaloid eutypellazine E (117) isolated form a South Atlantic deep sea (5610 m) sediment-derived fungus Eutypella sp. MCCC 3A00281 moderately inhibited in vitro HIV-1 infection [115]; a tricyclic anthraquinone ω-hydroxyemodin (118) isolated from the Red Sea brown alga Padina pavonica-derived fungus Fusarium equiseti moderately inhibited hepatitis C virus NS3/4A serine protease in vitro [116]; the known polyketide malformin C (119) isolated from the marine-derived fungus Aspergillus niger SCSIO JcswF30 potently inhibited HIV-1 infection in vitro [117]; the known marine β-carboline alkaloid manzamine A (120) isolated from the Indo-Pacific sponge Acanthostrongylophora sp. potently inhibited HSV-1 replication and release in vitro, observing “that manzamine A had optimal structure features for anti-HSV-1 activity” [118]; the carotenoid peridin (106) isolated from the Taiwanese zoanthid Palythoa mutuki moderately inhibited all serotypes of dengue virus in vitro [119]; a new meroterpenoid stachybonoid A (121) isolated from the Chinese marine crinoid Himerometra magnipinna-derived fungus Stachybotrys chartarum 952, which evidenced moderate activity against denge virus replication and expression of prM protein [120], and two new pregnane-type steroids subergorgol T and U (122, 123) isolated from a South China Sea gorgonian coral Subergorgia suberosa that moderately inhibited influenza virus strain A/WSN/33 (H1N1) in vitro [121].

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

Table 2 presents the 2016–2017 preclinical pharmacology of marine chemicals (124234), which demonstrated either antidiabetic or anti-inflammatory activity, as well as affected the immune or nervous system, and whose structures are depicted in Figure 2.
Table 2. Marine pharmacology in 2016–2017: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Table 2. Marine pharmacology in 2016–2017: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Drug ClassCompound/
Organism a
ChemistryPharmacological ActivityIC50 bMMOA cCountry dReferences
Antidiabeticagelasine G (124)/spongeAlkaloid-terpenoid fPTP1B inhibition15 µM *Akt insulin pathway increaseIDN, JAP[151]
AntidiabeticBDDE (125)/algaShikimate hDecrease glucose levels in vivo10 mg/kg **PTP1B expression inhibitionCHN[152]
Antidiabeticdieckol (126)/algaShikimate hDecrease in glucose levels1 µg/g **Akt insulin pathway increaseS. KOR[153]
Antidiabeticgombasterol E (127)/spongeTerpenoid f Enhanced glucose uptake in vitro 20 µM *AMPK phosphorylation increaseS. KOR[154]
Antidiabeticleptolide (128)/soft coralTerpenoid fMurine glucose tolerance and insulin sensitivity increased 0.1 mg/kg **PKB phosphorylationESP[155]
Antidiabeticnectriacids B and C (129, 130)/fungusPolyketides dα-glucosidase inhibition23.5, 42.3 µMC-12 carboxyl esterification requiredCHN[156]
Antidiabeticpenicilliumin B (131)/fungusTerpenoid fGlomerular mesangial cells fibrogenic inhibition0.5 µM *NADPH oxidase inhibitionCHN[157]
Antidiabeticwailupemycin I (132)/bacteriumPolyketide dα-glucosidase inhibition8.3 µMCompetitive inhibitionCHN[158]
Antidiabeticasperentin B (133)/fungusPolyketide dPTP1B inhibition2 μMUndeterminedDEU[159]
Antidiabeticlasiodiplactone A (134)/fungusPolyketide dα-glucosidase inhibition29.4 μMUndeterminedCHN[160]
Antidiabetic7-hydroxy-de-O-methyllasiodiplodin (135)/bacteriumPolyketide dα-glucosidase inhibition25.8 μMUndeterminedCHN[161]
Antidiabeticsescandelin B (136)/fungusPolyketide d α-glucosidase inhibition17.2 μMUndeterminedCHN[162]
Anti-inflammatoryAMT-E (127)/algaTerpenoid fMurine colitis inhibition 10 mg/kg **Inhibition of TNF-α, IL-6ESP, MAR[163]
Anti-inflammatoryBacillus sp. diketopiperazines (138140)/bacteriumPeptide gTGFBIp inhibition in vivo5 μM **Septic responses inhibitionS. KOR[164,165]
Anti-inflammatory6-bromoisatin (141)/molluscAlkaloid gLung inflammation inhibition in vivo0.05 mg/g **Inhibition of TNF-α, IL-6AUS[166]
Anti-inflammatoryceylonamide A (142)/spongeTerpenoid f Macrophage RANKL inhibition13 μM *SAR completedIDN, JPN, NLD[167]
Anti-inflammatorycitrinin H1 (143)/fungusPolyketide eMicroglia NO and PGE2 release inhibition8 μMNF-κB inhibitionS. KOR, VNM[168]
Anti-inflammatory nonenolide derivative (144)/algaPolyketide eBMDC cytokine release inhibition7.6–10.9 μMJNK, ERK, AP-1, NF-κB inhibitionS. KOR[169]
Anti-inflammatorycucumarioside A2-2 (145)/sea cucumberTerpenoid fBinding of macrophage P2X purinergic receptors0.02 μM *Induction Ca2+ oscillationsRUS[170]
Anti-inflammatorycurvularin derivative (146)/fungusPolyketide eMacrophage PGE2 and NO release inhibition1.9–2.7 μMNF-κB signaling inhibition S. KOR[171]
Anti-inflammatory9,11-dihydrogracilin A (147)/spongeTerpenoid fPBMC proliferation inhibition3 μM *IL-6 and IL-10 inhibitionITA[172]
Anti-inflammatorydysivillosin A (148)/spongeTerpenoid fBasophil β-hexosaminidase inhibition8.2 μMIL-4 and LTB4 inhibitionCHN[173]
Anti-inflammatoryepinecidin-1 (149)/fishPeptide gInhibition of MyD88 protein levels6 µg/mL *Proteasome degradation requiredTWN[174]
Anti-inflammatoryexcavatolide B (150)/soft coralTerpenoid fAttenuation of rat arthritis [175]2.5, 5 mg/kg **Decreased MMP-2, MMP-9, CD11b in tissuesCHN, TWN[175]
Anti-inflammatoryfucoxanthin (151)/algaTerpenoid fDecreased mice paw edema, adipogenesis and ear inflammation4 mg/kg **Modulation of iNOS, PLA2, COX-2, ACC, IL-6 and Nrf2 expression JPN, S. KOR, MEX[176,177,178,179]
Anti-inflammatoryH. crispa peptide (152)/sea anemonePeptide gMacrophage histamine receptor inhibition10 μM *Intracellular Ca2+ increase inhibitionRUS[180]
Anti-inflammatoryhipposponlachnin B (153)/spongeTerpenoid f Basophil β-hexosaminidase inhibition24 μMIL-4 and LTB4 inhibitionCHN[181]
Anti-inflammatoryogipeptins A-D (154157)/bacteriumPeptide gMacrophage TNF-α production inhibition1 µM *Block LPS binding to CD14 JPN[182]
Anti-inflammatoryoscarellin (158)/spongeAlkaloid gMacrophage TNF-α and IL-6 expression inhibition>10 μMJNK, ERK, AP-1, NF-κB inhibitionS. KOR, USA[183]
Anti-inflammatorypseudane-VIII (159)/bacteriumAlkaloid g Macrophage NO release inhibition6 µM *iNOS and IL-1β inhibitionS. KOR[184]
Anti-inflammatoryacremeremophilane B (160)/fungusTerpenoid fMacrophage NO release inhibition8 µMUndeterminedCHN, DEU[185]
Anti-inflammatoryactinoquinolines A and B (161, 162)/bacteriumAlkaloid gCOX-1 and -2 inhibition1.4–7.6 μMUndeterminedEGY, USA[186]
Anti-inflammatoryanthenoside O (163)/starfishTerpenoid fMacrophage SOX inhibition>10 µM *UndeterminedRUS, VNM[187]
Anti-inflammatoryaurasperone C (164)/fungusPolyketide eCOX-2 inhibition4.2 μMUndeterminedCHN[188]
Anti-inflammatorybriarenolides M and N (165, 166)/soft coralTerpenoid fMacrophage iNOS expression inhibition10 μM *UndeterminedTWN[189]
Anti-inflammatorybriarenolides ZII and ZVI (167, 168)/soft coralTerpenoid fMacrophage iNOS and COX-2 expression inhibition10 μM *UndeterminedTWN[190]
Anti-inflammatorydihydrobipolaroxin (169)/fungusTerpenoid fMacrophage NO release inhibition>12.5 μM *UndeterminedCHN[191]
Anti-inflammatoryechinulin (170)/fungusAlkaloid gMicroglia NO release inhibition4.6 μMUndeterminedCHN, S. KOR[192]
Anti-inflammatory5α-iodozoanthenamine (171)/zoanthidAlkaloid gNeutrophil SOX and elastase inhibition>10 μM *UndeterminedTWN[193]
Anti-inflammatoryklyflaccisteroid J and K (172, 173)/soft coralTerpenoid fNeutrophil SOX and elastase inhibition1.5–5.8 μM UndeterminedTWN[194,195]
Anti-inflammatoryL. varium diterpenoid (174)/soft coralTerpenoid fNeutrophil elastase inhibition>10 μM *UndeterminedEGY, SAU, TWN[196]
Anti-inflammatorypetasitosterones B and C (175, 176)/soft coralTerpenoid fNeutrophil SOX and elastase inhibition2.7–4.4 µMUndeterminedTWN[197]
Anti-inflammatoryPinnnigorgia sp. sterols (177, 178)/soft coralTerpenoid fMacrophage COX-2 and iNOS expression inhibition10 µM *UndeterminedTWN[198]
Anti-inflammatorypinnigorgiol A (179)/soft coralTerpenoid fNeutrophil SOX and elastase inhibition4, 5 µMUndeterminedTWN [199]
Anti-inflammatorypinnigorgiol E (180)/soft coralTerpenoid fNeutrophil SOX and elastase inhibition1.6, 3.9 µMUndeterminedTWN[200]
Anti-inflammatorypinnisterols A and H (181, 182)/soft coral Terpenoid fNeutrophil SOX and elastase inhibition2.3–3.3 µMUndeterminedTWN[201,202]
Anti-inflammatoryplancipyrroside B (183)/starfishTerpenoid fMacrophage iNOS expression inhibition5.9 µMUndeterminedRUS, VNM[203]
Anti-inflammatoryprotolinckioside A (184)/starfishTerpenoid fMacrophage SOX inhibition10 µMUndeterminedIND, RUS[204]
Anti-inflammatorysarcophytonolide O (185)/soft coralTerpenoid fMacrophage iNOS expression inhibition8 μMUndeterminedCHN, USA[205]
Anti-inflammatorysinularectols A and B (186, 187)/soft coralTerpenoid fNeutrophil SOX and elastase inhibition0.9–8.5 μMUndeterminedTWN[206]
Anti-inflammatorysinubrasolides A and D (188, 189)/soft coralTerpenoid fNeutrophil SOX and elastase inhibition1.4–8 μMUndeterminedCAN, SAU, TWN[207,208]
Anti-inflammatoryuprolides N, O and P (190192)/soft coralTerpenoid fMacrophage TNF-α and IL-6 release inhibition1.4–4.2 μMUndeterminedIND, PAN, USA[209]
Immune systemcucumarioside A2-2 (145)/sea cucumberTerpenoid fIncrease in spleen white pulp and macrophage activation3 mg/kg **Increased B cell PCNA and M1 macrophages RUS, TWN[210,211]
Immune systemgracilins A, H and L
(193195)/sponge
Terpenoid fCD147 receptor modulation and T-cell IL-2 release inhibition1 μM *Hypersensitivity and NFATc inhibitionESP, GBR[212,213]
Immune systemshinorine and porphyra-334 (196, 197)/algaPeptide gNF-κB stimulation50 µg/mL *Tryptophan metabolism modulationAUT[214]
Immune systemsinulariolide (198)/soft coralTerpenoid fDendritic cell maturation suppression25 µg/mL *IL-6, IL-12 and NO inhibitionTWN[215]
Immune systemCDMW-3 (199)/fungusPeptide gPCA inhibition in vivo20 mg/kg **Mast cell histamine and cytokine release inhibitionCHN[216]
Immune systemchrysamide C (200)/fungusAlkaloid gIL-17 inhibition>1 μM *UndeterminedCHN[217]
Immune systemcocosolide (201)/cyanobacteriumPolyketide dIL-2 inhibition2.5 μM *UndeterminedCHN, USA[218]
Immune systemmyxillin A (202)/spongeAlkaloid gIL-12p40 release inhibition10 µg/mL *UndeterminedDNK, ISL[219]
Immune systempectinioside A (203)/starfishTerpenoid fIncrease OVA-specific IgG1 in vivo25 µg *UndeterminedJPN[220]
Immune systempeniphenone (204)/fungusPolyketide dLymphocyte immune suppression8.1–9.3 µg/mLUndeterminedCHN[221]
Immune systemUSF-19A (205)/bacteriumPeptide gSplenocyte IL-5 release inhibition0.57 μM UndeterminedCHN[222]
Nervous systemAPETx4 (206)/sea anemonePeptide gKv10.1 potassium channel inhibition1.1 µMBinds channel in closed stateBEL, DEU [223]
Nervous systemastaxanthin (207)/shrimpTerpenoid fPenitrem A toxicity reversal 20 µM *Block BK channel EGY, USA[224]
Nervous systemcrambescidin 816 (208)/spongeAlkaloid gCortical neurons cytosolic Ca2+ increase10 µM *AMPA and NMDA receptors involvedESP, FRA, IRL[225]
Nervous systemC. generalis O-conotoxin (209)/cone snailPeptide gΑ9α10 nACh receptor inhibition16.2 nMNon-competitive inhibitionAUS, CHN[226]
Nervous systemC. princeps PiVIIA peptide (210)/cone snailPeptide gNeuronal Ca2+ current increase3 µM *Potentiates two types Ca2+ channelsCUB, MEX[227]
Nervous systemconorphin T (211)/cone snailPeptide gKOR agonist9.8 µMIn vivo colonic receptor inhibitionAUS[228]
Nervous system11-dehydrosinulariolide (212)/soft coralTerpenoid fAmelioration PD and spinal cord injury attenuation 5 µg/rat **DJ-1 expression upregulation and microglia activationTWN[229,230]
Nervous systemdiscorhabdin G (213)/spongeAlkaloid g Eel and human AChE inhibition1.3 µMReversible competitive inhibitionDEU, ITA, SVN[231]
Nervous systemfucoxanthin (151)/algaTerpenoid fBACE1 inhibition5.3 µMMixed inhibitionGBR, S. KOR[232]
Nervous systemfucoxanthin (151)/algaTerpenoid fReversal BDNF expression50 mg/kg **Reversed AChE activityCHN[233]
Nervous systemfucoxanthin (151)/algaTerpenoid fNeuroprotection after TBI-induced brain injury100 mg/kg **Nrf2-ARE pathway modulationCHN[234]
Nervous system5-hydroxycyclopenicillone (214)/fungusPolyketide dH2O2-induced neuronal death protection30 µM *DHHP free radical inhibitionCHN, USA[235]
Nervous systemmaitotoxin (215)/alga Polyketide dActivation of NSCC 10 pM *TRPC1 inhibitionMEX[236]
Nervous systemmakaluvamine J (216)/spongeAlkaloid gReduction of mitochondrial damage0.1–1 µM *Nrf2 activationESP, FJI, GBR[237]
Nervous systemMEC-1 (217)/spongePolyketide dAChE inhibition20.9 µMDocking studiesEGY[238]
Nervous systemmellpaladine A (218)/ascidianAlkaloid gIn vivo behavior modulation8 nM/mouse **Serotonin receptor affinityJPN[239]
Nervous systemMs 9a-1 peptide (219)/sea anemonePeptide gDecrease in nociceptive and inflammatory response in vivo0.3 mg/kg **TRPA1 modulationNOR, RUS[240]
Nervous tissuephlorofucofuroeckol-A (220)/algaPolyketide dGlutamate-induced neurotoxicity inhibition10 µM *Intracellular and mitochondrial ROS inhibtionS. KOR[241]
Nervous systempiloquinone (221)/bacteriumPolyketide dMAO-B inhibition1.2 µMReversible competitive inhibitionS. KOR, USA[242]
Nervous systempseudopterosin A (222)/soft coralTerpenoid fSynaptic transmission alteration1 µM *Extensive brain distribution USA[243]
Nervous systemsqualamine (223)/sharkTerpenoid fReduction of α-synuclein aggregation in vivo50 µM **α-synuclein displaced from lipid membranesESP, GBR, ITA, NLD, USA[244]
Nervous systemstryphnusin (224)/spongeAlkaloid gEel AChR inhibition232 µMReversible competitive inhibitionHRV, NOR, SVN, SWE[245]
Nervous systemxyloketal B (225)/fungusPolyketide dCerebral infarction modulation50 mg/kg **Decreased ROS and cytokinesCAN, CHN, USA[246]
Nervous systemaraplysillin X (226)/spongeAlkaloid gBACE1 inhibition31.4 µMUndeterminedNZL, USA[247]
Nervous systemcaracolamide A (227)/cyanobacterium Alkaloid gCa2+ channel modulation10 pM *UndeterminedBRA, JOR, PAN, USA, [248]
Nervous systemconorfamide-Sr3 (228)/snailPeptide gBlocks volatage-gated K+ channel2.7 µM ***Shaker channel specificMEX[249]
Nervous systemcontryphan-Bt (229)/cone snailPeptide gStiff-tail syndrome in vivo5 ng/mouse **UndeterminedCHN[250]
Nervous systemdehydroaustin (230)/fungusMeroterpenoid fAChE inhibition0.4 µMUndeterminedCHN[251]
Nervous systemhymenidin (231)/spongeAlkaloid gKv1.3- Kv1.6 K+ channel inhibition2.5–7.6 µMUndeterminedBEL, GBR, SVN[252]
Nervous systempsammaplysene A (232)/spongeAlkaloid gBinding to RNA-binding protein HNRNPK86.2 µM ***UndeterminedUSA[253]
Nervous systemterreulactone C (233)/fungusMeroterpenoid fAChE inhibition28 nMUndeterminedCHN[254]
Nervous systemturripeptide (234)/turrid snailPeptide gΑ9α10 nAChR inhibition10.2 µMUndeterminedAUS, KAS, MEX, PHL, USA[255]
a Organism:Kingdom Animalia: shrimp (Phylum Arthropoda); ascidian, fish (Phylum Chordata; coral, sea anemone and zoanthid (Phylum Cnidaria); sea cucumber, starfish (Phylum Echinodermata); cone snail, turrid snail (Phylum Mollusca); sponge (Phylum Porifera); Kingdom Fungi: fungus; Kingdom Plantae: alga; diatoms; Kingdom Monera: bacterium; b IC50: concentration of a compound required for 50% inhibition, *: apparent IC50, ** in vivo study; ***: Ki: concentration needed to reduce the activity of an enzyme by half; c MMOA: molecular mechanism of action; d Country: AUS: Australia; AUT: Austria; BEL: Belgium; BRA: Brazil; CAN: Canada; CHN: China; CUB: Cuba; DEU: Germany; DNK: Denmark; EGY: Egypt; ESP: Spain; FJI: Fiji; FRA: France; GBR: United Kingdom; HRV: Croatia; IDN: Indonesia; IND, India; ISL: Iceland; IRL: Ireland; ITA: Italy; JOR: Jordan; JPN: Japan; KAS: Kazakhstan; MAR: Morocco; MEX: Mexico; NZL: New Zealand; NLD: Netherlands; NOR: Norway; PAN: Panama; PHL: Philippines; RUS: Russian Federation; SAU: Saudi Arabia; S. KOR: South Korea; SVN: Slovenia; SWE: Sweden; TWN: Taiwan; VNM: Vietnam; Chemistry: e Polyketide; f Terpene; g Nitrogen-containing compound; h Shikimate. Abbreviations: ACC: acetyl-CoA carboxylase; Ach: acetylcholine; AChE: acetylcholinesterase; Akt: also known as protein kinase B is a serine/threonine protein kinase; AMPA: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; AMPK: AMP-activated protein kinase; AMT-E: 11-hydroxy-1′-O-methylamentadione; AP-1: activator protein-1; BACE1: beta secretase aspartic protease; BDDE: Bis (2,3-dibromo-4,5-dihydroxybenzyl) ether; BK: voltage-gated potassium channels; BMDC: bone marrow-derived dendritic cells; CD14: cluster of differentiation 14; COX: cyclooxygenase; DHHP: α, α-diphenyl-β-picrylhydrazyl; DJ-1: cysteine protease encoded by PARK7 gene; ERK: extracellular signal-regulated kinase; HNRNPK: heterogenous nuclear ribonucleoprotein K; IL: interleukin; iNOS: inducible nitric oxide synthase; JNK: c-Jun NH2-terminal kinase; LPS: Lipopolysaccharide; MAO-B: monoamine oxidase B; MyD88: myeloid differentiation primary response protein 88; iNOS: inducible nitric oxide synthase; KOR: κ-opioid receptor; Kv current: voltage-gated K+ current; nAChR: nicotinic acetylcholine receptor; NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells; NMDA: N-methyl-d-aspartate receptor; NO: nitric oxide; nAChR: nicotinic acetylcholine receptor; Nrf2-ARE: nuclear transcription factor E2-related factor antioxidant response element; NSCC: non-selective cation channel; PBMC: human peripheral blood mononuclear cells; PCA: passive cutaneous anaphylaxis; PCNA: proliferating cell nuclear antigen; PD: Parkinson’s disease; PKB: protein kinase B; PLA2: phospholipase A2; PTP1B: tyrosine protein; phosphatase 1B; RANKL: receptor activator of nuclear factor-κB ligand; ROS: reactive oxygen species; SOX: superoxide; SQDC: sulfoquinovosyl diacylglycerols; TGFBIp: transforming growth factor β- induced protein; TNF-α: tumor necrosis factor-α; TRPA1: transient receptor potential ankyrin-repeat 1 receptor; TRPC1: transient receptor potential canonical type 1.
Figure 2. Marine pharmacology in 2016–2017: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Figure 2. Marine pharmacology in 2016–2017: marine compounds with antidiabetic and anti-inflammatory activity; and affecting the immune and nervous system.
Marinedrugs 19 00049 g002aMarinedrugs 19 00049 g002bMarinedrugs 19 00049 g002cMarinedrugs 19 00049 g002dMarinedrugs 19 00049 g002eMarinedrugs 19 00049 g002fMarinedrugs 19 00049 g002gMarinedrugs 19 00049 g002h

3.1. Antidiabetic Activity

As shown in Table 2 and Figure 2, twelve publications reported on the mode of action of marine-derived antidiabetic compounds (124132) during 2016–2017. Yamazaki and colleagues contributed to the pharmacology of diabetes by noting that the diterpene marine alkaloid agelasine G (124) isolated from the Japanese marine sponge Agelas nakamurai selectively inhibited protein tyrosine phosphase B (PTP1B) and enhanced insulin-stimulated phosphorylation of serine/threonine protein kinase B or Akt in vitro, noting that further studies may “provide a candidate for anti-diabetes therapeutic agents” [151]. Xu and colleagues observed that a novel bromophenol bis (2,3-dibromo-4,5-dihydroxybenzyl) ether (BDDE) (125) isolated from the red alga Odonthalia corymbifera, increased glucose uptake in vitro, decreased the expression of protein tyrosine phosphatase 1B, activated the insulin signaling pathway and in mice “significantly decreased the blood glucose”, thus, suggesting BDDE might be a “treatment of type-2 diabetes” [152]. Kim and colleagues reported that the marine algal polyphenol dieckol (126), isolated from the marine brown alga Ecklonia cava attenuated blood glucose levels in the zebrafish model of hyperglycemia, an in vivo paradigm used to study chronic diseases, such as diabetes, by a mechanism that stimulated the protein kinase B (Akt) pathway, thus demonstrating the “anti-diabetic effect” of this compound [153]. Woo and colleagues showed that the novel polyoxygenated steroid gombasterol E (127) isolated from the Korean marine sponge Clathria gombawuiensis moderately enhanced both glucose uptake in adipocytes and phosphorylation of AMP-activated protein kinase and acetyl-CoA carboxylase in mouse skeletal myoblasts [154]. Villa-Pérez and colleagues reported that the furanocembranolide diterpenoid leptolide (128) isolated from soft coral Leptogorgia alba improved insulin sensitivity by increasing intracellular insulin signaling in both liver and skeletal muscle tissues of a diet-induced obese mice model, a “preclinical model of insulin resistance”, concluding that “furanocembranolides as a new therapeutic class to treat Type 2 diabetes” [155]. Cui and colleagues demonstrated that two new polyketides nectriacids B and C (129, 130) isolated from a South China Sea mangrove Sonneratia ovata-derived endophytic fungus Nectria sp. HN001 moderately inhibited α-glucosidase, a significant finding because this enzyme prevents “breaking down complex carbohydrates for absorption” [156]. Lin and colleagues discovered a new sesquiterpene penicilliumin B (131) isolated from South China Sea deep sea (1300 m) sediment-derived Penicillium strain F00120 that potently inhibited kidney fibrogenic action of high glucose in vitro through oxidative stress disruption, thus, suggesting this compound had potential for “therapy of diabetic nephropathy” [157]. Chen and colleagues studied a polyketide wailupemycin I (132) isolated from a Chinese marine alga Enteromorpha prolifera-derived Streptomyces sp. OUCMDZ-3434 that moderately inhibited α-glucosidase by competitive inhibition of the enzyme [158].
Moreover, four marine natural products (133136) listed in Table 2 and shown Figure 2, demonstrated antidiabetic activity during 2016–2017, but the mechanism of action of these compounds remained undetermined at the time of publication: a new polyketide asperentin B (133) isolated from a deep (2769 m) Mediterranean Sea sediment-derived Aspergillus sydowii which “strongly” inhibited human protein tyrosine phosphatase 1B, an important “target for the treatment of type 2 diabetes” [159]; a novel lactone lasiodiplactone (134) isolated from a South China Sea mangrove Acanthus ilicifolius-derived endophytic fungus Lasiodiplodia theobromae ZJ-HQ1 that exhibited α-glucosidase inhibitory activity that was better “than the clinical α-glucosidase inhibitor acarbose” [160]; a new de-O-methyllasiodiplodin (135) isolated from the co-cultivation of the Chinese mangrove Clerodendrum inerme-derived endophytic fungus Trichoderma sp. 307 and the aquatic pathogenic bacterium Acinetobacter johnsonii B2 that inhibited α-glucosidase better “than the positive control acarbose” [161]; and a known isocoumarin analogue sescandelin B (136) isolated form the Chinese Kandelia obovata-derived endophytic fungus Talaromyces amesolkiae, which showed moderate inhibition of α-glucosidase, “which was much better than acarbose” [162].

3.2. Anti-Inflammatory Activity

As shown in Table 2 and Figure 2, there was a remarkable increase in anti-inflammatory pharmacology of marine compounds (127, 138192) during 2016–2017. The molecular mechanism of action of anti-inflammatory marine natural products (127, 138159) was assessed in both in vitro and in vivo preclinical pharmacological studies in twenty one papers, which used several in vitro and in vivo models of inflammation.
Zbakh and colleagues evaluated the anti-inflammatory properties of the meroterpene 11-hydroxy-1′-O-methylamentadione (AMT-E) (127) isolated from the brown alga Cystoseira usneoides in a murine model of experimental colitis, observing that the levels of myeloperoxidase, cytokines and the expression of the pro-inflammatory genes nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2) were reduced, concluding that AMG-W might be a candidate for “prevention/treatment of inflammatory bowel disease” [163]. Jung and colleagues investigated three known diketopiperazines (138140) isolated from a Korean marine sediment-derived bacteria Bacillus sp. HC001 and Piscicoccus sp. 12L081 and observed that they effectively inhibited LPS-induced release and expression of transforming growth factor β-induced protein in human endothelial cells, as well as protected against cecal ligation and puncture-induced lethality and organ damage in a murine model of experimental sepsis [164,165]. Ahmad and colleagues reported that the orally administered brominated indol 6-bromoisatin (141) isolated form the Australian marine gastropod mollusc Dicathais orbita inhibited inflammation in a murine model of lipopolysaccharide-induced acute lung injury by significantly reducing pro-inflammatory tumor necrosis-α (TNF-α) and interleukin-1 β (IL-1β) production and associated lung damage, thus, concluding this compound “provides a lead for the development of safer anti-inflammatory drugs” [166]. El-Desoky and colleagues studied a new nitrogenous diterpene ceylonamide A (142), isolated from the Indonesian marine sponge Spongia ceylonensis and determined that they exhibited moderate in vitro anti-inflammatory activity of receptor activator of nuclear factor-κB ligand (RANKL)-induced osteoclastogenesis in RAW264 macrophages with structure-activity relationships studies. This revealed that the “position of the carbonyl group of the γ-lactam ring …at its nitrogen atom were important for inhibitory activity” [167]. Ngan and colleagues isolated a known polyketide citrinin H1 (143) from the Korean sand-derived bacterium Penicillium sp. SF-5629 and observed moderate inhibition E. coli LPS-activated BV-2 microglia in vitro nitric oxide (NO) and prostaglandin E2 (PGE2) release with concomitant downregulation of the inducible enzymes nitric oxide synthase and cyclooxygenase-2 attributed to “inhibition of NF-κB and p38 MAPK signaling pathways” [168]. Manzoor and colleagues observed that 4-hydroxy-2,3-dimethyl-2-nonen-4-olide (144) extracted from Korean marine alga Ulva pertusa moderately inhibited pro-inflammatory cytokines IL-12 p40 and IL-6 release from bone marrow-derived dendritic cells, as well as signal transduction by inhibiting phosphorylation of nuclear factor -κB (NF-κB), and thus, warranted further study to evaluate its potential as a “therapeutic agent for inflammation-associated maladies” [169]. Aminin and colleagues assessed the anti-inflammatory pharmacology of the triterpene glycoside cucumarioside A2-2 (145) isolated from the Russian edible sea cucumber Cucumaria japonica by demonstrating that this compound potently interacted with P2X purinergic receptors, in particular P2X4, on mature macrophage membranes and enhanced reversible ATP-dependent Ca2+ signaling thus leading to “activation of cellular immunity” [170]. Ha and colleagues studied the anti-inflammatory effects of a known curvularin-type metabolite (146) derived from an Antarctic Ross Sea sponge-derived fungus Penicillium sp. SF-5850 that potently inhibited LPS-induced RAW264.7 macrophages production of pro-inflammatory NO and PGE2 as well as the expression of iNOS and COX-2 by inhibition of the NF-κB, a “transcriptional factor(s) involved in inflammation-related disorders” [171]. Ciaglia and colleagues characterized the anti-inflammatory effects of the terpenoid 9,11-dihydrogracilin A (147) isolated from the Antarctic marine sponge Dendrilla membranosa and observed that, in human peripheral blood mononuclear cells, in vitro, it downregulated NF-κB as well as inhibition of IL-10, while in vivo it significantly attenuated mouse ear edema and inflammation, thus supporting the use of this compound “in inflammatory cutaneous diseases [172]. Jiao and colleagues investigated the “terpene-polyketide-pyridine hybrid” dysivillosin A (148) isolated from the South China Sea marine sponge Dysidea villosa that inhibited RBL-2H3 mast cell release of β-hexosaminidase, a marker of degranulation, as well as pro-inflammatory leukotriene B4 and IL-4 by suppressing the IgE/Syk signaling pathway, thus suggesting a “new chemotherapeutic scaffold targeting Syk-associated allergy” [173]. Su and colleagues extended the pharmacology of the known antimicrobial peptide epinecidin-1 (149) isolated from the orange-spotted grouper Epinephelus coioides by showing that its anti-inflamatory activity resulted from degradation of the Toll-like receptor signaling adaptor protein MyD88 in RAW 264.7 mouse macrophages and concomitant activation of the Smurf E3 ligase proteasome degradation pathway [174]. Lin and colleagues continued the evaluation of the terpenoid excavatolide (150) isolated from the Taiwanese gorgonia coral Briareum excavatum, observing that it inhibited LPS-induced osteoclast-like cell formation and tartrate-resistant acid phosphatase (TRAP) expression, in vitro, while also significantly reducing paw oedema and TRAP-positive multinucleated cell formation in two rat models of experimental arthritis [175]. Several studies extended the anti-inflammatory mechanisms of the marine carotenoid fucoxanthin (151) isolated from the edible brown alga Undaria pinnatifida: Choe and colleagues reported that fucoxanthin inhibited paw edema in an experimental in vivo model of inflammation by reducing activation of iNOS, COX-2 and NF-κB [176]; Grassa-López and colleagues observed that fucoxanthin ameliorated lipogenesis, decreased insulin resistance as well as biomarkers of both inflammation and cardiovascular disfunction in an experimental rat obesity model [177]; Sugiura and colleagues determined that in vitro fucoxanthin suppressed PLA2 and COX-2 expression in a rat basophilic leukemia-2H3 cells, while in vivo inhibiting PLA2, COX-2 and hyaluronidase in several ICR mouse ear models of inflammation [178]; Taira and colleagues observed that concentration-dependent cytoprotection or apoptosis in RAW264.7 macrophage cells by the carotenoids fucoxanthinol and fucoxanthin resulted from activation of the Nrf2-ARE signaling pathway [179]. Sintsova and colleagues characterized a novel Kunitz-type peptide (152) isolated from the sea anemone Heteractis crispa that significantly decreased intracellular Ca2+ in histamine-treated murine bone marrow-derived macrophages thus suggesting the involvement of “H1-type histamine receptor blockage” in the anti-inflammatory mechanism of action of this peptide [180]. Hong and colleagues determined that the diterpenoid hipposponlachnin B (153) isolated from the South China Sea marine sponge Hippospongia lachne decreased production of β-hexosaminidase, a degranulation biomarker and pro-inflammatory mediators IL-4 and LTB4 by RBL-2H3 cells, suggesting a possible therapeutic use for “the treatment of allergy” [181]. Kozuma and colleagues isolated new cyclic peptides ogipeptins A-D (154157) from the culture broth of the Japanese marine Gram-negative bacterium Pseudoalteromonas sp. SANK 71903 that blocked the binding of LPS to the cluster of differentiation 14 (CD14) in vitro and decreased TNF-α release by human U937 monocytic cells, concluding that these peptides could be developed for use “as anti-LPS drugs against LPS-associated diseases” [182]. Kwon and colleagues reported a new anthranilic acid derivative oscarellin (158) isolated from a Philippine sponge Oscarella stillans that strongly inhibited LPS-induced TNF-αand IL-6 production in murine macrophage RAW 264.7 macrophages by a mechanism “associated with inactivation of c-Jun NH2-terminal kinase (JNK), extracellular signal-regulated kinase (ERK), activator protein-1 (AP-1), and NF-κB and activation of activating transcription factor-3 (ATF-3)” [183]. Kim and colleagues observed that the novel alkaloid pseudane-VII (159) isolated from the marine bacterium Pseudolateromonas sp. M2 modulated release of pro-inflammatory mediators (NO, IL-1β and IL-6) by LPS-treated RAW 264.7 murine macrophage cell line in vitro via inhibition of “MAPK phosphorylation and NF-κB translocation”, while in vivo treatment with this compound suppressed production of IL-1β and iNOS expression, suggesting “pseudane VII has potential as treatment for inflammatory responses and diseases” [184].
In contrast, for marine compounds (160192) shown in Table 2 and Figure 2, only the anti-inflammatory activity (IC50) was reported, no molecular mechanism of action was reported at the time of publication: A new eremophilane-sesquiterpene acremeremophilane B (160) isolated from a South Atlantic deep sea (2869 m)-sediment derived fungus Acremonium sp. [185]; two new quinoline alkaloids actinoquinolines A and B (161, 162) isolated from a Californian marine-derived Streptomyces sp. strain CNP975 [186]; a new polyhydroxysteroidal glycoside anthenoside O (163) isolated from the tropical South China Sea starfish Anthenea aspera [187]; a new polyketide aurasperone C (164), isolated from a South China Sea brown alga Sargassum sp.-derived Aspergillus niger SCSIO Jcsw6F30 [188]; two new 9-hydroxybriarane diterpenoids briarenolides ZII and ZVI (167, 168) isolated from the Taiwanese gorgonian coral Briareum sp. [190]; a known eremophilane sesquiterpene dihydrobipolaroxin (169) isolated from a South China Sea deep (2439 m) marine sediment-derived Aspergillus sp. SCSIOW2 [191]; a known alkaloid echinulin (170) isolated from an Antarctic Ross Sea “unidentified” marine organism-derived Aspergillus sp. SF-5976 [192]; a novel halogenated 5α-iodozoanthenamine (171) isolated from the colorful Taiwanese zoanthid Zoanthus kuroshio [193]; two novel steroids klyflaccisteroid J and K (172, 173) isolated from the Taiwanese soft coral Klyxum flaccidum [194,195]; a known lobane diterpenoid (174) isolated from a Taiwanese soft coral Lobophytum varium [196]; two new bioactive steroids petasitosterones B and C (175, 176) isolated from the Taiwanese soft coral Umbellulifera petasites [197]; two new sterols (177, 178) [198], and novel 9,11-secosterols pinnigorgiol A (179) [199], pinnigorgiol E (180) [200], pinnisterol A (181) [201] and pinnisterol H (182) [202] isolated from the Taiwanese gorgonian coral Pinnigorgia sp.; a new pyrrole oligoglycoside plancipyrroside B (183) isolated from an extract of the Vietnamese starfish Acanthaster planci [203]; a novel steroidal glycoside protolinckioside A (184) isolated from an Arabian sea starfish Protoreaster lincki [204]; a known terpenoid sarcophytonolide O (185) isolated from the Chinese soft coral Lobophytum crassum [205]; two novel norcembranoids sinulerectols A and B (186, 187) isolated from South China Sea soft coral Sinularia erecta [206]; a known terpenoid sinubrasolide A (188) and the steroid sinubrasone D (189) isolated from the Taiwanese “cultivation pool” soft coral Sinularia brassica [207,208]; and the novel diterpenes uprolides N, O and P (190192) isolated from the Panamanian octocoral Eunicea succinea [209].

3.3. Marine Compounds with Activity on the Immune System

In 2016–2017, the preclinical pharmacology of marine compounds that were reported to affect the immune system is shown in Table 2 and Figure 2. The molecular mechanism of action of marine natural products that affected the immune system (145, 193199) was described in seven papers which used several in vitro and in vivo models. Pislyagin and colleagues reported that the triterpene glycoside cucumarioside A2-2 (145) isolated from the Far Eastern sea cucumber Cucumaria japonica caused changes in mouse spleen morphology, proliferative activity “predominantly in B-cells” and macrophage activation in mouse spleen, concomitant with increase in IL-1β, iNOs, ROS and NO production [210,211]. Sánchez and colleagues investigated three terpenoids gracillin A, H and L (193195) isolated from the marine sponge Spongionella gracillis that inhibited human T lymphocyte IL-2 production, as well as CD147 expression by reducing calcineurin phosphatase activity, thus concluding that gracilins are a “valuable option for synthetic drug development” [212,213]. Kicha and colleagues determined the immunomodulatory effects of the mycosporine-like amino acids (MAA) shinorine and porphyra-334 (196, 197) isolated from the red alga Porphyra sp. and reported they both induced NF-κB activity and moderated tryptophan metabolism in human myelomonocytic cell line THP-1, thus recommending “a more detailed risk-benefit assessment” before MAA are used “in daily care products” [214]. Chung and colleagues studied the cembrane-type diterpenoid sinulariolide (198), isolated from the Taiwanese cultured soft coral Sinularia flexibilis and observed suppression of LPS-phenotypic maturation, cytokine and NO production and co-stimulatory molecule expression of murine bone marrow-derived dendritic cells, as well as signaling pathways, concluding “that sinulariolide may be utilized in the treatment of autoimmune andinflammatory disorders” [215]. Gao and colleagues determined significant effects of several peptides (199) isolated from Indian Ocean deep (1654 m) sediment-derived actinomycete Williamsia sp. MCCC 1A11233 on both IgE-sensitized RBL-2H3 mast cell histamine and proinflammatory cytokine release, and a murine model of passive cutaneous anaphylaxis, thus observing that “the three compounds have the potential to prevent or treat IgE-sensitized allergic disorders” [216].
In contrast, for the marine compounds (200205) listed in Table 2 and depicted in Figure 2, only the immune bioactivity (IC50) was reported, but no molecular mechanism of action had been described by the time of publication: A novel dimeric nitrophenyl trans-epoxyamide chrysamide C (200) produced by a Indian Ocean deep (3386 m) sediment-derived fungus Penicillium chrysogenum SCSIO41001 [217]; a new dimeric macrolide xylopyranoside cocosolide (201) isolated from the Guamanian marine cyanobacterium Symploca sp. [218]; a new N-acyl dopamine glycoside myxillin A (202) isolated from the Icelandic hydrothermal vent-derived sponge Myxilla incrustans [219]; the known steroidal saponin pectinioside A (203) isolated from the starfish Patiria pectinifera [220]; a new benzophenone polyketide peniphenone (204) obtained from the mangrove Sonneratia apetala endophytic fungus Penicillium sp. ZJ-SY2 [221]; and a known analog of the antimycin-type depsipeptide somalimycin, named USF-19A (205) isolated from the South China Sea deep (3536 m) sediment-derived Streptomyces somaliensis SCSIO ZH66 [222].

3.4. Marine Compounds Affecting the Nervous System

As shown in Table 2 and Figure 2, in 2016–2017, the preclinical nervous system pharmacology of marine compounds (206225) reported several mechanisms of action involving potassium (K+) channels, nicotinic acetylcholine, calcium (Ca2+) and serotonin receptors, as well as in vivo models of antinociception and neuroprotection.
Five marine compounds were shown to bind K+ channels (206, 207), calcium channels (208), nicotinic acetylcholine receptors (nACHR) (209) and serotonin receptor 5-HT7 (218). Moreels and colleagues electrophysiologically demonstrated that the novel peptide APETx4 (206) isolated from the marine sea anemone Anthopleura elegantissima inhibited the K+ channel Kv10.1 by keeping it in a closed state and thus, because the Kv10.1 is overexpressed in a human tumors, APETx4 could become a “scaffold for design and synthesis of more potent and safer anticancer drugs” [223]. Goda and colleagues conducted detailed studies that determined that the marine xanthophyll carotenoid astaxanthin (207) reversed the toxicity of the calcium-dependent Maxi-K (BK) K+ channel antagonist food mycotoxin penitrem A (PA) “by restoring normal levels of the targeted neurotransmitters” both in Schwann cells CRL-2765 in vitro and in both the nematode Caenorhabditis elegans and rat in vivo models thus concluding astaxanthin might be useful “for in vivo prevention of PA-induced brain toxicity [224]. Mendez and colleagues investigated the marine guanidine alkaloid crambrescidin 816 (208) produced by the Mediterranean sponge Crambe crambe and known to block voltage dependent L-type Ca2+ channels, and observed that cytotoxicity was concomitant with an increase of cytosolic Ca2+ in a primary culture of cortical neurons by a mechanism that “was mediated by both NMDA and AMPA glutamate receptor subtypes” activation [225]. Jiang and colleagues discovered a novel O-conotoxin GeXXVIIA linear peptide (209) from the venom of the South China sea cone snail Conus generalis that potently inhibited the human α9α10 nicotinic acetylcholine receptor (nACh), a nAChR receptor expressed in both the nervous system, as well as in other non-neuronal cells, by a non-competitive and voltage-independent mechanism. The data suggest that this new inhibitor “would facilitate unraveling the functions of this nAChR subtype” [226]. Uchimasu and colleagues reported a novel guanidine alkaloid mellpaladine A (218) isolated from a Palauan Didemnidae tunicate that modulated mice behavioral profiles after an intracerebroventricular injection with particular selectivity for the serotonin receptor 5-HT7, thus, supporting the notion that “marine alkaloids are unique…source of neuroactive compounds” [239].
Two additional studies extended the pharmacology of conopeptides (210211). Bernáldez and colleagues discovered a novel 25-mer peptide member of the γ-conotoxin family PiVIIA (210) in the venom of the Mexican marine snail Conus princepts that increased Ca2+ currents in dorsal root ganglion neurons without affecting the Na+, K+ or acid sensing ionic channel currents, further studies being proposed to “define its potential use as a positive modulator of neuronal activity” [227]. Brust and colleagues identified a previously uncharacterized nine amino acid conorphin-T (211) from a Conus textile venom peptide library with selectivity for κ-opioid receptors (KOR) that led to the development of several novel KOR agonists, which potently inhibited splanchnic colonic nociceptors, thus, becoming promising leads “for the development of irritable bowel syndrome treatments” [228].
One study contributed to nociceptive pharmacology. Logashina and colleagues reported that the novel peptide Ms 9a-1 (219), isolated from the venom of the sea anemone Metridium senile, produced significant analgesic and anti-inflammatory effects in a murine thermal hyperalgesia model by a mechanism that potentiated the response of transient receptor potential ankyrin-repeat 1 receptor (TRPA1) and was followed by loss of TRPA1-expressing neurons [240].
The neuroprotective activity of marine compounds (151, 212, 214, 216, 220, 223, 225) was reported in ten studies. Three studies reported on the neuroprotective pharmacology of the marine carotenoid fucoxanthin (151) isolated from several brown algae: Undaria pinnatifida, Ecklonia stolonifera and Sargassum horneri: Jung and colleagues studied β-site amyloid precursor protein cleaving enzyme 1 (BACE-1), which is strongly correlated with Alzheimer’s disease, and observed that fucoxanthin inhibited BACE1 activity exhibiting mixed-type inhibition in vitro, while molecular docking simulations showed that two fucoxanthin hydroxyl groups interacted with two BACE1 residues (Gly11 and Ala127), mechanistic studies that suggest fucoxanthin “may be a good template for anti-AD drugs” [232]. Lin and colleagues demonstrated that in mice fucoxanthin reversed scopolamine-induced cognitive impairments by inhibiting brain acetylcholinesterase by a non-competitive mechanism, as well as increasing choline acetyltransferase activity and brain-derived neurotrophic factor (BDNF) expression in hippocampus and cortex. The studies appear to anticipate fucoxanthin’s “therapeutic efficacy for the treatment of AD by acting on multiple targets” [233]. Zhang and colleagues discovered that fucoxanthin provided neuroprotection in an experimental murine model of traumatic brain injury (TBI), “a major public health problem” by alleviating TBI-induced secondary brain injury, cerebral edema and lesions, while also demonstrating that it attenuated TBI-induced apoptosis and oxidative stress “at least partly” via regulation of the nuclear factor erythroid 2-related factor (Nrf2)-antioxidant-response element (ARE) and Nrf2-autophagy pathways, thus, making fucoxanthin “an attractive therapeutic agent in the treatment of TBI in the future” [234]. Two studies were reported by the same research group describing the neuroprotective and anti-inflammatory effects of the cembranolide analog 11-dehydrosinulariolide (11-de) (212) isolated from the soft coral Sinularia flexibilis: Feng and colleagues observed that 11-de increased the expression of BDNF in a neuroblastoma cell line in vitro and had a protective effect in both an in vivo zebrafish and rat Parkinson’s disease model, results that the investigators hoped would “help treat patients diagnosed with Parkinson’s disease” [229]. Chen and colleagues reported that 11-de improved the functional recovery in a rat thoracic spinal cord contusion injury experimental model with an antiapoptotic and anti-inflammatory mechanism that attenuated iNOS and TNF-α, thus, proposing “this compound may be a promising therapeutic agent for spinal cord injury” [230]. Fang and colleagues showed that a new cyclopentenone 5-hydroxycyclopenicillone (214) isolated from a culture of the marine sponge Hymeniacidon perleve-derived fungus Trichoderma sp. HPQJ-34 acted as a moderate free radical scavenger, and had anti-Aβ fibrillization and neuroprotective properties, concluding that this compound “might be of interest to neuropharmacology research and anti-AD drug discovery programs” [235]. Alonso and colleagues evaluated the pyrroloiminoquinone makaluvamine J (216), isolated from a Fijian marine sponge Zyzzya sp. and showed that it “provided full neuroprotection”, as it potently reduced mitochondrial damage by reactive oxygen species, as well as improved endogenous glutathione and catalase in mouse and human neuronal models, thus, potentially contributing to “antioxidant therapies in neurodegenerative diseases” [237]. Kim and colleagues found that the phlorotannin phlorofucofuroeckol (220) isolated from the brown seaweed Ecklonia cava, increased cell viability in glutamate-treated rat adrenal phaeochromocytoma PC12 cells by inhibiting apoptotic cell death as well as mitochondrial reactive oxygen species generation, results, which taken together, suggest PFF may be developed as “ a neuroprotective agent in ischemic stroke” [241]. Perni and colleagues discovered that the aminosterol squalamine (223) isolated from the dogfish shark Squalus acanthias displaced the “intrinsically disordered” protein α-synuclein, associated with Parkinsons’s disease, from lipid vesicles and membranes by competitively binding at the surfaces, as well as suppressed muscle paralysis in a nematode worm Caenorhabditis elegans strain oversexpressing α-synuclein, suggesting squalamine “could be a means for a therapeutic intervention in Parkinsons’s disease” [244]. Pan and colleagues extended the pharmacology of xyloketal B (225) isolated from the mangrove fungus Xylaria sp. by investigating effects of xyloketal in an adult mice stroke model, observing that pre- and post-treatment treatment reduced both brain infarct volume and the generation of ROS and pro-inflammatory cytokines by suppression of ROS/TLR4/NF-κB inflammatory signaling pathway, thus providing evidence for “potential application of xyloketal B in stroke therapy” [246].
As shown in Table 2, five marine compounds were shown to modulate other molecular targets, i.e., the acetylcholinesterases (213, 217, 224), endogenous transient receptor potential canonical type 1 channel (TRPC1) (215), human monoamine oxidase B enzyme (221), and prolonged synaptic transmission (222). Botic Lee and colleagues discovered that the brominated alkaloid discorhabdin G (213) isolated from the Antarctic Latrunculia biformis sponge inhibited electric eel and human acetylcholinesterases by a reversible and competitive mechanism, observations that could potentially lead to new Alzheimer’s disease “cholinesterase inhibitors based on the scaffold of discorhabdins” [231]. Adelhameed and colleagues reported the isolation and structure elucidation of a new phytoceramide MEC-1-4 (217) from the Egyptian Red Sea sponge Mycale euplectellioides that moderately inhibited acetylcholinesterase by interacting with the enzyme “via hydrogen bonding, hydrophobic contacts and hydrophilic-hydrophobic interactions”, thus suggesting MEC-1-4 might become a “valuable lead compound for AD management” [238]. Moodie and colleagues evaluated the known brominated phenethylamine derivative stryphnusin (224) isolated from the Norwegian sponge Stryphus fortis that moderately inhibited electric eel acetylcholinesterase by a reversible competitive mechanism and with no effect on muscle function or neuromuscular transmission thus contributing to novel and promising approaches for “symptomatic treatment of AD” [245]. Flores and colleagues investigated the marine polyether toxin maitotoxin (215) produced by the dinoflagellate Gambierdiscus toxicus that is responsible for ciguatera fish poisoning, demonstrating that its mechanism of action involves activation of the non-voltage-gated cation TRPC1 in X. laevis oocytes and thus proposing this toxin as a “useful tool for further studies of TRPC1 channels” [236]. Lee and colleagues reported that the polyketide piloquinone (221) isolated from a Californian marine sediment-derived Streptomyces sp. CNQ-027 potently inhibited recombinant human monoamine oxidase B enzyme considered a “target in AD and Parkinson’s disease”, by a competitive and reversible mechanism which highlighted the importance of the “ester functionality in the ring system for bioactivity”, thus possibly becoming “a new potential lead compound for the development of MAO inhibitors” [242]. Caplan and colleagues extended the pharmacology of the marine diterpene glycoside pseudopterosin A (222) isolated from the Bahamanian gorgonian soft coral Pseudopterogorgia elisabethae by demonstrating it prolonged synaptic transmission in an experimental oxidative stress model, as well as extensively distributed in murine brain, findings that suggested a “potential as a novel neuromodulatory agent” [243].
Finally, and as shown in Table 2, several marine compounds (226234) affected the nervous system, but their detailed molecular mechanisms of action remained undetermined at the time of publication: a novel Indonesian Aplysinellidae sponge-derived bromotyrosine araplysillin X (226) that inhibited the aspartic protease BACE1 involved in AD [247]; a novel Panamanian marine cyanobacterium cf. Symploca sp.-derived terpene caracolamide A (227) with in vitro calcium influx and calcium channel oscillation modulatory activity [248]; a Mexican marine cone snail Conus spurius peptide conorfamide-Sr3 (228) shown to block Shaker subtype voltage-gated potassium channels [249]; a new conopeptide contryphan-Bt (229) isolated from the South China Sea cone snail Conus betulinus shown to be neurotoxic to mice [250]; acetylcholinesterase inhibitory activity in two know meroterpenoids, dehydroaustin (230) isolated from the Chinese mangrove endophytic fungus Aspergillus sp. 16-5c [251], and terreulactone C (233) isolated from the South China Sea mangrove-derived endophytic fungus Penicillium sp. SK5GW1L [254]; a known alkaloid hymenidin (231), originally isolated from the Caribbean marine sponge Agelas citrina that inhibited voltage-gated potassium channels [252]; a neuroprotective and known bromotyrosine alkaloid psammaplysene A (232) originally isolated from the marine sponge Psammaplysilla sp. [253]; and inhibition of α9α10 nicotinic acetylcholine receptor by turripeptide (234) isolated from the Philippine marine gastropod Unedogemmula bisaya venom [255].

4. Marine Compounds with Miscellaneous Mechanisms of Action

Further 2016–2017 preclinical pharmacology for marine compounds (49, 51) as well as that of 79 compounds (235313) with miscellaneous mechanisms of action is shown in Table 3, with their corresponding structures, presented in Figure 3. Given that, at the time of publication, a comprehensive pharmacological characterization of these compounds remained unavailable, their assignment to a particular drug class will probably require further investigation into their molecular mechanism of action.
Table 3 presents not only the pharmacological activity (an IC50), but also the molecular mechanism of action of the following marine natural compounds: Fungus Acremonium sp. (F9A015) polyketide acredinone C (235) [256]; algal terpenoid astaxanthin (207) [257]; sponge Axinyssa sp. bisabolene sesquiterpene (236) [258]; sponge Carteriospongia sp. scalarane sesterterpenoid (237) [259]; fungus Aspergillus ungui NKH-007 depsidone 7-chlorofolipastatin (238) [260]; sponge Acanthostrongylophora ingens halogenated alkaloid chloromethylhalicyclamine B (239) [261]; cyanobacterium Leptolyngbya sp. depsipeptide coibamide A (240) [262]; sponge Theonella aff. swinhoei cyclic peptide cyclotheonellazole A (241) [263]; bacterium SNA-024 N6,N6-dimethyladenosine (242) [264]; gorgonian Briareum excavatum briarane-type diterpene excavatolide B (150) [265]; brown algae Undaria pinnatifida, Ecklonia stolonifera and/or Sargassum horneri marine carotenoid fucoxanthin (51) [266]; sponge Phorbas sp. diterpenoid gagunin D (243) [267]; bacterium Bacillus sp. strain SCO-147 (-)-4-hydroxysattabacin (244) [268]; cyanobacteria Lyngbya majuscula and Tolypothrix sp. γ-pyrone kalkipyrone (245) [269]; sponge Stylissacarteri sp. alkaloid latonduine A (246) [270]; sponge Negombata sp. alkaloid latrunculin A (247) [271]; bacterium Streptomyces sp. polyketide nahuoic acid A (248) [272]; bacterium Streptomyces sp. YP127 napyradiomycin A1 (49) [273]; sponge Leucetta microraphis alkaloid leucettamine B-related synthetic analogue polyandrocarpamine A (249) [274]; sponge Cacospongia sp. terpenoid scalaradial (250) [275]; fungus Stachybotrys sp. KCB13F013 meroterpenoid stachybotrysin (251) [276]; soft coral Clavularia sp. terpenoid stolonidiol (252) [277]; cyanobacterium Lyngbya sp. peptide tasiamide F (253) [278]; sponge Theonella sp. cyclic peptide theonellamide A (254) [279]; fungus Penicillium sp. HL-85-ALS5-R004 polyketide toluquinol (255) [280]; fungus Penicillium janthinellum alkaloid N-Me-trichodermamide B (256) [281]; gorgonian-derived fungus Aspergillus versicolor cyclopeptides versicotides D-F (257259) [282]; sponges Ircinia and Spongia spp. furanoterpene (7E, 12E, 20Z, 18S)-variabilin (260) [283]; and fungus Xylaria sp. terpenoid xyloketal B (225) [284].
Also presented in Table 3 is the pharmacological activity (IC50 for enzyme or receptor inhibition) of marine-derived compounds (261313), although their respective mechanisms of action remained undetermined and will require further investigation: sponge Aaptos aaptos alkaloid 9-methoxyaaptamine (261) [285]; mangrove-derived fungus Ascomycota sp. SK2YWS-L polyketide ascomindone A (262) [286]; mudflat-derived fungus Aspergillus niger polyketide aurasperone B (263) [287]; alga-derived fungi Penicillium thomii and Penicillium lividum meroterpenoid austalide H acid ethyl ester (264) [288]; fungus Biscogniauxia mediterranea peptide biscogniauxone (265) [289]; fungus Aspergillus ustus sesterpenoid cerebroside D (266) [290]; sponge Spongia ceylonensis diterpene ceylonin A (267) [291]; sponge-derived fungus Stachybotrys chartarum sesquiterpene chartarene D (268) [292]; sponge-associated fungus Talaromyces stipitatus KUFA 0207 anthraquinone citreorosein (269) [293]; fungus Aspergillus sp. SCSIOW3 polyketide cordyol (270) [294]; fungus Chaetomium cristatum dioxopiperazine alkaloid cristazine (271) [295]; sponge-associated fungus Emericella variecolor polyketides diasteltoxins A-C (272274) [296]; sponge Latrunculia sp. pyrroloiminoquinone alkaloid discorhabdin L (275) [297]; sponge Dysidea sp. polybrominated diphenyl ether 3,4,5-tribromo-2-(2′,4′-dibromophenoxy)-phenol (276) [298]; sponge Dysidea sp. tetracyclic meroterpene dysiherbol A (277) [299]; sand dollar Scaphechinus mirabilis aminonaphthoquinone echinamine B (278) [300]; sponge derived fungus Stachylidium sp. tetrapeptide endolide B (279) [301]; mangrove-derived fungus Eurotium rubrum MA-150 (±)-europhenol A (280) [302]; sponge Fascaplysinopsis sp. bis-indole alkaloid fascaplysin (281) [303]; nudibranch Jorunna funebris tetrahydroisoquinolinequinone alkaloid fennebricin A (282) [304]; sponge Petrosia sp. depsipeptide halicylindramide A (283) [305]; sponge Halichondria cf. panicea polyacetylene isopetrosynol (284) [306]; cyanobacterium Leptolyngbya sp. macrolide polyketide leptolyngbyolide B (285) [307]; ascidian Lissoclinum mandelai macrocyclic polyketide mandelalide C (286) [308]; sponge Hyrtios digitatus tetracyclic merosesquiterpene 19-methoxy-9,15-ene-puupehenol (287) [309]; sponge Monanchora pulchra cyclic guanidine alkaloid monanchomycalin B (288) [310]; sponge Plakortis simplex polyketide monotriajaponide A (289) [311]; sponge Mycale lissochela terpenoid mycalenitrile-15 (290) [312]; sponge Hyrtios sp. meroterpenoid nakijinol G (291) [313]; sponge Theonella swinhoei cyclic pentapeptide nazumazole D (292) [314]; fungus Aspergillus sp. LF660 benzocoumarin polyketide pannorin (293) [315]; fungus Alternaria sp. NH-F6 polyketides perylenequinones (294, 295) [316]; sponge Petrosia alfiani xestoquinone derivatives petroquinones A and B (296, 297) [317]; fungus Peyronellaea glomerata isocoumarin derivatives peyroisocoumarins B and D (298, 299) [318]; fungus Phoma sp. NTOU4195 polyketide phomaketide A (300) [319]; ascidian Sidnyum elegans phosphorylated polyketide phosphoeleganin (301) [320]; soft coral Pseudopterogorgia rigida sesquiterpenes (302, 303) [321]; sponge Spongia ceylonensis diterpene ent-13-norisocopalen-15-al-18-oic acid (304) [322]; bryozoan Schizomavella mamillata 5-alkylresorcinol derivatives schizols A and B (305, 306) [323]; fungus Stachybotrys longispora FG216 isoindolinone derivatives FGFC6 and FGFC7 (307, 308) [324]; sponge Spongia pertusa Esper sesquiterpene (309) [325]; sponge Psammocinia sp. furanosesterterpene tetronic acid sulawesin A (310) [326]; cyanobacterium Okeania sp. cyclic depsipeptide urumamide (311) [327]; gorgonian-derived fungus Aspergillus versicolor LZD-14-1 alkaloid versiquinazoline B (312) [328]; and sponge Xestospongia testudinaria new bioactive steroidal ketone (313) [329].
Table 3. Marine pharmacology in 2016–2017: marine compounds with miscellaneous mechanisms of action.
Table 3. Marine pharmacology in 2016–2017: marine compounds with miscellaneous mechanisms of action.
Compound/Organism aChemistryPharmacological ActivityIC50 bMMOA cCountry dReferences
acredinone C (235)/fungusPolyketide eOsteoclast differentiation induction inhibition10 μM *NFATc1 transcription inhibitionS. KOR[256]
astaxanthin (207)/shrimpTerpenoid fHepatic stellate cell activation inhibition10 μM *Decreased ROS and NOX2 expression reductionUSA[257]
Axinyssa sp. bisabolene (236)/spongeTerpenoid fPTP1B inhibition1.9 μMAkt phosphorylationJPN[258]
Carteriospongia sp. terpenoid (237)/spongeTerpenoid fApoptosis induction0.06 μg/mL *Topoisomerase IIα and Hsp90 inhibitionEGY, SWE, TWN[259]
7-chlorofolipastatin (238)/fungusPolyketide eMacrophage SOAT 1 inhibition6.8 μMSOAT 1 and 2 inhibition in vitroJPN[260]
chloromethylhalicyclamine B (239)/spongeAlkaloid gProtein kinase CK1δ/ε inhibition6 μMATP-binding site dockingFRA, ITA, NLD[261]
coibamide A (240)/cyanobacteriumPeptide fVEGFA secretion inhibition<5 nMAntiangiogenic propertiesUSA[262]
cyclotheonellazole A (241)/spongePeptide fchymotrypsin and elastase inhibition0.034–0.62 nMEnzyme S2 subsite bindingBEL, ISR, NLD[263]
N6,N6-dimethyladenosine (242)/bacteriumAlkaloid gAKT phosphorylation inhibition5 μM *S473 site inhibitionUSA[264]
excavatolide B (150)/soft coralTerpenoid fModulation of atrial myocytes10 μM *Ca2+ homeostasis modulationTWN[265]
fucoxanthin (51)/algaTerpenoid fLung fibrosis attenuation10 mg/kg ***Type 1 collagen expression decreaseS. KOR[266]
gagunin D (243)/spongeTerpenoid fMelanin synthesis inhibition12.7 µMTyrosinase expression inhibitionS. KOR[267]
(−)-4-hydroxysattabacin (244)/bacteriumPolyketide eMelanin synthesis inhibition25 μg/mL *Tyrosinase, TRP-1 and TRP-2 expression inhibitionS. KOR[268]
kalkipyrone (245)/cyanobacteriumPolyketide eAdipose tissue suppression5 mg/kg ***Enhance LA plasma levelsJPN[269]
latonduine A (246)/spongeAlkaloid gCTFR inhibition62 nMPARP isozymes inhibitionCAN, GBR[270]
latrunculin A (247)/spongeAlkaloid gECFC tube inhibition0.043 µMSpecific kinases inhibitionUSA[271]
nahuoic acid (248)/bacteriumPolyketide eSETD8 inhibition6.5 µMCompetitive inhibition of SAM bindingCAN, PNG[272]
napyradiomycin A1 (49)/bacteriumTerpenoid fAngiogenesis inhibition10 µMVE-cadherin inhibitionS. KOR[273]
polyandrocarpamine A (249)/spongeAlkaloid gDYRK and CLK selective inhibition0.17–0.93 µMCyclin D1 phosphorylation inhibitionAUS, BRA, DEU, FRA[274]
scalaradial (250)/spongeTerpenoid fTRPM2 ion channel inhibition0.2 µMLack of PLA2 inhibitionJPN, NZL, USA[275]
stachybotrysin (251)/fungusTerpenoid fOsteoclast differentiation inhibition5 μg/mL *MAPK kinase pathway inhibitionJPN, S. KOR[276]
stolonidiol (252)/soft coralTerpenoid fPKCα membrane translocation5 µM *Increased ChAT activityUSA[277]
tasiamide F (253)/cyanobacteriumPeptide gCathepsin D and E inhibition23–57 nMDocking studies completedUSA[278]
theonellamide A (254)/spongePeptide gBilipid membrane disruption20 µM *Binding to sterolsJPN[279]
toluquinol (255)/fungusPolyketide eLymphangiogenesis inhibition6.2 µMSuppression of Akt and ERK ½ phosphorylationBEL[280]
N-Me-trichodermamide B (256)/fungusAlkaloid gH2O2 oxidative damage inhibition5 µM *Nrf2-signaling regulationCHN[281]
versicotides D–F (257259)/fungusPeptide gFoam cell formation inhibition10 µM *Cholesterol influx inhibitionCHN[282]
variabilin (260)/spongeTerpenoid fPTP1B inhibition1.5 μMTCPTP inhibitionIND, JPN[283]
xyloketal B (225)/fungusTerpenoid fNAFLD attenuation5 mg/kg ***SREBP-1c expression inhibitionCHN[284]
9-methoxyaaptamine (261)/spongeAlkaloid gPPRE activation0.039 μg/mL *UndeterminedIDN, MYS[285]
ascomindone A (262)/fungusPolyketide eDPPH radical scavenging inhibition18.1 μMUndeterminedCHN[286]
aurasperone B (263)/fungusPolyketide eDPPH radical scavenging inhibition0.01 μMUndeterminedS. KOR[287]
austalide H acid ethy ester (264)/ fungusTerpenoid fEndo-1,3-β-D-glucanase inhibition0.2 μMUndeterminedRUS[288]
B. mediterranea cyclopentapeptide (265)/fungusPeptide gGSK-3β inhibition8.04 μMUndeterminedCHN, DEU[289]
cerebroside (266)/fungusPolyketide eSpermatozoa inhibition8 μMUndeterminedRUS[290]
ceylonin A (267)/spongeTerpenoid fOsteoclast inhibition<50 μM *UndeterminedNLD, JPN
chartarene D (268)/fungusTerpenoid fTyrosine kinases inhibition0.1–0.8 μMUndeterminedCHN, DEU[292]
citreorosein (269)/fungusPolyketide eAnti-obesity activity0.17 μMUndeterminedGBR, PRT, THAI[293]
cordyol C (270)/fungusPolyketide eErythrocyte biomembrane protection4.9 μMUndeterminedCHN[294]
cristazine (271)/fungusAlkaloid gDHHP radical scavenging19 μMUndeterminedS. KOR[295]
diasteltoxins A–C (272274)/fungusPolyketide eThioredoxin reductase inhibition7.2–12.8 μMUndeterminedCHN, DEU[296]
discorhabdin L (275)/spongeAlkaloid gHIF-1α transcription inhibition0.73 μMUndeterminedNZL, USA[297]
Dysidea sp. diphenyl ether (276)/spongePolyketide eMitochondrial complex II inhibition6.4 nMUndeterminedJPN[298]
dysiherbol A (277)/spongeTerpenoid fNF-κB inhibition0.49 μMUndeterminedAUS, CHN[299]
echinamine B (278)/sea urchinPolyketide eDHHP radical scavenging6.5 μMUndeterminedRUS[300]
endolide B (279)/fungusPeptide gSerotonin receptor 5HT2b inhibition0.77 μM **UndeterminedDEU[301]
europhenol A (280)/fungusPolyketide eDHHP radical scavenging1.23 μg/mLUndeterminedCHN, HUN[302]
fascaplysin (281)/spongeAlkaloid gP-glycoprotein induction1 μM *UndeterminedIDN[303]
fennebricin A (282)/nudibranchAlkaloid gNF-κB inhibition1 μMUndeterminedCHN, HUN[304]
halicylindramide A(283)/spongePeptide gFXR receptor inhibition0.5 μMUndeterminedAUS, S.KOR[305]
isopetrosynol (284)/spongePolyketide ePTP1B inhibition8.2 μMUndeterminedIDN, JPN[306]
leptolyngbyolide B (285)/
cyanobacterium
Polyketide eF-actin depolymerization11.6 μMUndeterminedJPN[307]
mandelalide C (286)/ascidianPolyketide eMitochondrial complex V inhibition3.4 μMUndeterminedUSA[308]
19-methoxy-9,15-ene-puupehenol (287)/spongeTerpenoid fSR-B1 receptor activation1.78 μMUndeterminedAUS, MYS[309]
monanchomycalin B (288)/spongeAlkaloid gTRPV1, 2 and 3 receptor inhibition2.8–6.0 μMUndeterminedRUS, S. KOR[310]
monotriajaponide A (289)/spongePolyketide ePPAR-α and -β activation12.5 μM *UndeterminedCHN, ITA, USA[311]
mycalenitrile-15 (290)/spongeTerpenoid fPTP1B inhibition8.6 μM *UndeterminedCHN, ITA[312]
nakijinol G (291)/spongeTerpenoid fPTP1B inhibition4.8 μMUndeterminedCHN[313]
nazumazole D (292)/spongePeptide gchymotrypsin activity inhibition2 μMUndeterminedJPN[314]
pannorin (293)/fungusPolyketide eGSK-3β inhibition0.35 µMUndeterminedDEU[315]
perylenequinones (294, 295)/fungusPolyketide eBRD4 protein inhibition10 μM *UndeterminedCHN[316]
petroquinones A and B (296, 297)/spongePolyketide eUSP7 inhibition0.13–2.0 μMUndeterminedIND, JPN, NLD[317]
peyroisocoumarins B and D (298, 299)/fungusPolyketide eARE expression induction10 μM *UndeterminedCHN, DEU[318]
phomaketide A (300)/fungusPolyketide eAngiogenesis inhibition8.1 μMUndeterminedTWN[319]
phosphoeleganin (301)/ascidianPolyketide ePTP1B inhibition11 μMUndeterminedCHN, ITA[320]
P. rigida sesquiterpenes (302, 303)/soft coralTerpenoid fCDC25 phosphatases12–3.4 μMUndeterminedGRC, FRA[321]
S. ceylonensis diterpene (304)/spongeTerpenoid fUSP7 inhibition8.2 µMUndeterminedEGY, JPN[322]
schizols A and B (305, 306)/bryozoaPolyketide eABTS cation radical inhibition6.2–7.6 µMUndeterminedESP[323]
S. longispora isoindolinones (307, 308)/fungusAlkaloid gFibrinolytic activity25 μg/mL *UndeterminedCHN[324]
S. pertusa quinone (309)/spongeTerpenoid fCDK-2 inhibition4.8 µM **UndeterminedCHN[325]
sulawesin A (310)/spongeTerpenoid fUSP7 inhibition2.8 μMUndeterminedEGY, IDN, JPN, NLD[326]
urumamide (311)/cyanobacteriumPeptide gChymotrypsin inhibition33 μMUndeterminedJPN[327]
versiquinazoline B (312)/fungusAlkaloid gThioredoxin reductase inhibition12 µMUndeterminedCHN, DEU[328]
X. testudinaria steroidal ketone (313)/ spongeTerpenoid fPTP1B inhibition4.27 μMUndeterminedCHN[329]
a Organism, Kingdom Animalia: shrimp (Phylum Arthropoda); bryozoan (Phylum Bryozoa); ascidian (Phylum Chordata), hydroids, soft corals (Phylum Cnidaria), sea urchin (Phylum Echinodermata), sponge (Phylum Porifera); Kingdom Fungi: fungus; Kingdom Plantae: alga; Kingdom Monera: bacterium; b IC50: concentration of a compound required for 50% inhibition in vitro; *: estimated IC50; ** Kd: equilibrium constant ratio of forward and reverse constants; Ki: concentration needed to reduce the activity of an enzyme by half.*** 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; HUN: Hungary; IDN: Indonesia; ISR: Israel; ITA: Italy; JPN: Japan; MYS: Malaysia; NLD: The Netherlands; NZL: New Zealand; PNG: Papua New Guinea; PRT: Portugal; RUS: Russian Federation; S. KOR: South Korea; SWE: Sweden; THAI: Thailand; TWN: Taiwan; Chemistry: e Polyketide; f Terpene; g Nitrogen-containing compound; Abbreviations: ABTS: 2,2′-azinobis(3-ethylbenzothiazoline-6-sulphonic acid); Akt: protein kinase B; ARE: antioxidant response element; BRD4: bromodomain-containing protein 4; CDK: cyclin-dependent kinase; ChAT: choline acetyltansferase; CTFR: cystic fibrosis transmembrane conductance regulator; CLK: cdc2-like kinases; DDYRK: dual-specificity, tyrosine phosphorylation regulated kinase; DPPH: α, α-diphenyl-β-picrylhydrazyl; ECFC: endothelial colony-forming cell; ERK: extracellular signal-regulated kinase; FXR: farnesoid X receptor; GSK-3β: glycogen synthase kinase 3; Hsp90: heat shock protein 90; MAPK: mitogen-activated protein kinase; NAFLD: nonalcoholic fatty liver disease; NFATc1: nuclear factor of activated T cells, cytoplasmic 1; NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells; NO: nitric oxide; NOX2: NADPH oxidase 2; Nrf2:nuclear factor-erythroid 2-related factor 2; PARP: poly-ADP ribose polymerase; PKC: protein kinase C; PPAR: peroxisome proliferator-activated receptor; PTP1B: protein tyrosine phosphatase 1B; PPRE: peroxisome proliferator activated receptor response element; ROS: reactive oxygen species; SERCA2A: SR Ca2+ ATPase 2A; SETD8: lysine histone methyltransferase 5A; SOAT: sterol O-acyltransferase; SREBP-1c: sterol regulatory element-binding protein-1c;TLR5: Toll-like receptor 5; TCPTP: T-cell protein tyrosine phosphatase;TRP: tyrosinase-related protein; TRPM2: melastatin-like transient receptor potential ion channel; USP7: ubiquitin-specific protease 7; VEGF: vascular endothelial growth factor.
Figure 3. Marine pharmacology in 2016–2017: Marine compounds with miscellaneous mechanisms of action.
Figure 3. Marine pharmacology in 2016–2017: Marine compounds with miscellaneous mechanisms of action.
Marinedrugs 19 00049 g003aMarinedrugs 19 00049 g003bMarinedrugs 19 00049 g003cMarinedrugs 19 00049 g003dMarinedrugs 19 00049 g003eMarinedrugs 19 00049 g003fMarinedrugs 19 00049 g003gMarinedrugs 19 00049 g003hMarinedrugs 19 00049 g003i

5. Reviews on Marine Pharmacology and Pharmaceuticals

In 2016–2017 several reviews covered 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 2016 and 2017 [330,331]; chemistry and biology of guanidine natural products [332,333]; biological properties of secondary metabolites from sea hares of Aplysia genus [334]; alkynyl-containing peptides of marine cyanobacteria and molluscs [335]; bioactive cyanobacterial secondary metabolites for health [336]; biological active metabolites from marine-derived myxobacteria [337]; antimicrobial metabolites from the marine bacteria genus Pseudoalteromonas [338]; marine natural products from marine-derived Penicillium fungi [339]; biological activity of secondary metabolites from marine-algal-derived endophytic fungi [340]; pharmacological potential of fucosterol from marine algae [341]; pharmacological activities of Antarctic marine natural products [342]; bioactive acetylated triterpene glycosides from Holothuroidea in the past six decades [343]; terpenoids from octocorals of the genus Pachyclavularia [344]; bioactive marine natural products from sponges of the genus Hyrtios [345]; secondary metabolites from the marine sponge genus Phyllospongia [346]; discovery strategies of bioactive compounds synthetized by nonribosomal peptide synthetases and type-1 polyketide synthase derived from marine microbiomes [347]; developing natural product drugs: supply problems and how they have been overcome [348]; the global marine pharmaceutical pipeline in 2020: 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: antimycobacterial metabolites from marine invertebrates [349]; antimicrobials from cnidarians [350]; (c) antiprotozoal and antimalarial marine pharmacology: natural products in drug discovery against neglected tropical diseases [351]; antimycobacterial natural products from marine Pseudopterogorgia elisabethae [352]; (d) immuno- and anti-inflammatory marine pharmacology: marine natural products inhibitors of neutrophil-associated inflammation [353]; (e) cardiovascular and antidiabetic marine pharmacology: bioactive components from fish for dyslipidemia and cardiovascular risk reduction [354]; (f) nervous system marine pharmacology: marine natural products from sponges with neuroprotective activity [355]; a transcriptomic survey of ion channel-based conotoxin in the Chinese cone snail Conus betulinus [356]; dinoflagellate cyclic imine toxins as potent antagonists of nicotinic acetylcholine receptors [357]; inhibition of nociception and pain transmission by analgesic conopeptides ion channel inhibition by targeting G protein-coupled receptors [358]; (g) miscellaneous molecular targets and uses: ichthyotoxicity evaluation of marine natural products [359]; pharmacological potential of non-ribosomal peptides from marine sponges and tunicates [360]; aeroplysinin-1 as a multi-targeted bioactive sponge-derived natural product [361]; therapeutic potential of the phycotoxin yessotoxin [362], and new modalities for challenging drug targets in pharmaceutical discovery [363].

6. Conclusions

This marine pharmacology 2016–2017 review is the eleventh contribution to the marine preclinical pharmacology pipeline review series that was initiated by Alejandro M. S. Mayer (AMSM) in 1998 [1,2,3,4,5,6,7,8,9,10], with the aim of providing a systematic overview of selected peer-reviewed preclinical marine pharmacological literature. The global preclinical marine pharmacology research highlighted in this review involved chemists and pharmacologists from 54 countries, a remarkable increase from our last review, namely: Australia, Austria, Bangladesh, Belgium, Brazil, Canada, China, Costa Rica, Croatia, Cuba, Denmark, Egypt, Fiji, France, Germany, Greece, Hungary, Iceland, India, Indonesia, Ireland, Israel, Italy, Japan, Jordan, Kazakhstan, Malaysia, Mexico, Morocco, Myanmar, the Netherlands, Nigeria, New Zealand, Norway, Oman, Panama, Papua New Guinea, Philippines, Poland, Portugal, Russian Federation, Saudi Arabia, Serbia, Slovenia, South Korea, Spain, Sweden, Switzerland, Taiwan, Thailand, Turkey, United Kingdom, United States and Vietnam. Thus, during 2016–2017 the marine preclinical pharmaceutical pipeline continued to provide novel pharmacology and potential new leads for the marine clinical pharmaceutical pipeline. As shown at the marine preclinical and clinical pharmaceutical pipeline website: https://www.midwestern.edu/departments/marinepharmacology.xml there are currently 13 marine-derived pharmaceuticals approved by the U.S. Food and Drug Administration and 1 by Australia, and more than 23 marine-derived compounds in Phases I, II and III of global clinical pharmaceutical development.

Author Contributions

Marine pharmacology review conceptualization, A.M.S.M.; 2016–2017 literature analysis and investigation, A.M.S.M. and Aimee J. Guerrero (A.J.G.); writing—original draft preparation, A.M.S.M., A.J.G., Abimael Rodríguez (A.D.R.), Orazio Taglialatela-Scafati (O.T.-S.), Nobuhiro Fusetani (N.F.) and Fumiaki Nakamura (F.N.); writing—review and editing, A.M.S.M., A.J.G., A.D.R., O.T.-S., N.F. and F.N. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge financial support from Midwestern University to AMSM; and NIH-SC1 Award (Grant 1SC1GM086271-01A1) of the University of Puerto Rico to ADR, and a grant from Regione Campania-POR Campania FESR 2014/2020 “Combattere la resistenza tumorale: piattaforma integrata multidisciplinare per un approccio tecnologico innovativo alle oncoterapie-Campania Oncoterapie” (Project N. B61G18000470007) 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.

Acknowledgments

We thank the contributions of Kelly Le, Pharm. D., Jefferson Trieu, Pharm. D., Linda Huang, Pharm. D from the Chicago College of Pharmacy, Midwestern University for 2016–2017 literature retrieval, and Colleen Bannon, Midwestern University Library, for Endnote database assistance. We particularly wish to acknowledge Mary Hall’s tireless and careful review of the pharmacology data presented in 940 peer-reviewed articles retrieved from the global marine preclinical pharmacology 2016–2017 literature. We are grateful for the secretarial assistance of Victoria Sears and Laura Phelps, Pharmacology Department, College of Graduate Studies.

Conflicts of Interest

The authors declare no conflict 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.; Hamann, M.T. Marine pharmacology in 1999: Compounds with antibacterial, anticoagulant, antifungal, 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. Part 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]
  4. Mayer, A.M.; 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. Phys. C Toxicol. Pharmacol. 2005, 140, 265–286. [Google Scholar]
  5. Mayer, A.M.; Rodriguez, A.D.; Berlinck, R.G.; Hamann, M.T. Marine pharmacology in 2003–2004: 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. Phys. C Toxicol. Pharmacol. 2007, 145, 553–581. [Google Scholar]
  6. Mayer, A.M.; Rodriguez, A.D.; Berlinck, R.G.; Hamann, M.T. Marine pharmacology in 2005–2006: 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]
  7. Mayer, A.M.; Rodriguez, A.D.; Berlinck, R.G.; Fusetani, N. Marine pharmacology in 2007–2008: 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. Phys. C Toxicol. Pharmacol. 2011, 153, 191–222. [Google Scholar]
  8. Mayer, A.M.; 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]
  9. Mayer, A.M.S.; Rodriguez, 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. [Google Scholar]
  10. Mayer, A.M.S.; Guerrero, A.J.; Rodriguez, A.D.; Taglialatela-Scafati, O.; Nakamura, F.; Fusetani, N. Marine Pharmacology in 2014–2015: Marine Compounds with Antibacterial, Antidiabetic, Antifungal, Anti-Inflammatory, Antiprotozoal, Antituberculosis, Antiviral, and Anthelmintic Activities; Affecting the Immune and Nervous Systems, and Other Miscellaneous Mechanisms of Action. Mar. Drugs 2019, 18, 5. [Google Scholar]
  11. 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; Volume 1, pp. 197–308. [Google Scholar]
  12. Beesoo, R.; Bhagooli, R.; Neergheen-Bhujun, V.S.; Li, W.W.; Kagansky, A.; Bahorun, T. Antibacterial and antibiotic potentiating activities of tropical marine sponge extracts. Comp. Biochem. Phys. C Toxicol. Pharmacol. 2017, 196, 81–90. [Google Scholar] [CrossRef] [PubMed]
  13. Fernando, I.P.S.; Sanjeewa, K.K.A.; Kim, H.S.; Kim, S.Y.; Lee, S.H.; Lee, W.W.; Jeon, Y.J. Identification of sterols from the soft coral Dendronephthya gigantea and their anti-inflammatory potential. Environ. Toxicol. Pharmacol. 2017, 55, 37–43. [Google Scholar] [CrossRef] [PubMed]
  14. Joseph, F.R.S.; Iniyan, A.M.; Vincent, S.G.P. HR-LC-MS based analysis of two antibacterial metabolites from a marine sponge symbiont Streptomyces pharmamarensis ICN40. Microb. Pathog. 2017, 111, 450–457. [Google Scholar] [CrossRef]
  15. Al-Amoudi, S.; Essack, M.; Simoes, M.F.; Bougouffa, S.; Soloviev, I.; Archer, J.A.; Lafi, F.F.; Bajic, V.B. Bioprospecting Red Sea Coastal Ecosystems for Culturable Microorganisms and Their Antimicrobial Potential. Mar. Drugs 2016, 14, 165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Iniguez-Martinez, A.M.; Cardoso-Martinez, F.; de la Rosa, J.M.; Cueto, M.; Diaz-Marrero, A.R.; Darias, J.; Becerril-Espinosa, A.; Plata-Rosas, L.; Soria-Mercado, I.E. Compounds isolated from Salinispora arenicola of the Gulf of California, Mexico. Rev. Biol. Mar. Oceanogr. 2016, 51, 161–170. [Google Scholar] [CrossRef] [Green Version]
  17. Mayer, A.M.; 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] [Green Version]
  18. Swanson-Mungerson, M.; Incrocci, R.; Subramaniam, V.; Williams, P.; Hall, M.L.; Mayer, A.M.S. Effects of cyanobacteria Oscillatoria sp. lipopolysaccharide on B cell activation and Toll-like receptor 4 signaling. Toxicol. Lett. 2017, 275, 101–107. [Google Scholar] [CrossRef]
  19. Di, X.; Oskarsson, J.T.; Omarsdottir, S.; Freysdottir, J.; Hardardottir, I. Lipophilic fractions from the marine sponge Halichondria sitiens decrease secretion of pro-inflammatory cytokines by dendritic cells and decrease their ability to induce a Th1 type response by allogeneic CD4+ T cells. Pharm. Biol. 2017, 55, 2116–2122. [Google Scholar] [CrossRef] [Green Version]
  20. Cheng, C.; Othman, E.M.; Reimer, A.; Grune, M.; Kozjak-Pavlovic, V.; Stopper, H.; Hentschel, U.; Abdelmohsen, U.R. Ageloline A, new antioxidant and antichlamydial quinolone from the marine sponge-derived bacterium Streptomyces sp. SBT345. Tetrahedrom Lett. 2016, 57, 2786–2789. [Google Scholar] [CrossRef]
  21. Davison, J.R.; Lohith, K.M.; Wang, X.; Bobyk, K.; Mandadapu, S.R.; Lee, S.L.; Cencic, R.; Nelson, J.; Simpkins, S.; Frank, K.M.; et al. A New Natural Product Analog of Blasticidin S Reveals Cellular Uptake Facilitated by the NorA Multidrug Transporter. Antimicrob. Agents Chemother. 2017, 61. [Google Scholar] [CrossRef] [Green Version]
  22. Le, N.T.; Niemann, H.; Proksch, P.; Tait, K.; Linossier, I.; Rehel, K.; Hellio, C.; Fay, F. Sponge-Inspired Dibromohemibastadin Prevents and Disrupts Bacterial Biofilms without Toxicity. Mar. Drugs 2017, 15. [Google Scholar] [CrossRef] [Green Version]
  23. Wyche, T.P.; Alvarenga, R.F.R.; Piotrowski, J.S.; Duster, M.N.; Warrack, S.R.; Cornilescu, G.; De Wolfe, T.J.; Hou, Y.; Braun, D.R.; Ellis, G.A.; et al. Chemical Genomics, Structure Elucidation, and in Vivo Studies of the Marine-Derived Anticlostridial Ecteinamycin. ACS Chem. Biol. 2017, 12, 222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Tseng, S.P.; Hung, W.C.; Huang, C.Y.; Lin, Y.S.; Chan, M.Y.; Lu, P.L.; Lin, L.; Sheu, J.H. 5-Episinuleptolide Decreases the Expression of the Extracellular Matrix in Early Biofilm Formation of Multi-Drug Resistant Acinetobacter baumannii. Mar. Drugs 2016, 14, 143. [Google Scholar] [CrossRef] [PubMed]
  25. Solstad, R.G.; Li, C.; Isaksson, J.; Johansen, J.; Svenson, J.; Stensvag, K.; Haug, T. Novel Antimicrobial Peptides EeCentrocins 1, 2 and EeStrongylocin 2 from the Edible Sea Urchin Echinus esculentus Have 6-Br-Trp Post-Translational Modifications. PLoS ONE 2016, 11, e0151820. [Google Scholar] [CrossRef] [Green Version]
  26. Sung, A.A.; Gromek, S.M.; Balunas, M.J. Upregulation and Identification of Antibiotic Activity of a Marine-Derived Streptomyces sp. via Co-Cultures with Human Pathogens. Mar. Drugs 2017, 15, 250. [Google Scholar] [CrossRef] [Green Version]
  27. Adnani, N.; Chevrette, M.G.; Adibhatla, S.N.; Zhang, F.; Yu, Q.; Braun, D.R.; Nelson, J.; Simpkins, S.W.; McDonald, B.R.; Myers, C.L.; et al. Coculture of Marine Invertebrate-Associated Bacteria and Interdisciplinary Technologies Enable Biosynthesis and Discovery of a New Antibiotic, Keyicin. ACS Chem. Biol. 2017, 12, 3093–3102. [Google Scholar] [CrossRef]
  28. Mokhlesi, A.; Stuhldreier, F.; Wex, K.W.; Berscheid, A.; Hartmann, R.; Rehberg, N.; Sureechatchaiyan, P.; Chaidir, C.; Kassack, M.U.; Kalscheuer, R.; et al. Cyclic Cystine-Bridged Peptides from the Marine Sponge Clathria basilana Induce Apoptosis in Tumor Cells and Depolarize the Bacterial Cytoplasmic Membrane. J. Nat. Prod. 2017, 80, 2941–2952. [Google Scholar] [CrossRef]
  29. Leoni, G.; De, P.A.; Mardirossian, M.; Gambato, S.; Florian, F.; Venier, P.; Wilson, D.N.; Tossi, A.; Pallavicini, A.; Gerdol, M. Myticalins: A Novel Multigenic Family of Linear, Cationic Antimicrobial Peptides from Marine Mussels (Mytilus spp.). Mar. Drugs 2017, 15, 261. [Google Scholar] [CrossRef] [Green Version]
  30. Costantino, V.; Della, S.G.; Saurav, K.; Teta, R.; Bar-Shalom, R.; Mangoni, A.; Steindler, L. Plakofuranolactone as a Quorum Quenching Agent from the Indonesian Sponge Plakortis cf. lita. Mar. Drugs 2017, 15, 59. [Google Scholar] [CrossRef] [Green Version]
  31. Lee, B.C.; Lee, A.; Jung, J.H.; Choi, S.H.; Kim, T.S. In vitro and in vivo anti-Vibrio vulnificus activity of psammaplin A, a natural marine compound. Mol. Med. Rep 2016, 14, 2691–2696. [Google Scholar] [CrossRef]
  32. Song, Y.; Li, Q.; Qin, F.; Sun, C.; Liang, H.; Wei, X.; Wong, N.K.; Ye, L.; Zhang, Y.; Shao, M.; et al. Neoabyssomicins A–C, polycyclic macrolactones from the deep-sea derived Streptomyces koyangensis SCSIO 5802. Tetrahedron 2017, 73, 5366–5372. [Google Scholar] [CrossRef]
  33. Zhang, X.; Ye, X.; Chai, W.; Lian, X.Y.; Zhang, Z. New Metabolites and Bioactive Actinomycins from Marine-Derived Streptomyces sp. ZZ338. Mar. Drugs 2016, 14, 181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wang, D.; Wang, C.; Gui, P.; Liu, H.; Khalaf, S.M.H.; Elsayed, E.A.; Wadaan, M.A.M.; Hozzein, W.N.; Zhu, W. Identification, Bioactivity, and Productivity of Actinomycins from the Marine-Derived Streptomyces heliomycini. Front. Microbiol. 2017, 8, 1147. [Google Scholar] [CrossRef] [PubMed]
  35. Balan, S.S.; Kumar, C.G.; Jayalakshmi, S. Aneurinifactin, a new lipopeptide biosurfactant produced by a marine Aneurinibacillus aneurinilyticus SBP-11 isolated from Gulf of Mannar: Purification, characterization and its biological evaluation. Microbiol. Res. 2017, 194, 1–9. [Google Scholar] [CrossRef]
  36. Li, X.D.; Li, X.M.; Li, X.; Xu, G.M.; Liu, Y.; Wang, B.G. Aspewentins D-H, 20-Nor-isopimarane Derivatives from the Deep Sea Sediment-Derived Fungus Aspergillus wentii SD-310. J. Nat. Prod. 2016, 79, 1347–1353. [Google Scholar] [CrossRef]
  37. Chakraborty, K.; Thilakan, B.; Raola, V.K. Antimicrobial polyketide furanoterpenoids from seaweed-associated heterotrophic bacterium Bacillus subtilis MTCC 10403. Phytochemistry 2017, 142, 112–125. [Google Scholar] [CrossRef]
  38. Huang, H.; Liu, T.; Wu, X.; Guo, J.; Lan, X.; Zhu, Q.; Zheng, X.; Zhang, K. A new antibacterial chromone derivative from mangrove-derived fungus Penicillium aculeatum (No. 9EB). Nat. Prod. Res. 2017, 31, 2593–2598. [Google Scholar] [CrossRef]
  39. Tareq, F.S.; Shin, H.J. Bacilotetrins A and B, Anti-Staphylococcal Cyclic-Lipotetrapeptides from a Marine-Derived Bacillus subtilis. J. Nat. Prod. 2017, 80, 2889–2892. [Google Scholar] [CrossRef]
  40. Brana, A.F.; Sarmiento-Vizcaino, A.; Perez-Victoria, I.; Otero, L.; Fernandez, J.; Palacios, J.J.; Martin, J.; de la Cruz, M.; Diaz, C.; Vicente, F.; et al. Branimycins B and C, Antibiotics Produced by the Abyssal Actinobacterium Pseudonocardia carboxydivorans M-227. J. Nat. Prod. 2017, 80, 569–573. [Google Scholar] [CrossRef]
  41. Meng, L.H.; Wang, C.Y.; Mandi, A.; Li, X.M.; Hu, X.Y.; Kassack, M.U.; Kurtan, T.; Wang, B.G. Three Diketopiperazine Alkaloids with Spirocyclic Skeletons and One Bisthiodiketopiperazine Derivative from the Mangrove-Derived Endophytic Fungus Penicillium brocae MA-231. Org. Lett. 2016, 18, 5304–5307. [Google Scholar] [CrossRef]
  42. Wang, W.; Kim, H.; Patil, R.S.; Giri, A.G.; Won, D.H.; Hahn, D.; Sung, Y.; Lee, J.; Choi, H.; Nam, S.J.; et al. Cadiolides J-M, antibacterial polyphenyl butenolides from the Korean tunicate Pseudodistoma antinboja. Bioorg. Med. Chem. Lett. 2017, 27, 574–577. [Google Scholar] [CrossRef] [PubMed]
  43. Park, S.R.; Tripathi, A.; Wu, J.; Schultz, P.J.; Yim, I.; McQuade, T.J.; Yu, F.; Arevang, C.J.; Mensah, A.Y.; Tamayo-Castillo, G.; et al. Discovery of cahuitamycins as biofilm inhibitors derived from a convergent biosynthetic pathway. Nat. Commun. 2016, 7, 10710. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, C.X.; Ding, R.; Jiang, S.T.; Tang, J.S.; Hu, D.; Chen, G.D.; Lin, F.; Hong, K.; Yao, X.S.; Gao, H. Aldgamycins J-O, 16-Membered Macrolides with a Branched Octose Unit from Streptomycetes sp. and Their Antibacterial Activities. J. Nat. Prod. 2016, 79, 2446–2454. [Google Scholar] [CrossRef] [PubMed]
  45. Jiang, S.; Zhang, L.; Pei, X.; Deng, F.; Hu, D.; Chen, G.; Wang, C.; Hong, K.; Yao, X.; Gao, A.H. Chalcomycins from Marine-Derived Streptomyces sp. and Their Antimicrobial Activities. Mar. Drugs 2017, 15, 153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Liu, H.; Li, X.M.; Liu, Y.; Zhang, P.; Wang, J.N.; Wang, B.G. Chermesins A-D: Meroterpenoids with a Drimane-Type Spirosesquiterpene Skeleton from the Marine Algal-Derived Endophytic Fungus Penicillium chermesinum EN-480. J. Nat. Prod. 2016, 79, 806–811. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, S.; Chen, D.; Cai, R.; Cui, H.; Long, Y.; Lu, Y.; Li, C.; She, Z. Cytotoxic and Antibacterial Preussomerins from the Mangrove Endophytic Fungus Lasiodiplodia theobromae ZJ-HQ1. J. Nat. Prod. 2016, 79, 2397–2402. [Google Scholar] [CrossRef]
  48. Lee, J.H.; Kim, E.; Choi, H.; Lee, J. Collismycin C from the Micronesian Marine Bacterium Streptomyces sp. MC025 Inhibits Staphylococcus aureus Biofilm Formation. Mar. Drugs 2017, 15, 387. [Google Scholar] [CrossRef] [Green Version]
  49. Wu, B.; Wiese, J.; Wenzel-Storjohann, A.; Malien, S.; Schmaljohann, R.; Imhoff, J.F. Engyodontochones, Antibiotic Polyketides from the Marine Fungus Engyodontium album Strain LF069. Chemistry 2016, 22, 7452–7462. [Google Scholar] [CrossRef]
  50. Li, H.L.; Li, X.M.; Li, X.; Wang, C.Y.; Liu, H.; Kassack, M.U.; Meng, L.H.; Wang, B.G. Antioxidant Hydroanthraquinones from the Marine Algal-Derived Endophytic Fungus Talaromyces islandicus EN-501. J. Nat. Prod. 2017, 80, 162–168. [Google Scholar] [CrossRef]
  51. Nguyen, H.M.; Ito, T.; Kurimoto, S.I.; Ogawa, M.; Win, N.N.; Hung, V.Q.; Nguyen, H.T.; Kubota, T.; Kobayashi, J.; Morita, H. New merosesquiterpenes from a Vietnamese marine sponge of Spongia sp. and their biological activities. Bioorg. Med. Chem. Lett. 2017, 27, 3043–3047. [Google Scholar] [CrossRef]
  52. Lee, J.; Shin, A.; Lee, H. Isolation and synthesis of luffariellolide derivatives and evaluation of antibacterial activities against gram-negative bacteria. Bull. Korean Chem. Soc. 2017, 38, 804–807. [Google Scholar] [CrossRef]
  53. Kim, C.K.; Riswanto, R.; Won, T.H.; Kim, H.; Elya, B.; Sim, C.J.; Oh, D.C.; Oh, K.B.; Shin, J. Manzamine Alkaloids from an Acanthostrongylophora sp. Sponge. J. Nat. Prod. 2017, 80, 1575–1583. [Google Scholar] [CrossRef] [PubMed]
  54. Lacret, R.; Perez-Victoria, I.; Oves-Costales, D.; de la Cruz, M.; Domingo, E.; Martin, J.; Diaz, C.; Vicente, F.; Genilloud, O.; Reyes, F. MDN-0170, a New Napyradiomycin from Streptomyces sp. Strain CA-271078. Mar. Drugs 2016, 14, 188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Nenkep, V.; Yun, K.; Son, B.W. Oxysporizoline, an antibacterial polycyclic quinazoline alkaloid from the marine-mudflat-derived fungus Fusarium oxysporum. J. Antibiot. 2016, 69, 709–711. [Google Scholar] [CrossRef] [PubMed]
  56. Zheng, C.J.; Huang, G.L.; Xu, Y.; Song, X.M.; Yao, J.; Liu, H.; Wang, R.P.; Sun, X.P. A new benzopyrans derivatives from a mangrove-derived fungus Penicillium citrinum from the South China Sea. Nat. Prod. Res. 2016, 30, 821–825. [Google Scholar] [CrossRef]
  57. Auckloo, B.N.; Pan, C.; Akhter, N.; Wu, B.; Wu, X.; He, S. Stress-Driven Discovery of Novel Cryptic Antibiotics from a Marine Fungus Penicillium sp. BB1122. Front. Microbiol. 2017, 8, 1450. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, C.F.; Wang, Y.; Zhang, X.L.; Wei, M.Y.; Wang, C.Y.; Shao, C.L. Two Dichlorinated Benzophenone Derivatives from the Soft Coral-Derived Pestalotiopsis sp. Fungus and Their Antibacterial Activity. Chem. Nat. Compd. 2017, 53, 1174–1176. [Google Scholar] [CrossRef]
  59. Pan, C.; Shi, Y.; Auckloo, B.N.; Hassan, S.S.U.; Akhter, N.; Wang, K.; Ye, Y.; Arthur Chen, C.T.; Tao, X.; Wu, B. Isolation and Antibiotic Screening of Fungi from a Hydrothermal Vent Site and Characterization of Secondary Metabolites from a Penicillium Isolate. Mar. Biotechnol. 2017, 19, 469–479. [Google Scholar] [CrossRef]
  60. Shi, T.; Qi, J.; Shao, C.L.; Zhao, D.L.; Hou, X.M.; Wang, C.Y. Bioactive Diphenyl Ethers and Isocoumarin Derivatives from a Gorgonian-Derived Fungus Phoma sp. (TA07-1). Mar. Drugs 2017, 15, 146. [Google Scholar] [CrossRef] [Green Version]
  61. Wang, W.; Liao, Y.; Tang, C.; Huang, X.; Luo, Z.; Chen, J.; Cai, P. Cytotoxic and Antibacterial Compounds from the Coral-Derived Fungus Aspergillus tritici SP2-8-1. Mar. Drugs 2017, 15, 348. [Google Scholar] [CrossRef] [Green Version]
  62. Tedesco, P.; Maida, I.; Palma, E.F.; Tortorella, E.; Subko, K.; Ezeofor, C.C.; Zhang, Y.; Tabudravu, J.; Jaspars, M.; Fani, R.; et al. Antimicrobial Activity of Monoramnholipids Produced by Bacterial Strains Isolated from the Ross Sea (Antarctica). Mar. Drugs 2016, 14, 83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zhang, X.; Xu, H.Y.; Huang, A.M.; Wang, L.; Wang, Q.; Cao, P.Y.; Yang, P.M. Antibacterial Meroterpenoids from the South China Sea Sponge Dysidea sp. Chem. Pharm. Bull. 2016, 64, 1036–1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Mangamuri, U.; Muvva, V.; Poda, S.; Naragani, K.; Munaganti, R.K.; Chitturi, B.; Yenamandra, V. Bioactive metabolites produced by Streptomyces Cheonanensis VUK-A from Coringa mangrove sediments: Isolation, structure elucidation and bioactivity. 3 Biotech. 2016, 6, 63. [Google Scholar] [CrossRef] [PubMed]
  65. Williams, D.E.; Dalisay, D.S.; Chen, J.; Polishchuck, E.A.; Patrick, B.O.; Narula, G.; Ko, M.; Av-Gay, Y.; Li, H.; Magarvey, N.; et al. Aminorifamycins and Sporalactams Produced in Culture by a Micromonospora sp. Isolated from a Northeastern-Pacific Marine Sediment Are Potent Antibiotics. Org. Lett. 2017, 19, 766–769. [Google Scholar] [CrossRef] [PubMed]
  66. Tan, Y.; Hu, Y.; Wang, Q.; Zhou, H.; Wang, Y.; Gan, M. Tetrocarcins N and O, glycosidic spirotetronates from a marine-derived Micromonospora sp. identified by PCR-based screening. RSC Adv. 2016, 6, 91773–91778. [Google Scholar] [CrossRef]
  67. Gui, C.; Zhang, S.; Zhu, X.; Ding, W.; Huang, H.; Gu, Y.C.; Duan, Y.; Ju, J. Antimicrobial Spirotetronate Metabolites from Marine-Derived Micromonospora harpali SCSIO GJ089. J. Nat. Prod. 2017, 80, 1594–1603. [Google Scholar] [CrossRef]
  68. Okada, M.; Sugita, T.; Wong, C.P.; Wakimoto, T.; Abe, I. Identification of Pyridinium with Three Indole Moieties as an Antimicrobial Agent. J. Nat. Prod. 2017, 80, 1205–1209. [Google Scholar] [CrossRef]
  69. Zubair, M.S.; Alarif, W.M.; Al-Footy, K.O.; Ph, M.; Ali, M.; Basaif, S.A.; Al-Lihaibi, S.S.; Ayyad, S.E. New antimicrobial biscembrane hydrocarbon and cembranoid diterpenes from the soft coral Sarcophyton trocheliophorum. Turk. J. Chem. 2016, 40, 385–392. [Google Scholar] [CrossRef]
  70. Liu, H.B.; Lauro, G.; O’Connor, R.D.; Lohith, K.; Kelly, M.; Colin, P.; Bifulco, G.; Bewley, C.A. Tulongicin, an Antibacterial Tri-Indole Alkaloid from a Deep-Water Topsentia sp. Sponge. J. Nat. Prod. 2017, 80, 2556–2560. [Google Scholar] [CrossRef]
  71. Hu, Z.; Qin, L.; Wang, Q.; Ding, W.; Chen, Z.; Ma, Z. Angucycline antibiotics and its derivatives from marine-derived actinomycete Streptomyces sp. A6H. Nat. Prod. Res. 2016, 30, 2551–2558. [Google Scholar] [CrossRef]
  72. Espiritu, R.A. Membrane permeabilizing action of amphidinol 3 and theonellamide A in raft-forming lipid mixtures. Z. Naturforsch. C 2017, 72, 43–48. [Google Scholar] [CrossRef] [PubMed]
  73. Iwamoto, M.; Sumino, A.; Shimada, E.; Kinoshita, M.; Matsumori, N.; Oiki, S. Channel Formation and Membrane Deformation via Sterol-Aided Polymorphism of Amphidinol 3. Sci. Rep 2017, 7, 10782. [Google Scholar] [CrossRef] [PubMed]
  74. Pejin, B.; Ciric, A.; Markovic, D.; Tommonaro, G.; Sokovic, M. In vitro avarol does affect the growth of Candida sp. Nat. Prod. Res. 2016, 30, 1956–1960. [Google Scholar] [CrossRef] [PubMed]
  75. Quezada, M.; Licona-Cassani, C.; Cruz-Morales, P.; Salim, A.A.; Marcellin, E.; Capon, R.J.; Barona-Gomez, F. Diverse Cone-Snail Species Harbor Closely Related Streptomyces Species with Conserved Chemical and Genetic Profiles, Including Polycyclic Tetramic Acid Macrolactams. Front. Microbiol. 2017, 8, 2305. [Google Scholar] [CrossRef] [PubMed]
  76. Jiao, W.H.; Hong, L.L.; Sun, J.B.; Piao, S.J.; Chen, G.D.; Deng, H.; Wang, S.P.; Yang, F.; Lin, H.W. (±)-Hippolide J—A pair of unusual antifungal enantiomeric sesterterpenoids from the marine sponge Hippospongia lachne. Eur. J. Org. Chem 2017, 2017, 3421–3426. [Google Scholar] [CrossRef]
  77. Kildgaard, S.; Subko, K.; Phillips, E.; Goidts, V.; de la Cruz, M.; Diaz, C.; Gotfredsen, C.H.; Andersen, B.; Frisvad, J.C.; Nielsen, K.F.; et al. A Dereplication and Bioguided Discovery Approach to Reveal New Compounds from a Marine-Derived Fungus Stilbella fimetaria. Mar. Drugs 2017, 15, 253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Son, S.; Ko, S.K.; Jang, M.; Kim, J.W.; Kim, G.S.; Lee, J.K.; Jeon, E.S.; Futamura, Y.; Ryoo, I.J.; Lee, J.S.; et al. New Cyclic Lipopeptides of the Iturin Class Produced by Saltern-Derived Bacillus sp. KCB14S006. Mar. Drugs 2016, 14, 72. [Google Scholar] [CrossRef] [Green Version]
  79. Okabe, M.; Sugita, T.; Kinoshita, K.; Koyama, K. Macrolides from a Marine-Derived Fungus, Penicillium meleagrinum var. viridiflavum, Showing Synergistic Effects with Fluconazole against Azole-Resistant Candida albicans. J. Nat. Prod. 2016, 79, 1208–1212. [Google Scholar] [CrossRef]
  80. Jamison, M.T.; Dalisay, D.S.; Molinski, T.F. Peroxide Natural Products from Plakortis zyggompha and the Sponge Association Plakortis halichondrioides-Xestospongia deweerdtae: Antifungal Activity against Cryptococcus gattii. J. Nat. Prod. 2016, 79, 555–563. [Google Scholar] [CrossRef]
  81. Calabro, K.; Kalahroodi, E.L.; Rodrigues, D.; Diaz, C.; Cruz, M.; Cautain, B.; Laville, R.; Reyes, F.; Perez, T.; Soussi, B.; et al. Poecillastrosides, Steroidal Saponins from the Mediterranean Deep-Sea Sponge Poecillastra compressa (Bowerbank, 1866). Mar. Drugs 2017, 15, 199. [Google Scholar] [CrossRef] [Green Version]
  82. Aksoy, S.C.; Uzel, A.; Bedir, E. Cytosine-type nucleosides from marine-derived Streptomyces rochei 06CM016. J. Antibiot. 2016, 69, 51–56. [Google Scholar] [CrossRef] [PubMed]
  83. Yang, F.; Wang, R.P.; Xu, B.; Yu, H.B.; Ma, G.Y.; Wang, G.F.; Dai, S.W.; Zhang, W.; Jiao, W.H.; Song, S.J.; et al. New antimalarial norterpene cyclic peroxides from Xisha Islands sponge Diacarnus megaspinorhabdosa. Bioorg. Med. Chem. Lett. 2016, 26, 2084–2087. [Google Scholar] [CrossRef] [PubMed]
  84. Almaliti, J.; Malloy, K.L.; Glukhov, E.; Spadafora, C.; Gutierrez, M.; Gerwick, W.H. Dudawalamides A-D, Antiparasitic Cyclic Depsipeptides from the Marine Cyanobacterium Moorea producens. J. Nat. Prod. 2017, 80, 1827–1836. [Google Scholar] [CrossRef] [PubMed]
  85. Chan, S.T.; Nani, R.R.; Schauer, E.A.; Martin, G.E.; Williamson, R.T.; Sauri, J.; Buevich, A.V.; Schafer, W.A.; Joyce, L.A.; Goey, A.K.; et al. Characterization and Synthesis of Eudistidine C, a Bioactive Marine Alkaloid with an Intriguing Molecular Scaffold. J. Org. Chem. 2016, 81, 10631–10640. [Google Scholar] [CrossRef] [PubMed]
  86. Buedenbender, L.; Grkovic, T.; Duffy, S.; Kurtboke, D.I.; Avery, V.K.; Carroll, A.R. Naseseazine C, a new anti-plasmodial dimeric diketopiperazine from a marine sediment derived Streptomyces sp. Tetrahedrom Lett. 2016, 57, 5893–5895. [Google Scholar] [CrossRef]
  87. Wang, J.; Pearce, A.N.; Chan, S.T.; Taylor, R.B.; Page, M.J.; Valentin, A.; Bourguet-Kondracki, M.L.; Dalton, J.P.; Wiles, S.; Copp, B.R. Biologically Active Acetylenic Amino Alcohol and N-Hydroxylated 1,2,3,4-Tetrahydro-beta-carboline Constituents of the New Zealand Ascidian Pseudodistoma opacum. J. Nat. Prod. 2016, 79, 607–610. [Google Scholar] [CrossRef] [PubMed]
  88. Campos, P.E.; Wolfender, J.L.; Queiroz, E.F.; Marcourt, L.; Al-Mourabit, A.; Frederich, M.; Bordignon, A.; De Voogd, N.; Illien, B.; Gauvin-Bialecki, A. Unguiculin A and Ptilomycalins E-H, Antimalarial Guanidine Alkaloids from the Marine Sponge Monanchora unguiculata. J. Nat. Prod. 2017, 80, 1404–1410. [Google Scholar] [CrossRef]
  89. White, A.M.; Dao, K.; Vrubliauskas, D.; Konst, Z.A.; Pierens, G.K.; Mandi, A.; Andrews, K.T.; Skinner-Adams, T.S.; Clarke, M.E.; Narbutas, P.T.; et al. Catalyst-Controlled Stereoselective Synthesis Secures the Structure of the Antimalarial Isocyanoterpene Pustulosaisonitrile-1. J. Org. Chem. 2017, 82, 13313–13323. [Google Scholar] [CrossRef]
  90. Carballeira, N.M.; Montano, N.; Amador, L.A.; Rodriguez, A.D.; Golovko, M.Y.; Golovko, S.A.; Reguera, R.M.; Alvarez-Velilla, R.; Balana-Fouce, R. Novel Very Long-Chain alpha-Methoxylated Delta5,9 Fatty Acids from the Sponge Asteropus niger Are Effective Inhibitors of Topoisomerases IB. Lipids 2016, 51, 245–256. [Google Scholar] [CrossRef] [Green Version]
  91. Cardoso-Martinez, F.; de la Rosa, J.M.; Diaz-Marrero, A.R.; Darias, J.; D’Croz, L.; Jimenez-Anton, M.D.; Corral, M.J.; Garcia, R.; Alunda, J.M.; Cueto, M. Oxysterols from an octocoral of the genus Gorgonia from the eastern Pacific of Panama. RSC Adv. 2016, 6, 38579–38591. [Google Scholar] [CrossRef] [Green Version]
  92. Chianese, G.; Silber, J.; Luciano, P.; Merten, C.; Erpenbeck, D.; Topaloglu, B.; Kaiser, M.; Tasdemir, D. Antiprotozoal Linear Furanosesterterpenoids from the Marine Sponge Ircinia oros. J. Nat. Prod. 2017, 80, 2566–2571. [Google Scholar] [CrossRef] [PubMed]
  93. Ogawa, H.; Iwasaki, A.; Sumimoto, S.; Kanamori, Y.; Ohno, O.; Iwatsuki, M.; Ishiyama, A.; Hokari, R.; Otoguro, K.; Omura, S.; et al. Janadolide, a Cyclic Polyketide-Peptide Hybrid Possessing a tert-Butyl Group from an Okeania sp. Marine Cyanobacterium. J. Nat. Prod. 2016, 79, 1862–1866. [Google Scholar] [CrossRef] [PubMed]
  94. Notarte, K.; Nakao, Y.; Yaguchi, T.; Bungihan, M.; Suganuma, K.; Delacruz, T. Trypanocidal activity, cytotoxicity and histone modifications induced by malformin A1 isolated from the marine-derived fungus Aspergillus tubingensis IFM 63452. Mycosphere 2017, 8, 111–120. [Google Scholar] [CrossRef]
  95. Elsayed, Y.; Refaat, J.; Abdelmohsen, U.; Ahmed, S.; Fouad, M. Rhodozepinone, a new antitrypanosomal azepino-diindole alkaloid from the marine sponge-derived bacterium Rhodococcus sp. UA13. Med. Chem. Res. 2017, 26, 2751–2760. [Google Scholar] [CrossRef]
  96. Arai, M.; Yamano, Y.; Kamiya, K.; Setiawan, A.; Kobayashi, M. Anti-dormant mycobacterial activity and target molecule of melophlins, tetramic acid derivatives isolated from a marine sponge of Melophlus sp. J. Nat. Med. 2016, 70, 467–475. [Google Scholar] [CrossRef] [PubMed]
  97. Rodrigues, F.C.; Gupta, R.; Geden, S.; Roberts, J.; Winder, P.; Pomponi, S.A.; Diaz, M.C.; Reed, J.K.; Wright, A.E.; Rohde, K.H. Selective Killing of Dormant Mycobacterium tuberculosis by Marine Natural Products. Antimicrob. Agents Chemother. 2017, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Luo, X.; Zhou, X.; Lin, X.; Qin, X.; Zhang, T.; Wang, J.; Tu, Z.; Yang, B.; Liao, S.; Tian, Y.; et al. Antituberculosis compounds from a deep-sea-derived fungus Aspergillus sp. SCSIO Ind09F01. Nat. Prod. Res. 2017, 31, 1958–1962. [Google Scholar] [CrossRef] [PubMed]
  99. Huang, P.; Xie, F.; Ren, B.; Wang, Q.; Wang, J.; Wang, Q.; Abdel-Mageed, W.M.; Liu, M.; Han, J.; Oyeleye, A.; et al. Anti-MRSA and anti-TB metabolites from marine-derived Verrucosispora sp. MS100047. Appl. Microbiol. Biotechnol. 2016, 100, 7437–7447. [Google Scholar] [CrossRef]
  100. Jimenez-Romero, C.; Rode, J.E.; Perez, Y.M.; Franzblau, S.G.; Rodriguez, A.D. Exploring the Sponge Consortium Plakortis symbiotica-Xestospongia deweerdtae as a Potential Source of Antimicrobial Compounds and Probing the Pharmacophore for Antituberculosis Activity of Smenothiazole A by Diverted Total Synthesis. J. Nat. Prod. 2017, 80, 2295–2303. [Google Scholar] [CrossRef]
  101. Chen, S.; He, L.; Chen, D.; Cai, R.; Long, Y.; Lu, Y.; She, Z. Talaramide A, an unusual alkaloid from the mangrove endophytic fungus Talaromyces sp. (HZ-YX1) as an inhibitor of mycobacterial PknG. New J. Chem. 2017, 41, 4273–4276. [Google Scholar] [CrossRef]
  102. Kamiya, K.A.; Setiawan, M.; Kobayashi, A.M. Anti-dormant mycobacterial activity of viomellein and xanthomegnin, naphthoquinone dimers produced by marine-derived Aspergillus sp. Nat. Prod. Commun. 2017, 12, 579–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. O’Rourke, A.; Kremb, S.; Bader, T.M.; Helfer, M.; Schmitt-Kopplin, P.; Gerwick, W.H.; Brack-Werner, R.; Voolstra, C.R. Alkaloids from the Sponge Stylissa carteri Present Prospective Scaffolds for the Inhibition of Human Immunodeficiency Virus 1 (HIV-1). Mar. Drugs 2016, 14, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yamashita, A.; Tamaki, M.; Kasai, H.; Tanaka, T.; Otoguro, T.; Ryo, A.; Maekawa, S.; Enomoto, N.; de Voogd, N.J.; Tanaka, J.; et al. Inhibitory effects of metachromin A on hepatitis B virus production via impairment of the viral promoter activity. Antiviral. Res. 2017, 145, 136–145. [Google Scholar] [CrossRef] [PubMed]
  105. Ishikawa, C.; Jomori, T.; Tanaka, J.; Senba, M.; Mori, N. Peridinin, a carotenoid, inhibits proliferation and survival of HTLV-1-infected T-cell lines. Int. J. Oncol. 2016, 49, 1713–1721. [Google Scholar] [CrossRef] [Green Version]
  106. Niu, S.; Si, L.; Liu, D.; Zhou, A.; Zhang, Z.; Shao, Z.; Wang, S.; Zhang, L.; Zhou, D.; Lin, W. Spiromastilactones: A new class of influenza virus inhibitors from deep-sea fungus. Eur. J. Med. Chem. 2016, 108, 229–244. [Google Scholar] [CrossRef]
  107. Kim, S.H.; Ha, T.K.; Oh, W.K.; Shin, J.; Oh, D.C. Antiviral Indolosesquiterpenoid Xiamycins C-E from a Halophilic Actinomycete. J. Nat. Prod. 2016, 79, 51–58. [Google Scholar] [CrossRef]
  108. Cheng, Y.B.; Lee, J.C.; Lo, I.W.; Chen, S.R.; Hu, H.C.; Wu, Y.H.; Wu, Y.C.; Chang, F.R. Ecdysones from Zoanthus spp. with inhibitory activity against dengue virus 2. Bioorg. Med. Chem. Lett. 2016, 26, 2344–2348. [Google Scholar] [CrossRef] [Green Version]
  109. Hermawan, I.; Furuta, A.; Higashi, M.; Fujita, Y.; Akimitsu, N.; Yamashita, A.; Moriishi, K.; Tsuneda, S.; Tani, H.; Nakakoshi, M.; et al. Four Aromatic Sulfates with an Inhibitory Effect against HCV NS3 Helicase from the Crinoid Alloeocomatella polycladia. Mar. Drugs 2017, 15, 117. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, M.; Tietjen, I.; Chen, M.; Williams, D.E.; Daoust, J.; Brockman, M.A.; Andersen, R.J. Sesterterpenoids Isolated from the Sponge Phorbas sp. Activate Latent HIV-1 Provirus Expression. J. Org. Chem. 2016, 81, 11324–11334. [Google Scholar] [CrossRef]
  111. Ma, X.; Nong, X.H.; Ren, Z.; Wang, J.; Liang, X.; Wang, L.; Qi, S.H. Antiviral peptides from marine gorgonian-derived fungus Aspergillus sp. SCSIO 41501. Tetrahedron 2017, 58, 1151–1155. [Google Scholar] [CrossRef]
  112. Huang, Z.; Nong, X.; Ren, Z.; Wang, J.; Zhang, X.; Qi, S. Anti-HSV-1, antioxidant and antifouling phenolic compounds from the deep-sea-derived fungus Aspergillus versicolor SCSIO 41502. Bioorg. Med. Chem. Lett. 2017, 27, 787–791. [Google Scholar] [CrossRef] [PubMed]
  113. Tian, Y.Q.; Lin, X.P.; Wang, Z.; Zhou, X.F.; Qin, X.C.; Kaliyaperumal, K.; Zhang, T.Y.; Tu, Z.C.; Liu, Y. Asteltoxins with Antiviral Activities from the Marine Sponge-Derived Fungus Aspergillus sp. SCSIO XWS02F40. Molecules 2016, 21, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Yuan, W.; Cheng, S.; Fu, W.; Zhao, M.; Li, X.; Cai, Y.; Dong, J.; Huang, K.; Gustafson, K.R.; Yan, P. Structurally Diverse Metabolites from the Soft Coral Sinularia verruca Collected in the South China Sea. J. Nat. Prod. 2016, 79, 1124–1131. [Google Scholar] [CrossRef] [PubMed]
  115. Niu, S.; Liu, D.; Shao, Z.; Proksch, P.; Lin, W. Eutypellazines A-M, thiodiketopiperazine-type alkaloids from deep sea derived fungus Eutypella sp. MCCC 3A00281. RSC Adv. 2017, 7, 33580–33590. [Google Scholar] [CrossRef] [Green Version]
  116. Hawas, U.W.; Al-Farawati, R.; Abou El-Kassem, L.T.; Turki, A.J. Different Culture Metabolites of the Red Sea Fungus Fusarium equiseti Optimize the Inhibition of Hepatitis C Virus NS3/4A Protease (HCV PR). Mar. Drugs 2016, 14, 190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Zhou, X.; Fang, W.; Tan, S.; Lin, X.; Xun, T.; Yang, B.; Liu, S.; Liu, Y. Aspernigrins with anti-HIV-1 activities from the marine-derived fungus Aspergillus niger SCSIO Jcsw6F30. Bioorg. Med. Chem. Lett. 2016, 26, 361–365. [Google Scholar] [CrossRef]
  118. Palem, J.R.; Mudit, M.; Hsia, S.V.; Sayed, K.A. Discovery and preliminary structure-activity relationship of the marine natural product manzamines as herpes simplex virus type-1 inhibitors. Z. Naturforsch. C 2017, 72, 49–54. [Google Scholar] [CrossRef]
  119. Lee, J.C.; Chang, F.R.; Chen, S.R.; Wu, Y.H.; Hu, H.C.; Wu, Y.C.; Backlund, A.; Cheng, Y.B. Anti-Dengue Virus Constituents from Formosan Zoanthid Palythoa mutuki. Mar. Drugs 2016, 14, 151. [Google Scholar] [CrossRef] [Green Version]
  120. Zhang, P.; Li, Y.; Jia, C.; Lang, J.; Niaz, S.I.; Yuan, J.; Yu, J.; Chen, S.; Liu, L. Antiviral and anti-inflammatory meroterpenoids: Stachybonoids A–F from the crinoid-derived fungus Stachybotrys chartarum 952. RSC Adv. 2017, 7, 49910–49916. [Google Scholar] [CrossRef] [Green Version]
  121. Cheng, W.; Ren, J.; Huang, Q.; Long, H.; Jin, H.; Zhang, L.; Liu, H.; Van, O.L.; Lin, W. Pregnane steroids from a gorgonian coral Subergorgia suberosa with anti-flu virus effects. Steroids 2016, 108, 99–104. [Google Scholar] [CrossRef] [PubMed]
  122. Abdjul, D.B.; Yamazaki, H.; Kanno, S.; Takahashi, O.; Kirikoshi, R.; Ukai, K.; Namikoshi, M. Haliclonadiamine Derivatives and 6-epi-Monanchorin from the Marine Sponge Halichondria panicea Collected at Iriomote Island. J. Nat. Prod. 2016, 79, 1149–1154. [Google Scholar] [CrossRef] [PubMed]
  123. Al-Footy, K.O.; Alarif, W.M.; Zubair, M.S.; Ghandourah, M.A.; Aly, M.M. Antibacterial and cytotoxic properties of isoprenoids from the red sea soft coral, Lobophytum sp. Trop. J. Pharm. Res. 2016, 15, 1431–1438. [Google Scholar] [CrossRef] [Green Version]
  124. Angulo-Preckler, C.; Genta-Jouve, G.; Mahajan, N.; de la Cruz, M.; de Pedro, N.; Reyes, F.; Iken, K.; Avila, C.; Thomas, O.P. Gersemiols A-C and Eunicellol A, Diterpenoids from the Arctic Soft Coral Gersemia fruticosa. J. Nat. Prod. 2016, 79, 1132–1136. [Google Scholar] [CrossRef] [PubMed]
  125. Kamada, T.; Phan, C.S.; Tin, H.S.; Vairappan, C.S.; Muhammad, T.S.T. 16-Hydroxycembra-1,3,7,11-tetraene, a new Cembrane Diterpene from Malaysian Soft Coral Genus Sarcophyton. Nat. Prod. Commun. 2016, 11, 1077–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Kubota, T.; Nakamura, K.; Sakai, K.; Fromont, J.; Gonoi, T.; Kobayashi, J. Hyrtinadines C and D, New Azepinoindole-Type Alkaloids from a Marine Sponge Hyrtios sp. Chem. Pharm. Bull. 2016, 64, 975–978. [Google Scholar] [CrossRef] [Green Version]
  127. Li, C.; La, M.P.; Tang, H.; Sun, P.; Liu, B.S.; Zhuang, C.L.; Yi, Y.H.; Zhang, W. Chemistry and Bioactivity of Briaranes from the South China Sea Gorgonian Dichotella gemmacea. Mar. Drugs 2016, 14, 201. [Google Scholar] [CrossRef] [Green Version]
  128. Nguyen, H.M.; Ito, T.; Win, N.N.; Kodama, T.; Hung, V.Q.; Nguyen, H.T.; Morita, H. New antibacterial sesquiterpene aminoquinones from a Vietnamese marine sponge of Spongia sp. Phytochem. Lett. 2016, 17, 288–292. [Google Scholar] [CrossRef]
  129. Phan, C.S.; Ng, S.Y.; Kamada, T.; Vairappan, C.S. Two New Lobane Diterpenes from a Bornean Soft Coral Sinularia sp. Nat. Prod. Commun. 2016, 11, 899–900. [Google Scholar] [CrossRef] [Green Version]
  130. Von Salm, J.L.; Witowski, C.G.; Fleeman, R.M.; McClintock, J.B.; Amsler, C.D.; Shaw, L.N.; Baker, B.J. Darwinolide, a New Diterpene Scaffold That Inhibits Methicillin-Resistant Staphylococcus aureus Biofilm from the Antarctic Sponge Dendrilla membranosa. Org. Lett. 2016, 18, 2596–2599. [Google Scholar] [CrossRef]
  131. Brasseur, L.; Hennebert, E.; Fievez, L.; Caulier, G.; Bureau, F.; Tafforeau, L.; Flammang, P.; Gerbaux, P.; Eeckhaut, I. The Roles of Spinochromes in Four Shallow Water Tropical Sea Urchins and Their Potential as Bioactive Pharmacological Agents. Mar. Drugs 2017, 15, 179. [Google Scholar] [CrossRef] [Green Version]
  132. Kong, F.D.; Ma, Q.Y.; Huang, S.Z.; Wang, P.; Wang, J.F.; Zhou, L.M.; Yuan, J.Z.; Dai, H.F.; Zhao, Y.X. Chrodrimanins K-N and Related Meroterpenoids from the Fungus Penicillium sp. SCS-KFD09 Isolated from a Marine Worm, Sipunculus nudus. J. Nat. Prod. 2017, 80, 1039–1047. [Google Scholar] [CrossRef] [PubMed]
  133. Mohammed, R.; Radwan, M.; Ma, G.; Mohamed, T.A.; Seliem, M.A.; Thabet, M.; ElSohly, M.A. Bioactive sterols and sesquiterpenes from the Red Sea soft coral Sinularia terspilli. Med. Chem. Res. 2017, 26, 1647–1652. [Google Scholar] [CrossRef]
  134. Liang, X.R.; Miao, F.P.; Song, Y.P.; Liu, X.H.; Ji, N.Y. Citrinovirin with a new norditerpene skeleton from the marine algicolous fungus Trichoderma citrinoviride. Bioorg. Med. Chem. Lett. 2016, 26, 5029–5031. [Google Scholar] [CrossRef] [PubMed]
  135. Liu, S.; Dai, H.; Konuklugil, B.; Orfali, R.S.; Lin, W.; Kalscheuer, R.; Liu, Z.; Proksch, P. Phenolic bisabolanes from the sponge-derived fungus Aspergillus sp. Phytochem. Lett. 2016, 18, 187–191. [Google Scholar] [CrossRef]
  136. Mohamed-Benkada, M.; Francois, P.Y.; Verite, P.; Pagniez, F.; Caroff, N.; Ruiz, N. Identification and Biological Activities of Long-Chain Peptaibols Produced by a Marine-Derived Strain of Trichoderma longibrachiatum. Chem. Biodivers. 2016, 13, 521–530. [Google Scholar] [CrossRef]
  137. Sun, Y.; Feng, Z.; Tomura, T.; Suzuki, A.; Miyano, S.; Tsuge, T.; Mori, H.; Suh, J.W.; Iizuka, T.; Fudou, R.; et al. Heterologous Production of the Marine Myxobacterial Antibiotic Haliangicin and Its Unnatural Analogues Generated by Engineering of the Biochemical Pathway. Sci. Rep. 2016, 6, 22091. [Google Scholar] [CrossRef] [Green Version]
  138. Ye, X.; Anjum, K.; Song, T.; Wang, W.; Yu, S.; Huang, H.; Lian, X.Y.; Zhang, Z. A new curvularin glycoside and its cytotoxic and antibacterial analogues from marine actinomycete Pseudonocardia sp. HS7. Nat. Prod. Res. 2016, 30, 1156–1161. [Google Scholar] [CrossRef]
  139. Zhang, Y.; Adnani, N.; Braun, D.R.; Ellis, G.A.; Barns, K.J.; Parker-Nance, S.; Guzei, I.A.; Bugni, T.S. Micromonohalimanes A and B: Antibacterial Halimane-Type Diterpenoids from a Marine Micromonospora Species. J. Nat. Prod. 2016, 79, 2968–2972. [Google Scholar] [CrossRef] [Green Version]
  140. Zhou, S.L.; Wang, M.; Zhao, H.G.; Huang, Y.H.; Lin, Y.Y.; Tan, G.H.; Chen, S.L. Penicilazaphilone C, a new antineoplastic and antibacterial azaphilone from the Marine Fungus Penicillium sclerotiorum. Arch. Pharm. Res. 2016, 39, 1621–1627. [Google Scholar] [CrossRef]
  141. He, K.Y.; Zhang, C.; Duan, Y.R.; Huang, G.L.; Yang, C.Y.; Lu, X.R.; Zheng, C.J.; Chen, G.Y. New chlorinated xanthone and anthraquinone produced by a mangrove-derived fungus Penicillium citrinum HL-5126. J. Antibiot. 2017, 70, 823–827. [Google Scholar] [CrossRef]
  142. Kim, J.; Shin, D.; Kim, S.H.; Park, W.; Shin, Y.; Kim, W.K.; Lee, S.K.; Oh, K.B.; Shin, J.; Oh, D.C. Borrelidins C-E: New Antibacterial Macrolides from a Saltern-Derived Halophilic Nocardiopsis sp. Mar. Drugs 2017, 15, 166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Li, J.L.; Chen, D.; Huang, L.; Ni, M.; Zhao, Y.; Fan, H.; Bao, X. Antichlamydial Dimeric Indole Derivatives from Marine Actinomycete Rubrobacter radiotolerans. Planta Med. 2017, 83, 805–811. [Google Scholar] [CrossRef] [PubMed]
  144. Zin, W.W.M.; Buttachon, S.; Dethoup, T.; Pereira, J.A.; Gales, L.; Inacio, A.; Costa, P.M.; Lee, M.; Sekeroglu, N.; Silva, A.M.S.; et al. Antibacterial and antibiofilm activities of the metabolites isolated from the culture of the mangrove-derived endophytic fungus Eurotium chevalieri KUFA 0006. Phytochemistry 2017, 141, 86–97. [Google Scholar] [CrossRef]
  145. Niu, S.; Liu, D.; Shao, Z.; Proksch, P.; Lin, W. Eutypellazines N-S, new thiodiketopiperazines from a deep sea sediment derived fungus Eutypella sp. with anti-VRE activities. Tetrahedron Lett. 2017, 58, 3695–3699. [Google Scholar] [CrossRef]
  146. Teta, R.; Marteinsson, V.T.; Longeon, A.; Klonowski, A.M.; Groben, R.; Bourguet-Kondracki, M.L.; Costantino, V.; Mangoni, A. Thermoactinoamide A, an Antibiotic Lipophilic Cyclopeptide from the Icelandic Thermophilic Bacterium Thermoactinomyces vulgaris. J. Nat. Prod. 2017, 80, 2530–2535. [Google Scholar] [CrossRef]
  147. Wang, Q.; Zhang, Y.; Wang, M.; Tan, Y.; Hu, X.; He, H.; Xiao, C.; You, X.; Wang, Y.; Gan, M. Neo-actinomycins A and B, natural actinomycins bearing the 5H-oxazolo[4,5-b]phenoxazine chromophore, from the marine-derived Streptomyces sp. IMB094. Sci. Rep. 2017, 7, 3591. [Google Scholar] [CrossRef] [Green Version]
  148. Jamison, M.T.; Macho, J.; Molinski, T.F. Structure-activity of antifungal compounds inspired by aminobisabolenes from the sponge Halichondria sp. Bioorg. Med. Chem. Lett. 2016, 26, 5244–5246. [Google Scholar] [CrossRef]
  149. Zhu, Y.; Wang, Y.; Gu, B.; Yang, F.; Jiao, W.H.; Hu, G.; Yu, H.B.; Han, B.N.; Zhang, W.; Shen, Y.; et al. Antifungal bromopyrrole alkaloids from the South China Sea sponge Agelas sp. Tetrahedron 2016, 72, 2964–2971. [Google Scholar] [CrossRef]
  150. Bertin, M.J.; Demirkiran, O.; Navarro, G.; Moss, N.A.; Lee, J.; Goldgof, G.M.; Vigil, E.; Winzeler, E.A.; Valeriote, F.A.; Gerwick, W.H. Kalkipyrone B, a marine cyanobacterial gamma-pyrone possessing cytotoxic and anti-fungal activities. Phytochemistry 2016, 122, 113–118. [Google Scholar] [CrossRef] [Green Version]
  151. Yamazaki, H.; Kanno, S.I.; Abdjul, D.B.; Namikoshi, M. A bromopyrrole-containing diterpene alkaloid from the Okinawan marine sponge Agelas nakamurai activates the insulin pathway in Huh-7 human hepatoma cells by inhibiting protein tyrosine phosphatase 1B. Bioorg. Med. Chem. Lett. 2017, 27, 2207–2209. [Google Scholar] [CrossRef]
  152. Xu, F.; Wang, F.; Wang, Z.; Lv, W.; Wang, W.; Wang, Y. Glucose Uptake Activities of Bis (2, 3-Dibromo-4, 5-Dihydroxybenzyl) Ether, a Novel Marine Natural Product from Red Alga Odonthaliacorymbifera with Protein Tyrosine Phosphatase 1B Inhibition, In Vitro and In Vivo. PLoS ONE 2016, 11, e0147748. [Google Scholar] [CrossRef] [PubMed]
  153. Kim, E.A.; Lee, S.H.; Lee, J.H.; Kang, N.; Oh, J.Y.; Seun-heui, S.; Ahn, G.; Ko, S.C.; Fernando, S.P.; Kim, S.Y.; et al. A marine algal polyphenol, dieckol, attenuates blood glucose levels by Akt pathway in alloxan induced hyperglycemia zebrafish model. RSC Adv. 2016, 6, 78570–78575. [Google Scholar] [CrossRef]
  154. Woo, J.K.; Ha, T.K.Q.; Oh, D.C.; Oh, W.K.; Oh, K.B.; Shin, J. Polyoxygenated Steroids from the Sponge Clathria gombawuiensis. J. Nat. Prod. 2017, 80, 3224–3233. [Google Scholar] [CrossRef]
  155. Villa-Perez, P.; Cueto, M.; Diaz-Marrero, A.R.; Lobaton, C.D.; Moreno, A.; Perdomo, G.; Cozar-Castellano, I. Leptolide Improves Insulin Resistance in Diet-Induced Obese Mice. Mar. Drugs 2017, 15, 289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Cui, H.; Liu, Y.; Nie, Y.; Liu, Z.; Chen, S.; Zhang, Z.; Lu, Y.; He, L.; Huang, X.; She, Z. Polyketides from the Mangrove-Derived Endophytic Fungus Nectria sp. HN001 and Their alpha-Glucosidase Inhibitory Activity. Mar. Drugs 2016, 14, 86. [Google Scholar] [CrossRef] [Green Version]
  157. Lin, X.; Wu, Q.; Yu, Y.; Liang, Z.; Liu, Y.; Zhou, L.; Tang, L.; Zhou, X. Penicilliumin B, a novel sesquiterpene methylcyclopentenedione from a deep sea-derived Penicillium strain with renoprotective activities. Sci. Rep. 2017, 7, 10757. [Google Scholar] [CrossRef] [Green Version]
  158. Chen, Z.; Hao, J.; Wang, L.; Wang, Y.; Kong, F.; Zhu, W. New alpha-glucosidase inhibitors from marine algae-derived Streptomyces sp. OUCMDZ-3434. Sci. Rep. 2016, 6, 20004. [Google Scholar] [CrossRef]
  159. Wiese, J.; Aldemir, H.; Schmaljohann, R.; Gulder, T.A.M.; Imhoff, J.F. Asperentin B, a New Inhibitor of the Protein Tyrosine Phosphatase 1B. Mar. Drugs 2017, 15, 191. [Google Scholar] [CrossRef]
  160. Chen, S.; Liu, Z.; Liu, H.; Long, Y.; Chen, D.; Lu, Y.; She, Z. Lasiodiplactone A, a novel lactone from the mangrove endophytic fungus Lasiodiplodia theobromae ZJ-HQ1. Org. Biomol. Chem. 2017, 15, 6338–6341. [Google Scholar] [CrossRef]
  161. Zhang, L.; Niaz, S.I.; Khan, D.; Wang, Z.; Zhu, Y.; Zhou, H.; Lin, Y.; Li, J.; Liu, L. Induction of Diverse Bioactive Secondary Metabolites from the Mangrove Endophytic Fungus Trichoderma sp. (Strain 307) by Co-Cultivation with Acinetobacter johnsonii (Strain B2). Mar. Drugs 2017, 15, 35. [Google Scholar] [CrossRef] [Green Version]
  162. Chen, S.; Liu, Y.; Liu, Z.; Cai, R.; Lu, Y.; Huang, X.; She, Z. Isocoumarins and benzofurans from the mangrove endophytic fungus Talaromyces amestolkiae possess a-glucosidase inhibitory and antibacterial activities. RSC Adv. 2016, 6, 26412–26420. [Google Scholar] [CrossRef]
  163. Zbakh, H.; Talero, E.; Avila, J.; Alcaide, A.; de Los, R.C.; Zubia, E.; Motilva, V. The Algal Meroterpene 11-Hydroxy-1′-O-Methylamentadione Ameloriates Dextran Sulfate Sodium-Induced Colitis in Mice. Mar. Drugs 2016, 14, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Jung, B.; Ku, S.K.; Gao, M.; Kim, K.M.; Han, M.S.; Choi, H.; Bae, J.S. Suppressive effects of three diketopiperazines from marine-derived bacteria on TGFBIp-mediated septic responses in human endothelial cells and mice. Arch. Pharm. Res. 2016, 39, 843–854. [Google Scholar] [CrossRef] [PubMed]
  165. Kang, H.; Ku, S.K.; Choi, H.; Bae, J.S. Three diketopiperazines from marine-derived bacteria inhibit LPS-induced endothelial inflammatory responses. Bioorg. Med. Chem. Lett. 2016, 26, 1873–1876. [Google Scholar] [CrossRef]
  166. Ahmad, T.B.; Rudd, D.; Benkendorff, K.; Mahdi, L.K.; Pratt, K.A.; Dooley, L.; Wei, C.; Kotiw, M. Brominated indoles from a marine mollusc inhibit inflammation in a murine model of acute lung injury. PLoS ONE 2017, 12, e0186904. [Google Scholar] [CrossRef] [Green Version]
  167. El-Desoky, A.H.; Kato, H.; Angkouw, E.D.; Mangindaan, R.E.; de Voogd, N.J.; Tsukamoto, S. Ceylonamides A-F, Nitrogenous Spongian Diterpenes That Inhibit RANKL-Induced Osteoclastogenesis, from the Marine Sponge Spongia ceylonensis. J. Nat. Prod. 2016, 79, 1922–1928. [Google Scholar] [CrossRef]
  168. Ngan, N.T.; Quang, T.H.; Kim, K.W.; Kim, H.J.; Sohn, J.H.; Kang, D.G.; Lee, H.S.; Kim, Y.C.; Oh, H. Anti-inflammatory effects of secondary metabolites isolated from the marine-derived fungal strain Penicillium sp. SF-5629. Arch. Pharm. Res. 2017, 40, 328–337. [Google Scholar] [CrossRef]
  169. Manzoor, Z.; Koo, J.E.; Ali, I.; Kim, J.E.; Byeon, S.H.; Yoo, E.S.; Kang, H.K.; Hyun, J.W.; Lee, N.H.; Koh, Y.S. 4-Hydroxy-2,3-Dimethyl-2-Nonen-4-Olide Has an Inhibitory Effect on Pro-Inflammatory Cytokine Production in CpG-Stimulated Bone Marrow-Derived Dendritic Cells. Mar. Drugs 2016, 14, 88. [Google Scholar] [CrossRef] [Green Version]
  170. Aminin, D.; Pislyagin, E.; Astashev, M.; Es’kov, A.; Kozhemyako, V.; Avilov, S.; Zelepuga, E.; Yurchenko, E.; Kaluzhskiy, L.; Kozlovskaya, E.; et al. Glycosides from edible sea cucumbers stimulate macrophages via purinergic receptors. Sci. Rep. 2016, 6, 39683. [Google Scholar] [CrossRef] [Green Version]
  171. Ha, T.M.; Ko, W.; Lee, S.J.; Kim, Y.C.; Son, J.Y.; Sohn, J.H.; Yim, J.H.; Oh, H. Anti-Inflammatory Effects of Curvularin-Type Metabolites from a Marine-Derived Fungal Strain Penicillium sp. SF-5859 in Lipopolysaccharide-Induced RAW264.7 Macrophages. Mar. Drugs 2017, 15, 282. [Google Scholar] [CrossRef] [Green Version]
  172. Ciaglia, E.; Malfitano, A.M.; Laezza, C.; Fontana, A.; Nuzzo, G.; Cutignano, A.; Abate, M.; Pelin, M.; Sosa, S.; Bifulco, M.; et al. Immuno-Modulatory and Anti-Inflammatory Effects of Dihydrogracilin A, a Terpene Derived from the Marine Sponge Dendrilla membranosa. Int. J. Mol. Sci. 2017, 18, 1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Jiao, W.H.; Cheng, B.H.; Shi, G.H.; Chen, G.D.; Gu, B.B.; Zhou, Y.J.; Hong, L.L.; Yang, F.; Liu, Z.Q.; Qiu, S.Q.; et al. Dysivillosins A-D, Unusual Anti-allergic Meroterpenoids from the Marine Sponge Dysidea villosa. Sci. Rep. 2017, 7, 8947. [Google Scholar] [CrossRef] [PubMed]
  174. Su, B.C.; Chen, J.Y. Antimicrobial Peptide Epinecidin-1 Modulates MyD88 Protein Levels via the Proteasome Degradation Pathway. Mar. Drugs 2017, 15, 362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Lin, Y.Y.; Jean, Y.H.; Lee, H.P.; Lin, S.C.; Pan, C.Y.; Chen, W.F.; Wu, S.F.; Su, J.H.; Tsui, K.H.; Sheu, J.H.; et al. Excavatolide B Attenuates Rheumatoid Arthritis through the Inhibition of Osteoclastogenesis. Mar. Drugs 2017, 15, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Choi, J.H.; Kim, N.H.; Kim, S.J.; Lee, H.J.; Kim, S. Fucoxanthin Inhibits the Inflammation Response in Paw Edema Model through Suppressing MAPKs, Akt, and NFkappaB. J. Biochem. Mol. Toxicol. 2016, 30, 111–119. [Google Scholar] [CrossRef] [PubMed]
  177. Grasa-Lopez, A.; Miliar-Garcia, A.; Quevedo-Corona, L.; Paniagua-Castro, N.; Escalona-Cardoso, G.; Reyes-Maldonado, E.; Jaramillo-Flores, M.E. Undaria pinnatifida and Fucoxanthin Ameliorate Lipogenesis and Markers of Both Inflammation and Cardiovascular Dysfunction in an Animal Model of Diet-Induced Obesity. Mar. Drugs 2016, 14, 148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Sugiura, Y.; Kinoshita, Y.; USUI, M.; Tanaka, R.; Matsushita, T.; Masaaki, M. The Suppressive Effect of a Marine Carotenoid, Fucoxanthin, on Mouse Ear Swelling through Regulation of Activities and mRNA Expression of Inflammation-associated Enzymes. Food Sci. Technol. Res. 2016, 22, 227–234. [Google Scholar] [CrossRef] [Green Version]
  179. Taira, J.; Sonamoto, M.; Uehara, M. Dual Biological Functions of a Cytoprotective Effect and Apoptosis Induction by Bioavailable Marine Carotenoid Fucoxanthinol through Modulation of the Nrf2 Activation in RAW264.7 Macrophage Cells. Mar. Drugs 2017, 15, 305. [Google Scholar] [CrossRef] [Green Version]
  180. Sinstova, O.V.; Pislygin, E.A.; Gladkikh, I.N.; Monastyrnaya, M.M.; Menchinskaya, E.S.; Lynchenko, E.V.; Aminin, D.L.; Kozlovskaya, E.P. Kunitz-type peptides of the sea anemone Heteractis crispa: Potential anti-inflammatory compounds. Russ. J. Bioorg. Chem. 2017, 43, 91–97. [Google Scholar]
  181. Hong, L.L.; Yu, H.B.; Wang, J.; Jiao, W.H.; Cheng, B.H.; Yang, F.; Zhou, Y.J.; Gu, B.B.; Song, S.J.; Lin, H.W. Unusual Anti-allergic Diterpenoids from the Marine Sponge Hippospongia lachne. Sci. Rep. 2017, 7, 43138. [Google Scholar] [CrossRef] [Green Version]
  182. Kozuma, S.; Hirota-Takahata, Y.; Fukuda, D.; Kuraya, N.; Nakajima, M.; Ando, O. Identification and biological activity of ogipeptins, novel LPS inhibitors produced by marine bacterium. J. Antibiot. 2017, 70, 79–83. [Google Scholar] [CrossRef]
  183. Kwon, I.S.; Kwak, J.H.; Pyo, S.; Lee, H.W.; Kim, A.; Schmitz, F.J. Oscarellin, an Anthranilic Acid Derivative from a Philippine Sponge, Oscarella stillans, as an Inhibitor of Inflammatory Cytokines in Macrophages. J. Nat. Prod. 2017, 80, 149–155. [Google Scholar] [CrossRef] [PubMed]
  184. Kim, M.E.; Jung, I.; Lee, J.S.; Na, J.Y.; Kim, W.J.; Kim, Y.O.; Park, Y.D.; Lee, J.S. Pseudane-VII Isolated from Pseudoalteromonas sp. M2 Ameliorates LPS-Induced Inflammatory Response In Vitro and In Vivo. Mar. Drugs 2017, 15, 336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Cheng, Z.; Zhao, J.; Proksch, P.; Zhao, Z.; Lin, W. Eremophilane-Type Sesquiterpenoids from an Acremonium sp. Fungus Isolated from Deep-Sea Sediments. J. Nat. Prod. 2016, 79, 1035–1047. [Google Scholar] [CrossRef] [PubMed]
  186. Hassan, H.M.; Boonlarppradab, C.; Fenical, W. Actinoquinolines A and B, anti-inflammatory quinoline alkaloids from a marine-derived Streptomyces sp., strain CNP975. J. Antibiot. 2016, 69, 511–514. [Google Scholar] [CrossRef] [PubMed]
  187. Malyarenko, T.V.; Kharchenko, S.D.; Kicha, A.A.; Ivanchina, N.V.; Dmitrenok, P.S.; Chingizova, E.A.; Pislyagin, E.A.; Evtushenko, E.V.; Antokhina, T.I.; Minh, C.V.; et al. Anthenosides L-U, Steroidal Glycosides with Unusual Structural Features from the Starfish Anthenea aspera. J. Nat. Prod. 2016, 79, 3047–3056. [Google Scholar] [CrossRef]
  188. Fang, W.; Lin, X.; Wang, J.; Liu, Y.; Tao, H.; Zhou, X. Asperpyrone-Type Bis-Naphtho-gamma-Pyrones with COX-2-Inhibitory Activities from Marine-Derived Fungus Aspergillus niger. Molecules 2016, 21, 941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Su, Y.D.; Wen, Z.H.; Wu, Y.C.; Fang, L.S.; Chen, Y.H.; Chang, Y.C.; Sheu, J.H.; Sung, P.J. Briarenolides M-T, new briarane diterpenoids from a Formosan octocoral Briareum sp. Tetrahedron 2016, 72, 944–951. [Google Scholar] [CrossRef]
  190. Su, Y.D.; Sung, C.S.; Wen, Z.H.; Chen, Y.H.; Chang, Y.C.; Chen, J.J.; Fang, L.S.; Wu, Y.C.; Sheu, J.H.; Sung, P.J. New 9-Hydroxybriarane Diterpenoids from a Gorgonian Coral Briareum sp. (Briareidae). Int. J. Mol. Sci. 2016, 17, 79. [Google Scholar] [CrossRef] [Green Version]
  191. Wang, L.; Li, M.; Tang, J.; Li, X. Eremophilane Sesquiterpenes from a Deep Marine-Derived Fungus, Aspergillus sp. SCSIOW2, Cultivated in the Presence of Epigenetic Modifying Agents. Molecules 2016, 21, 473. [Google Scholar] [CrossRef] [Green Version]
  192. Kwon, J.; Lee, H.; Ko, W.; Kim, D.; Kim, K.; Kwon, H.C. Chemical constituents isolated from Antarctic marine-derived Aspergillus sp. SF-5976 and their anti-inflammatory effects in LPS-stimulated RAW 264.7 and BV2 cells. Tetrahedron 2017, 73, 3905–3912. [Google Scholar] [CrossRef]
  193. Hsu, Y.M.; Chang, F.R.; Lo, I.W.; Lai, K.H.; El-Shazly, M.; Wu, T.Y.; Du, Y.C.; Hwang, T.L.; Cheng, Y.B.; Wu, Y.C. Zoanthamine-Type Alkaloids from the Zoanthid Zoanthus kuroshio Collected in Taiwan and Their Effects on Inflammation. J. Nat. Prod. 2016, 79, 2674–2680. [Google Scholar] [CrossRef] [PubMed]
  194. Tseng, W.R.; Huang, C.Y.; Tsai, Y.Y.; Lin, Y.S.; Hwang, T.L.; Su, J.H.; Sung, P.J.; Dai, C.F.; Sheu, J.H. New cytotoxic and anti-inflammatory steroids from the soft coral Klyxum flaccidum. Bioorg. Med. Chem. Lett. 2016, 26, 3253–3257. [Google Scholar] [CrossRef] [PubMed]
  195. Tsai, Y.Y.; Huang, C.Y.; Tseng, W.R.; Chiang, P.L.; Hwang, T.L.; Su, J.H.; Sung, P.J.; Dai, C.F.; Sheu, J.H. Klyflaccisteroids K-M, bioactive steroidal derivatives from a soft coral Klyxum flaccidum. Bioorg. Med. Chem. Lett. 2017, 27, 1220–1224. [Google Scholar] [CrossRef]
  196. Ahmed, A.F.; Teng, W.T.; Huang, C.Y.; Dai, C.F.; Hwang, T.L.; Sheu, J.H. Anti-Inflammatory Lobane and Prenyleudesmane Diterpenoids from the Soft Coral Lobophytum varium. Mar. Drugs 2017, 15, 300. [Google Scholar] [CrossRef] [Green Version]
  197. Huang, C.Y.; Chang, C.W.; Tseng, Y.J.; Lee, J.; Sung, P.J.; Su, J.H.; Hwang, T.L.; Dai, C.F.; Wang, H.C.; Sheu, J.H. Bioactive Steroids from the Formosan Soft Coral Umbellulifera petasites. Mar. Drugs 2016, 14, 180. [Google Scholar] [CrossRef] [Green Version]
  198. Su, Y.D.; Cheng, C.H.; Wen, Z.H.; Wu, Y.C.; Sung, P.J. New anti-inflammatory sterols from a gorgonian Pinnigorgia sp. Bioorg. Med. Chem. Lett. 2016, 26, 3060–3063. [Google Scholar] [CrossRef]
  199. Chang, Y.C.; Kuo, L.M.; Su, J.H.; Hwang, T.L.; Kuo, Y.H.; Lin, C.S.; Wu, Y.C.; Sheu, J.H.; Sung, P.J. Pinnigorgiols A-C, 9,11-secosterols with a rare ring arrangement from a gorgonian coral Pinnigorgia sp. Tetrahedron 2016, 72, 999–1004. [Google Scholar] [CrossRef]
  200. Chang, Y.C.; Hwang, T.L.; Sheu, J.H.; Wu, Y.C.; Sung, P.J. New Anti-Inflammatory 9,11-Secosterols with a Rare Tricyclo[5,2,1,1]decane Ring from a Formosan Gorgonian Pinnigorgia sp. Mar. Drugs 2016, 14, 218. [Google Scholar] [CrossRef] [Green Version]
  201. Chang, Y.C.; Kuo, L.M.; Hwang, T.L.; Yeh, J.; Wen, Z.H.; Fang, L.S.; Wu, Y.C.; Lin, C.S.; Sheu, J.H.; Sung, P.J. Pinnisterols A-C, New 9,11-Secosterols from a Gorgonian Pinnigorgia sp. Mar. Drugs 2016, 14, 12. [Google Scholar] [CrossRef] [Green Version]
  202. Chang, Y.C.; Hwang, T.L.; Kuo, L.M.; Sung, P.J. Pinnisterols D-J, New 11-Acetoxy-9,11-secosterols with a 1,4-Quinone Moiety from Formosan Gorgonian Coral Pinnigorgia sp. (Gorgoniidae). Mar. Drugs 2017, 15, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Vien, L.T.; Hanh, T.T.; Huong, P.T.; Dang, N.H.; Thanh, N.V.; Lyakhova, E.; Cuong, N.X.; Nam, N.H.; Kiem, P.V.; Kicha, A.; et al. Pyrrole Oligoglycosides from the Starfish Acanthaster planci Suppress Lipopolysaccharide-Induced Nitric Oxide Production in RAW264.7 Macrophages. Chem. Pharm. Bull. 2016, 64, 1654–1657. [Google Scholar] [CrossRef] [Green Version]
  204. Malyarenko, T.V.; Kicha, A.A.; Kalinovsky, A.I.; Ivanchina, N.V.; Popov, R.S.; Pislyagin, E.A.; Menchinskaya, E.S.; Padmakumar, K.P.; Stonik, V.A. Four New Steroidal Glycosides, Protolinckiosides A-D, from the Starfish Protoreaster lincki. Chem. Biodivers. 2016, 13, 998–1007. [Google Scholar] [CrossRef]
  205. Zhao, M.; Cheng, S.; Yuan, W.; Xi, Y.; Li, X.; Dong, J.; Huang, K.; Gustafson, K.R.; Yan, P. Cembranoids from a Chinese Collection of the Soft Coral Lobophytum crassum. Mar. Drugs 2016, 14, 111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Huang, C.Y.; Tseng, Y.J.; Chokkalingam, U.; Hwang, T.L.; Hsu, C.H.; Dai, C.F.; Sung, P.J.; Sheu, J.H. Bioactive Isoprenoid-Derived Natural Products from a Dongsha Atoll Soft Coral Sinularia erecta. J. Nat. Prod. 2016, 79, 1339–1346. [Google Scholar] [CrossRef] [PubMed]
  207. Huang, C.Y.; Ahmed, A.F.; Su, J.H.; Sung, P.J.; Hwang, T.L.; Chiang, P.L.; Dai, C.F.; Liaw, C.C.; Sheu, J.H. Bioactive new withanolides from the cultured soft coral Sinularia brassica. Bioorg. Med. Chem. Lett. 2017, 27, 3267–3271. [Google Scholar] [CrossRef]
  208. Huang, C.Y.; Su, J.H.; Liaw, C.C.; Sung, P.J.; Chiang, P.L.; Hwang, T.L.; Dai, C.F.; Sheu, J.H. Bioactive Steroids with Methyl Ester Group in the Side Chain from a Reef Soft Coral Sinularia brassica Cultured in a Tank. Mar. Drugs 2017, 15, 280. [Google Scholar] [CrossRef] [Green Version]
  209. Torres-Mendoza, D.; Gonzalez, Y.; Gomez-Reyes, J.F.; Guzman, H.M.; Lopez-Perez, J.L.; Gerwick, W.H.; Fernandez, P.L.; Gutierrez, M. Uprolides N, O and P from the Panamanian Octocoral Eunicea succinea. Molecules 2016, 21, 819. [Google Scholar] [CrossRef] [Green Version]
  210. Pislyagin, E.A.; Manzhulo, I.V.; Dmitrenok, P.S.; Aminin, D.L. Cucumarioside A2-2 causes changes in the morphology and proliferative activity in mouse spleen. Acta Histochem. 2016, 118, 387–392. [Google Scholar] [CrossRef]
  211. Pislyagin, E.A.; Manzhulo, I.V.; Gorpenchenko, T.Y.; Dmitrenok, P.S.; Avilov, S.A.; Silchenko, A.S.; Wang, Y.M.; Aminin, D.L. Cucumarioside A(2)-2 Causes Macrophage Activation in Mouse Spleen. Mar. Drugs 2017, 15, 341. [Google Scholar] [CrossRef] [Green Version]
  212. Sanchez, J.A.; Alfonso, A.; Rodriguez, I.; Alonso, E.; Cifuentes, J.M.; Bermudez, R.; Rateb, M.E.; Jaspars, M.; Houssen, W.E.; Ebel, R.; et al. Spongionella Secondary Metabolites, Promising Modulators of Immune Response through CD147 Receptor Modulation. Front. Immunol. 2016, 7, 452. [Google Scholar] [CrossRef] [Green Version]
  213. Sanchez, J.A.; Alfonso, A.; Leiros, M.; Alonso, E.; Rateb, M.E.; Jaspars, M.; Houssen, W.E.; Ebel, R.; Tabudravu, J.; Botana, L.M. Identification of Spongionella compounds as cyclosporine A mimics. Pharmacol. Res. 2016, 107, 407–414. [Google Scholar] [CrossRef]
  214. Becker, K.; Hartmann, A.; Ganzera, M.; Fuchs, D.; Gostner, J.M. Immunomodulatory Effects of the Mycosporine-Like Amino Acids Shinorine and Porphyra-334. Mar. Drugs 2016, 14, 119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Chung, T.W.; Li, Y.R.; Huang, W.Y.; Su, J.H.; Chan, H.L.; Lin, S.H.; Liu, C.S.; Lin, S.C.; Lin, C.C.; Lin, C.H. Sinulariolide suppresses LPSinduced phenotypic and functional maturation of dendritic cells. Mol. Med. Rep. 2017, 16, 6992–7000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Gao, Y.Y.; Liu, Q.M.; Liu, B.; Xie, C.L.; Cao, M.J.; Yang, X.W.; Liu, G.M. Inhibitory Activities of Compounds from the Marine Actinomycete Williamsia sp. MCCC 1A11233 Variant on IgE-Mediated Mast Cells and Passive Cutaneous Anaphylaxis. J. Agric. Food Chem. 2017, 65, 10749–10756. [Google Scholar] [CrossRef] [PubMed]
  217. Chen, S.; Wang, J.; Lin, X.; Zhao, B.; Wei, X.; Li, G.; Kaliaperumal, K.; Liao, S.; Yang, B.; Zhou, X.; et al. Chrysamides A-C, Three Dimeric Nitrophenyl trans-Epoxyamides Produced by the Deep-Sea-Derived Fungus Penicillium chrysogenum SCSIO41001. Org. Lett. 2016, 18, 3650–3653. [Google Scholar] [CrossRef] [PubMed]
  218. Gunasekera, S.P.; Li, Y.; Ratnayake, R.; Luo, D.; Lo, J.; Reibenspies, J.H.; Xu, Z.; Clare-Salzler, M.J.; Ye, T.; Paul, V.J.; et al. Discovery, Total Synthesis and Key Structural Elements for the Immunosuppressive Activity of Cocosolide, a Symmetrical Glycosylated Macrolide Dimer from Marine Cyanobacteria. Chemistry 2016, 22, 8158–8166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Einarsdottir, E.; Liu, H.B.; Freysdottir, J.; Gotfredsen, C.H.; Omarsdottir, S. Immunomodulatory N-acyl Dopamine Glycosides from the Icelandic Marine Sponge Myxilla incrustans Collected at a Hydrothermal Vent Site. Planta Med. 2016, 82, 903–909. [Google Scholar] [CrossRef] [Green Version]
  220. Kawase, O.; Ohno, O.; Suenaga, K.; Xuan, X. Immunological Adjuvant Activity of Pectinioside A, the Steroidal Saponin from the Starfish Patiria pectinifera. Nat. Prod. Commun. 2016, 11, 605–606. [Google Scholar] [CrossRef]
  221. Liu, H.; Chen, S.; Liu, W.; Liu, Y.; Huang, X.; She, Z. Polyketides with Immunosuppressive Activities from Mangrove Endophytic Fungus Penicillium sp. ZJ-SY(2). Mar. Drugs 2016, 14, 217. [Google Scholar] [CrossRef] [Green Version]
  222. Li, H.; Huang, H.; Hou, L.; Ju, J.; Li, W. Discovery of Antimycin-Type Depsipeptides from a wbl Gene Mutant Strain of Deepsea-Derived Streptomyces somaliensis SCSIO ZH66 and Their Effects on Pro-inflammatory Cytokine Production. Front. Microbiol. 2017, 8, 678. [Google Scholar] [CrossRef]
  223. Moreels, L.; Peigneur, S.; Galan, D.T.; De, P.E.; Beress, L.; Waelkens, E.; Pardo, L.A.; Quinton, L.; Tytgat, J. APETx4, a Novel Sea Anemone Toxin and a Modulator of the Cancer-Relevant Potassium Channel KV10.1. Mar. Drugs 2017, 15, 287. [Google Scholar] [CrossRef]
  224. Goda, A.A.; Naguib, K.M.; Mohamed, M.M.; Amra, H.A.; Nada, S.A.; Abdel-Ghaffar, A.B.; Gissendanner, C.R.; El Sayed, K.A. Astaxanthin and Docosahexaenoic Acid Reverse the Toxicity of the Maxi-K (BK) Channel Antagonist Mycotoxin Penitrem A. Mar. Drugs 2016, 14, 208. [Google Scholar] [CrossRef] [Green Version]
  225. Mendez, A.G.; Juncal, A.B.; Silva, S.B.L.; Thomas, O.P.; Martin, V.V.; Alfonso, A.; Vieytes, M.R.; Vale, C.; Botana, L.M. The Marine Guanidine Alkaloid Crambescidin 816 Induces Calcium Influx and Cytotoxicity in Primary Cultures of Cortical Neurons through Glutamate Receptors. ACS Chem. Neurosci. 2017, 8, 1609–1617. [Google Scholar] [CrossRef] [Green Version]
  226. Jiang, S.; Tae, H.S.; Xu, S.; Shao, X.; Adams, D.J.; Wang, C. Identification of a Novel O-Conotoxin Reveals an Unusual and Potent Inhibitor of the Human alpha9alpha10 Nicotinic Acetylcholine Receptor. Mar. Drugs 2017, 15, 170. [Google Scholar] [CrossRef] [Green Version]
  227. Bernaldez, J.; Jimenez, S.; Gonzalez, L.J.; Ferro, J.N.; Soto, E.; Salceda, E.; Chavez, D.; Aguilar, M.B.; Licea-Navarro, A. A New Member of Gamma-Conotoxin Family Isolated from Conus princeps Displays a Novel Molecular Target. Toxins 2016, 8, 39. [Google Scholar] [CrossRef] [Green Version]
  228. Brust, A.; Croker, D.E.; Colless, B.; Ragnarsson, L.; Andersson, A.; Jain, K.; Garcia-Caraballo, S.; Castro, J.; Brierley, S.M.; Alewood, P.F.; et al. Conopeptide-Derived kappa-Opioid Agonists (Conorphins): Potent, Selective, and Metabolic Stable Dynorphin A Mimetics with Antinociceptive Properties. J. Med. Chem. 2016, 59, 2381–2395. [Google Scholar] [CrossRef]
  229. Feng, C.W.; Hung, H.C.; Huang, S.Y.; Chen, C.H.; Chen, Y.R.; Chen, C.Y.; Yang, S.N.; Wang, H.D.; Sung, P.J.; Sheu, J.H.; et al. Neuroprotective Effect of the Marine-Derived Compound 11-Dehydrosinulariolide through DJ-1-Related Pathway in In Vitro and In Vivo Models of Parkinson’s Disease. Mar. Drugs 2016, 14, 187. [Google Scholar] [CrossRef]
  230. Chen, C.H.; Chen, N.F.; Feng, C.W.; Cheng, S.Y.; Hung, H.C.; Tsui, K.H.; Hsu, C.H.; Sung, P.J.; Chen, W.F.; Wen, Z.H. A Coral-Derived Compound Improves Functional Recovery after Spinal Cord Injury through Its Antiapoptotic and Anti-Inflammatory Effects. Mar. Drugs 2016, 14, 160. [Google Scholar] [CrossRef] [Green Version]
  231. Botic, T.; Defant, A.; Zanini, P.; Zuzek, M.C.; Frangez, R.; Janussen, D.; Kersken, D.; Knez, Z.; Mancini, I.; Sepcic, K. Discorhabdin alkaloids from Antarctic Latrunculia spp. sponges as a new class of cholinesterase inhibitors. Eur. J. Med. Chem. 2017, 136, 294–304. [Google Scholar] [CrossRef]
  232. Jung, H.A.; Ali, M.Y.; Choi, R.J.; Jeong, H.O.; Chung, H.Y.; Choi, J.S. Kinetics and molecular docking studies of fucosterol and fucoxanthin, BACE1 inhibitors from brown algae Undaria pinnatifida and Ecklonia stolonifera. Food Chem. Toxicol. 2016, 89, 104–111. [Google Scholar] [CrossRef] [PubMed]
  233. Lin, J.; Huang, L.; Yu, J.; Xiang, S.; Wang, J.; Zhang, J.; Yan, X.; Cui, W.; He, S.; Wang, Q. Fucoxanthin, a Marine Carotenoid, Reverses Scopolamine-Induced Cognitive Impairments in Mice and Inhibits Acetylcholinesterase in Vitro. Mar. Drugs 2016, 14, 67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Zhang, L.; Wang, H.; Fan, Y.; Gao, Y.; Li, X.; Hu, Z.; Ding, K.; Wang, Y.; Wang, X. Fucoxanthin provides neuroprotection in models of traumatic brain injury via the Nrf2-ARE and Nrf2-autophagy pathways. Sci. Rep. 2017, 7, 46763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Fang, F.; Zhao, J.; Ding, L.; Huang, C.; Naman, C.B.; He, S.; Wu, B.; Zhu, P.; Luo, Q.; Gerwick, W.H.; et al. 5-Hydroxycyclopenicillone, a New beta-Amyloid Fibrillization Inhibitor from a Sponge-Derived Fungus Trichoderma sp. HPQJ-34. Mar. Drugs 2017, 15, 260. [Google Scholar] [CrossRef] [Green Version]
  236. Flores, P.L.; Rodriguez, E.; Zapata, E.; Carbo, R.; Farias, J.M.; Martinez, M. Maitotoxin Is a Potential Selective Activator of the Endogenous Transient Receptor Potential Canonical Type 1 Channel in Xenopus laevis Oocytes. Mar. Drugs 2017, 15, 198. [Google Scholar] [CrossRef] [Green Version]
  237. Alonso, E.; Alvarino, R.; Leiros, M.; Tabudravu, J.N.; Feussner, K.; Dam, M.A.; Rateb, M.E.; Jaspars, M.; Botana, L.M. Evaluation of the Antioxidant Activity of the Marine Pyrroloiminoquinone Makaluvamines. Mar. Drugs 2016, 14, 197. [Google Scholar] [CrossRef] [Green Version]
  238. Abdelhameed, R.; Elgawish, M.; Mira, A.; Ibrahim, A.; Ahmed, S.; Shimizu, K.; Yamada, K. Anti-choline esterase activity of ceramides from the Red Sea marine sponge Mycale euplectellioides. RSC Adv. 2016, 6, 20422–20430. [Google Scholar] [CrossRef]
  239. Uchimasu, H.; Matsumura, M.; Tsuda, M.; Kumagai, K.; Akakabe, M.; Fujieda, M.J.; Sakai, R. Mellpaladines and dopargimine, novel neuroactive guanidine alkaloids from a Palauan Didemnidae tunicate. Tetrahedron 2016, 72, 7185–7193. [Google Scholar] [CrossRef]
  240. Logashina, Y.A.; Mosharova, I.V.; Korolkova, Y.V.; Shelukhina, I.V.; Dyachenko, I.A.; Palikov, V.A.; Palikova, Y.A.; Murashev, A.N.; Kozlov, S.A.; Stensvag, K.; et al. Peptide from Sea Anemone Metridium senile Affects Transient Receptor Potential Ankyrin-repeat 1 (TRPA1) Function and Produces Analgesic Effect. J. Biol. Chem. 2017, 292, 2992–3004. [Google Scholar] [CrossRef] [Green Version]
  241. Kim, J.J.; Kang, Y.J.; Shin, S.A.; Bak, D.H.; Lee, J.W.; Lee, K.B.; Yoo, Y.C.; Kim, D.K.; Lee, B.H.; Kim, D.W.; et al. Phlorofucofuroeckol Improves Glutamate-Induced Neurotoxicity through Modulation of Oxidative Stress-Mediated Mitochondrial Dysfunction in PC12 Cells. PLoS ONE 2016, 11, e0163433. [Google Scholar] [CrossRef]
  242. Lee, H.W.; Choi, H.; Nam, S.J.; Fenical, W.; Kim, H. Potent Inhibition of Monoamine Oxidase B by a Piloquinone from Marine-Derived Streptomyces sp. CNQ-027. J. Microbiol. Biotechnol. 2017, 27, 785–790. [Google Scholar] [CrossRef] [Green Version]
  243. Caplan, S.L.; Zheng, B.; Dawson-Scully, K.; White, C.A.; West, L.M. Pseudopterosin A: Protection of Synaptic Function and Potential as a Neuromodulatory Agent. Mar. Drugs 2016, 14, 55. [Google Scholar] [CrossRef] [Green Version]
  244. Perni, M.; Galvagnion, C.; Maltsev, A.; Meisl, G.; Muller, M.B.; Challa, P.K.; Kirkegaard, J.B.; Flagmeier, P.; Cohen, S.I.; Cascella, R.; et al. A natural product inhibits the initiation of alpha-synuclein aggregation and suppresses its toxicity. Proc. Natl. Acad. Sci. USA 2017, 114, E1009–E1017. [Google Scholar] [CrossRef] [Green Version]
  245. Moodie, L.W.; Zuzek, M.C.; Frangez, R.; Andersen, J.H.; Hansen, E.; Olsen, E.K.; Cergolj, M.; Sepcic, K.; Hansen, K.O.; Svenson, J. Synthetic analogs of stryphnusin isolated from the marine sponge Stryphnus fortis inhibit acetylcholinesterase with no effect on muscle function or neuromuscular transmission. Org. Biomol. Chem. 2016, 14, 11220–11229. [Google Scholar] [CrossRef] [Green Version]
  246. Pan, N.; Lu, L.Y.; Li, M.; Wang, G.H.; Sun, F.Y.; Sun, H.S.; Wen, X.J.; Cheng, J.D.; Chen, J.W.; Pang, J.Y.; et al. Xyloketal B alleviates cerebral infarction and neurologic deficits in a mouse stroke model by suppressing the ROS/TLR4/NF-kappaB inflammatory signaling pathway. Acta Pharmacol. Sin. 2017, 38, 1236–1247. [Google Scholar] [CrossRef]
  247. Dai, J.; Parrish, S.M.; Yoshida, W.Y.; Yip, M.L.R.; Turkson, J.; Kelly, M.; Williams, P. Bromotyrosine-derived metabolites from an Indonesian marine sponge in the family Aplysinellidae (Order Verongiida). Bioorg. Med. Chem. Lett. 2016, 26, 499–504. [Google Scholar] [CrossRef] [Green Version]
  248. Naman, C.B.; Almaliti, J.; Armstrong, L.; Caro-Diaz, E.J.; Pierce, M.L.; Glukhov, E.; Fenner, A.; Spadafora, C.; Debonsi, H.M.; Dorrestein, P.C.; et al. Discovery and Synthesis of Caracolamide A, an Ion Channel Modulating Dichlorovinylidene Containing Phenethylamide from a Panamanian Marine Cyanobacterium cf. Symploca Species. J. Nat. Prod. 2017, 80, 2328–2334. [Google Scholar] [CrossRef]
  249. Campos-Lira, E.; Carrillo, E.; Aguilar, M.B.; Gajewiak, J.; Gomez-Lagunas, F.; Lopez-Vera, E. Conorfamide-Sr3, a structurally novel specific inhibitor of the Shaker K(+) channel. Toxicon 2017, 138, 53–58. [Google Scholar] [CrossRef]
  250. Han, P.; Cao, Y.; Liu, S.; Dai, X.; Yao, G.; Fan, C.; Wu, W.; Chen, J. Contryphan-Bt: A pyroglutamic acid containing conopeptide isolated from the venom of Conus betulinus. Toxicon 2017, 135, 17–23. [Google Scholar] [CrossRef]
  251. Long, Y.; Cui, H.; Liu, X.; Xiao, Z.; WEN, S.; She, Z.; Huang, X. Acetylcholinesterase Inhibitory Meroterpenoid from a Mangrove Endophytic Fungus Aspergillus sp. 16-5c. Molecules 2017, 22, 727. [Google Scholar] [CrossRef] [Green Version]
  252. Zidar, N.; Zula, A.; Tomasic, T.; Rogers, M.; Kirby, R.W.; Tytgat, J.; Peigneur, S.; Kikelj, D.; Ilas, J.; Masic, L.P. Clathrodin, hymenidin and oroidin, and their synthetic analogues as inhibitors of the voltage-gated potassium channels. Eur. J. Med. Chem. 2017, 139, 232–241. [Google Scholar] [CrossRef] [PubMed]
  253. Boccitto, M.; Lee, N.; Sakamoto, S.; Spruce, L.A.; Handa, H.; Clardy, J.; Seeholzer, S.H.; Kalb, R.G. The Neuroprotective Marine Compound Psammaplysene A Binds the RNA-Binding Protein HNRNPK. Mar. Drugs 2017, 15, 246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Ding, B.; Wang, Z.; Huang, X.; Liu, Y.; Chen, W.; She, Z. Bioactive alpha-pyrone meroterpenoids from mangrove endophytic fungus Penicillium sp. Nat. Prod. Res. 2016, 30, 2805–2812. [Google Scholar] [CrossRef] [PubMed]
  255. Omaga, C.A.; Carpio, L.D.; Imperial, J.S.; Daly, N.L.; Gajewiak, J.; Flores, M.S.; Espino, S.S.; Christensen, S.; Filchakova, O.M.; Lopez-Vera, E.; et al. Structure and Biological Activity of a Turripeptide from Unedogemmula bisaya Venom. Biochemistry 2017, 56, 6051–6060. [Google Scholar] [CrossRef] [PubMed]
  256. Yeon, J.T.; Kim, H.; Kim, K.J.; Lee, J.; Won, D.H.; Nam, S.J.; Kim, S.H.; Kang, H.; Son, Y.J. Acredinone C and the Effect of Acredinones on Osteoclastogenic and Osteoblastogenic Activity. J. Nat. Prod. 2016, 79, 1730–1736. [Google Scholar] [CrossRef] [PubMed]
  257. Yang, Y.; Bae, M.; Kim, B.; Park, Y.K.; Koo, S.I.; Lee, J.Y. Astaxanthin prevents and reverses the activation of mouse primary hepatic stellate cells. J. Nutr. Biochem. 2016, 29, 21–26. [Google Scholar] [CrossRef] [Green Version]
  258. Abdjul, D.B.; Kanno, S.I.; Yamazaki, H.; Ukai, K.; Namikoshi, M. A dimeric urea of the bisabolene sesquiterpene from the Okinawan marine sponge Axinyssa sp. inhibits protein tyrosine phosphatase 1B activity in Huh-7 human hepatoma cells. Bioorg. Med. Chem. Lett. 2016, 26, 315–317. [Google Scholar] [CrossRef]
  259. Lai, K.H.; Liu, Y.C.; Su, J.H.; El-Shazly, M.; Wu, C.F.; Du, Y.C.; Hsu, Y.M.; Yang, J.C.; Weng, M.K.; Chou, C.H.; et al. Antileukemic Scalarane Sesterterpenoids and Meroditerpenoid from Carteriospongia Phyllospongia sp., Induce Apoptosis via Dual Inhibitory Effects on Topoisomerase II and Hsp90. Sci. Rep. 2016, 6, 36170. [Google Scholar] [CrossRef]
  260. Uchida, R.; Nakajyo, K.; Kobayashi, K.; Ohshiro, T.; Terahara, T.; Imada, C.; Tomoda, H. 7-Chlorofolipastatin, an inhibitor of sterol O-acyltransferase, produced by marine-derived Aspergillus ungui NKH-007. J. Antibiot. 2016, 69, 647–651. [Google Scholar] [CrossRef]
  261. Esposito, G.; Bourguet-Kondracki, M.L.; Mai, L.H.; Longeon, A.; Teta, R.; Meijer, L.; Van, S.R.; Mangoni, A.; Costantino, V. Chloromethylhalicyclamine B, a Marine-Derived Protein Kinase CK1delta/epsilon Inhibitor. J. Nat. Prod. 2016, 79, 2953–2960. [Google Scholar] [CrossRef]
  262. Serrill, J.D.; Wan, X.; Hau, A.M.; Jang, H.S.; Coleman, D.J.; Indra, A.K.; Alani, A.W.; McPhail, K.L.; Ishmael, J.E. Coibamide A, a natural lariat depsipeptide, inhibits VEGFA/VEGFR2 expression and suppresses tumor growth in glioblastoma xenografts. Invest. New Drugs 2016, 34, 24–40. [Google Scholar] [CrossRef] [PubMed]
  263. Issac, M.; Aknin, M.; Gauvin-Bialecki, A.; De Voogd, N.; Ledoux, A.; Frederich, M.; Kashman, Y.; Carmeli, S. Cyclotheonellazoles A-C, Potent Protease Inhibitors from the Marine Sponge Theonella aff. swinhoei. J. Nat. Prod. 2017, 80, 1110–1116. [Google Scholar] [CrossRef] [PubMed]
  264. Vaden, R.M.; Oswald, N.W.; Potts, M.B.; MacMillan, J.B.; White, M.A. FUSION-Guided Hypothesis Development Leads to the Identification of N(6),N(6)-Dimethyladenosine, a Marine-Derived AKT Pathway Inhibitor. Mar. Drugs 2017, 15, 75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Hwang, H.R.; Tai, B.Y.; Cheng, P.Y.; Chen, P.N.; Sung, P.J.; Wen, Z.H.; Hsu, C.H. Excavatolide B Modulates the Electrophysiological Characteristics and Calcium Homeostasis of Atrial Myocytes. Mar. Drugs 2017, 15, 25. [Google Scholar] [CrossRef] [Green Version]
  266. Ma, S.Y.; Park, W.S.; Lee, D.S.; Choi, G.; Yim, M.J.; Lee, J.M.; Jung, W.K.; Park, S.G.; Seo, S.K.; Park, S.J.; et al. Fucoxanthin inhibits profibrotic protein expression in vitro and attenuates bleomycin-induced lung fibrosis in vivo. Eur. J. Pharmacol. 2017, 811, 199–207. [Google Scholar] [CrossRef]
  267. Lee, H.Y.; Jang, E.J.; Bae, S.Y.; Jeon, J.E.; Park, H.J.; Shin, J.; Lee, S.K. Anti-Melanogenic Activity of Gagunin D, a Highly Oxygenated Diterpenoid from the Marine Sponge Phorbas sp., via Modulating Tyrosinase Expression and Degradation. Mar. Drugs 2016, 14, 212. [Google Scholar] [CrossRef] [Green Version]
  268. Kim, K.; Leutou, A.S.; Jeong, H.; Kim, D.; Seong, C.N.; Nam, S.J.; Lim, K.M. Anti-Pigmentary Effect of (-)-4-Hydroxysattabacin from the Marine-Derived Bacterium Bacillus sp. Mar. Drugs 2017, 15, 138. [Google Scholar] [CrossRef] [Green Version]
  269. Koyama, T.; Kawazoe, Y.; Iwasaki, A.; Ohno, O.; Suenaga, K.; Uemura, D. Anti-obesity activities of the yoshinone A and the related marine gamma-pyrone compounds. J. Antibiot. 2016, 69, 348–351. [Google Scholar] [CrossRef]
  270. Carlile, G.W.; Robert, R.; Matthes, E.; Yang, Q.; Solari, R.; Hatley, R.; Edge, C.M.; Hanrahan, J.W.; Andersen, R.; Thomas, D.Y.; et al. Latonduine Analogs Restore F508del-Cystic Fibrosis Transmembrane Conductance Regulator Trafficking through the Modulation of Poly-ADP Ribose Polymerase 3 and Poly-ADP Ribose Polymerase 16 Activity. Mol. Pharmacol. 2016, 90, 65–79. [Google Scholar] [CrossRef] [Green Version]
  271. Ebrahim, H.Y.; El Sayed, K.A. Discovery of Novel Antiangiogenic Marine Natural Product Scaffolds. Mar. Drugs 2016, 14, 57. [Google Scholar] [CrossRef] [Green Version]
  272. Williams, D.E.; Izard, F.; Arnould, S.; Dalisay, D.S.; Tantapakul, C.; Maneerat, W.; Matainaho, T.; Julien, E.; Andersen, R.J. Structures of Nahuoic Acids B-E Produced in Culture by a Streptomyces sp. Isolated from a Marine Sediment and Evidence for the Inhibition of the Histone Methyl Transferase SETD8 in Human Cancer Cells by Nahuoic Acid A. J. Org. Chem. 2016, 81, 1324–1332. [Google Scholar] [CrossRef] [PubMed]
  273. Hwang, J.S.; Kim, G.J.; Choi, H.G.; Kim, M.C.; Hahn, D.; Nam, J.W.; Nam, S.J.; Kwon, H.C.; Chin, J.; Cho, S.J.; et al. Identification of Antiangiogenic Potential and Cellular Mechanisms of Napyradiomycin A1 Isolated from the Marine-Derived Streptomyces sp. YP127. J. Nat. Prod. 2017, 80, 2269–2275. [Google Scholar] [CrossRef] [PubMed]
  274. Loaec, N.; Attanasio, E.; Villiers, B.; Durieu, E.; Tahtouh, T.; Cam, M.; Davis, R.A.; Alencar, A.; Roue, M.; Bourguet-Kondracki, M.L.; et al. Marine-Derived 2-Aminoimidazolone Alkaloids. Leucettamine B-Related Polyandrocarpamines Inhibit Mammalian and Protozoan DYRK & CLK Kinases. Mar. Drugs 2017, 15, 316. [Google Scholar]
  275. Starkus, J.G.; Poerzgen, P.; Layugan, K.; Kawabata, K.G.; Goto, J.I.; Suzuki, S.; Myers, G.; Kelly, M.; Penner, R.; Fleig, A.; et al. Scalaradial Is a Potent Inhibitor of Transient Receptor Potential Melastatin 2 (TRPM2) Ion Channels. J. Nat. Prod. 2017, 80, 2741–2750. [Google Scholar] [CrossRef] [PubMed]
  276. Kim, J.W.; Ko, S.K.; Kim, H.M.; Kim, G.H.; Son, S.; Kim, G.S.; Hwang, G.J.; Jeon, E.S.; Shin, K.S.; Ryoo, I.J.; et al. Stachybotrysin, an Osteoclast Differentiation Inhibitor from the Marine-Derived Fungus Stachybotrys sp. KCB13F013. J. Nat. Prod. 2016, 79, 2703–2708. [Google Scholar] [CrossRef]
  277. Mason, J.W.; Schmid, C.L.; Bohn, L.M.; Roush, W.R. Stolonidiol: Synthesis, Target Identification, and Mechanism for Choline Acetyltransferase Activation. J. Am. Chem. Soc. 2017, 139, 5865–5869. [Google Scholar] [CrossRef]
  278. Al-Awadhi, F.H.; Ratnayake, R.; Paul, V.J.; Luesch, H. Tasiamide F, a potent inhibitor of cathepsins D and E from a marine cyanobacterium. Bioorg. Med. Chem. 2016, 24, 3276–3282. [Google Scholar] [CrossRef] [Green Version]
  279. Espiritu, R.A.; Cornelio, K.; Kinoshita, M.; Matsumori, N.; Murata, M.; Nishimura, S.; Kakeya, H.; Yoshida, M.; Matsunaga, S. Marine sponge cyclic peptide theonellamide A disrupts lipid bilayer integrity without forming distinct membrane pores. Biochim. Biophys. Acta 2016, 1858, 1373–1379. [Google Scholar] [CrossRef]
  280. Garcia-Caballero, M.; Blacher, S.; Paupert, J.; Quesada, A.R.; Medina, M.A.; Noel, A. Novel application assigned to toluquinol: Inhibition of lymphangiogenesis by interfering with VEGF-C/VEGFR-3 signalling pathway. Br. J. Pharmacol. 2016, 173, 1966–1987. [Google Scholar] [CrossRef]
  281. Yang, Z.; Zhu, M.L.; Li, D.H.; Zeng, R.; Han, B.N. N-Me-trichodermamide B isolated from Penicillium janthinellum, with antioxidant properties through Nrf2-mediated signaling pathway. Bioorg. Med. Chem. 2017, 25, 6614–6622. [Google Scholar] [CrossRef]
  282. Chen, R.; Cheng, Z.; Huang, J.; Liu, D.; Wu, C.; Guo, P.; Lin, W. Versicotides D–F, new cyclopeptides with lipid-lowering activities. RSC Adv. 2017, 7, 49235–49243. [Google Scholar] [CrossRef] [Green Version]
  283. Abdjul, D.B.; Yamazaki, H.; Kanno, S.I.; Wewengkang, D.S.; Rotinsulu, H.; Sumilat, D.A.; Ukai, K.; Kapojos, M.M.; Namikoshi, M. Furanoterpenes, new types of protein tyrosine phosphatase 1B inhibitors, from two Indonesian marine sponges, Ircinia and Spongia spp. Bioorg. Med. Chem. Lett. 2017, 27, 1159–1161. [Google Scholar] [CrossRef] [PubMed]
  284. Zhang, Y.; Meng, T.; Zuo, L.; Bei, Y.; Zhang, Q.; Su, Z.; Huang, Y.; Pang, J.; Xiang, Q.; Yang, H. Xyloketal B Attenuates Fatty Acid-Induced Lipid Accumulation via the SREBP-1c Pathway in NAFLD Models. Mar. Drugs 2017, 15, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Mohamad, H.; Muhammad, R.; Andriani, Y.; Bakar, K.; Ismail, N.; Saidin, J.; Latip, J.; Musa, N.; Parenrengi, A. Potential secondary metabolites from marine sponge Aaptos aaptos for atherosclerosis and vibriosis treatments. Nat. Prod. Commun. 2017, 12, 1227–1230. [Google Scholar] [CrossRef] [Green Version]
  286. Tan, C.; Liu, Z.; Chen, S.; Huang, X.; Cui, H.; Long, Y.; Lu, Y.; She, Z. Antioxidative Polyketones from the Mangrove-Derived Fungus Ascomycota sp. SK2YWS-L. Sci. Rep. 2016, 6, 36609. [Google Scholar] [CrossRef] [Green Version]
  287. Leutou, A.S.; Yun, K.; Son, B.W. Induced production of 6,9-dibromoflavasperone, a new radical scavenging naphthopyranone in the marine-mudflat-derived fungus Aspergillus niger. Arch. Pharm. Res. 2016, 39, 806–810. [Google Scholar] [CrossRef]
  288. Sobolevskaya, M.P.; Zhuravleva, O.I.; Leshchenko, E.V.; Zakharenko, A.M.; Denisenko, V.A.; Kirichuk, N.N.; Popov, R.S.; Berdyshev, D.V.; Pislyagin, E.A.; Pivkin, M.V.; et al. New metabolites from the alga-derived fungi Penicillium thomii Maire and Penicillium lividum Westling. Phytochem. Lett. 2016, 15, 7–12. [Google Scholar] [CrossRef]
  289. Wu, B.; Wiese, J.; Schmaljohann, R.; Imhoff, J.F. Biscogniauxone, a New Isopyrrolonaphthoquinone Compound from the Fungus Biscogniauxia mediterranea Isolated from Deep-Sea Sediments. Mar. Drugs 2016, 14, 204. [Google Scholar] [CrossRef] [Green Version]
  290. Oleinikova, G.K.; Denisenko, V.A.; Berdyshev, D.V.; Pushilin, M.A.; Kirichuk, N.N.; Menzorova, N.I.; Kuzmich, A.S.; Yurchenko, E.A.; Zhuravleva, O.I. Two new sesterterpenoids, terretonins H and I, from the marine-derived fungus Aspergillus ustus. Phytochem. Lett. 2016, 17, 135–139. [Google Scholar] [CrossRef]
  291. El-Desoky, A.H.; Kato, H.; Kagiyama, I.; Hitora, Y.; Losung, F.; Mangindaan, R.E.; de Voogd, N.J.; Tsukamoto, S. Ceylonins A-F, Spongian Diterpene Derivatives That Inhibit RANKL-Induced Formation of Multinuclear Osteoclasts, from the Marine Sponge Spongia ceylonensis. J. Nat. Prod. 2017, 80, 90–95. [Google Scholar] [CrossRef]
  292. Li, Y.; Liu, D.; Cheng, Z.; Proksch, P.; Lin, W. Cytotoxic trichothecene-type sesquiterpenes from the sponge-derived fungus Stachybotrys chartarum with tyrosine kinase inhibition. RSC Adv. 2017, 7, 7259–7267. [Google Scholar] [CrossRef] [Green Version]
  293. Noinart, J.; Buttachon, S.; Dethoup, T.; Gales, L.; Pereira, J.A.; Urbatzka, R.; Freitas, S.; Lee, M.; Silva, A.M.S.; Pinto, M.M.M.; et al. A New Ergosterol Analog, a New Bis-Anthraquinone and Anti-Obesity Activity of Anthraquinones from the Marine Sponge-Associated Fungus Talaromyces stipitatus KUFA 0207. Mar. Drugs 2017, 15, 139. [Google Scholar] [CrossRef] [Green Version]
  294. Li, X.; Xia, Z.; Tang, J.; Wu, J.; Tong, J.; Li, M.; Ju, J.; Chen, H.; Wang, L. Identification and Biological Evaluation of Secondary Metabolites from Marine Derived Fungi-Aspergillus sp. SCSIOW3, Cultivated in the Presence of Epigenetic Modifying Agents. Molecules 2017, 22, 1302. [Google Scholar] [CrossRef] [Green Version]
  295. Yun, K.; Khong, T.T.; Leutou, A.S.; Kim, G.D.; Hong, J.; Lee, C.H.; Son, B.W. Cristazine, a New Cytotoxic Dioxopiperazine Alkaloid from the Mudflat-Sediment-Derived Fungus Chaetomium cristatum. Chem. Pharm. Bull. 2016, 64, 59–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Long, H.; Cheng, Z.; Huang, W.; Wu, Q.; Li, X.; Cui, J.; Proksch, P.; Lin, W. Diasteltoxins A-C, Asteltoxin-Based Dimers from a Mutant of the Sponge-Associated Emericella variecolor Fungus. Org. Lett. 2016, 18, 4678–4681. [Google Scholar] [CrossRef] [PubMed]
  297. Goey, A.K.; Chau, C.H.; Sissung, T.M.; Cook, K.M.; Venzon, D.J.; Castro, A.; Ransom, T.R.; Henrich, C.J.; McKee, T.C.; McMahon, J.B.; et al. Screening and Biological Effects of Marine Pyrroloiminoquinone Alkaloids: Potential Inhibitors of the HIF-1alpha/p300 Interaction. J. Nat. Prod. 2016, 79, 1267–1275. [Google Scholar] [CrossRef] [PubMed]
  298. Arai, M.; Shin, D.; Kamiya, K.; Ishida, R.; Setiawan, A.; Kotoku, N.; Kobayashi, M. Marine spongean polybrominated diphenyl ethers, selective growth inhibitors against the cancer cells adapted to glucose starvation, inhibits mitochondrial complex II. J. Nat. Med. 2017, 71, 44–49. [Google Scholar] [CrossRef] [PubMed]
  299. Jiao, W.H.; Shi, G.H.; Xu, T.T.; Chen, G.D.; Gu, B.B.; Wang, Z.; Peng, S.; Wang, S.P.; Li, J.; Han, B.N.; et al. Dysiherbols A-C and Dysideanone E, Cytotoxic and NF-kappaB Inhibitory Tetracyclic Meroterpenes from a Dysidea sp. Marine Sponge. J. Nat. Prod. 2016, 79, 406–411. [Google Scholar] [CrossRef] [PubMed]
  300. Vasileva, E.A.; Mishchenko, N.P.; Zadorozhny, P.A.; Fedoreyev, S.A. New Aminonaphthoquinone from the Sea Urchins Strongylocentrotus pallidus and Mesocentrotus nudus. Nat. Prod. Commun. 2016, 11, 821–824. [Google Scholar] [CrossRef] [Green Version]
  301. Almeida, C.; El, M.F.; Kehraus, S.; Schnakenburg, G.; Konig, G.M. Endolides A and B, Vasopressin and Serotonin-Receptor Interacting N-Methylated Peptides from the Sponge-Derived Fungus Stachylidium sp. Org. Lett. 2016, 18, 528–531. [Google Scholar] [CrossRef]
  302. Meng, L.H.; Mandi, A.; Li, X.M.; Liu, Y.; Kurtan, T.; Wang, B.G. Isolation, Stereochemical Study, and Antioxidant Activity of Benzofuranone Derivatives from a Mangrove-derived Fungus Eurotium rubrum MA-150. Chirality 2016, 28, 581–584. [Google Scholar] [CrossRef] [PubMed]
  303. Manda, S.; Sharma, S.; Wani, A.; Joshi, P.; Kumar, V.; Guru, S.K.; Bharate, S.S.; Bhushan, S.; Vishwakarma, R.A.; Kumar, A.; et al. Discovery of a marine-derived bis-indole alkaloid fascaplysin, as a new class of potent P-glycoprotein inducer and establishment of its structure-activity relationship. Eur. J. Med. Chem. 2016, 107, 1–11. [Google Scholar] [CrossRef] [PubMed]
  304. Huang, R.Y.; Chen, W.T.; Kurtan, T.; Mandi, A.; Ding, J.; Li, J.; Li, X.W.; Guo, Y.W. Bioactive isoquinolinequinone alkaloids from the South China Sea nudibranch Jorunna funebris and its sponge-prey Xestospongia sp. Future Med. Chem. 2016, 8, 17–27. [Google Scholar] [CrossRef] [PubMed]
  305. Hahn, D.; Kim, H.; Yang, I.; Chin, J.; Hwang, H.; Won, D.H.; Lee, B.; Nam, S.J.; Ekins, M.; Choi, H.; et al. The Halicylindramides, Farnesoid X Receptor Antagonizing Depsipeptides from a Petrosia sp. Marine Sponge Collected in Korea. J. Nat. Prod. 2016, 79, 499–506. [Google Scholar] [CrossRef]
  306. Abdjul, D.B.; Yamazaki, H.; Takahashi, O.; Kirikoshi, R.; Ukai, K.; Namikoshi, M. Isopetrosynol, a New Protein Tyrosine Phosphatase 1B Inhibitor, from the Marine Sponge Halichondria cf. panicea Collected at Iriomote Island. Chem. Pharm. Bull. 2016, 64, 733–736. [Google Scholar] [CrossRef]
  307. Cui, J.; Morita, M.; Ohno, O.; Kimura, T.; Teruya, T.; Watanabe, T.; Suenaga, K.; Shibasaki, M. Leptolyngbyolides, Cytotoxic Macrolides from the Marine Cyanobacterium Leptolyngbya sp.: Isolation, Biological Activity, and Catalytic Asymmetric Total Synthesis. Chemistry 2017, 23, 8500–8509. [Google Scholar] [CrossRef]
  308. Nazari, M.; Serrill, J.D.; Wan, X.; Nguyen, M.H.; Anklin, C.; Gallegos, D.A.; Smith, A.B., III; Ishmael, J.E.; McPhail, K.L. New Mandelalides Expand a Macrolide Series of Mitochondrial Inhibitors. J. Med. Chem. 2017, 60, 7850–7862. [Google Scholar] [CrossRef]
  309. Wahab, H.A.; Pham, N.B.; Muhammad, T.S.; Hooper, J.N.; Quinn, R.J. Merosesquiterpene Congeners from the Australian Sponge Hyrtios digitatus as Potential Drug Leads for Atherosclerosis Disease. Mar. Drugs 2016, 15, 6. [Google Scholar] [CrossRef] [Green Version]
  310. Korolkova, Y.; Makarieva, T.; Tabakmakher, K.; Shubina, L.; Kudryashova, E.; Andreev, Y.; Mosharova, I.; Lee, H.S.; Lee, Y.J.; Kozlov, S. Marine Cyclic Guanidine Alkaloids Monanchomycalin B and Urupocidin A Act as Inhibitors of TRPV1, TRPV2 and TRPV3, but not TRPA1 Receptors. Mar. Drugs 2017, 15, 87. [Google Scholar] [CrossRef] [Green Version]
  311. Chianese, G.; Yu, H.B.; Yang, F.; Sirignano, C.; Luciano, P.; Han, B.N.; Khan, S.; Lin, H.W.; Taglialatela-Scafati, O. PPAR Modulating Polyketides from a Chinese Plakortis simplex and Clues on the Origin of Their Chemodiversity. J. Org. Chem. 2016, 81, 5135–5143. [Google Scholar] [CrossRef] [Green Version]
  312. Xue, D.Q.; Liu, H.L.; Chen, S.H.; Mollo, E.; Gavagnin, M.; Li, J.; Li, X.W.; Guo, Y.W. 5-Alkylpyrrole-2-carboxaldehyde derivatives from the Chinese sponge Mycale lissochela and their PTP1B inhibitory activities. Chinese Chem. Lett. 2017, 28, 1190–1193. [Google Scholar] [CrossRef]
  313. Wang, J.; Mu, F.R.; Jiao, W.H.; Huang, J.; Hong, L.L.; Yang, F.; Xu, Y.; Wang, S.P.; Sun, F.; Lin, H.W. Meroterpenoids with Protein Tyrosine Phosphatase 1B Inhibitory Activity from a Hyrtios sp. Marine Sponge. J. Nat. Prod. 2017, 80, 2509–2514. [Google Scholar] [CrossRef] [PubMed]
  314. Fukuhara, K.; Takada, K.; Okada, S.; Matsunaga, S. Nazumazoles D-F, Cyclic Pentapeptides That Inhibit Chymotrypsin, from the Marine Sponge Theonella swinhoei. J. Nat. Prod. 2016, 79, 1694–1697. [Google Scholar] [CrossRef] [PubMed]
  315. Wiese, J.; Imhoff, J.F.; Gulder, T.A.; Labes, A.; Schmaljohann, R. Marine Fungi as Producers of Benzocoumarins, a New Class of Inhibitors of Glycogen-Synthase-Kinase 3beta. Mar. Drugs 2016, 14, 200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Ding, H.; Zhang, D.; Zhou, B.; Ma, Z. Inhibitors of BRD4 Protein from a Marine-Derived Fungus Alternaria sp. NH-F6. Mar. Drugs 2017, 15, 76. [Google Scholar] [CrossRef] [Green Version]
  317. Tanokashira, N.; Kukita, S.K.; Kato, H.; Nehira, T.; Angkouw, E.D.; Mangindaan, R.E.; Devoogd, N.J.; Tsukamoto, S. Petroquinones: Trimeric and dimeric xestoquinone derivatives isolated from the marine sponge Petrosia alfiani. Tetrahedron 2016, 72, 5530–5540. [Google Scholar] [CrossRef]
  318. Zhao, Y.; Liu, D.; Proksch, P.; Yu, S.; Lin, W. Isocoumarin Derivatives from the Sponge-Associated Fungus Peyronellaea glomerata with Antioxidant Activities. Chem. Biodivers. 2016, 13, 1186–1193. [Google Scholar] [CrossRef]
  319. Lee, M.S.; Wang, S.W.; Wang, G.J.; Pang, K.L.; Lee, C.K.; Kuo, Y.H.; Cha, H.J.; Lin, R.K.; Lee, T.H. Angiogenesis Inhibitors and Anti-Inflammatory Agents from Phoma sp. NTOU4195. J. Nat. Prod. 2016, 79, 2983–2990. [Google Scholar] [CrossRef]
  320. Imperatore, C.; Luciano, P.; Aiello, A.; Vitalone, R.; Irace, C.; Santamaria, R.; Li, J.; Guo, Y.W.; Menna, M. Structure and Configuration of Phosphoeleganin, a Protein Tyrosine Phosphatase 1B Inhibitor from the Mediterranean Ascidian Sidnyum elegans. J. Nat. Prod. 2016, 79, 1144–1148. [Google Scholar] [CrossRef]
  321. Georgantea, P.; Ioannou, E.; Evain-Bana, E.; Bagrel, D.; Martinet, N.; Vagias, C.; Roussis, V. Sesquiterpenes with inhibitory activity against CDC25 phosphatases from the soft coral Pseudopterogorgia rigida. Tetrahedron 2016, 72, 3262–3269. [Google Scholar] [CrossRef]
  322. El-Desoky, A.H.; Kato, H.; Tsukamoto, S. Ceylonins G-I: Spongian diterpenes from the marine sponge Spongia ceylonensis. J. Nat. Med. 2017, 71, 765–769. [Google Scholar] [CrossRef]
  323. Ortega, M.J.; Pantoja, J.J.; de Los, R.C.; Zubia, E. 5-Alkylresorcinol Derivatives from the Bryozoan Schizomavella mamillata: Isolation, Synthesis, and Antioxidant Activity. Mar. Drugs 2017, 15, 344. [Google Scholar] [CrossRef] [Green Version]
  324. Yin, Y.; Fu, Q.; Wu, W.; Cai, M.; Zhou, X.; Zhang, Y. Producing Novel Fibrinolytic Isoindolinone Derivatives in Marine Fungus Stachybotrys longispora FG216 by the Rational Supply of Amino Compounds According to Its Biosynthesis Pathway. Mar. Drugs 2017, 15, 214. [Google Scholar] [CrossRef]
  325. Li, J.; Gu, B.B.; Sun, F.; Xu, J.R.; Jiao, W.H.; Yu, H.B.; Han, B.N.; Yang, F.; Zhang, X.C.; Lin, H.W. Sesquiterpene Quinones/Hydroquinones from the Marine Sponge Spongia pertusa Esper. J. Nat. Prod. 2017, 80, 1436–1445. [Google Scholar] [CrossRef]
  326. Afifi, A.H.; Kagiyama, I.; El-Desoky, A.H.; Kato, H.; Mangindaan, R.E.P.; de Voogd, N.J.; Ammar, N.M.; Hifnawy, M.S.; Tsukamoto, S. Sulawesins A-C, Furanosesterterpene Tetronic Acids That Inhibit USP7, from a Psammocinia sp. Marine Sponge. J. Nat. Prod. 2017, 80, 2045–2050. [Google Scholar] [CrossRef]
  327. Kanamori, Y.; Iwasaki, A.; Sumimoto, S.; Suenaga, K. Urumamide, a novel chymotrypsin inhibitor with a ß-amino acid from a marine cyanobacterium Okeania sp. Tetrahedrom Lett. 2016, 57, 4213–4216. [Google Scholar] [CrossRef]
  328. Cheng, Z.; Lou, L.; Liu, D.; Li, X.; Proksch, P.; Yin, S.; Lin, W. Versiquinazolines A-K, Fumiquinazoline-Type Alkaloids from the Gorgonian-Derived Fungus Aspergillus versicolor LZD-14-1. J. Nat. Prod. 2016, 79, 2941–2952. [Google Scholar] [CrossRef]
  329. He, W.F.; Xue, D.Q.; Yao, L.G.; Li, J.; Liu, H.L.; Guo, Y.W. A new bioactive steroidal ketone from the South China Sea sponge Xestospongia testudinaria. J. Asian Nat. Prod. Res 2016, 18, 195–199. [Google Scholar] [CrossRef]
  330. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2016, 33, 382–431. [Google Scholar] [CrossRef] [Green Version]
  331. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2017, 34, 235–294. [Google Scholar] [CrossRef] [Green Version]
  332. Berlinck, R.G.; Romminger, S. The chemistry and biology of guanidine natural products. Nat. Prod. Rep. 2016, 33, 456–490. [Google Scholar] [CrossRef]
  333. Shi, Y.; Moazami, Y.; Pierce, J.G. Structure, synthesis and biological properties of the pentacyclic guanidinium alkaloids. Bioorg. Med. Chem. 2017, 25, 2817–2824. [Google Scholar] [CrossRef]
  334. Pereira, R.B.; Andrade, P.B.; Valentao, P. Chemical Diversity and Biological Properties of Secondary Metabolites from Sea Hares of Aplysia Genus. Mar. Drugs 2016, 14, 39. [Google Scholar] [CrossRef] [Green Version]
  335. Chai, Q.Y.; Yang, Z.; Lin, H.W.; Han, B.N. Alkynyl-Containing Peptides of Marine Origin: A Review. Mar. Drugs 2016, 14, 216. [Google Scholar] [CrossRef] [Green Version]
  336. Mazard, S.; Penesyan, A.; Ostrowski, M.; Paulsen, I.T.; Egan, S. Tiny Microbes with a Big Impact: The Role of Cyanobacteria and Their Metabolites in Shaping Our Future. Mar. Drugs 2016, 14, 97. [Google Scholar] [CrossRef]
  337. Davila-Cespedes, A.; Hufendiek, P.; Crusemann, M.; Schaberle, T.F.; Konig, G.M. Marine-derived myxobacteria of the suborder Nannocystineae: An underexplored source of structurally intriguing and biologically active metabolites. Beilstein. J. Org. Chem 2016, 12, 969–984. [Google Scholar] [CrossRef]
  338. Offret, C.; Desriac, F.; Le, C.P.; Mounier, J.; Jegou, C.; Fleury, Y. Spotlight on Antimicrobial Metabolites from the Marine Bacteria Pseudoalteromonas: Chemodiversity and Ecological Significance. Mar. Drugs 2016, 14, 129. [Google Scholar] [CrossRef] [Green Version]
  339. Ma, H.G.; Liu, Q.; Zhu, G.L.; Liu, H.S.; Zhu, W.M. Marine natural products sourced from marine-derived Penicillium fungi. J. Asian Nat. Prod. Res. 2016, 18, 92–115. [Google Scholar] [CrossRef]
  340. Zhang, P.; Li, X.; Wang, B.G. Secondary Metabolites from the Marine Algal-Derived Endophytic Fungi: Chemical Diversity and Biological Activity. Planta Med. 2016, 82, 832–842. [Google Scholar] [CrossRef] [Green Version]
  341. Abdul, Q.A.; Choi, R.J.; Jung, H.A.; Choi, J.S. Health benefit of fucosterol from marine algae: A review. J. Sci. Food Agric. 2016, 96, 1856–1866. [Google Scholar] [CrossRef]
  342. Avila, C. Ecological and Pharmacological Activities of Antarctic Marine Natural Products. Planta Med. 2016, 82, 767–774. [Google Scholar] [CrossRef] [PubMed]
  343. Bahrami, Y.; Franco, C.M. Acetylated Triterpene Glycosides and Their Biological Activity from Holothuroidea Reported in the Past Six Decades. Mar. Drugs 2016, 14, 147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Chang, Y.C.; Sheu, J.H.; Wu, Y.C.; Sung, P.J. Terpenoids from Octocorals of the Genus Pachyclavularia. Mar. Drugs 2017, 15, 382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Shady, N.H.; El-Hossary, E.M.; Fouad, M.A.; Gulder, T.A.M.; Kamel, M.S.; Abdelmohsen, U.R. Bioactive Natural Products of Marine Sponges from the Genus Hyrtios. Molecules 2017, 22, 781. [Google Scholar] [CrossRef] [PubMed]
  346. Zhang, H.; Dong, M.; Wang, H.; Crews, P. Secondary Metabolites from the Marine Sponge Genus Phyllospongia. Mar. Drugs 2017, 15, 12. [Google Scholar] [CrossRef] [Green Version]
  347. Amoutzias, G.D.; Chaliotis, A.; Mossialos, D. Discovery Strategies of Bioactive Compounds Synthesized by Nonribosomal Peptide Synthetases and Type-I Polyketide Synthases Derived from Marine Microbiomes. Mar. Drugs 2016, 14, 80. [Google Scholar] [CrossRef] [Green Version]
  348. Newman, D.J. Developing natural product drugs: Supply problems and how they have been overcome. Pharmacol. Ther. 2016, 162, 1–9. [Google Scholar] [CrossRef]
  349. Daletos, G.; Ebrahim, W.; Ancheeva, E.; El-Neketi, M.; Song, W.; Lin, W.; Proksch, P. Natural Products from Deep-Sea-Derived Fungi A New Source of Novel Bioactive Compounds? Curr. Med. Chem. 2018, 25, 186–207. [Google Scholar] [CrossRef]
  350. Mariottini, G.L.; Grice, I.D. Antimicrobials from Cnidarians. A New Perspective for Anti-Infective Therapy? Mar. Drugs 2016, 14, 48. [Google Scholar] [CrossRef] [Green Version]
  351. Cheuka, P.M.; Mayoka, G.; Mutai, P.; Chibale, K. The Role of Natural Products in Drug Discovery and Development against Neglected Tropical Diseases. Molecules 2016, 22, 58. [Google Scholar] [CrossRef] [Green Version]
  352. Sansinenea, E.; Ortiz, A. Antimycobacterial Natural Products from Marine Pseudopterogorgia elisabethae. Curr. Org. Synth. 2016, 13, 556–568. [Google Scholar] [CrossRef]
  353. Chen, C.Y.; Tsai, Y.F.; Chang, W.Y.; Yang, S.C.; Hwang, T.L. Marine Natural Product Inhibitors of Neutrophil-Associated Inflammation. Mar. Drugs 2016, 14, 141. [Google Scholar] [CrossRef] [PubMed]
  354. Chiesa, G.; Busnelli, M.; Manzini, S.; Parolini, C. Nutraceuticals and Bioactive Components from Fish for Dyslipidemia and Cardiovascular Risk Reduction. Mar. Drugs 2016, 14, 113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Alghazwi, M.; Kan, Y.Q.; Zhang, W.; Gai, W.P.; Yan, X.X. Neuroprotective Activities of Marine Natural Products from Marine Sponges. Curr. Med. Chem. 2016, 23, 360–382. [Google Scholar] [CrossRef]
  356. Huang, Y.; Peng, C.; Yi, Y.; Gao, B.; Shi, Q. A Transcriptomic Survey of Ion Channel-Based Conotoxins in the Chinese Tubular Cone Snail (Conus betulinus). Mar. Drugs 2017, 15, 228. [Google Scholar] [CrossRef] [PubMed]
  357. Molgo, J.; Marchot, P.; Araoz, R.; Benoit, E.; Iorga, B.I.; Zakarian, A.; Taylor, P.; Bourne, Y.; Servent, D. Cyclic imine toxins from dinoflagellates: A growing family of potent antagonists of the nicotinic acetylcholine receptors. J. Neurochem. 2017, 142 (Suppl. 2), 41–51. [Google Scholar] [CrossRef]
  358. Sadeghi, M.; McArthur, J.R.; Finol-Urdaneta, R.K.; Adams, D.J. Analgesic conopeptides targeting G protein-coupled receptors reduce excitability of sensory neurons. Neuropharmacology 2017, 127, 116–123. [Google Scholar] [CrossRef] [Green Version]
  359. Bai, H.; Kong, W.W.; Shao, C.L.; Li, Y.; Liu, Y.Z.; Liu, M.; Guan, F.F.; Wang, C.Y. Zebrafish Embryo Toxicity Microscale Model for Ichthyotoxicity Evaluation of Marine Natural Products. Mar. Biotechnol. 2016, 18, 264–270. [Google Scholar] [CrossRef]
  360. Agrawal, S.; Adholeya, A.; Deshmukh, S.K. The Pharmacological Potential of Non-ribosomal Peptides from Marine Sponge and Tunicates. Front. Pharmacol. 2016, 7, 333. [Google Scholar] [CrossRef] [Green Version]
  361. Garcia-Vilas, J.A.; Martinez-Poveda, B.; Quesada, A.R.; Medina, M.A. Aeroplysinin-1, a Sponge-Derived Multi-Targeted Bioactive Marine Drug. Mar. Drugs 2016, 14, 1. [Google Scholar] [CrossRef] [Green Version]
  362. Alfonso, A.; Vieytes, M.R.; Botana, L.M. Yessotoxin, a Promising Therapeutic Tool. Mar. Drugs 2016, 14, 30. [Google Scholar] [CrossRef] [Green Version]
  363. Valeur, E.; Gueret, S.; Adihou, H.; Gopalakrishnan, R.; Lemurell, M.; Waldmann, H.; Grossman, T.; Plowright, A. New Modalities for Challenging Targets in Drug Discovery. Angew. Chem. Int. Ed. 2017, 56, 10294–10323. [Google Scholar] [CrossRef] [PubMed]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mayer, A.M.S.; Guerrero, A.J.; Rodríguez, A.D.; Taglialatela-Scafati, O.; Nakamura, F.; Fusetani, N. Marine Pharmacology in 2016–2017: 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 2021, 19, 49. https://doi.org/10.3390/md19020049

AMA Style

Mayer AMS, Guerrero AJ, Rodríguez AD, Taglialatela-Scafati O, Nakamura F, Fusetani N. Marine Pharmacology in 2016–2017: 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. 2021; 19(2):49. https://doi.org/10.3390/md19020049

Chicago/Turabian Style

Mayer, Alejandro M. S., Aimee J. Guerrero, Abimael D. Rodríguez, Orazio Taglialatela-Scafati, Fumiaki Nakamura, and Nobuhiro Fusetani. 2021. "Marine Pharmacology in 2016–2017: 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 19, no. 2: 49. https://doi.org/10.3390/md19020049

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