Next Article in Journal
A Computational Inter-Species Study on Safrole Phase I Metabolism-Dependent Bioactivation: A Mechanistic Insight into the Study of Possible Differences among Species
Previous Article in Journal
A Comprehensive Structural Analysis of Clostridium botulinum Neurotoxin A Cell-Binding Domain from Different Subtypes
Previous Article in Special Issue
Cytotoxic Effects of Cannabidiol on Neonatal Rat Cortical Neurons and Astrocytes: Potential Danger to Brain Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antifungal and Antibacterial Activities of Isolated Marine Compounds

1
General Safety Section, General Services Department, Birzeit University, Ramallah 00972, Palestine
2
Faculty of Pharmacy, Nursing and Health Professions, Birzeit University, Ramallah 00972, Palestine
3
Pharmaceutical Sciences Department, Faculty of Pharmacy, Al-Quds University, Jerusalem 20002, Palestine
4
Department of Sciences, University of Basilicata, Via dell’Ateneo Lucano 10, 85100 Potenza, Italy
5
Department of Geography, Environmental Management and Energy Studies, University of Johannesburg, Auckland Park Kingsway Campus, Johannesburg 2092, South Africa
6
Department of European and Mediterranean Cultures, University of Basilicata, Via Lanera 20, 75100 Matera, Italy
*
Author to whom correspondence should be addressed.
Toxins 2023, 15(2), 93; https://doi.org/10.3390/toxins15020093
Submission received: 11 December 2022 / Revised: 7 January 2023 / Accepted: 7 January 2023 / Published: 18 January 2023
(This article belongs to the Collection Toxic and Pharmacological Effect of Plant Toxins)

Abstract

:
To combat the ineffectiveness of currently available pharmaceutical medications, caused by the emergence of increasingly resistant bacterial and fungal strains, novel antibacterial and antifungal medications are urgently needed. Novel natural compounds with antimicrobial activities can be obtained by exploring underexplored habitats such as the world’s oceans. The oceans represent the largest ecosystem on earth, with a high diversity of organisms. Oceans have received some attention in the past few years, and promising compounds with antimicrobial activities were isolated from marine organisms such as bacteria, fungi, algae, sea cucumbers, sea sponges, etc. This review covers 56 antifungal and 40 antibacterial compounds from marine organisms. These compounds are categorized according to their chemical structure groups, including polyketides, alkaloids, ribosomal peptides, and terpenes, and their organismal origin. The review provides the minimum inhibitory concentration MIC values and the bacterial/fungal strains against which these chemical compounds show activity. This study shows strong potential for witnessing the development of new novel antimicrobial drugs from these natural compounds isolated and evaluated for their antimicrobial activities.
Key Contribution: This review was conducted to summarize and categorize compounds isolated from marine organisms and showed significant antifungal and antimicrobial activities.

1. Introduction

Antibiotics are one of the most powerful medications developed to fight against dangerous infections. Their discovery has greatly improved human and animal health. Unfortunately, we are now witnessing a period in which people are dying from untreatable infections. The particular reason for these circumstances is the emergence and spread of antibiotic-resistant microorganisms. Antibiotic-resistant infections can be difficult, and sometimes impossible, to treat, resulting in mortality cases. The center for disease control and prevention CDC’s 2019 Antibiotic Resistance (AR) Threats Report mentions that antimicrobial resistance is an urgent global public health threat, killing at least 1.27 million people worldwide [1]. The report adds that more than 2.8 million antimicrobial-resistant infections occur each year in the U.S.A., causing the death of more than 35,000 people.
Antibacterial resistance is rapidly developing in bacteria as a result of the incorrect and excessive use of antibacterial medications among healthcare professionals and patients [2,3,4]. Therefore, there is a need to improve the appropriate use of antibiotics and reduce unnecessary use. Parallel to that, discovering and developing new antimicrobial products is of great importance to human, animal, and agricultural health [5].
The majority of new antibacterial agents developed in the pharmaceutical industry have been semisynthetic modifications of the original natural product discovered more than 50 years ago. In fact, beta-lactams, macrolides, and quinolones accounted for more than 70% of antibacterial medications approved between 1981 and 2005 [2]. Existing natural products that have been modified have given rise to compounds that momentarily defeat the resistance mechanisms. Eventually, bacterial resistance will be overcome only through the development of completely new natural compounds.
Over the past 50 years, most antibiotics have been discovered in terrestrial species. The ability of aquatic animals to create antibacterial chemicals has received little attention. The marine ecosystem is the largest and most important ecosystem on earth. It includes a great diversity of different groups of organisms that range in size from nanoscale microorganisms to whales [6]. Therefore, the ocean represents a potential source for the development of new antibiotics, and research into uncharted marine ecosystems is necessary to meet the pressing demand for new effective antibiotics. This variety provides a wealth of sources for the exact purpose of identifying novel medications that may be effective against particular diseases. There is a substantially higher chance of finding new antibacterial drug leads in marine environments than in terrestrial environments [7].
A variety of marine organisms, including bacteria, fungi, seaweeds, corals, sea cucumbers, sponges, and others, have been used to isolate antifungal and antibacterial biological compounds that fall into the following chemical groups: peptides, terpenoids, diacylglycerols, steroids, polysaccharides, polyketides, alkaloids, and others [8,9,10,11,12,13].
Although utilizing the reservoir of marine species freely for bioassays and therapy is difficult due to the relatively low availability of biologically active compounds, these difficulties can be overcome by several methods, such as mariculture (the cultivation of marine sponges), sponge-bioreactor specimen creation, sponge-cell culture systems (perimorph culture), genetic modification, and synthesis. The preferred options among them are still chemical synthesis and semi-synthesis. To investigate structure-activity relationships (SAR), synthetic organic chemistry can offer extensive biological screening and access to synthetic analogs [14,15].
The present study aims to investigate and summarize the antifungal and antibacterial activities of different chemical compounds isolated from marine organisms.

2. Isolated Marine Compounds with Antifungal Activity

2.1. Antifungal Compounds Isolated from Marine Bacteria

Marine microbes, frequently referred to as chemical gold, are considered to be a great source of novel treatments [16,17]. Bacteria are ubiquitous throughout the marine ecosystem. They can adapt to and change for any challenging environment. Therefore, marine bacteria are generally more effective than terrestrial bacteria in the bioremediation of toxic, heavy metals, hydrocarbon, and xenobiotics, as well as many other recalcitrant compounds. This is attributed to the production of extracellular polymeric substances (EPS) and the formation of biofilms [18].
Ieodoglucomide C (1 in Scheme 1) and ieodoglycolipid (2 in Scheme 1) are two glycolipids which are both isolated from the aquatic bacterium Bacillus licheniformis. It was found that they both have potent antifungal activity, with MIC values of 0.02–0.03 µM against the human pathogens Candida albicans, Colletotrichum acutatum, Botrytis cinerea, Rhizoctonia solani, and Aspergillus niger [19,20].
Hedaya48, which was synthesized by the Aplysina fistularis sponge when subjected to various UV radiation dosages, 5,7-dimethoxy-4-p-methoxyl phenyl coumarin (3 in Scheme 1), and saadamycin (4 in Scheme 1) were all new antimycotic substances identified from endophytic Streptomyces sp. The MIC value of saadamycin was reported to be 1–5.16 µg/mL, whereas 7.5–100 µg/mL was observed for 5,7-dimethoxy-4-p-methoxyl phenyl coumarin against dermatophytes as well as other fungi, including Cryptococcus humicolus, Fusarium oxysporum, Aspergillus fumigatus, A. niger, and Microsporum gypseum [21,22].
Actinomycetes were used to create the new and superior antifungal drug caerulomycin A. (5 in Scheme 1). Actinomycete strain PM0525875 for extraction was obtained from a marine invertebrate. Actinomycetes extracts showed strong effectiveness against drug-resistant fungus strains in in vitro investigations. The fluconazole-resistant Candida glabrata, C. albicans, C. albicans CO9, and Candida krusei were the pathogenic fungal test strains used to determine the MIC value of caerulomycin A. The MIC values reported ranged between 0.39 and 1.56 µg/mL [23,24].
The secondary metabolite, pedein A (6 in Scheme 1), was isolated from the cell mass of the myxobacterium Chondromyces pediculatus. Pedein A inhibited the growth of a broad spectrum of yeasts and fungi, whereas Gram-positive and Gram-negative bacteria such as Bacillus subtilis, Brevibacterium ammoniagenes, Corynebacterium fascians, Micrococcus luteus, Staphylococcus aureus, Escherichia coli, Enterobacter aerogenes, Pseudomonas aeruginosa, and Salmonella typhimurium were not sensitive to the antibiotic. MIC value for Rhodotorula glutinis was reported to be 0.6 µg/mL, and an MIC value of 1.6 µg/mL was reported for both Saccharomyces cerevisiae and Candida albicans. Furthermore, pedein A showed inhibitory activity against the growth of some filamentous fungi with a zone diameter range of 22–35 mm for Botrytis cinerea, Gibberella fujikuroi, Pythium debaryanum, Rhizopus arrhizus, Trichoderma koningii, and Ustilago maydis [20,21].
Other important isolated antifungal compounds, their marine sources, and their activities are listed in Table 1.

2.2. Antifungal Compounds Isolated from Marine Fungi

From the ocean’s surface to its deepest parts, fungi have been discovered to exist in almost every aquatic habitat studied [25]. As a result of marine fungi’s superior biological characteristics to terrestrial fungi and their ability to adapt to extreme pH, temperature, and salinity, a wider range of biotechnological applications of marine fungi are possible [26].
In Greenland, Trichoderma sp. strain MF106 was the source of pyridoxatin (15 in Scheme 1), demonstrating antifungal activity with IC50 values of 1.07 ± 0.34 µM against Trichophyton rubrum and 6.9 ± 0.04 µM against C. Albicans [27,28]. The Japanese isolated diketopiperazine (16 in Scheme 1) demonstrated growth inhibition against P. oryzae. and P. yezoensis with an IC50 value of 350 nM [29,30].
Didymellamide A (17 in Scheme 1), isolated from the fungus Stagonosporopsis cucurbitacearum, reduced the growth of C. albicans, Candida glabrata, and Cryptococcus neoformans strains at doses of 1.6–3.1 µg/mL [31,32]. Additionally, the Aspergillus sclerotiorum PT06-1 isolates of sclerotide B (18 in Scheme 1) and sclerotide B (19 in Scheme 1) both exhibited activity against Candida albicans, with MIC values of 7.0 and 3.5 µM, respectively [33,34].
The plant pathogenic fungus Fusarium graminearum, Alternaria brassicae, and Colletotrichum gloeosporioides were all inhibited by varioxepine A, which has an MIC value of 4 µg/mL; peniciadametizine A, which has an MIC value of 4 µg/mL; and penicibilaenes A, which has an MIC value of 1.0 µg/mL (2022 in Scheme 1), respectively. They were extracted from Paecilomyces variotii, Phoma sp. Q60596, and Penicillium bilaiae MA-267 fungus [35,36,37]. On the other hand, penicibrocazines B and E (23,24 in Scheme 1), which were isolated from Penicillium brocae MA-231 (Avicennia marina culture extract), showed activity against the plant pathogen Gaeumannomyces graminis, with a 0.25 µg/mL MIC value for both [38].

2.3. Antifungal Compounds Isolated from Marine Algae

Caulerprenylol B (25 in Scheme 1), which was obtained from Chinese alga Caulerpa racemosa, has excellent antifungal activity against T. rubrum fungus, which causes two of the most common fungal infections, known as ’athlete’s foot’ and ’jock itch’, with an MIC80 value of 16 µg/mL [39].
Lobophorolide (26 in Scheme 1), isolated from Lobophora variegata (marine brown alga) of the Bahamas and Egypt, has excellent activity against the pathogenic ascomycete Lindra thalassiae and the saprophytic deuteromycete Dendryphiella salina, with IC50 values of 0.135 and 0.034 µg/mL, respectively. Further, it showed antifungal activity against C. albicans wild and amphotericin-resistant strains, with IC50 values of 1.3 and 0.5 µg/mL [24,40].
The isolated isolauraldehyde (27 in Scheme 1) showed antifungal activity against C. albicans, A. fumigatus, and A. flavus with MIC values of 70, 100, and 1000 µg/mL, respectively. The organic extract of isolauraldehyde was obtained from the red alga Laurencia obtuse [41,42].
The growth of Mycobacterium smegmatis and Neurospora crassa could be inhibited by the ethanolic extract of Gracilaria domigensis [43]. Gracilaria sjoestedii and Gracilaria debilis ethanolic extract had antifungal activity against C. albicans [44].

2.4. Antifungal Compounds Isolated from Sea Cucumbers

Sea cucumbers are animals with long bodies and leathery skin. They contain several antifungal compounds, such as variegatuside D. (28 in Scheme 1), which was isolated from Stichopus variegates and which showed antifungal activity against Microsporum gypseum, C. albicans, C. pseudotropicalis, and C. parapsilosis, all of which have 3.4 µg/mL MIC80 value [45,46].
Scabraside A (29 in Scheme 1) isolated from Holothuria scabra exhibited antifungal activities against A. fumigatus, C. pseudotropicalis, M. gypseum, T. rubrum, and C. albicans, with MIC values of 2, 4, 4, 8, and 8 µg/mL, respectively [47].
Antifungal activity against C. tropicalis and M. gypseum 31388 with MIC80 values of 1.4–5.7 µM were reported for holotoxin D1 (30 in Scheme 1) and stichloroside C1 (31, in Scheme 1), which were isolated from Apostichopus japonicus Selenka [48].
The growth of Cryptococcus neoformans, Richophyton rubrum, C. albicans, C. tropicalis, A. fumigatus, and C. krusei could be inhibited with MIC80 values ranging from 0.7 to 2.81 µM by marmoratoside A, impatient side A, and bivittoside D (3234 in Scheme 1) isolated from Bohadschia marmorata Jaeger [49,50].

2.5. Antifungal Compounds Isolated from Sea Sponges

Sponges are elementary multi-cellular animals with dense skeleton muscles. They have a vast repertoire of antifungal compounds, which are useful in cases of resistance to amphotericin B and fluconazole [51].
The growth of C. albicans was inhibited by the isolated epiplakinic acid F (35, in Scheme 1) and agelasidine F and C (36,37 in Scheme 1), which have MIC values of 3.1, 4, and 0.5 µg/mL, respectively. Epiplakinic acid F was extracted from the Seychelles sponge genus Plakinastrella. Agelasidine F and C were obtained from Agelas citrina (Caribbean sponge) [51,52,53]. Table 2 lists other isolated compounds from sea sponges that exhibit antifungal activity against C. albicans.
The highly oxygenated alkaloid massadine (38 in Scheme 1), which was isolated from the marine sponge Stylissa aff. massa, inhibited Geranylgeranyltransferase-I from C. albicans with an IC50 value of 3.9 µM. Moreover, massadine inhibited the growth of Cryptococcus neoformans with an MIC value of 32 µM, but it did not inhibit the growth of C. albicans at a concentration of 64 µM [53].
Haliscosamine (50 in Scheme 1), plakortide F acid (51 in Scheme 1) and simplexolide E (52 in Scheme 1) showed antifungal effectiveness against C. neoformans with MIC values ranged 0.2–3.66 μg/mL, respectively.
Haliscosamine was obtained from Haliclona viscosa (Moroccan sponge), plakortide F acid from Plakortis halichondrioides sponge, and simplexolide E from the sponge Plakortis simplex found in China [62,63,64].
Puupehenone (53 in Scheme 1) (isolated from Hyrtios sp. sponge) showed antifungal activity with MIC values of 1.25 µg/mL and 2.50 µg/mL against C. neoformans and C. krusei, respectively [65]. The isolated Chinese Hippolachnin A (54 in Scheme 1) from Hippospongia lachne sponge showed antifungal activity against C. neoformans, T. rubrum, and M. gypseum, with MIC values of 0.41 μM for each fungus [66]. Furthermore, with MIC values ranging between 1.9 and 7.8 µg/mL, the Brazilian batzelladine L (55 in Scheme 1) isolated from the Monanchora arbuscular sponge exhibited activity against A. flavus strains [67].
A reasonably new nematicide (a substance active against nematode worms), onnamide F (56 in Scheme 1), which was isolated from Trachycladus laevispirulifer, is helpful in Saccharomyces cerevisiae or baker’s yeast infections. It has an LD99 (dosage required to kill 99% of the fungi population) of 1.4 μg/mL [68].
Fluconazole resistance has been increasing recently, specifically in immunocompromised individuals such as HIV patients prescribed fluconazole prophylactically. Because of that, other antifungal compounds have been screened for efficacy in resistant strains. Geodisterol-3-O-sulfite and 29-demethylgeodisterol-3-O-sulfite, active constituents of Topsentia sp. extracts, have been used in fluconazole-resistant strains. Many Saccharomyces cerevisiae strains can overexpress the MDR1 efflux pump (a pump responsible for pumping out toxic substances such as fluconazole). Hence, these two compounds have been used in reverse [69].

3. Isolated Marine Compounds with Antibacterial Activity

3.1. Ribosomal Peptides—Antimicrobial Peptides

Antimicrobial peptides (AMPs) are large, amphipathic molecules synthesized by ribosomes using 12–45 amino acids, which typically have a tertiary structure (conformation). Due to their broad-spectrum antibacterial properties, they are suited for targeting prokaryotic cell membranes. AMPs are different from the adaptable lymphocyte-based immunity that characterizes higher vertebrates. AMPs that are produced by bacteria are named bacteriocins. In multicellular organisms, AMPs are found on the external surfaces (skin) or within the neutrophils. Marine invertebrates have their AMPs in cells that are similar to neutrophiles called hemocytes. Due to the presence of a good amount of lysine, arginine, and histidine and a low amount of acidic and neutral amino acids, AMPs are highly cationic at physiological pH. In addition to possessing phospholipids with no net charge, this cationic nature gives AMPs selectivity and selective toxicity towards bacterium cells, and their amphipathic nature may help explain their antibacterial effect [5,70,71].
Arenicin-1 is a peptide that is purified from the hemocytes of lugworm Arenicola marina [72]. The linear sequence polypeptide is composed of RWSIVYAYVRVRGVLVRYRRSIW, with a positive 6 net charge (57 in Scheme 1) [73]. Arenicin-1 (1) inhibited Gram-negative bacteria such as Escherichia coli and Proteus mirabilis and Gram-positive bacteria such as Staphylococcus aureus with MIC values of 0.8, 2, and 6 µg/mL, respectively [74,75].
Halocidin is derived from the tunicate Halocynthia aurantium’s hemocytes [76]. The translated peptide consists of a halosidine segment, a single glycine residue, an N-terminal signal peptide, a C-terminal anion extension, and a single glycine residue (58 in Scheme 1) [77]. On the other hand, the modified active peptide comprises two peptides, one with 15 amino acids and the other with 18, joined by a disulfide bond. Halocidin congeners, called Khal, appeared to have potent antibacterial action against methicillin-resistant Staphylococcus aureus (MRSA), vancomycin-resistant Enterococcus (VRE), and segregates of polyresistant Pseudomonas aeruginosa with MICs within the range of 2–4 μg/mL. One of two derivatives showed promising results in an animal model of Listeria monocytogenes infection [78].
Hedistin (59 in Scheme 1) is an amphipathic antibacterial polypeptide obtained from Nereis diversicolor coelomocytes, a marine annelid worm [79]. This peptide had a high MIC of 1–2 μg/mL against Micrococcus luteus and Micrococcus nishinomiyaensis, indicating that it was effective against Gram-positive bacteria. The synthesized peptide was effective against S. aureus, with MIC values ranging from 8 to 15 μg/mL, as well as other Staphylococcus species [80].
Clavanin A (60 in Scheme 1) peptide that isolated from the hemocytes of the Styela clava. [81,82] Clavanin A is rich in phenylalanine amino acids, which can replace it with other hydrophobicity amino acids without losing antibacterial activity [83]. Clavanin A showed potent antibacterial activity against Gram-positive as well as Gram-negative bacteria [84]. Clavanin’s MIC against S. aureus, including methicillin-resistant S. aureus strains, equals 1.4 to 3.8 μg/mL. Three strains of Enterococcus faecium had MIC values ranging from 0.1 to 1.1μg/mL. Strains of E. coli with an MIC value ranging from 0.4 to 2.3 μg/mL and three strains of P. aeruginosa with an MIC value ranging from 0.4 to 0.8 μg/mL were likewise susceptible to clavanin A [85,86].

3.2. Nonribosomal Peptides

Large multifunctional protein complexes called nonribosomal peptide synthetases (NRPSs) produce nonribosomal peptides [87]. The DNA does not encode many nonproteinogenic amino acids in these peptides. The most common nonribosomal peptides with antibacterial activity include bogorol A (61 in Scheme 1), which was isolated from Bacillus laterosporus bacterium. It was confirmed that bogorol A showed activity in opposition to Methicillin-resistant S. aureus with an MIC of 2.5 μg/mL as well as vancomycin-resistant Enterococcus with an MIC of 9 μg/mL [88]. The cationicity of bogorol A has a significant role in targeting the bacterial membrane.
With an MIC value of 2.3 μg/mL, the isolated emericellamide A (62 in Scheme 1) from the marine fungus Emericella sp. exhibited antibacterial activity against S. aureus [89]. Furthermore, thiocoraline (63 in Scheme 1), which was isolated from Actinomycete micromonospora, showed activity against S. aureus, M. luteus, and B. subtilis with MIC values of 0.03–0.05 μg/mL [90,91].
Bleich et al. [92] described YM-266183 (64 in Scheme 1) as an antibacterial peptide. It was produced by Bacillus cereus isolated from a marine sponge Halichondria japonica. The peptide is highly active against Gram-positive bacteria, including S. aureus and Enterococci, with MIC values of 0.68 μg/mL and 0.025 μg/mL, respectively [92,93].

3.3. Polyketides

Bisanthraquinone metabolites BE-43472B and BE-43472A (65,66 in Scheme 1) isolated from a marine streptomycete showed biological activities against clinically derived isolates of E. faecium as well as S. aureus. The most potent activity displayed MIC values of 0.23 and 0.90 µg/mL against a panel (n = 25 each) of clinical MRSA and VRE, respectively [10,94].
The obtained Abyssomicin C AB 18-032 (67 in Scheme 1) from marine actinomycete Verrucosispora sp. [10] exhibits potent antibiotic activity against Gram-positive bacteria, including pathogenic S. aureus strains, with an MIC value of 4 μg/mL [95].
Pestalone (68 in Scheme 1), which is produced by a cultured marine fungus isolated from the brown alga Rosenvingea sp., showed potent antibiotic activity against methicillin-resistant S. aureus as well as vancomycin-resistant bacteria with MIC values of 37 µg/mL and 78 µg/mL, respectively [96,97]. Table 3 lists some other isolated polyketide compounds.

3.4. Alkaloids

An aaptamine—that is, a 1H-benzo[de][1,6]-naphthyridine alkaloid, Isoaaptamine (74 in Scheme 1)—was isolated from the marine sponge Aaptos aaptos and was evaluated as a potent inhibitor with an IC50 value of 3.7 ± 0.2 µg/mL against S. aureus [105].
The bromopyrrole alkaloid nagelamides G (75 in Scheme 1), which was isolated from the Okinawan marine sponge Agelas sp., exhibited antibacterial activity against Gram-positive bacteria M. luteus and B. subtilis with MIC values of 2.08 and 16.7 µg/mL, respectively [106].
With MIC values of 0.625 and 1.25 µg/mL, ambiguine H isonitrile (76 in Scheme 1) obtained from Fischerella sp. showed activity against Scaphirhynchus albus and B. subtilis, respectively. Furthermore, ambiguine I isonitrile (77 in Scheme 1) exhibited antibacterial inhibitory activities against the same bacterial strains with MIC values of 0.078 and 0.312 µg/mL, respectively [107]. Furthermore, cribrostain 6 (78 in Scheme 1), which was isolated from the blue marine sponge Cribrochalina sp., showed an antibacterial activity against the same bacterial strain with MIC values of 16 and 2 µg/mL, respectively [108].
Staphylococcus aureus (methicillin-resistant S. aureus) and Mycobacterium intracellulare were both inhibited with MIC values of 5 and 10 µg/mL, respectively, by batzelladine M (79 in Scheme 1). Additionally, antibacterial inhibitory activities against P. aeruginosa were also shown by batzelladine L (80 in Scheme 1) with MIC values ranging from 0.31 to 20 µg/mL. Batzelladine L and batzelladine M are two polycyclic guanidine alkaloids extracted from the Jamaican sponge Monanchora unguifera [109]. Antibacterial inhibitory activities were also ascertained for lynamicins A-D (81 in Scheme 1), which were isolated from a marine actinomycete, NPS12745, with MIC values ranging between 1.8 and 9.5 µg/mL [110].
Marinopyrroles A and B (82 in Scheme 1), which were both isolated from marine Streptomyces strain bacterium, exhibited strong antibiotic activities against MRSA, with an MIC value range of 0.31–0.61 µg/mL [111].

3.5. Terpenes

Terpenes are a diverse class of natural products composed of repeating isoprene units. They include hemiterpenes (C5), di-unit monoterpenes (C10), tri-unit sesquiterpenes (C15), tetra-unit diterpenes (C20), penta-unit sesterterpenes (C25), and so on. They are created when mono-isoprene is broken down one unit at a time. Skeletal rearrangements, which frequently take place, alter the normal head-to-tail orientation of the isoprene units and add variety to the terpenoid structures [112].
The isolated xeniolide I diterpenes (83 in Scheme 1) from soft coral Xenia novaebrittanniae shows activity against B. subtilis and E. coli with MIC values of 1.00 and 1.25 µg/mL, respectively [113,114].
The acyclic diterpene, crinitol (84 in Scheme 1), which was obtained from Sargassum tortile alga, exhibits antibacterial activity against Propionibacterium acnes, B. subtilis, and Streptococcus mutans with MIC values of 25, 50, and 50 µg/mL, respectively [115].
More terpenes extracted from marine organisms are summarized in Table 4.

4. Conclusions

To sum up everything that has been stated so far, this review shed light on 96 compounds isolated from a variety of marine organisms and showed promising activities against bacteria and fungi. These compounds show great potential for the development of novel antibiotic drugs that can help overcome the problem of antibiotic resistance and have the potential to decrease treatment failures in humans, as many of these compounds showed powerful activities against antibiotic-resistant strains of bacteria and fungi such as methicillin-resistant Staphylococcus aureus, vancomycin-resistant bacteria, and many others. Currently, a significant portion of the novel antifungal and antibacterial drugs in clinical trials was derived from marine species, particularly bacteria as well as sponges. Given the vast number of undiscovered compounds in the oceans, all the new compounds identified might only be the tip of the iceberg, which is quite significant.

Author Contributions

A.M.T., Z.S., M.A., W.J., M.Y., S.A.B., R.K. and L.S. wrote and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors acknowledge support by MDPI, and thank Birzeit University, Al-Quds University, and Basilicata University for supporting the present study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Centers for Disease Control and Prevention. Antibiotic Resistance Threats in the United States, 2019; US Department of Health and Human Services, CDC: Atlanta, GA, USA, 2019.
  2. Okada, B.K.; Seyedsayamdost, M.R. Antibiotic dialogues: Induction of silent biosynthetic gene clusters by exogenous small molecules. FEMS Microbiol. Rev. 2017, 41, 19–33. [Google Scholar] [CrossRef] [Green Version]
  3. Farha, M.A.; Brown, E.D. Strategies for target identification of antimicrobial natural products. Nat. Prod. Rep. 2016, 33, 668–680. [Google Scholar] [CrossRef] [PubMed]
  4. Bengtsson-Palme, J.; Kristiansson, E.; Larsson, D.J. Environmental factors influencing the development and spread of antibiotic resistance. FEMS Microbiol. Rev. 2018, 42, fux053. [Google Scholar] [CrossRef] [PubMed]
  5. Fair, R.J.; Tor, Y. Antibiotics and bacterial resistance in the 21st century. Perspect. Med. Chem. 2014, 6, S14459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Velmurugan, P.; Venil, C.K.; Veera Ravi, A.; Dufossé, L. Marine bacteria are the cell factory to produce bioactive pigments: A prospective pigment source in the ocean. Front. Sustain. Food Syst. 2020, 4, 589655. [Google Scholar] [CrossRef]
  7. Gomes, N.G.; Dasari, R.; Chandra, S.; Kiss, R.; Kornienko, A. Marine invertebrate metabolites with anticancer activities: Solutions to the “supply problem”. Mar. Drugs 2016, 14, 98. [Google Scholar] [CrossRef] [Green Version]
  8. Fan, Y.; Wang, Y.; Liu, P.; Fu, P.; Zhu, T.; Wang, W.; Zhu, W. Indole-diterpenoids with anti-H1N1 activity from the aciduric fungus Penicillium camemberti OUCMDZ-1492. J. Nat. Prod. 2013, 76, 1328–1336. [Google Scholar] [CrossRef]
  9. Gong, K.K.; Tang, X.L.; Zhang, G.; Cheng, C.L.; Zhang, X.W.; Li, P.L.; Li, G.Q. Polyhydroxylated steroids from the South China Sea soft coral Sarcophyton sp. and their cytotoxic and antiviral activities. Mar. Drugs 2013, 11, 4788–4798. [Google Scholar] [CrossRef] [Green Version]
  10. Manivasagan, P.; Venkatesan, J.; Sivakumar, K.; Kim, S.K. Pharmaceutically active secondary metabolites of marine actinobacteria. Microbiol. Res. 2014, 169, 262–278. [Google Scholar] [CrossRef]
  11. Plouguerné, E.; De Souza, L.M.; Sassaki, G.L.; Cavalcanti, J.F.; Romanos, M.T.V.; Da Gama, B.A.; Barreto-Bergter, E. Antiviral sulfoquinovosyldiacylglycerols (SQDGs) from the Brazilian brown seaweed Sargassum vulgare. Mar. Drugs 2013, 11, 4628–4640. [Google Scholar] [CrossRef]
  12. Tajima, H.; Wakimoto, T.; Takada, K.; Ise, Y.; Abe, I. Revised structure of cyclolithistide A, a cyclic depsipeptide from the marine sponge Discodermia japonica. J. Nat. Prod. 2014, 77, 154–158. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, W.; Wang, S.X.; Guan, H.S. The antiviral activities and mechanisms of marine polysaccharides: An overview. Mar. Drugs 2012, 10, 2795–2816. [Google Scholar] [CrossRef] [PubMed]
  14. Malve, H. Exploring the ocean for new drug developments: Marine pharmacology. J. Pharm. Bioallied Sci. 2016, 8, 83–91. [Google Scholar] [CrossRef] [PubMed]
  15. Venkatesan, J.; Anil, S.; Kim, S.K.; Shim, M.S. Marine fish proteins and peptides for cosmeceuticals: A review. Mar. Drugs 2017, 15, 143. [Google Scholar] [CrossRef]
  16. Williams, P.G. Panning for chemical gold: Marine bacteria as a source of new therapeutics. Trends Biotechnol. 2009, 27, 45–52. [Google Scholar] [CrossRef]
  17. Gulder, T.A.; Moore, B.S. Chasing the treasures of the sea—Bacterial marine natural products. Curr. Opin. Microbiol. 2009, 12, 252–260. [Google Scholar] [CrossRef] [Green Version]
  18. Joseph, A. Oceans: Abode of Nutraceuticals, Pharmaceuticals, and Biotoxins. In Investigating Seafloors and Oceans; Joseph, A., Ed.; Candice Janco: Goa, India, 2016; pp. 493–554. ISBN 9780128093573. [Google Scholar]
  19. Tareq, F.S.; Lee, H.S.; Lee, Y.J.; Lee, J.S.; Shin, H.J. Ieodoglucomide C and Ieodoglycolipid, New Glycolipids from a Marine-Derived Bacterium Bacillus licheniformis 09IDYM23. Lipids 2015, 50, 513–519. [Google Scholar] [CrossRef]
  20. Choudhary, A.; Naughton, L.M.; Montánchez, I.; Dobson, A.D.; Rai, D.K. Current status and future prospects of marine natural products (MNPs) as antimicrobials. Mar. Drugs 2017, 15, 272. [Google Scholar] [CrossRef] [PubMed]
  21. El-Gendy, M.M.; El-Bondkly, A.M. Production and genetic improvement of a novel antimycotic agent, saadamycin, against dermatophytes and other clinical fungi from endophytic Streptomyces sp. Hedaya48. J. Ind. Microbiol. Biotechnol. 2010, 37, 831–841. [Google Scholar] [CrossRef] [PubMed]
  22. Gouda, S.; Das, G.; Sen, S.K.; Shin, H.S.; Patra, J.K. Endophytes: A treasure house of bioactive compounds of medicinal importance. Front. Microbiol. 2016, 7, 1538. [Google Scholar] [CrossRef]
  23. Ambavane, V.; Tokdar, P.; Parab, R.; Sreekumar, E.S.; Mahajan, G.B.; Mishra, P.D.; Ranadive, P. Caerulomycin A—An antifungal compound isolated from marine actinomycetes. Adv. Microbiol. 2014, 4, 567–578. [Google Scholar] [CrossRef]
  24. El-Hossary, E.M.; Cheng, C.; Hamed, M.M.; Hamed, A.N.E.S.; Ohlsen, K.; Hentschel, U.; Abdelmohsen, U.R. Antifungal potential of marine natural products. Eur. J. Med. Chem. 2017, 126, 631–651. [Google Scholar] [CrossRef] [PubMed]
  25. Shin, H.J. Natural products from marine fungi. Mar. Drugs 2020, 18, 230. [Google Scholar] [CrossRef] [PubMed]
  26. Kumar, V.; Sarma, V.V.; Thambugala, K.M.; Huang, J.J.; Li, X.Y.; Hao, G.F. Ecology and evolution of marine fungi with their adaptation to climate change. Front. Microbiol. 2021, 12, 719000. [Google Scholar] [CrossRef] [PubMed]
  27. Wu, B.; Oesker, V.; Wiese, J.; Schmaljohann, R.; Imhoff, J.F. Two new antibiotic pyridones produced by a marine fungus, Trichoderma sp. strain MF106. Mar. Drugs 2014, 12, 1208–1219. [Google Scholar] [CrossRef] [Green Version]
  28. Imhoff, J.F. Natural products from marine fungi—Still an underrepresented resource. Mar. Drugs 2016, 14, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Borthwick, A.D. 2, 5-Diketopiperazines: Synthesis, reactions, medicinal chemistry, and bioactive natural products. Chem. Rev. 2012, 112, 3641–3716. [Google Scholar] [CrossRef]
  30. Hu, J.; Li, Z.; Gao, J.; He, H.; Dai, H.; Xia, X.; Liu, C.; Zhang, L.; Song, F. New diketopiperazines from a marine-derived fungus strain Aspergillus versicolor MF180151. Mar. Drugs 2019, 17, 262. [Google Scholar] [CrossRef] [Green Version]
  31. Xu, L.; Meng, W.; Cao, C.; Wang, J.; Shan, W.; Wang, Q. Antibacterial and antifungal compounds from marine fungi. Mar. Drugs 2015, 13, 3479–3513. [Google Scholar] [CrossRef] [Green Version]
  32. Haga, A.; Tamoto, H.; Ishino, M.; Kimura, E.; Sugita, T.; Kinoshita, K.; Koyama, K. Pyridone alkaloids from a marine-derived fungus, Stagonosporopsis cucurbitacearum, and their activities against azole-resistant Candida albicans. J. Nat. Prod. 2013, 76, 750–754. [Google Scholar] [CrossRef]
  33. Sun, C.; Zhang, Z.; Ren, Z.; Yu, L.; Zhou, H.; Han, Y.; Shah, M.; Che, Q.; Zhang, G.; Li, D.; et al. Antibacterial cyclic tripeptides from Antarctica-sponge-derived fungus Aspergillus insulicola HDN151418. Mar. Drugs 2020, 18, 532. [Google Scholar] [CrossRef]
  34. Liu, J.; Gu, B.; Yang, L.; Yang, F.; Lin, H. New anti-inflammatory cyclopeptides from a sponge-derived fungus Aspergillus violaceofuscus. Front. Chem. 2018, 6, 226. [Google Scholar] [CrossRef]
  35. Jin, L.; Quan, C.; Hou, X.; Fan, S. Potential pharmacological resources: Natural bioactive compounds from marine-derived fungi. Mar. Drugs 2016, 14, 76. [Google Scholar] [CrossRef] [Green Version]
  36. Liu, Y.; Mándi, A.; Li, X.M.; Meng, L.H.; Kurtán, T.; Wang, B.G. Peniciadametizine A, a dithiodiketopiperazine with a unique spiro [furan-2,7′-pyrazino [1,2-b][1,2] oxazine] skeleton, and a related analogue, Peniciadametizine B, from the marine sponge-derived fungus Penicillium adametzioides. Mar. Drugs 2015, 13, 3640–3652. [Google Scholar] [CrossRef] [Green Version]
  37. Meng, L.H.; Li, X.M.; Liu, Y.; Wang, B.G. Penicibilaenes A and B, sesquiterpenes with a tricyclo [6.3. 1.01, 5] dodecane skeleton from the marine isolate of Penicillium bilaiae MA-267. Org. Lett. 2014, 16, 6052–6055. [Google Scholar] [CrossRef]
  38. Meng, L.H.; Zhang, P.; Li, X.M.; Wang, B.G. Penicibrocazines A–E, five new sulfide diketopiperazines from the marine-derived endophytic fungus Penicillium brocae. Mar. Drugs 2015, 13, 276–287. [Google Scholar] [CrossRef] [Green Version]
  39. Mehner, T.; Krienitz, L. Encyclopedia of Inland Waters; Likens, G.E., Ed.; Academic Press: Cambridge, MA, USA, 2009; pp. 103–113. [Google Scholar]
  40. Zerrifi, S.E.A.; El Khalloufi, F.; Oudra, B.; Vasconcelos, V. Seaweed bioactive compounds against pathogens and microalgae: Potential uses on pharmacology and harmful algae bloom control. Mar. Drugs 2018, 16, 55. [Google Scholar] [CrossRef] [Green Version]
  41. Alarif, W.M.; Al-Lihaibi, S.S.; Ayyad, S.E.N.; Abdel-Rhman, M.H.; Badria, F.A. Laurene-type sesquiterpenes from the Red Sea red alga Laurencia obtusa as potential antitumor–antimicrobial agents. Eur. J. Med. Chem. 2012, 55, 462–466. [Google Scholar] [CrossRef]
  42. Raeesossadati, M.J.; Ahmadzadeh, H.; McHenry, M.P.; Moheimani, N.R. CO2 bioremediation by microalgae in photobioreactors: Impacts of biomass and CO2 concentrations, light, and temperature. Algal Res. 2014, 6, 78–85. [Google Scholar] [CrossRef]
  43. De Almeida, C.L.F.; Falcão, H.D.S.; Lima, G.R.D.M.; Montenegro, C.D.A.; Lira, N.S.; de Athayde-Filho, P.F.; Rodrigues, L.C.; Souza, M.D.F.V.D.; Barbosa-Filho, J.M.; Batista, L.M. Bioactivities from marine algae of the genus Gracilaria. Int. J. Mol. Sci. 2011, 12, 4550–4573. [Google Scholar] [CrossRef]
  44. Lee, J.C.; Hou, M.F.; Huang, H.W.; Chang, F.R.; Yeh, C.C.; Tang, J.Y.; Chang, H.W. Marine algal natural products with anti-oxidative, anti-inflammatory, and anti-cancer properties. Cancer Cell Int. 2013, 13, 55–57. [Google Scholar] [CrossRef] [Green Version]
  45. Wang, X.H.; Zou, Z.R.; Yi, Y.H.; Han, H.; Li, L.; Pan, M.X. Variegatusides: New non-sulphated triterpene glycosides from the sea cucumber Stichopus variegates semper. Mar. Drugs 2014, 12, 2004–2018. [Google Scholar] [CrossRef] [Green Version]
  46. 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] [Green Version]
  47. Hua, H.A.N.; Ling, L.I.; Yi, Y.H.; Wang, X.H.; Pan, M.X. Triterpene glycosides from sea cucumber Holothuria scabra with cytotoxic activity. Chin. Herb. Med. 2012, 4, 183–188. [Google Scholar]
  48. Wang, Z.N.; Yuan, X. Concurrent effects of hot streak and gas species concentration on aerothermal characteristics in a turbine stage. In Turbo Expo: Power for Land, Sea, and Air; American Society of Mechanical Engineers: New York, NY, USA, 2012; Volume 44748, pp. 1431–1441. [Google Scholar]
  49. Elbandy, M.; Rho, J.R.; Afifi, R. Analysis of saponins as bioactive zoochemicals from the marine functional food sea cucumber Bohadschia cousteaui. Eur. Food Res. Technol. 2014, 238, 937–955. [Google Scholar] [CrossRef]
  50. Bordbar, S.; Anwar, F.; Saari, N. High-value components and bioactives from sea cucumbers for functional foods—A review. Mar. Drugs 2011, 9, 1761–1805. [Google Scholar] [CrossRef] [Green Version]
  51. 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]
  52. Stout, E.P.; Yu, L.C.; Molinski, T.F. Antifungal diterpene alkaloids from the Caribbean sponge Agelas citrina: Unified configurational assignments of agelasidines and agelasines. Eur. J. Org. Chem. 2012, 2012, 5131–5135. [Google Scholar] [CrossRef] [Green Version]
  53. Zhou, X.; Hartman, S.V.; Born, E.J.; Smits, J.P.; Holstein, S.A.; Wiemer, D.F. Triazole-based inhibitors of geranylgeranyltransferase II. Bioorganic Med. Chem. Lett. 2013, 23, 764–766. [Google Scholar] [CrossRef] [Green Version]
  54. Gotsbacher, M.P.; Karuso, P. New antimicrobial bromotyrosine analogues from the sponge Pseudoceratina purpurea and its predator Tylodina corticalis. Mar. Drugs 2015, 13, 1389–1409. [Google Scholar] [CrossRef] [Green Version]
  55. Olatunji, O.J. Bromotyrosines from the sponges Acanthodendrilla sp. and Pseudoceratina cf. Ph.D. Thesis, Prince of Songkla University, Hat Yai, Thailand, 2014. [Google Scholar]
  56. Fuwa, H. Contemporary strategies for the synthesis of tetrahydropyran derivatives: Application to total synthesis of neopeltolide, a marine macrolide natural product. Mar. Drugs 2016, 14, 65. [Google Scholar] [CrossRef] [PubMed]
  57. Harvey, A.L.; Edrada-Ebel, R.; Quinn, R.J. The re-emergence of natural products for drug discovery in the genomics era. Nat. Rev. Drug Discov. 2015, 14, 111–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Youssef, D.T.; Shaala, L.A.; Mohamed, G.A.; Badr, J.M.; Bamanie, F.H.; Ibrahim, S.R. Theonellamide G, a potent antifungal and cytotoxic bicyclic glycopeptide from the Red Sea marine sponge Theonella swinhoei. Mar. Drugs 2014, 12, 1911–1923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Kumar, R.; Subramani, R.; Feussner, K.D.; Aalbersberg, W. Aurantoside K, a new antifungal tetramic acid glycoside from a Fijian marine sponge of the genus Melophlus. Mar. Drugs 2012, 10, 200–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Kikuchi, M.; Nosaka, K.; Akaji, K.; Konno, H. Solid phase total synthesis of callipeltin isolated from marine sponge Latrunculia sp. Tetrahedron Lett. 2011, 52, 3872–3875. [Google Scholar] [CrossRef]
  61. Stierhof, M.; Hansen, K.Ø.; Sharma, M.; Feussner, K.; Subko, K.; Díaz-Rullo, F.F.; Isaksson, J.; Pérez-Victoria, I.; Clarke, D.; Hansen, E.; et al. New cytotoxic callipeltins from the Solomon Island marine sponge Asteropus sp. Tetrahedron 2016, 72, 6929–6934. [Google Scholar] [CrossRef]
  62. El-Amraoui, B.; Biard, J.F.; Fassouane, A. Haliscosamine: A new antifungal sphingosine derivative from the Moroccan marine sponge Haliclona viscosa. Springerplus 2013, 2, 252. [Google Scholar] [CrossRef] [Green Version]
  63. Xu, T.; Feng, Q.; Jacob, M.R.; Avula, B.; Mask, M.M.; Baerson, S.R.; Tripathi, S.K.; Mohammed, R.; Hamann, M.T.; Khan, I.A.; et al. The marine sponge-derived polyketide endoperoxide plakortide F acid mediates its antifungal activity by interfering with calcium homeostasis. Antimicrob. Agents Chemother. 2011, 55, 1611–1621. [Google Scholar] [CrossRef] [Green Version]
  64. Liu, X.F.; Shen, Y.; Yang, F.; Hamann, M.T.; Jiao, W.H.; Zhang, H.J.; Chen, W.-S.; Lin, H.W. Simplexolides A–E and plakorfuran A, six butyrate derived polyketides from the marine sponge Plakortis simplex. Tetrahedron 2012, 68, 4635–4640. [Google Scholar] [CrossRef] [Green Version]
  65. Tripathi, S.K.; Feng, Q.; Liu, L.; Levin, D.E.; Roy, K.K.; Doerksen, R.J.; Baerson, S.R.; Shi, X.; Pan, X.; Xu, W.-H.; et al. Puupehenone, a marine-sponge-derived sesquiterpene quinone, potentiates the antifungal drug caspofungin by disrupting Hsp90 activity and the cell wall integrity pathway. Msphere 2020, 5, e00818-19. [Google Scholar] [CrossRef] [Green Version]
  66. Piao, S.J.; Song, Y.L.; Jiao, W.H.; Yang, F.; Liu, X.F.; Chen, W.S.; Han, B.-N.; Lin, H.W. Hippolachnin A, a new antifungal polyketide from the South China sea sponge Hippospongia lachne. Org. Lett. 2013, 15, 3526–3529. [Google Scholar] [CrossRef]
  67. Arevabini, C.; Crivelenti, Y.D.; de Abreu, M.H.; Bitencourt, T.A.; Santos, M.F.; Berlinck, R.G.; Marins, M. Antifungal activity of metabolites from the marine sponges Amphimedon sp. and Monanchora arbuscula against Aspergillus flavus strains isolated from peanuts (Arachis hypogaea). Nat. Prod. Commun. 2014, 9, 33–36. [Google Scholar] [CrossRef] [Green Version]
  68. Mosey, R.A.; Floreancig, P.E. Isolation, biological activity, synthesis, and medicinal chemistry of the pederin/mycalamide family of natural products. Nat. Prod. Rep. 2012, 29, 980–995. [Google Scholar] [CrossRef]
  69. Abdelmohsen, U.R.; Balasubramanian, S.; Oelschlaeger, T.A.; Grkovic, T.; Pham, N.B.; Quinn, R.J.; Hentschel, U. Potential of marine natural products against drug-resistant fungal, viral, and parasitic infections. Lancet Infect. Dis. 2017, 17, e30–e41. [Google Scholar] [CrossRef]
  70. Martins, A.; Vieira, H.; Gaspar, H.; Santos, S. Marketed marine natural products in the pharmaceutical and cosmeceutical industries: Tips for success. Mar. Drugs 2014, 12, 1066–1101. [Google Scholar] [CrossRef] [Green Version]
  71. Aslam, B.; Wang, W.; Arshad, M.I.; Khurshid, M.; Muzammil, S.; Nisar, M.A.; Alvi, R.F.; Aslam, M.A.; Qamar, M.U.; Salamat, M.K.F.; et al. Antibiotic resistance: A rundown of a global crisis. Infect. Drug Resist. 2018, 11, 1645. [Google Scholar] [CrossRef] [Green Version]
  72. Travkova, O.G.; Moehwald, H.; Brezesinski, G. The interaction of antimicrobial peptides with membranes. Adv. Colloid Interface Sci. 2017, 247, 521–532. [Google Scholar] [CrossRef]
  73. Cho, J.; Lee, D.G. The antimicrobial peptide arenicin-1 promotes generation of reactive oxygen species and induction of apoptosis. Biochim. Et Biophys. Acta (BBA)-Gen. Subj. 2011, 1810, 1246–1251. [Google Scholar] [CrossRef]
  74. Choi, H.; Lee, D.G. Synergistic effect of antimicrobial peptide arenicin-1 in combination with antibiotics against pathogenic bacteria. Res. Microbiol. 2012, 163, 479–486. [Google Scholar] [CrossRef]
  75. Panteleev, P.V.; Bolosov, I.A.; Balandin, S.V.; Ovchinnikova, T.V. Design of antimicrobial peptide arenicin analogs with improved therapeutic indices. J. Pept. Sci. 2015, 21, 105–113. [Google Scholar] [CrossRef] [PubMed]
  76. Han, J.; Jyoti, M.A.; Song, H.Y.; Jang, W.S. Antifungal activity and action mechanism of histatin 5-halocidin hybrid peptides against Candida ssp. PLoS ONE 2016, 11, e0150196. [Google Scholar] [CrossRef] [PubMed]
  77. Shin, S.H.; Lee, Y.S.; Shin, Y.P.; Kim, B.; Kim, M.H.; Chang, H.R.; Jang, W.S.; Lee, I.H. Therapeutic efficacy of halocidin-derived peptide HG1 in a mouse model of Candida albicans oral infection. J. Antimicrob. Chemother. 2013, 68, 1152–1160. [Google Scholar] [CrossRef] [PubMed]
  78. Jeong, J.E.; Kang, S.W.; Shin, Y.K.; Jun, J.C.; Kim, Y.O.; Hur, Y.B.; Kim, J.-H.; Chae, S.-H.; Lee, J.-S.; Choi, I.H.; et al. Comparative analysis of expressed sequence tags (ESTs) between normal group and softness syndrome group in Halocynthia roretzi. Mol. Cell. Toxicol. 2011, 7, 357–365. [Google Scholar] [CrossRef]
  79. Cuvillier-Hot, V.; Boidin-Wichlacz, C.; Tasiemski, A. Polychaetes as annelid models to study ecoimmunology of marine organisms. J. Mar. Sci. Technol. 2014, 22, 9–14. [Google Scholar] [CrossRef]
  80. Rajanbabu, V.; Chen, J.Y.; Wu, J.L. Antimicrobial peptides from marine organisms. In Springer Handbook of Marine Biotechnology; Springer: Berlin/Heidelberg, Germany, 2015; pp. 747–758. [Google Scholar]
  81. de Miranda, J.L.; Oliveira, M.D.; Oliveira, I.S.; Frias, I.A.; Franco, O.L.; Andrade, C.A. A simple nanostructured biosensor based on clavanin A antimicrobial peptide for gram-negative bacteria detection. Biochem. Eng. J. 2017, 124, 108–114. [Google Scholar] [CrossRef]
  82. Silva, O.N.; Fensterseifer, I.C.; Rodrigues, E.A.; Holanda, H.H.; Novaes, N.R.; Cunha, J.P.; Rezende, T.M.B.; Magalhães, K.G.; Moreno, S.E.; Jerônimo, M.S.; et al. Clavanin A improves outcome of complications from different bacterial infections. Antimicrob. Agents Chemother. 2015, 59, 1620–1626. [Google Scholar] [CrossRef] [Green Version]
  83. Andrade, C.A.; Nascimento, J.M.; Oliveira, I.S.; de Oliveira, C.V.; de Melo, C.P.; Franco, O.L.; Oliveira, M.D. Nanostructured sensor based on carbon nanotubes and clavanin A for bacterial detection. Colloids Surf. BBiointerfaces 2015, 135, 833–839. [Google Scholar] [CrossRef]
  84. Nguyen, L.T.; Haney, E.F.; Vogel, H.J. The expanding scope of antimicrobial peptide structures and their modes of action. Trends Biotechnol. 2011, 29, 464–472. [Google Scholar] [CrossRef] [PubMed]
  85. Miller, A.; Matera-Witkiewicz, A.; Mikołajczyk, A.; Wieczorek, R.; Rowinska-Zyrek, M. Chemical “butterfly effect” explaining the coordination chemistry and antimicrobial properties of clavanin complexes. Inorg. Chem. 2021, 60, 12730–12734. [Google Scholar] [CrossRef]
  86. Saúde, A.C.; Ombredane, A.S.; Silva, O.N.; Barbosa, J.A.; Moreno, S.E.; Araujo, A.C.G.; Franco, O.L. Clavanin bacterial sepsis control using a novel methacrylate nanocarrier. Int. J. Nanomed. 2014, 9, 5055–5069. [Google Scholar]
  87. Hur, G.H.; Vickery, C.R.; Burkart, M.D. Explorations of catalytic domains in non-ribosomal peptide synthetase enzymology. Nat. Prod. Rep. 2012, 29, 1074–1098. [Google Scholar] [CrossRef] [Green Version]
  88. Liu, Y.; Ding, S.; Shen, J.; Zhu, K. Nonribosomal antibacterial peptides that target multidrug-resistant bacteria. Nat. Prod. Rep. 2019, 36, 573–592. [Google Scholar] [CrossRef]
  89. Aldholmi, M.; Wilkinson, B.; Ganesan, A. Epigenetic modulation of secondary metabolite profiles in Aspergillus calidoustus and Aspergillus westerdijkiae through histone deacetylase (HDAC) inhibition by vorinostat. J. Antibiot. 2020, 73, 410–413. [Google Scholar] [CrossRef] [PubMed]
  90. Lukassen, M.B.; Saei, W.; Sondergaard, T.E.; Tamminen, A.; Kumar, A.; Kempken, F.; Wiebe, M.G.; Sørensen, J.L. Identification of the scopularide biosynthetic gene cluster in Scopulariopsis brevicaulis. Mar. Drugs 2015, 13, 4331–4343. [Google Scholar] [CrossRef] [Green Version]
  91. Pradhan, T.K.; Reddy, K.M.; Ghosh, S. Total synthesis of emericellamides A and B. TetrahedronAsymmetry 2013, 24, 1042–1051. [Google Scholar] [CrossRef]
  92. Bleich, R.; Watrous, J.D.; Dorrestein, P.C.; Bowers, A.A.; Shank, E.A. Thiopeptide antibiotics stimulate biofilm formation in Bacillus subtilis. Proc. Natl. Acad. Sci. USA 2015, 112, 3086–3091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Indraningrat, A.A.G.; Smidt, H.; Sipkema, D. Bioprospecting sponge-associated microbes for antimicrobial compounds. Mar. Drugs 2016, 14, 87. [Google Scholar] [CrossRef] [PubMed]
  94. Yamashita, Y.; Hirano, Y.; Takada, A.; Takikawa, H.; Suzuki, K. Total Synthesis of Bis-anthraquinone Antibiotic BE-43472B. Synthesis 2018, 50, 2490–2515. [Google Scholar] [CrossRef]
  95. Wang, Q.; Song, F.; Xiao, X.; Huang, P.; Li, L.; Monte, A.; Zhang, L. Abyssomicins from the South China Sea deep-sea sediment Verrucosispora sp.: Natural thioether Michael addition adducts as antitubercular prodrugs. Angew. Chem. 2013, 125, 1269–1272. [Google Scholar] [CrossRef] [Green Version]
  96. Augner, D.; Krut, O.; Slavov, N.; Gerbino, D.C.; Sahl, H.G.; Benting, J.; Nising, C.F.; Hillebrand, S.; Krönke, M.; Schmalz, H.-G. On the antibiotic and antifungal activity of pestalone, pestalachloride A, and structurally related compounds. J. Nat. Prod. 2013, 76, 1519–1522. [Google Scholar] [CrossRef]
  97. Liu, S.; Dai, H.; Makhloufi, G.; Heering, C.; Janiak, C.; Hartmann, R.; Mándi, A.; Kurtán, T.; Müller, W.E.G.; Kassack, M.U.; et al. Cytotoxic 14-membered macrolides from a mangrove-derived endophytic fungus, Pestalotiopsis microspora. J. Nat. Prod. 2016, 79, 2332–2340. [Google Scholar] [CrossRef] [PubMed]
  98. Linares-Otoya, L.; Linares-Otoya, V.; Armas-Mantilla, L.; Blanco-Olano, C.; Crüsemann, M.; Ganoza-Yupanqui, M.L.; Campos-Florian, J.; König, G.M.; Schäberle, T.F. Diversity and antimicrobial potential of predatory bacteria from the Peruvian coastline. Mar. Drugs 2017, 15, 308. [Google Scholar] [CrossRef]
  99. Mayer, A.M.; Rodríguez, A.D.; Berlinck, R.G.; Fusetani, N. Marine pharmacology in 2007–8: Marine compounds with antibacterial, anticoagulant, antifungal, anti-inflammatory, antimalarial, antiprotozoal, antituberculosis, and antiviral activities; affecting the immune and nervous system, and other miscellaneous mechanisms of action. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2011, 153, 191–222. [Google Scholar]
  100. Rateb, M.E.; Ebel, R. Secondary metabolites of fungi from marine habitats. Nat. Prod. Rep. 2011, 28, 290–344. [Google Scholar] [CrossRef]
  101. Wang, T.; Yang, S.; Li, H.; Lu, A.; Wang, Z.; Yao, Y.; Wang, Q. Discovery, structural optimization, and mode of action of essramycin alkaloid and its derivatives as anti-tobacco mosaic virus and anti-phytopathogenic fungus agents. J. Agric. Food Chem. 2019, 68, 471–484. [Google Scholar] [CrossRef] [PubMed]
  102. Wang, H.; Hesek, D.; Lee, M.; Lastochkin, E.; Oliver, A.G.; Chang, M.; Mobashery, S. The natural product essramycin and three of its isomers are devoid of antibacterial activity. J. Nat. Prod. 2016, 79, 1219–1222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Mohapatra, D.K.; Reddy, D.P.; Dash, U.; Yadav, J.S. Total synthesis of Z-isomer of phomolide B. Tetrahedron Lett. 2011, 52, 151–154. [Google Scholar] [CrossRef]
  104. Arunpanichlert, J.; Rukachaisirikul, V.; Sukpondma, Y.; Phongpaichit, S.; Supaphon, O.; Sakayaroj, J. A β-resorcylic macrolide from the seagrass-derived fungus Fusarium sp. PSU-ES73. Arch. Pharmacal Res. 2011, 34, 1633–1637. [Google Scholar] [CrossRef]
  105. Thawabteh, A.; Juma, S.; Bader, M.; Karaman, D.; Scrano, L.; Bufo, S.A.; Karaman, R. The biological activity of natural alkaloids against herbivores, cancerous cells and pathogens. Toxins 2019, 11, 656. [Google Scholar] [CrossRef] [Green Version]
  106. Das, J.; Bhandari, M.; Lovely, C.J. Isolation, Bioactivity, and Synthesis of Nagelamides. Stud. Nat. Prod. Chem. 2016, 50, 341–371. [Google Scholar]
  107. Swain, S.S.; Paidesetty, S.K.; Padhy, R.N. Antibacterial, antifungal and antimycobacterial compounds from cyanobacteria. Biomed. Pharmacother. 2017, 90, 760–776. [Google Scholar] [CrossRef]
  108. Bian, M.; Li, L.; Ding, H. Recent Advances on the Application of Electrocyclic Reactions in Complex Natural Product Synthesis. Synthesis 2017, 49, 4383–4413. [Google Scholar] [CrossRef]
  109. Pessoa, C.; dos Santos, M.F.C.; Berlinck, R.G.S.; Ferreira, P.M.P.; Cavalcanti, B.C. Cytotoxic batzelladine L from the Brazilian marine sponge Monanchora arbuscula. Planta Med. 2013, 79, PK6. [Google Scholar] [CrossRef]
  110. Saurav, K.; Zhang, W.; Saha, S.; Zhang, H.; Li, S.; Zhang, Q.; Wu, Z.; Zhang, G.; Zhu, Y.; Verma, G. In silico molecular docking, preclinical evaluation of spiroindimicins AD, lynamicin A and D isolated from deep marine sea derived Streptomyces sp. SCSIO 03032. Interdiscip. Sci. Comput. Life Sci. 2014, 6, 187–196. [Google Scholar] [CrossRef] [PubMed]
  111. Clive, D.L.; Cheng, P. The marinopyrroles. Tetrahedron 2013, 69, 5067–5078. [Google Scholar] [CrossRef]
  112. Thomson, R.H. (Ed.) The Chemistry of Natural Products; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  113. Thawabteh, A.M.; Thawabteh, A.; Lelario, F.; Bufo, S.A.; Scrano, L. Classification, toxicity and bioactivity of natural diterpenoid alkaloids. Molecules 2021, 26, 4103. [Google Scholar] [CrossRef] [PubMed]
  114. Rocha, J.; Peixe, L.; Gomes, N.C.; Calado, R. Cnidarians as a source of new marine bioactive compounds—An overview of the last decade and future steps for bioprospecting. Mar. Drugs 2011, 9, 1860–1886. [Google Scholar] [CrossRef] [PubMed]
  115. APA American Psychological Association. National Center for Biotechnology Information. Pubchem Compound Summary for CID 12699, N-Nitroso-N-methylurea. Retrieved 24. 2020. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/N-Nitroso-N-methylurea (accessed on 6 January 2023).
  116. Incerti-Pradillos, C.A.; Kabeshov, M.A.; O’Hora, P.S.; Shipilovskikh, S.A.; Rubtsov, A.E.; Drobkova, V.A.; Balandina, S.Y.; Malkov, A.V. Asymmetric Total Synthesis of (−)-Erogorgiaene and Its C-11 Epimer and Investigation of Their Antimycobacterial Activity. Chem. A Eur. J. 2016, 22, 14390–14396. [Google Scholar] [CrossRef] [Green Version]
  117. Pour, P.M.; Behzad, S.; Asgari, S.; Khankandi, H.P.; Farzaei, M.H. Sesterterpenoids. In Recent Advances in Natural Products Analysis; Elsevier: Amsterdam, The Netherlands, 2020; pp. 347–391. [Google Scholar]
  118. Chen, Y.; Zhao, J.; Li, S.; Xu, J. Total synthesis of sesterterpenoids. Nat. Prod. Rep. 2019, 36, 263–288. [Google Scholar] [CrossRef] [PubMed]
  119. McCulloch, M.W.; Haltli, B.; Marchbank, D.H.; Kerr, R.G. Evaluation of pseudopteroxazole and pseudopterosin derivatives against Mycobacterium tuberculosis and other pathogens. Mar. Drugs 2012, 10, 1711–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Ibañez, E.; Herrero, M.; Mendiola, J.A.; Castro-Puyana, M. Extraction and characterization of bioactive compounds with health benefits from marine resources: Macro and micro algae, cyanobacteria, and invertebrates. In Marine Bioactive Compounds; Springer: Berlin/Heidelberg, Germany, 2012; pp. 55–98. [Google Scholar]
  121. Dumas, F.; Kousara, M.; Chen, L.; Wei, L.; Le Bideau, F. Nonhalogenated heterotricyclic sesquiterpenes from marine origin i: Fused systems. Stud. Nat. Prod. Chem. 2017, 52, 269–302. [Google Scholar]
  122. Yende, S.R.; Harle, U.N.; Chaugule, B.B. Therapeutic potential and health benefits of Sargassum species. Pharmacogn. Rev. 2014, 8, 1–7. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Chemical structures of cited compounds that were isolated from marine organisms and that showed antimicrobial activities.
Scheme 1. Chemical structures of cited compounds that were isolated from marine organisms and that showed antimicrobial activities.
Toxins 15 00093 sch001aToxins 15 00093 sch001bToxins 15 00093 sch001cToxins 15 00093 sch001dToxins 15 00093 sch001eToxins 15 00093 sch001fToxins 15 00093 sch001gToxins 15 00093 sch001h
Table 1. Some antifungal compounds isolated from marine bacteria.
Table 1. Some antifungal compounds isolated from marine bacteria.
Isolated CompoundMarine SourcesActivitiesMIC
µg/mL
Basiliskamides A (7)
Basiliskamides B (8)
[22]
Bacillus laterosporusC. albicans
A. fumigatus
1.0–5.0
Mohangamide A (9)
[23]
Streptomyces sp.C. albicans4.14
Haliangicin (10)
[24,25]
Haliangium luteumBotrytis cinerea
Pythium ultimum
Saprolegnia parasitica
3.1
0.4
0.1
Gageostatin A (11)
[26,27]
B. subtilis 109GGC020Rhizoctonia solani
B. cinerea
4
Hassallidin A (12)
[28]
epilithic cyanobacteriaA. fumigatus
C. albicans
4.8
Macrolactins T (13)
Macrolactins B (14)
[29]
Bufo marinusPyricularia oryzae
Alternaria. solani
0.8–4.8
Compounds 714 as shown in Scheme 1
Table 2. Compounds with antifungal activity against C. albicans isolated from sea sponges.
Table 2. Compounds with antifungal activity against C. albicans isolated from sea sponges.
Isolated CompoundMarine SourcesRef.Conc. of Inhibition
Ceratinadins A (39)
Ceratinadins B (40)
Pseudoceratina sp.[54,55]MIC
2 µg/mL
4 µg/mL
Neopeltolide (41)Okinawan sponge[56,57]MIC
0.62 µg/mL
Theonellamide G (43)Theonella swinhoei
in the red sea
[58]IC50
4.49 µM
Aurantoside K (42) *Melophlus sp.[59]MIC
31.25 μg/mL
Callipeltins peptides F (44)
Callipeltins peptides G (45)
Callipeltins peptides H (46)
Callipeltins peptides I (47)
Callipeltins peptides J (48)
Callipeltins peptides K (49)
Latrunculia sp. sponge usually found in Vanuatu islands and South Pacific[60,61]MIC
100 µM
* Aurantoside K has activity against wild type C. albicans with MIC value of 1.95 μg/mL [59]. Compounds 3949 as shown in Scheme 1.
Table 3. Polyketide compounds which have antibacterial activity and which were isolated from marine organisms.
Table 3. Polyketide compounds which have antibacterial activity and which were isolated from marine organisms.
Isolated CompoundMarine SourcesActivitiesMIC
µg/mL
Ariakemicins A (69)
[98,99]
Rapidithrix sp. (marine gliding bacterium)Brevibacterium sp.
S. aureus
B. subtilis
830.4683
Ascochytatin (70)
[100]
marine-derived fungus, Ascochyta sp.B. subtilis4.8
Essramycin (71)
[101,102]
culture broth of the marine Streptomyces sp.E. coli
P. aeruginosa
B. subtilis
S. aureus
M. luteus
1–8
Phomolides B (72)
[103]
culture of Phomopsis sp.E. coli strains
B. subtilis
5–101.25
Sulfoalkylresorcinol (73)
[104]
marine-derived fungus Zygosporium sp.S. aureus12.5
Compounds 6973 as shown in Scheme 1.
Table 4. A list of some terpenes isolated from marine organisms.
Table 4. A list of some terpenes isolated from marine organisms.
Isolated CompoundMarine SourcesTerpenes ClassActivitiesMIC
µg/mL
Erogorgiaene (85) [116]Pseudopterogorgia elisabethaeSerrulatane DiterpenesM. tuberculosis12.5
22-deoxyvariabilin (86), [117,118]Sponge
Ircinia variabilis
SesterterpeneS. aureus
B. subtilis
50
100
Pseudopterosin P and Q, (87), [119]Pseudopterogorgia elisabethaediterpene glycosidesS. pyogenes
S. aureus
E. faecalis
0.8 and 1
2 and 2.3
3.5 and 3.6
Isojaspic acid (88), [120]sponge CacospongiameroditerpeneS. epidermis2.5
(–)-Microcionin, (89) [121]Fasciospongia sp.furanosesquiterpenesM. luteus6
Sargaquinoic acid, (90), [122]Sargassum sagamianumPlastoquinonesS. aureus2
Compounds 8590 as shown in Scheme 1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Thawabteh, A.M.; Swaileh, Z.; Ammar, M.; Jaghama, W.; Yousef, M.; Karaman, R.; A. Bufo, S.; Scrano, L. Antifungal and Antibacterial Activities of Isolated Marine Compounds. Toxins 2023, 15, 93. https://doi.org/10.3390/toxins15020093

AMA Style

Thawabteh AM, Swaileh Z, Ammar M, Jaghama W, Yousef M, Karaman R, A. Bufo S, Scrano L. Antifungal and Antibacterial Activities of Isolated Marine Compounds. Toxins. 2023; 15(2):93. https://doi.org/10.3390/toxins15020093

Chicago/Turabian Style

Thawabteh, Amin Mahmood, Zain Swaileh, Marwa Ammar, Weam Jaghama, Mai Yousef, Rafik Karaman, Sabino A. Bufo, and Laura Scrano. 2023. "Antifungal and Antibacterial Activities of Isolated Marine Compounds" Toxins 15, no. 2: 93. https://doi.org/10.3390/toxins15020093

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