Next Article in Journal
The Effects of Angiotensin II or Angiotensin 1-7 on Rat Pial Microcirculation during Hypoperfusion and Reperfusion Injury: Role of Redox Stress
Previous Article in Journal
AlphaFold-Predicted Structures of KCTD Proteins Unravel Previously Undetected Relationships among the Members of the Family
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Microbial Products and Their Perspective Application as Antimicrobial Agents

1
Department of Botany, School of Basic and Applied Sciences, Central University of Punjab, Bathinda 151401, India
2
Gujarat Pollution Control Board, Gandhinagar 382010, India
3
Department of Chemistry, Umeå University, 90187 Umeå, Sweden
4
Department of Biological Engineering, College of Engineering, Konkuk University, Seoul 05029, Korea
5
ENEA, Italian National Agency for New Technologies, Energy and Sustainable Economic Development, Department of Sustainability-CR Trisaia, SS Jonica 106, km 419 + 500, 75026 Rotondella, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2021, 11(12), 1860; https://doi.org/10.3390/biom11121860
Submission received: 17 October 2021 / Revised: 6 December 2021 / Accepted: 7 December 2021 / Published: 10 December 2021
(This article belongs to the Topic Compounds with Medicinal Value)

Abstract

:
Microorganisms including actinomycetes, archaea, bacteria, fungi, yeast, and microalgae are an auspicious source of vital bioactive compounds. In this review, the existing research regarding antimicrobial molecules from microorganisms is summarized. The potential antimicrobial compounds from actinomycetes, particularly Streptomyces spp.; archaea; fungi including endophytic, filamentous, and marine-derived fungi, mushroom; and microalgae are briefly described. Furthermore, this review briefly summarizes bacteriocins, halocins, sulfolobicin, etc., that target multiple-drug resistant pathogens and considers next-generation antibiotics. This review highlights the possibility of using microorganisms as an antimicrobial resource for biotechnological, nutraceutical, and pharmaceutical applications. However, more investigations are required to isolate, separate, purify, and characterize these bioactive compounds and transfer these primary drugs into clinically approved antibiotics.

1. Introduction

For the last few decades, antibiotics have saved millions of lives, but the prevalence of multidrug resistance (MDR) in microbial strains, nullifying the effects of antibiotics is an expected consequence of antibiotic abuse. The emergence and prevalence of antibiotic-resistant microbial strains remain one of the major health issues of the 21st century, creating pressure on natural microbiota. The ESKAPE pathogens (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter species) are one of the greatest challenges faced by medical practices as many of them are multidrug-resistant isolates [1]. The US Centres for Disease Control and Prevention (CDC) classified the most concerning antimicrobial resistance (AMR) threats, cataloguing carbapenem-resistant P. aeruginosa, Clostridium difficile, and A. baumannii; MDR Neisseria gonorrhoeae and carbapenem- and cephalosporin-resistant Enterobacteriaceae as “urgent” threats [2], requiring urgent measures to deal with the situation. Pendleton et al. [3] provide a contemporary summary and clinically relevant information on the ESKAPE pathogens. In contrast, a detailed description regarding the antimicrobial resistance mechanisms of ESKAPE pathogens was illustrated by Santajit and Indrawattana [1], and can be used as a tool and applied to emerging MDR pathogens. Mulani et al. [4] highlight the use of therapies, including the combination of antibiotics, bacteriophages, antimicrobial peptides, nanomedicines, and photodynamic light therapy to overcome the limitations of individual therapy. These advanced and combinatorial therapies could be used as an alternate solution to combat AMR.
Due to the increased consumption of livestock products in middle-income countries, antimicrobial consumption will increase up to 67%, and up to two-fold in India, Brazil, China, Russia, and South Africa, by 2030 [5]. The current review recapitulates the microbial metabolites, counting growth hormones, pigments, antibiotics, etc., that have become significant sources for life-saving drugs. Most microbial metabolites hold specific antimicrobial potential and act at particular target sites (Figure 1); thereby, they can be an attentive source for biotechnological applications, specifically for pharmaceuticals and nutraceuticals [6]. During the late 1980s, a shift from chemical synthesis in drug discovery from nature to the laboratory bench occurred, resulting in the discovery of approximately 50% natural drugs from 1981 to 2010 [7]. One of them was prodigiosin, an antimicrobial pigment produced by the marine bacterium Vibrio ruber, which induces autolytic activity in the Bacillus subtilis. Similarly, lantibiotics from Gram-positive bacteria were bioengineered to increase their effectiveness against a wide range of bacterial strains and to improve their stability while transmitting through the gastrointestinal (GI) tract making them protease-resistant [8].
Biofilms formed by bacteria are ubiquitous and are a part of their survival mechanisms. Biofilms have been involved in many clinical infections, such as atherosclerosis, pharyngitis, laryngitis, pertussis, bacterial vaginosis, etc. [9]. Although many bacterial strains are responsible for causing infections and diseases, bacterial antimicrobial compounds are reported as antifungal, antiviral, etc. We briefly highlighted the bacteriocins from lactic acid bacteria (LAB) which can disrupt the cell membrane integrity or inhibit cell wall synthesis, protein, and nucleic acid synthesis in pathogens. A recent study by Ting et al. [10] updated the epidemiology of the infectious keratitis (IK), the leading cause of corneal blindness; its causative microorganisms including bacteria, virus, fungi, parasites, and polymicrobial infections are major risk factors associated with the treatment of IK. Antimicrobial compounds such as vinaceuline, bafilomycin, antimycin, and other anti-methicillin-resistant S. aureus (MRSA) compounds synthesized by Streptomyces spp., which act antagonistically against different microbial strains are discussed in the present review. The current review also focuses on the halocins and sulfolobicin reported from archaea in a later section. Furthermore, the antimicrobials reported from endophytic, filamentous and marine-derived fungi along with mushrooms and microalgae are summarized. Microalgae act as a potential source of antimicrobial substances due to the synthesis of indoles, acetogenins, terpenes, phenols, and volatile halogenated hydrocarbons, which are also discussed. Hence, this review documents the potential antimicrobial compounds discovered from all possible microbial resources, including microorganisms inhabiting extreme habitats.
Figure 1. Antibiotics reported from different microorganisms with their target sites. Adapted from: [11].
Figure 1. Antibiotics reported from different microorganisms with their target sites. Adapted from: [11].
Biomolecules 11 01860 g001

2. Bacteria

Bacterial antimicrobial compounds have been used traditionally for numerous reasons, including delaying the spoilage of food material or crops by plant pathogens in agriculture, and extending the shelf life of products in the food industry [12]. Compared to terrestrial bacteria, marine bacteria have many unique secondary metabolites due to their more complex and biologically competitive environment and their unique pressure, temperature, salinity, oxygen, light, and pH conditions. These factors make them a rich source of effective antibiotics.
Many researchers have isolated and identified various antimicrobial compounds from marine bacteria against drug-resistant pathogens [13]. Marinomonas mediterranea, a marine bacterium isolated from the Mediterranean Sea at the Murcia coast has antagonistic activity against Pseudomonas sp. and S. aureus resistance to ceftazidime and meticillin antibiotics, respectively [14]. Ayuningrum et al. [15] isolated isatin from the marine bacterium Pseudoalteromonas rubra TKJD 22 associated with a marine tunicate, which has antibacterial activity against MDR pathogens including MDR E. coli, B. cereus, Micrococcus luteus, and B. megaterium. Ieodoglucomide and ieodoglycolipid isolated from the ethyl acetate extract of a marine-derived Bacillus licheniformis bears antifungal activity against the plant pathogens Colletotrichum acutatum and Botrytis cinerea, along with the human pathogen Candid albicans [16]. Similarly, janthinopolyenemycin A and B polyketides were isolated from the proteobacterium Janthinobacterium spp., strains ZZ145 and ZZ148, respectively, by Anjum et al. [17], who found that they hinder the growth of C. albicans. Schulze et al. [18] utilized a genome-assisted discovery strategy to isolate three macrolactams, lobosamides A, B, and C from Micromonospora sp. RL09-050-HVF-A. Among them, lobosamides A and B have antagonist activity against the microbial agent of African trypanosomiasis i.e., Trypanosoma brucei, whereas, lobosamide C has no bioactivity. Zhang et al. [19] reported Streptoseomycin, a macrolactone from the Streptomyces seoulensis A01, having specific activity against microaerophilic bacteria Helicobacter pylori. Bacicyclin, a cyclic peptide, was isolated from a Bacillus sp. strain BC028 associated with mussel (Mytilus edulis), it was found that it inhibits the growth of Enterococcus faecalis and S. aureus with minimal inhibitory concentration (MIC) values of 8 and 12 µM, respectively. In addition, it is used to design analogs with increased antibiotic efficacy [20].
Bacillus strains from both marine and terrestrial environments are widely known to produce extensive biocontrol metabolites, which include the ribosomally synthesized antimicrobial peptides (bacteriocins) [21], as well as non-ribosomally synthesized peptides (NRPs) and polyketides (PKs) [22].

2.1. Ribosomally Synthesized Antimicrobial Peptides (Bacteriocins) and Bacteriocin-Like Inhibitory Substances (BLIS)

Bacteriocins are antimicrobial ribosomal peptides reported from all major lineages of bacteria and some members of archaea. Gram-negative intestinal bacteria Escherichia coli produces bacteriocidal proteins, colicins, larger than 20 kDa which are antagonistic against zoonotic strains and might establish a defence line against multidrug-resistant strains. [23]. Bacteriocins have attracted increasing attention because of their use as a food preservative and therapeutic antibiotic. Furthermore, they have also received attention because they have a rapid-acting mechanism by forming pores in the membrane of target bacterial cells, even at very low concentrations (Figure 2). The recently reported bacteriocins along with their characteristics are presented in Table 1.
Hoyt et al. [24] isolated the first marine bacteriocin from Vibrio harveyi after screening 795 strains of Vibrio sp. from Galveston Island (Texas, USA). This laid the foundation for multiple studies focused on the identification and biochemical characterization of new bacteriocins and bacteriocin-like compounds. Genera of marine bacteria producing bacteriocins include Aeromonas, Bacillus, Burkholderia, Lactococcus, Pseudomonas, Photobacterium, Pediococcus, Enterococcus, Stenotrophomonas, Carnobacterium, Pseudoalteromonas, Streptomyces, etc. The major difference between marine and terrestrial bacteriocins is that marine bacteriocins are resistant to high and low temperatures, osmotic stress, and various proteolytic enzymes and organic solvents, whereas terrestrial bacteriocins are not [25].
Bacteriocins from lactic acid bacteria (LAB) have gained significant attention due to their food-grade quality and industrial significance. LAB are live microorganisms which when administered in acceptable quantities in a host, trigger a health benefit by promoting and sustaining a strong immune system. Thus, LAB benefits consumers nutritionally and acts as an immunity booster against diseases and infections. LAB and its by-products are generally regarded as safe (GRAS) as a human food component by the U.S. Food and Drug Administration (FDA). Hence it is safer to use LAB bacteriocin to constrain the growth of pathogenic/undesirable bacteria [26]. Lozo et al. [27] isolated the bacteria Lactobacillus paracasei from customarily homemade white-pickled cheese and reported that it produces bacteriocin 217 (Bac217), exhibiting antimicrobial activity against P. aeruginosa, Bacillus cereus, Salmonella sp., and S. aureus.
A study by Drissi et al. [28] suggests that bacteriocins are widespread across the human GI tract, with 317 microbial genomes encoding maximum bacteriocins of class I (44%) as compared to classes II (38.6%) and III (17.3%). Furthermore, they elaborated the bacteriocins produced by gut microbiota, i.e., class I bacteriocins display low antimicrobial activity, whereas maximum class II bacteriocins were reported from bacteria not occurring in the gut. Similarly, Leite et al. [29] described BLIS produced by B. cereus LFB-FIOCRUZ 1640 with activity against Listeria monocytoges and other Bacillus sp. in pineapple pulp and found that it can be used as a potential food bio preservative. Recently, Pircalabioru et al. [30] comprehensively reviewed bacteriocins’ potential as an antimicrobial agent against infections mainly due to resistant pathogens i.e., MRSA. In contrast, Jawan et al. [31] suggest that BLIS from L. lactis Gh1 inhibits the growth of L. monocytogenes and can be used in the food industry as functional foods for the preparation of starter culture and probiotic products. In addition, BLIS from B. subtilis BSC35 inhibits Clostridium perfringens; therefore, it can be used to control C. perfringens in fermented foods [32].
Table 1. List of recently reported Bacteriocins.
Table 1. List of recently reported Bacteriocins.
TypeCharacteristicsExampleProducerMode of ActionReferences
Bacteriocin type ILantibiotics, very small (<5 kDa) peptides containing lanthionine and
β-methyllanthionine
Nisin Z and Q, Enterocin W
Nukacin ISK-1
Lactococcus lactisMembrane permeabilization forming pore[33]
Bacteriocin type IISmall (<10 kDa), non-lanthionine-containing peptides
IIaheat-stable peptides synthesized as a precursor and processed after two glycine residues, antilisterial, bear consensus sequence
YGNGV-C at the N-terminal
Enterocin NKR-5-3C, Enterocin A,
Leucocin A, Munditicin
Pediococcuspentosaceus,
P. Acidilactici and L. sakei
Membrane permeabilization forming pore[34]
IIbTwo-component systems: two different peptides work together and generate an active poration complexLactococcin Q,
Enterocin NKR-5-3AZ,
Enterocin X
L. lactis sub sp. cremoris, L. plantarumMembrane permeabilization forming pore[30,35]
IIcN- and C- termini are covalently linked, generating a circular bacteriocinLactocyclicin Q, Leucocyclicin QL. gasseri,Enterococcusfaecalis, L. garvieaeMembrane permeabilization forming pore[36]
IIdOther class II bacteriocins, including unmodified,
sec-dependent bacteriocins and leaderless, non-pediocin-like bacteriocins
Lacticin Q and Z, Weissellicin Y and M, Leucocin Q and N, Bactofencin A, LsbBL. salivarius, L. lactis
Sub sp. Lactis
[30,37]
Bacteriocin type IIILarge peptides, sensitive to heat L. crispatus, L. helveticus,
E. faecalis
[38]
IIIa27 kDa, heat-labile proteinLysostaphin and enterolysin AS. simulans biovar Staphylolyticus, Enterococcus faecalisCell-wall degradation[39]
IIIb Helveticin J Lactobacillus helveticus Disrupt membrane potential, which causes ATP efflux[40]
Unfortunately, many factors cause a reduction in BLIS antimicrobial activity affecting the efficacy of bacteriocins. Such factors include the advent of bacteriocin-resistant strains, conditions that are destabilizing its biological activities such as oxidation processes, poor solubility, proteases or inactivation by other additives, and pH or temperature. Therefore, it is necessary to develop an antimicrobial system that minimizes these drawbacks and maximizes bacteriocins’ bioprotective potential.
Nisin belongs to type I bacteriocin and is the first antimicrobial peptide from Lactococcus and Streptococcus sp., it has been regarded as GRAS by both the FDA and the WHO [41]. Nisin has been used to inhibit microbial growth in beef, ground beef, sausages, liquid whole eggs, and poultry. It was reported that when nisin was crosslinked to chitosan, minimum inhibitory concentration (MIC) decreased from 48 μg/mL to 40 μg/mL for Staphylococcus aureus ATCC6538. The antimicrobial activity of nisin increased after crosslinking with a lesser concentration of chitosan i.e., the ratio of 200:1, thereby allowing better penetration into the lipid membrane [42]. The antibacterial constancy of nisin was successfully enhanced after its conjugation with gellan. Therefore, this conjugate can be an encouraging biomaterial for wound dressings and transplant coatings [43]. A study revealed the proficiency of nisin in combination with polymyxin in combating P. aeruginosa biofilms and reducing the dose of polymyxin required to interrupt P. aeruginosa biofilms [44]. Polymyxin might facilitate the transfer of nisin to its target. Along with nisin’s synergistic action with polymyxin and clarithromycin against P. aeruginosa and other non–β-lactam antibiotics against MRSA [45] and strains of vancomycin-resistant enterococci [46] were also reported. Webber et al. [47] embedded 0.89 µg cm−2 of positively charged nisin Z within polyelectrolyte multilayers (PEMs) i.e., nine layers of carrageenan (CAR) and chitosan (CS), forming a 4.5 bilayer film with antimicrobial activity against S. aureus and MRSA. Therefore, the antimicrobial potential of CAR/CS multilayers helps to realize its applicability within food, pharmaceutical, and biomedical industries [47]. Although nisin has a broad range of biomedical applications and is used in food bio preservation, further justification of nisin’s practicality and evaluation of its efficacy in biomedical fields will require in vivo and in vitro studies.
The mechanism of action of bacteriocins depends on the bacteriocin receptor molecules used. All bacteriocins from the same class do not follow a similar mode of action. Bacteriocins might utilize the same receptors but their mode of interaction with these receptor molecules and the aftereffect on the target cell could be quite different. For example, as described by Kjos et al. [48], bacteriocins lactococcin A and lactococcin Z share 55.6% similarity in the N-terminal region, both use IIC and IID components of mannose phosphotransferase system (man-PTS) as receptor molecules on target cells. As depicted in Figure 2, lactococcin A results in pore formation and dissipates the cell membrane potential, whereas lactococcin Z kills target cells without following any of these mechanisms. Lozo et al. [49] comprehensively represented the bacteriocins classification-based receptor molecules in target cells according to structural similarity within the same receptor molecule.
Figure 2. Mode of action of bacteriocins. Inhibition of cell wall synthesis: class II bacteriocins (e.g., lactococcin) cross the cell wall and bind with the pore-forming receptor in the mannose-phosphotransferase (man-PTS), resulting in the pore formation in the cell membrane. Pore formation: class I bacteriocins, (e.g., nisin) can follow both mechanisms. Nisin generated pores in the cell membrane resulting in the efflux of ions (K+ and Mg2+), amino acids (glutamic acid, lysin), generating proton motive force dissipation and ultimately causes cell death. Adapted from: [50,51].
Figure 2. Mode of action of bacteriocins. Inhibition of cell wall synthesis: class II bacteriocins (e.g., lactococcin) cross the cell wall and bind with the pore-forming receptor in the mannose-phosphotransferase (man-PTS), resulting in the pore formation in the cell membrane. Pore formation: class I bacteriocins, (e.g., nisin) can follow both mechanisms. Nisin generated pores in the cell membrane resulting in the efflux of ions (K+ and Mg2+), amino acids (glutamic acid, lysin), generating proton motive force dissipation and ultimately causes cell death. Adapted from: [50,51].
Biomolecules 11 01860 g002
Class I bacteriocins are cationic lantibiotics (e.g., nisin) that electrostatically bind with the negatively charged membrane phospholipids II, allowing further interaction of bacteriocin′s hydrophobic domain with the target cytoplasmic membrane (lipid II), thereby preventing the biosynthesis of peptidoglycan [30,52]. Similarly, class III bacteriocins, enterolysin A with N-terminal endopeptidase domain and a C-terminal substrate recognition domain, exhibit antimicrobial activity against streptococci by cleaving the peptidoglycan cross-links between l-alanine and d-glutamic acid of the stem peptide and between l-lysine of the stem peptide and d-aspartic acid of the interpeptide bridge of the target cell [53].

2.2. Non-Ribosomal Synthesized Peptides (NRPs) and Polyketides (PKs)

NRPs and PKs include a range of cyclic, linear, and branched compounds, synthesized by composite enzymes viz. non-ribosomal peptide synthetases (NRPS), polyketide synthetases (PKS), and hybrid of NRPS/PKS, respectively [22,54]. Lipopeptides (LPs) are NRPs produced by Bacillales; LPs have significant antimicrobial activity [55]. LAB is considered the primary producer of ribosomally synthesized antimicrobial peptides, as reviewed by Alvarez-Sieiro et al. [53] and Pircalabioru et al. [30]. However, the classification scheme for antimicrobial compounds produced by Bacillus is not explored in comparison to LAB. Caulier et al. [22] reviewed and updated the antimicrobial metabolites classification from the B. subtilis group based on biosynthetic pathway and chemical nature. Zhao et al. [21] acknowledged 31 types of PKs, NRPs, and NRPS/PKS hybrid synthesized antimicrobials using antiSMASH.

2.3. Lipopeptides (LPs)

LPs occur naturally and are of bacterial origin, contain a hydrophobic long alkyl chain that associates with a hydrophilic polypeptide, and they form a cyclic or linear structure. Traditional LPs including the iturins, surfactins, and fengycins (Table 2) produced from Bacillus species are homologs that differ in length, branching pattern, and saturation of their acyl chain. LPs comprise anionic (e.g., surfactin and daptomycin) or cationic (e.g., colistin and polymixin B) peptide motif, dictating the range of their activity. As demonstrated by Perez et al. [56] Bacillus sp. P5 synthesize LPs iturin A, bacteriocin subtilosin A, and surfactin exhibiting antimicrobial activity against L. monocytogenes and B. cereus, along with the antifungal activity. A study by Kourmentza et al. [57] reported that a mixture of mycosubtilin and mycosubtilin/surfactin LPs inhibit the growth of filamentous fungi Byssochlamys fulva and Paecilomyces variotti, with MICs of 1–16 mg/L and Candida krusei with MIC of 16–64 mg/L.
Table 2. Various types of LPs and their characteristics.
Table 2. Various types of LPs and their characteristics.
TypeCharacteristic FeaturesMolecular WeightChemical StructureProducerApplicabilityReferences
SurfactinCyclic heptapeptide is an antibiotic with seven amino acids i.e., Glu-Leu-Leu-Val-Asp-LeuLeu (ELLVDLL).
A, B, and C types varying according to their amino acid sequences.
~1.03 kDa Biomolecules 11 01860 i001B. subtilis MSH1 and B. amyloliquefaciens ES-2Antimicrobial, antifungal, insecticidal, antimycoplasma, hemolysis, and formation of ion channels in lipid membranes.[58]
IturinContains two major parts: a peptide part composed of 7 amino acid residues (Asn-Tyr-Asn-Gln-Pro-Asn-Ser) and 11-12 carbons hydrophobic tail. Example Iturin A, Bacillomycin D, Bacillomycin L, Mycosubtilin~1.04 kDa Biomolecules 11 01860 i002B. subtilis, B. amyloliquefaciens B128 and B. amyloliquefaciens BUZ-14Antimicrobial and antifungal activities. Disrupt the membrane of yeast cells by increasing the electrical conductance of bimolecular lipid membranes.[59]
FengycinAn array of 10 amino acids with a lactone ring and a ß-hydroxy fatty acid linked to the N-terminus of a decapeptide. Example Plipastatin A and B1463.7 g/mol Biomolecules 11 01860 i003B. subtilisAct as bioagents showing hypocholesterolemic activities, immuno-modulators; antibiotics, antiviral, and antitumor agents; toxins; and enzyme inhibitors[60]
Surfactins, a cyclic heptapeptide that formulates a lactone bridge with β-hydroxy fatty acids, are the most potent biosurfactant. They display an array of activities including hemolytic, antiviral, anti-mycoplasma, and antibacterial [61]. Surfactin WH1 fungin from Bacillus amyloliquefaciens WH1 is an antifungal inhibiting glucan synthase that reduces the synthesis of callose on the fungal cell wall and binds to ATPase on the mitochondrial membrane, ultimately inducing apoptotic markers to stimulate the extracellular apoptotic pathway [62]. Many researchers claim that after inserting into the lipid bilayers, surfactin acts by forming voltage-independent channels in biofilms, distorting the membrane integrity and permeability of ions, i.e., K+ and Ca2+, causing membrane disruption [63].
Iturins are comprised of A, C, D, and E isoforms, bacillomycin D, F and L, and mycosubtilin that inhibit bacterial growth in the same manner as class I and class II bacteriocins [64]. A marine-derived Bacillus velezensis 11-5 produced a cyclic lipopeptide (CLP) iturin A, which is considered an antagonist against Magnaporthe oryzae, a rice pathogen [65]. Fengycin, an anti-fungal lipopeptide, isolated from Bacillus sp. is also called plipastatin. Both iturins and fengycins act as biocontrol agents preventing plant diseases and inhibiting the progression of a wide variety of plant fungal pathogens including Aspergillus flavus, Rhizoctonia solani, Fusarium graminearum, Botritis cinerea, and Penicillium expansum [66]. However, there is no doubt that LPs are a novel class of antibiotics exhibiting a wide range of activities. Therefore, detailed structural and functional knowledge is required to exploit them as potent antimicrobials, feed additives, and drug delivery systems.

3. Actinomycetes

Approximately 75% of the known industrial antibiotics and economically important compounds were obtained from the Streptomyces species [67]. Actinomycetes can synthesize antifungal, antiviral, antitumor, anti-inflammatory, antioxidants, immunosuppressive, plant-growth-promoting, and herbicidal compounds [68]. Among actinomycetes, Streptomyces is the most dominant because of a broad range of bioactive metabolites. Genus Streptomyces is classified into the family Streptomycetaceae based on its morphology and cell wall chemotype. Streptomyces spp. have filamentous hyphae, allowing them to efficiently utilize nutrients in the rhizosphere, enabling them to colonize and carry out a complex life cycle. Streptomyces spp. catabolizes complex molecules and substances, such as cellulose, lignocellulose, xylan, lignin, etc. to produce well-known bioactive compounds. The genus Streptomyces alone contributes approximately 7500 of the 10,000 known compounds from actinomycetes, whereas the other genera including Actinomadura, Micromonospora, Nocardia, Saccharopolyspora, Actinoplanes and Streptosporangium contribute approximately 2500 compounds [69]. Marine or terrestrial actinomycetes utilize enzymes polyketide synthases (PKS) or non-ribosomal peptide synthetases (NRPS) for the synthesis of metabolic bioactive compounds [70].
Pacios et al. [71] reviewed the importance of the Streptomyces genus as a prodigious producer of bioactive metabolites that act as a biological control against phytopathogenic bacteria. Widowati et al. [72] reported a new strain of marine actinomycetes, NPS12745 associated with marine sediment from the coast of San Diego, California, and after using 16S rRNA gene sequencing, NPS12745 was confirmed to be a fruitful strain of genus Marinispora, which produced ample new chlorinated bisindole pyrroles and their derivatives, including chromopyrrolic acid, which was earlier isolated from Chromobacterium violaceum. The first halogenated bisindole derivative was lynamicins A–E, having activity against several Gram-positive and Gram-negative bacteria, i.e., MSSA (methicillin-susceptible Staphylococcus aureus), MRSA, S. epidermidis, and Enterococcus faecalis, signifying a possible cure for nosocomial infections [73]. Siddharth and Vittal [69] isolated Streptomyces sp. S2A from the Gulf of Mannar, which have antagonistic activity against bacterial (Micrococcus luteus, S. epidermidis, Klebsiella pneumoniae, Bacillus cereus, and S. aureus) and fungal (Fusarium moniliforme and Bipolaris maydis) pathogens. A unique prenylated-indole derivative known as 3-acetonylidene-7-prenylindolin-2-one, hybrid isoprenoids, 3-cyanomethyl-6-prenylindole, 7-isoprenylindole-3-carboxylic acid, and 6-isoprenylindole-3-carboxylic acid were extracted from the Streptomyces sp. neau-D50. These antifungal compounds prevent the growth of phytopathogenic fungi Corynespora cassiicola, Phytophthora capsica, Colletotrichum orbiculare, and Fusarium oxysporum [74]. Djinni et al. [75] described Streptomyces sundarbansensis WR1L1S8, an endophyte sequestered from brown algae, yields an innovative anti-MRSA compound, [2-hydroxy-5-((6-hydroxy-4-oxo-4H-pyran-2-yl)methyl)-2-ropylchroman-4-one] beside three already reported polyketides, namely phaeochromycin B, C, and E, which are active against Gram-positive pathogenic MRSA. Sebak et al. [76] isolated Streptomyces sp. MS. 10 from Egyptian soil and reported the presence of saturated fatty acid through 1H NMR spectroscopy, which bears broad-spectrum antimicrobial activity against MRSA. Recently, Qureshi et al. [77] identified compounds, Actinomycin X2 and D, from Streptomyces smyrnaeus UKAQ_23 collected from the mangrove-sediment, with MIC of 1.56–12.5 µg/mL for non-MRSA and 3.125–12.5 µg/mL for MRSA, respectively.
Yang et al. [78] isolate vinaceuline, a cyclopeptide, activity against bacteria, from the broth culture of endophytic Streptomyces sp. YIM64018 allied with Paraboea sinensis. The same team isolated a new benzamide, 2-amino-3, 4-dihydroxy-5-methoxybenzamide in 2015, from Streptomyces YIM67086 that attacks E. coli and Candida albicans (MICs of 64 and 32 μg/mL, respectively). Ding et al. [79] reported that 7, 3’-di-(c,c dimethylallyloxy)-5-hydroxy-40-methoxyflavone, an antifungal compound, from the broth culture of Streptomyces sp. MA-12 obstructs the growth of plant pathogens Penicillium citrinum, Gibberella zeae, and Colletotrichum musae.
Lee et al. [80] isolated 87 actinomycetes species including Streptomyces pluripotens MUSC135T, that inhibit MRSA. This antibacterial metabolite-producing ability was confirmed by PKS (polyketide synthetase) and NRPS (non-ribosomal polyketide synthetase) gene detection process. Streptomyces sp. colonizing on root tissues produce ample antifungal and antibacterial compounds i.e., antimycin A18, phaeochromycin B, C and E, diastaphenazine, 3-acetonylidene-7-prenylindolin-2-one, and staurosporine, some of which are represented in Table 3. Similarly, Jaroszewicz et al. [81] isolated Streptomyces sp. M4_24 and M5_8 strains and identified the presence of dichloranthrabenzoxocinone and 4,10- or 10,12-dichloro-3-O-methylanthrabenzoxocinone, which are putative antimicrobial compounds. A newly discovered lipopeptide NRPS/PKS-derived colibrimycins, from Streptomyces sp. CS147, isolated from Attini ant niche displayed antagonism against virus protease [82]. An endophytic actinomycetes, VITGV01, isolated from a farm tomato plant produced different antibiotics on different media which were active against Gram-positive and Gram-negative bacteria including B. subtilis, S. aureus, E. coli, and Klebsiella pneumoniae [83]. The unique properties of rhizospheric actinomycetes which allow them to produce a diverse range of bioactive metabolites with antagonistic outcomes toward pathogens have led them to be a potent agent ensuring plant health.
Cycloserin, an antibiotic produced by Streptomyces orchidaceus, blocks protein synthesis and is used to treat tuberculosis in conjunction with other drugs [84]. Robertsen and Musiol-Kroll [85] reviewed the actinomycetes-derived polyketide drugs, such as erythromycin A, tetracyclines, rifamycin, tylosin, monensin A, amphotericin B, etc. with antimicrobial activity, including the source of the compounds, their structure, the biosynthetic mechanisms, and mode of action. However, the increasing rate of MDR requires the rediscovery of compounds from potential producers. However, many organisms require special cultivation conditions, therefore, many strategies need to be developed in order to overcome such barriers. Hug et al. [86] described the strategies and innovative methods such as advanced cultivation methods, genomics, metabolomics, and metagenomics-based approaches used to explore the new reservoir of actinomycetes and improve the efficacy of antimicrobial compounds. Hence, it was concluded that Streptomyces spp. can be used as a promising candidate with the potential to be scaled up for industrial production, which could benefit both the agricultural and pharmaceutical industry.
Table 3. Bioactive compounds from endophytic actinomycetes.
Table 3. Bioactive compounds from endophytic actinomycetes.
Endophytic
Actinomycetes
HostBioactive CompoundsStructureBioactivityReferences
Streptomyces sp. YIM64018Paraboea sinensisVinaceuline-Antibacterial activity[87]
Streptomyces sp. neau-D50Glycine max3-acetonylidene-7-prenylindolin-2-one, 7-isoprenylindole-3-carboxylic acid Biomolecules 11 01860 i004Cytotoxic and antifungal activities[74]
Streptomyces sp. YIM56209Drymaria
cordata
Bafilomycin D, B1, B2, C1, C2, C1 amide and C2 amide Biomolecules 11 01860 i005Antibacterial, antifungal, insecticidal, antihelmintic and cytotoxic activity[88]
Streptomyces diastaticus
Sub sp.
ardesiacus
Artemisia
annua
Diastaphenazines-Antibacterial and antifungal activity[89]
Streptomyces sp. YIM67086Dysophylla stellata4-hydroxy-3-methoxybenzoic acid, p-hydroxytruxinic acid-Antifungal activity[78]
Microbispora sp. LGMB259Vochysia
divergens
β-carboline or 1-vinyl-β-carboline-3-carboxylic acid Biomolecules 11 01860 i006Antibacterial, antifungal and anticancer activity[90]
Streptomyces sp. YIM66017Alpinia
oxyphylla
Yangjinhualine A and 2,6-dimethoxy terephthalic acid Biomolecules 11 01860 i007Radical scavenging activity[91]
Streptomyces albidoflavus
07A-01824
Bruguiera gymnorrhizaAntimycin A18 Biomolecules 11 01860 i008Antifungal activity[92]
Streptomyces sp. VITMK1mangrove soilPyrrolopyrazines-Antimicrobial[93]
Streptomyces sp. Diketopiperzines-Anti-H1N1 activity[94]

4. Archaea

Archaeocins, is a proteinaceous antibiotic produced from archaea which mark the chronicled beginning in the series of antimicrobial compounds. The term “archaeocin” was used to differentiate the archaeal peptide and protein-based antibiotics from those produced by bacteria [95]. Only two phylogenetic groups have produced archaeocins (Table 4); one is euryarchaeal producing “halocins”, whereas the other group is crenarchaeal genus Sulfolobus producing “sulfolobicin” [96]. Valera et al. [97] reported halocins, the first proteinaceous antimicrobial compound from halophilic members of the archaeal domain. Archaeal protein VLL-28, from Sulfolobus islandicus, is the first archaeal antimicrobial peptide, possessing a broad-spectrum antibacterial and antifungal activity [98]. Until recently, very few reports were available on the characterization of antimicrobial compounds from archaea. Besse et al. [99] comprehensively reviewed the archaeocins and sulfolobicins antimicrobial peptides ribosomally-synthesized by archaea belonging to the order Halobacteriales and Sulfolobales, respectively. However, until recently halocin A4, G1, R1, H1 [100]; H2 [37]; H3, H5 [97]; H4 [101]; H6 [102]; C8 [103]; S8 [104]; HalR1 [105]; and Sech7a [106] have been considered up to their molecular level, however, their mode of action is not yet clearly understood [107]. Only some workers reported that halocins kill the indicator organisms by altering the cell permeability at membrane level followed by cell lysis. However, to date, only the mode of action mechanism of halocin H6/H7 produced by Haloferax gibbonsii was characterized. HalH6 specifically inhibits Na+/H+ antiporter and proton flux ultimately causing cell lysis and death [108].
H1 and H4 are proteinaceous halocins of roughly 30–40 kDa [109], whereas C8, H6, H7, R1, U1, and S8 are microhalocins which are smaller than 10kDa. Microhalocins are more vigorous than proteinaceous halocins since they are resistant to varying temperature, salinity, exposure to organic solvents, acids, and bases [109]. Halocins have wide-ranging activity against haloarchaea and members of the family Halobacteriaceae [110]. Mainly halocin production is prompted during the progression between exponential and stationary phases, with H1 being an exception, produced during the exponential phase of the growth cycle [111]. Recently, Sahli et al. [112] screened 81 halophilic strains collected from solar salterns of Algeria’s northern coast for the production of antimicrobial compounds, through partial 16S rRNA gene sequencing, these strains were recognized to belong to the Haloferax (Hfx) sp.
Table 4. Archaeocins reported from halobacteria.
Table 4. Archaeocins reported from halobacteria.
HalocinProducersSize (kDa)OriginActive AgainstMode of ActionReferences
HalH1Haloferaxmediterranei Xia331Solar salterns,
Alicante, Spain
Members of the
Halobacteriales
Alter membrane
permeability
[99]
HalH4Hfx. mediterranei R434.9Solar salterns, TunisiaMembers of the Halobacteriales,
Strains of Sulfolobus sp.
Alter macromolecular synthesis, cell wall conformation, and Na+/H+ antiport inhibitor[113]
HalH6Hfx. gibbonsii Ma2.3932Solar salterns,
Alicante, Spain
Members of the
Halobacteriales
Alter intracellular osmotic balance, Na +/H+ antiport inhibitor[99]
HalS8Haloarchaeal strain S8a, Halobacterium salinarum strain ETD53.58Great Salt Lake, (Utah, United States)Halobacterium salinarum NRC817, Hbt. sp. strain GRB and Hfx. gibbonsiiND[104,114]
HalC8Natrinema sp. AS70927.4Chaidan Salt Lake in Qinghai province, China ND[111,115]
HalR1Hbt. salinarum GN1013.8Guerrero Negro, MexicoMembers of the Halobacteriales,
Strains of Sulfolobus sp.,
Methanosarcina thermophile
ND[37,99]
SulfolobicinsSulfolobus
Islandicus HEN2/2
33.9 proprotein), 3.6 (mature)Solfataric fields, IcelandStrains of Sulfolobus sp.ND[116]
Note: ND: Note Detected or Not Reported.
Roscetto et al. [117] reported that VLL-28 damages the cell wall of Candida albicans and C. parapsilosis by binding to their cell surface. Kumar and Tiwari [118] purified halocin HA1 from Haloferax larsenii HA1 and HA3 from H. larsenii HA3; both were halocidal against H. larsenii HA10, instigating cellular distortion, releasing cell contents, and finally causing cell death. Because of these properties, it can be used for the preservation of leather hides and salted foods in the leather and food industries. Ghanmi et al. [114] isolated Halobacterium salinarum ETD5, H. salinarum ETD8, and Haloterrigena thermotolerans SS1R12 of the order Halobacteriales and reported that their antimicrobial activity is due to the production of a halocin, HalS8, a hydrophobic peptide. Quadri et al. [119] isolated archeal strain, Natrinema gari, the common producer of antimicrobial compounds, which after partial purification and characterization resembles the microhalocin HalC8. Besse et al. [115] confirmed that Natrinema sp. synthesizes Halocin C8, a 7.4 kDa peptide involving the genes halC8.
Although many studies characterized the synthesis of halocins, the research concerning their structure and mode of action is still far behind in comparison to the antibiotics produced by other domains. Nowadays, when archaea gain more attention, it becomes necessary to explore their metabolites’, biosynthetic pathways, mode of action, etc., using the latest available technology.

5. Fungi

In 1929, Alexander Fleming discovered the mold juice ‘Penicillin’ from Penicillium notatum fungus with an antibacterial activity [120]. Afterwards, several researchers started to search for a better strain to attain higher yields in easier growth conditions. After extensive research, Penicillium chrysogenum strains were considered for the commercial production of penicillin [121]. Revilla reported in 1986 the formation of the intermediate isopenicillin N in the course of penicillin G production in P. chrysogenum cultures [122], thereafter the formation of isopenicillin N/penicillin N and its late transformation to cephalosporin C in Acremonium chrysogenum [123]. Cephalosporins, a known antimicrobial agent, were purified from a marine fungus, Cephalosporium acremonium [124]. Recently, Li et al. [125] reported that pneumocandins, a lipohexapeptides of the echinocandin family, were produced by wild-type fungi Glarea lozoyensis and Pezicula (Cryptosporiopsis) species. Pneumocandins non-competitively bind to a catalytic unit of β-1,3-glucan synthase, resulting in osmotic uncertainty and cell lysis.

5.1. Endophytic Fungi

Huang et al. [126] discovered ten-membered lactones from endophytic fungus Phomopsis sp. YM 311483, with antifungal activity against Aspergillus niger, Fusarium, and Botrytis cinere. Endophytic Fusarium sp. from Selaginella pollescens collected from the Guanacaste conservation area of Costa Rica inhibit C. albicans [127]. The number of antimicrobial compounds were reported from the endophytic fungi, some of which are listed in Table 5.

5.2. Marine-Derived Fungi

In 2015 Meng et al. [133] discovered pyranonigrin F from fungus Penicillium brocae MA-231 allied with the Avicennia marina, a marine mangrove plant. Pyranonigrin F inhibits S. aureus (Gram-positive), Vibrio harveyi, and Vibrio parahemolyticus (Gram-negative bacteria), with considerably lower MIC values in comparison to the positive control (chloromycetin). Likewise, it is active against plant fungal pathogens Alternaria brassicae and Colletotrichum gloeosprioides, with improved MIC values compared to the positive control (bleomycin). Wu et al. [134] discovered Lindgomycin from Lindgomyces strains LF327 and KF970, reported from a sponge in the Baltic Sea, Germany, and Antarctica, respectively. Lindgomycin displayed antimicrobial activity against S. aureus, S. epidermidis, and methicillin-resistant S. epidermidis (MRSE). However, the inhibiting potential was two times less than the positive control chloramphenicol. It also constrains plant pathogenic bacterium Xanthomonas campestris. There is a never-ending list of antimicrobial compounds from marine fungi; a few of which are listed in Table 6, which displays their host, producer species, and bioactivity.
Peniciadametizine A and Peniciadametizine B derivative of thiolated diketopiperazine was isolated from sponges-associated Penicillium sp. viz. Penicillium adametzioides AS-53 and Penicillium sp. LS54, respectively. Both derivatives inhibit A. brassicae (pathogenic fungus) with a MIC of 4.0 µg/mL and 32.0 μg/mL, respectively [138]. Communol A, G, and F extracted from P. commune 518 displayed antibacterial activities against E. coli with MIC values of 4.1, 23.8, and 6.4µM, respectively, and also against E. aerogenes [136]. Pyrrospirones were produced by marine-derived fungus Penicillium sp. ZZ380, isolated from Pachygrapsus crassipes which is a wild crab found on the seaside rocks of Putuo Mountain (Zhoushan, China). Pyrrospirones C-F, H, and I inhibit MRSA and E. coli having MIC values of 2.0–19.0 μg/mL [155]. Song et al. [156], following the previous lead, separated penicipyrrodiether A from a cultured marine fungal strain Penicillium sp. ZZ380 which inhibits E. coli and S. aureus with MIC of 34.0 and 5.0 μg/mL, respectively. These laboratory studies need to be directed toward developing the efficiency and effectiveness of isolated compounds that could benefit society in the long-term.

5.3. Mushrooms

Mushrooms are colonizing fungi belonging to division Eumycota and subdivision Basidiomycetes, characterized by the formation of basidiospores. Most of these macrofungi are edible, with culinary, nutritional, and medicinal characteristics, but many of them are not palatable or are poisonous [157]. Besides the nutritional and culinary properties, their antimicrobial activities attracted researchers seeking natural solutions to deal with the urgent requirements of food safety. Mushrooms have been publicly consumed for thousands of years due to their medicinal and nutritional properties. Secondary metabolites and extracts from mushrooms have recently attained considerable attention due to their anti-cancer, antioxidant, anti-inflammatory, antimicrobial, antidiabetic, and immunomodulatory properties. Approximately 1069 mushroom species have been consumed by people [158]. To date, numerous antimicrobial peptides have been acknowledged from mushrooms. Plectasin (endogenous peptide antibiotics), an antibacterial peptide, was extracted from Pseudoplectania nigrella. Mygind et al. [159] demonstrated the potent activity of recombinant plectasin against some Gram-positive Streptococcus pneumoniae. Wong et al. [160] described an antifungal peptide, cordymin isolated from medicinal mushroom Cordyceps militaris, which repressed mycelial growth of Bipolaris maydis, Mycosphaerella arachidicola, Candida albicans, and Rhizoctonia solani with IC50 values of 50 μM, 10 μM, 0.75 mM, and 80 μM, respectively. They also reported the remarkable pH stability (pH 6–13), thermostability (100 °C), and metal ion stability (10 mM Mg2+ and 10 mM Zn2+) of cordymin. An investigation by Gebreyohannes et al. [161] revealed that chloroform, ethanol, and hot water extract of Auricularia and Termitomyces sp. promisingly inhibited E. coli, K. pneumoniae, C. parapsilosis, and S. aureus. Poompouang and Suksomtip, [162] isolated an antifungal compound of 17 kDa from fruiting bodies of edible mushroom, Lentinus squarrosulus, inhibiting Trichophyton mentagrophytes and T. rubrum, a human fungal pathogen. More recently, Irshad et al. [163] comprehensively reviewed the synthesis and mode of action of polysaccharides silver nanoparticles (NPs) from Pleurotus mushroom. They characterized the NPs through ultraviolet-visible (UV–Vis), Fourier transformation infrared spectroscopy (FT-IR), scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), transmission electron microscopy (TEM), etc., and disclosed their promising antimicrobial efficiency. However, further studies are required in order to fortify and test these extracts and NPs against human and plant pathogenic microorganisms coupled with the purification and characterization of the compounds from mushrooms.
Hamamoto et al. [164] screened the volatile compound, 3,4-dichloro-4-methoxy benzaldehyde (DCMB) from mycelia of Porostereum spadiceum. It remarkably inhibited the plant-pathogenic bacteria (Clavibacter michiganensis and Ralstonia solanacearum) and inhibited the conidial germination of plant-pathogenic fungi (Alternaria brassicicola and Colletotrichum orbiculare). However, further studies are essential to investigate its effects on plant-pathogens in vivo. Subrata et al. [165] reported that edible wild mushrooms’ methanolic extracts exhibited different levels of antimicrobial activities. A recent study by Sevindi [166] analysed the phenolic content of the wild edible mushroom Melanoleuca melaleuca (Pers.) Murrill had antimicrobial activities inhibiting Gram-negative E. coli, Pseudomonas aeruginosa, and Acinetobacter baumannii.

5.4. Filamentous Fungi

Yeasts mainly occur in milk, meat, food, and products such as fruit, yogurt, jams, sausage, and cheeses. Generally, antimicrobial compounds produced from yeasts inhibit the evolution/growth of pathogenic organisms (bacteria or molds) in food products. Some classes of yeasts secrete toxins, thereby naming them killer yeasts. Killer yeasts naturally occur in rotten vegetables and fruits and constrain the growth of other yeast strains and also inhibit microbial growth [167]. Saccharomyces cerevisiae (baker’s yeast), unicellular yeast, is the most widely studied microorganisms involved in many biotechnological practices because of its good fermentation capacity [168]. The inhibitory mechanism of S. cerevisiae killer strains was discovered in 1963 by Bevan and co-worker’s, and the phenomenon is related to the secretion of a protein toxin, k1, and k28 from the host that kills sensitive target pathogenic cells in a receptor-mediated approach without direct cell-to-cell contact [169]. Other genera producing killer toxins include Cryptococcus, Candida, Kluyveromyces, Williopsis, Pichia, Debaromyces, and Zygosaccharomyces [170]. The anti-bacterial capability of S. cerevisiae is attributed to:
(a)
Secretion of inhibitory proteins,
(b)
Production of extracellular protease,
(c)
Stimulation of immunoglobulin A,
(d)
Procurement and eradication of secreted toxins,
(e)
Killer toxins, sulfur dioxide, etc.
Sequential re-pitching of Saccharomyces biomass is a common process during brewing. Therefore, yeast is reused many times before its final dumping [171]. Hence, yeast develops an adaptive response against oxidative stress like that of human cells, leading to the accumulation of vitamins (B6 and B12) and minerals (enzyme co-factors including zinc, manganese, and copper) in the yeast cell. Phenolic compounds are also adsorbed by Saccharomyces from the exterior medium, which increases the phenolic content and antioxidant activity within yeast cells [172]. Efficient means are required to disrupt yeast cell walls and separate the products of interest, which are further used for food applications. However, increasing consumer’ fears regarding the toxicity of killer yeast strains present in food and milk products constitutes a direct risk to public health.

6. Microalgae

The antimicrobial activity of microalgae is due to the presence of phytochemicals, including indoles, acetogenins, terpenes, fatty acids, phenols, and volatile halogenated hydrocarbons (Table 7) [173]. Moreno et al. [174] reported that Chaetoceros muelleri extracts’ antimicrobial activity is due to their lipid configuration, whereas Dunaliella salina’s is attributed to the presence of β-cyclocitral, α and β-ionone, phytol, and neophytadiene. In natural environmental conditions, microalgal cells release fatty acids against predators and pathogenic bacteria. It is elucidated that these fatty acids act on bacterial cell membranes causing cell seepage, a decline in nutrient intake, and reduced cellular respiration, ultimately resulting in cell death [175].
Chlorellin, the first antibacterial compound from a microalga Chlorella, is composed of a mixture of fatty acid and was isolated by Pratt et al. [183]. Chlorellin was reported to inhibit the activity of both Gram-positive and Gram-negative bacteria. Arthrospira platensis, commercially known as Spirulina had MICs of 0.20% for L. innocua and P. fluorescens, and an MIC of 0.25% for Serratia, whereas minimal bactericidal concentration (MBC) value was 0.30% for all of these species [184]. HPTLC screening and GC–MS analyses were conducted to detect and screen the macroalgae’s antimicrobial compounds. Peptides, namely AQ-1756, AQ-1757, and AQ-1766 identified from Tetraselmis suecica exhibited an antibacterial activity resulting in decreasing cell viability (human embryonic kidney cells) (HEK293) up to 75% after 24 h of treatment. AQ-1766 was more active against Gram-positive than Gram-negative bacteria, with MBC values between 40 and 50 µM [185]. Mendiola et al. [186] demonstrated that lipid fractions obtained from Chaetoceros muelleri by the supercritical CO2 method have antibacterial activity against Staphyloccocus aureus and E. coli. In contrast, extraction via classic methods using hexane, dichloromethane, and methanol solvent did not result in any activity against E. coli. However, these studies were unable to elaborate the mode of action of these antibacterial compounds.
Axenic microalgae co-culture can produce compounds with potent activity against pathogenic bacteria. Kokou et al. [187] reported that axenic cultures of Tetraselmis chui, Chlorella minutissima, Isochrysis sp. and Nannochloropsis sp. inhibit Vibrio harveyi. The potent activity of microalgal compounds against microorganisms requires further development in the search for drugs and food preservatives. Therefore, the exploitation in medicine deserves to be further investigated.

7. Discussion and Future Prospects

One of the significant challenges healthcare services face worldwide is the excessive use of antibiotics in medicine and food production, leading to microbiome disruption. With the outburst of antimicrobial resistance strains, there is a continuous decline in the antimicrobial drug pipeline, and it has become necessary to discover and develop new agents/metabolites to tackle antibiotic resistance. Novel compounds that target microbial resistance can be developed to regulate the huge risk posed by multi-drug resistance. However, the production cost needs to be reduced by isolating these compounds from natural sources such as microorganisms and then synthesizing them or modifying derivative compounds. Along with this, further research into their toxicity against human cells, their mode of action, in vivo effects, and their interactions with commonly available antibiotics must be conducted. After the discovery of penicillin, many drug discoveries from microbial sources were reported. In addition, the advancement of techniques such as genetic engineering during the 1970s opened the door to the ignored source, i.e., microbial metabolites [188].
Ample research is being conducted to search for novel antimicrobial agents from biological sources, including bacteria, actinomycetes, fungi, yeast, etc. Table 8 depicts selected commercially available antimicrobial products alongside their uses.
LAB producing bacteriocins are a promising candidate for the food industry as they help to extend shelf life and safeguarde consumers’ health. Actinomycetes, particularly Streptomyces spp., exhibited effective antagonistic activity and played a significant role in drug discovery and development.
In the ongoing rearch for novel antibiotics, archaeocins have generally been overlooked, and further studies on purifications and characterizations of archaeocins and sulfolobicins are in progress, resulting in the economical production of bioactive compounds for pharmaceutical applications. It is desirable to expand our understanding of the effectiveness and use of other naturally occurring ribosomally-synthesized peptide antimicrobials to understand their implantation and survival strategies, and to quantitatively estimate their efficacy for future applications in the pharmaceutical and health care sectors.
Fungi synthesized small quantities of bioactive compounds in response to explicit environmental conditions which cannot be reproduced easily in the laboratory. Therefore, to develop new antimicrobial drugs from these fungal metabolites, commercial-scale synthesis must be accomplished potentially through strain improvement, optimizing growth conditions, and incorporating techniques, such as metabolomics, genomics, and pathway engineering. Endophytic, filamentous, and marine-derived fungi also offer a suitable substitute against toxic, ineffective, and expensive antimicrobial drugs because they act as a warehouse filled with bioactive compounds with endless potential for biological properties. Antimicrobials, isolated from mushrooms, act as essential substitutes to synthetic drugs and preservatives, whose protection and influence on the health of humans, animals, and food are still uncertain. Although there are many edible mushrooms, the mushroom species identified have antimicrobial properties which are quite small. The current review demonstrates potent bioactive substances with antimicrobial activities from edible mushrooms. Hence, they must not be considered only as a culinary delicacy, but also taken as therapeutic agents. However, methods for isolation, purification, identification, and characterization of antimicrobial compounds from mushrooms need to be developed.
Microalgae are a promising source of high-value products, and large-scale screening programs have been conducted to discover the antimicrobial potential of microalgal extracts against pathogenic and foodborne organisms. However, major antibacterial and antifungal activity reports were predominantly from the Chlorella sp. and Chlamydomonas sp. Many hurdles exist in developing the marine product, including resource supply issues, large-scale production, production cost, and determination of the efficacy target. These obstacles must be bypassed by optimizing mass culturing conditions, utilizing biotechnological techniques, etc. Along with these measures, extensive clinical trials will be needed to determine the in vivo fortune of antimicrobials from microbial extracts on mammalian cells.
Therefore, developing and using robust screening and high-throughput methods will be essential to study their antimicrobial activity, thereby increasing the chances of discovering and identifying new antibiotic molecules. To achieve this goal, the experimental design must include all possible variables, such as recovering both intra- and extracellular extracts produced by microorganisms under variable growth conditions, utilizing potential inducers of antimicrobial activity, and testing these compounds against a more significant number of targets. In recent years, nanoencapsulation has gained much attention. It is a technique used for formulating and stacking a compound in nanosized carriers that can carry and deliver the molecules to the targeted site. Nanoencapsulation allows the conservation and controlled release of bioactive compounds, followed by resistance to pH and temperature variations, lesser product contamination, economic viability, and stability. Chromatographic separation techniques were used recently for purifying antimicrobials, followed by their chemical characterization using spectroscopic techniques, and response surface methodology (RSM) to predict the yield of the crude antimicrobial extract. Detailed functional and structural knowledge would explain antimicrobials’ mode of action and performance at cellular and molecular levels. However, for this, a better understanding of the structure, function and, existing mode of action of newly identified antimicrobials is required.

8. Conclusions

In conclusion, microorganisms are probable sources of bioactive compounds, and this review has explored of microorganisms’ aptitudes to deliver novel bioactive compounds with potential pharmaceutical and nutraceutical applications. Microorganisms have fascinated many researchers due to the ease of growth and understanding of their chemical interactions. The development of new biotechnological tools and techniques contributes to the discovery of next-generation antimicrobial compounds. The application of microorganisms in human foods, animal feeds, agriculture, and an increased market demand motivates the research and development of novel antibiotics and preservatives. Furthermore, the molecular docking and structural analysis approaches can design potent pathogen-specific antimicrobial agents that exhibit lesser toxicity, higher selectivity, and biodegradability. Therefore, exploiting microbial biodiversity and biotechnological potential to discover novel bioactive compounds to treat life-threatening diseases and safeguard human health.

Author Contributions

Acquisition of data, analysis and original draft and table preparation, A.R.; formal analysis, reviewing and figure preparation, K.C.S. and S.V.; review and editing, F.B., S.M., S.K.B., N.S. and C.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research has received funding from the Swedish Research council FORMAS (grant no. 2019-00492).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A.R. acknowledges the Indian Council of Medical Research (ICMR), New Delhi, India for the SRF. K.C.S. gratefully acknowledges the Council of Scientific and Industrial Research (CSIR), New Delhi, India, for the financial support towards Ph.D. Both A.R. and K.C.S. Acknowledges the Central University of Punjab, for administrative and technical support. The author also would like to acknowledge the KU Research Professor Program of Konkuk University, Seoul, South Korea. S.M. and C.F. would like to acknowledge the Swedish Energy Agency (Grant no. 2018-017772, project: 48007-1), Vinnova (2017-03301), the NordForsk NCoE program “NordAqua” (Project no. 82845) and Umeå University, Sweden. Diagrams were created with BioRender.com and molecular structures were retrieved from PubChem (https://pubchem.ncbi.nlm.nih.gov/, accessed on 17 October 2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Santajit, S.; Indrawattana, N. Mechanisms of antimicrobial resistance in ESKAPE pathogens. Biomed Res. Int. 2016, 2016, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Prevention, C.f.D.C.a. Antibiotic Resistance Threats in the United States. Available online: https://www.cdc.gov/drugresistance/pdf/threats-report/2019-ar-threats-report-508.pd (accessed on 29 January 2021).
  3. Pendleton, J.N.; Gorman, S.P.; Gilmore, B.F. Clinical relevance of the ESKAPE pathogens. Expert. Rev. Anti. Infect. Ther. 2013, 11, 297–308. [Google Scholar] [CrossRef]
  4. Mulani, M.S.; Kamble, E.E.; Kumkar, S.N.; Tawre, M.S.; Pardesi, K.R. Emerging strategies to combat ESKAPE pathogens in the era of antimicrobial resistance: A review. Front. Microbiol. 2019, 10, 1–24. [Google Scholar] [CrossRef]
  5. Van Boeckel, T.P.; Brower, C.; Gilbert, M.; Grenfell, B.T.; Levin, S.A.; Robinson, T.P.; Teillant, A.; Laxminarayan, R. Global trends in antimicrobial use in food animals. Proc. Natl. Acad. Sci. USA 2015, 112, 5649–5654. [Google Scholar] [CrossRef] [Green Version]
  6. Rani, A.; Saini, K.C.; Bast, F.; Mehariya, S.; Bhatia, S.K.; Lavecchia, R.; Zuorro, A. Microorganisms: A potential source of bioactive molecules for antioxidant applications. Molecules 2021, 26, 1142. [Google Scholar] [CrossRef] [PubMed]
  7. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the 30 years from 1981 to 2010. J. Nat. Prod. 2012, 75, 311–335. [Google Scholar] [CrossRef] [Green Version]
  8. Mazzoli, R.; Riedel, K.; Pessione, E. Editorial: Bioactive compounds from microbes. Front. Microbiol. 2017, 8, 392. [Google Scholar] [CrossRef] [PubMed]
  9. Vestby, L.K.; Grønseth, T.; Simm, R.; Nesse, L.L. Bacterial biofilm and its role in the pathogenesis of disease. Antibiotics 2020, 9, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Ting, D.S.J.; Ho, C.S.; Deshmukh, R.; Said, D.G.; Dua, H.S. Infectious keratitis: An update on epidemiology, causative microorganisms, risk factors, and antimicrobial resistance. Eye 2021, 35, 1084–1101. [Google Scholar] [CrossRef] [PubMed]
  11. Sanseverino, I.; Navarro Cuenca, A.; Loos, R.; Marinov, D.; Lettieri, T. State of the Art on the Contribution of Water to Antimicrobial Resistance; Publications Office of the European Union: Luxembourg, 2018. [Google Scholar] [CrossRef]
  12. Sumi, C.D.; Yang, B.W.; Yeo, I.-C.; Hahm, Y.T. Antimicrobial peptides of the genus Bacillus: A new era for antibiotics. Can. J. Microbiol. 2015, 61, 93–103. [Google Scholar] [CrossRef] [PubMed]
  13. Hu, Y.; Chen, J.; Hu, G.; Yu, J.; Zhu, X.; Lin, Y.; Chen, S.; Yuan, J. Statistical research on the bioactivity of new marine natural products discovered during the 28 years from 1985 to 2012. Mar. Drugs 2015, 13, 202–221. [Google Scholar] [CrossRef]
  14. Lucas-Elio, P.; Hernandez, P.; Sanchez-Amat, A.; Solano, F. Purification and partial characterization of marinocine, a new broad-spectrum antibacterial protein produced by Marinomonas mediterranea. Biochim. Biophys. Acta Gen. Subj. 2005, 1721, 193–203. [Google Scholar] [CrossRef] [PubMed]
  15. Ayuningrum, D.; Liu, Y.; Riyanti; Sibero, M.T.; Kristiana, R.; Asagabaldan, M.A.; Wuisan, Z.G.; Trianto, A.; Radjasa, O.K.; Sabdono, A.; et al. Tunicate-associated bacteria show a great potential for the discovery of antimicrobial compounds. PLoS ONE 2019, 14, e0213797. [Google Scholar] [CrossRef]
  16. 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]
  17. Anjum, K.; Sadiq, I.; Chen, L.; Kaleem, S.; Li, X.-C.; Zhang, Z.; Lian, X.-Y. Novel antifungal janthinopolyenemycins A and B from a co-culture of marine-associated Janthinobacterium spp. ZZ145 and ZZ148. Tetrahedron Lett. 2018, 59, 3490–3494. [Google Scholar] [CrossRef]
  18. Schulze, C.J.; Donia, M.S.; Siqueira-Neto, J.L.; Ray, D.; Raskatov, J.A.; Green, R.E.; McKerrow, J.H.; Fischbach, M.A.; Linington, R.G. Genome-directed lead discovery: Biosynthesis, structure elucidation, and biological evaluation of two families of polyene macrolactams against Trypanosoma brucei. ACS Chem. Biol. 2015, 10, 2373–2381. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, B.; Wang, K.B.; Wang, W.; Bi, S.F.; Mei, Y.N.; Deng, X.Z.; Jiao, R.H.; Tan, R.X.; Ge, H.M. Discovery, biosynthesis, and heterologous production of streptoseomycin, an anti-microaerophilic bacteria macrodilactone. Org. Lett. 2018, 20, 2967–2971. [Google Scholar] [CrossRef]
  20. Wiese, J.; Abdelmohsen, U.R.; Motiei, A.; Humeida, U.H.; Imhoff, J.F. Bacicyclin, a new antibacterial cyclic hexapeptide from Bacillus sp. strain BC028 isolated from Mytilus edulis. Bioorganic. Med. Chem. Lett. 2018, 28, 558–561. [Google Scholar] [CrossRef] [PubMed]
  21. Zhao, X.; Kuipers, O.P. Identification and classification of known and putative antimicrobial compounds produced by a wide variety of Bacillales species. BMC Genomics 2016, 17, 882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Caulier, S.; Nannan, C.; Gillis, A.; Licciardi, F.; Bragard, C.; Mahillon, J. Overview of the antimicrobial compounds produced by members of the Bacillus subtilis group. Front. Microbiol. 2019, 10, 302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Praptiwi; Fathoni, A.; Putri, A.L.; Wulansari, D.; Agusta, A. Assessment of Actinomycetes Isolated from Soils on Simeuleu Island as Antibacterial and Antioxidant. In Proceedings of the AIP Conference, Malang, Indonesia, 13–14 March 2019; pp. 1–8. [Google Scholar]
  24. Hoyt, P.R.; Sizemore, R.K. Competitive dominance by a bacteriocin-producing Vibrio harveyi strain. Appl. Environ. Microbiol. 1982, 44, 653–658. [Google Scholar] [CrossRef] [Green Version]
  25. Zimina, M.; Babich, O.; Prosekov, A.; Sukhikh, S.; Ivanova, S.; Shevchenko, M.; Noskova, S. Overview of global trends in classification, methods of preparation and application of bacteriocins. Antibiotics 2020, 9, 553. [Google Scholar] [CrossRef]
  26. Register, F. Nisin preparation: Affirmation of GRAS status as a direct human food ingredient. Fed. Regist. 1988, 53, 11247–11251. [Google Scholar]
  27. Lozo, J.; Vukasinovic, M.; Strahinic, I.; Topisirovic, L. Characterization and antimicrobial activity of bacteriocin 217 produced by natural isolate Lactobacillus paracasei subsp. paracasei BGBUK2-16. J. Food. Prot. 2004, 67, 2727–2734. [Google Scholar] [CrossRef] [PubMed]
  28. Drissi, F.; Buffet, S.; Raoult, D.; Merhej, V. Common occurrence of antibacterial agents in human intestinal microbiota. Front. Microbiol. 2015, 6, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Leite, J.A.; Tulini, F.L.; Reis-Teixeira, F.B.d.; Rabinovitch, L.; Chaves, J.Q.; Rosa, N.G.; Cabral, H.; De Martinis, E.C.P. Bacteriocin-like inhibitory substances (BLIS) produced by Bacillus cereus: Preliminary characterization and application of partially purified extract containing BLIS for inhibiting Listeria monocytogenes in pineapple pulp. LWT Food Sci. Tech. 2016, 72, 261–266. [Google Scholar] [CrossRef]
  30. Gradisteanu Pircalabioru, G.; Popa, L.I.; Marutescu, L.; Gheorghe, I.; Popa, M.; Czobor Barbu, I.; Cristescu, R.; Chifiriuc, M.-C. Bacteriocins in the era of antibiotic resistance: Rising to the challenge. Pharmaceutics 2021, 13, 196. [Google Scholar] [CrossRef]
  31. Jawan, R.; Abbasiliasi, S.; Mustafa, S.; Kapri, M.R.; Halim, M.; Ariff, A.B. In vitro evaluation of potential probiotic strain Lactococcus lactis Gh1 and its bacteriocin-like inhibitory substances for potential use in the food industry. Probiotics Antimicrob. Proteins 2020, 13, 422–440. [Google Scholar] [CrossRef]
  32. Hyun, W.B.; Kang, H.S.; Lee, J.W.; Abraha, H.B.; Kim, K.P. A newly-isolated Bacillus subtilis BSC35 produces bacteriocin-like inhibitory substance with high potential to control Clostridium perfringens in food. LWT Food Sci. Tech. 2021, 138, 110625. [Google Scholar] [CrossRef]
  33. Garsa, A.K.; Choudhury, P.K.; Puniya, A.K.; Dhewa, T.; Malik, R.K.; Tomar, S.K. Bovicins: The bacteriocins of streptococci and their potential in methane mitigation. Probiotics Antimicrob. Proteins 2019, 11, 1403–1413. [Google Scholar] [CrossRef]
  34. Meade, E.; Slattery, M.A.; Garvey, M. Bacteriocins, potent antimicrobial peptides and the fight against multi drug resistant species: Resistance is futile? Antibiotics 2020, 9, 32. [Google Scholar] [CrossRef] [Green Version]
  35. Shand, R.F.; Price, L.B.; O’Connor, E.M. Halocins: Protein antibiotics from hypersaline environments. In Microbiology and Biogeochemistry of Hypersaline Environments; Oren, A., Ed.; CRC Press: Boca Raton, FL, USA, 1999; pp. 295–306. [Google Scholar]
  36. Ibrahim, O.O. Classification of antimicrobial peptides bacteriocins, and the nature of some bacteriocins with potential applications in food safety and bio-pharmaceuticals. EC Microbiol. 2019, 15, 591–608. [Google Scholar]
  37. O’connor, E.; Shand, R. Halocins and sulfolobicins: The emerging story of archaeal protein and peptide antibiotics. J. Ind. Microbiol. Biotechnol. 2002, 28, 23–31. [Google Scholar] [CrossRef] [PubMed]
  38. Garcia-Gutierrez, E.; Mayer, M.J.; Cotter, P.D.; Narbad, A. Gut microbiota as a source of novel antimicrobials. Gut Microbes 2019, 10, 1–21. [Google Scholar] [CrossRef] [Green Version]
  39. Newstead, L.L.; Varjonen, K.; Nuttall, T.; Paterson, G.K. Staphylococcal-produced bacteriocins and antimicrobial peptides: Their potential as alternative treatments for Staphylococcus aureus Infections. Antibiotics 2020, 9, 40. [Google Scholar] [CrossRef] [Green Version]
  40. Negash, A.W.; Tsehai, B.A. Current applications of bacteriocin. Int. J. Microbiol. 2020, 2020, 4374891. [Google Scholar] [CrossRef]
  41. Małaczewska, J.; Kaczorek-Łukowska, E. Nisin—A lantibiotic with immunomodulatory properties: A review. Peptides 2021, 137, 170479. [Google Scholar] [CrossRef]
  42. Yu, X.; Lu, N.; Wang, J.; Chen, Z.; Chen, C.; Mac Regenstein, J.; Zhou, P. Effect of N-terminal modification on the antimicrobial activity of nisin. Food Control 2020, 114, 107227. [Google Scholar] [CrossRef]
  43. Peng, X.; Zhu, L.; Wang, Z.; Zhan, X. Enhanced stability of the bactericidal activity of nisin through conjugation with gellan gum. Int. J. Biol. Macromol. 2020, 148, 525–532. [Google Scholar] [CrossRef] [PubMed]
  44. Field, D.; Seisling, N.; Cotter, P.D.; Ross, R.P.; Hill, C. Synergistic nisin-polymyxin combinations for the control of pseudomonas biofilm formation. Front. Microbiol. 2016, 7, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Alves, F.C.B.; Albano, M.; Andrade, B.F.M.T.; Chechi, J.L.; Pereira, A.F.M.; Furlanetto, A.; Rall, V.L.M.; Fernandes, A.A.H.; Dos Santos, L.D.; Barbosa, L.N. Comparative proteomics of methicillin-resistant Staphylococcus aureus subjected to synergistic effects of the lantibiotic nisin and oxacillin. Microb. Drug Resist. 2020, 26, 179–189. [Google Scholar] [CrossRef]
  46. El-Kazzaz, S.S.; Abou El-Khier, N.T. Effect of the lantibiotic nisin on inhibitory and bactericidal activities of antibiotics used against vancomycin-resistant enterococci. J. Glob. Antimicrob. Resist. 2020, 22, 263–269. [Google Scholar] [CrossRef]
  47. Webber, J.L.; Namivandi-Zangeneh, R.; Drozdek, S.; Wilk, K.A.; Boyer, C.; Wong, E.H.H.; Bradshaw-Hajek, B.H.; Krasowska, M.; Beattie, D.A. Incorporation and antimicrobial activity of nisin Z within carrageenan/chitosan multilayers. Sci. Rep. 2021, 11, 1690. [Google Scholar] [CrossRef] [PubMed]
  48. Kjos, M.; Salehian, Z.; Nes, I.F.; Diep, D.B. An extracellular loop of the mannose phosphotransferase system component IIC is responsible for specific targeting by class IIa bacteriocins. J. Bacteriol. 2010, 192, 5906–5913. [Google Scholar] [CrossRef] [Green Version]
  49. Lozo, J.; Topisirovic, L.; Kojic, M. Natural bacterial isolates as an inexhaustible source of new bacteriocins. Appl. Microbiol. Biotechnol. 2021, 105, 477–492. [Google Scholar] [CrossRef] [PubMed]
  50. Perez, R.H.; Perez, M.T.M.; Elegado, F.B. Bacteriocins from lactic acid bacteria: A review of biosynthesis, mode of action, fermentative production, uses, and prospects. Int. J. Phil. Sci. Tech. 2015, 8, 61–67. [Google Scholar] [CrossRef] [Green Version]
  51. Da Costa, R.J.; Voloski, F.L.; Mondadori, R.G.; Duval, E.H.; Fiorentini, Â.M. Preservation of meat products with bacteriocins produced by lactic acid bacteria isolated from meat. J. Food Qual. 2019, 2019, 1–12. [Google Scholar] [CrossRef] [Green Version]
  52. Shindo, K.; Takenaka, A.; Noguchi, T.; Hayakawa, Y.; Seto, H. Thiazostatin A and thiazostatin B, new antioxidants produced by Streptomyces tolurosus. J. Antibiot. 1989, 42, 1526–1529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Alvarez-Sieiro, P.; Montalbán-López, M.; Mu, D.; Kuipers, O.P. Bacteriocins of lactic acid bacteria: Extending the family. Appl. Microbiol. Biotechnol. 2016, 100, 2939–2951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Wang, H.; Fewer, D.P.; Holm, L.; Rouhiainen, L.; Sivonen, K. Atlas of nonribosomal peptide and polyketide biosynthetic pathways reveals common occurrence of nonmodular enzymes. Proc. Natl. Acad. Sci. USA 2014, 111, 9259–9264. [Google Scholar] [CrossRef] [Green Version]
  55. Aleti, G.; Sessitsch, A.; Brader, G. Genome mining: Prediction of lipopeptides and polyketides from Bacillus and related Firmicutes. Comput. Struct. Biotechnol. J. 2015, 13, 192–203. [Google Scholar] [CrossRef]
  56. Perez, K.J.; Viana, J.d.S.; Lopes, F.C.; Pereira, J.Q.; dos Santos, D.M.; Oliveira, J.S.; Velho, R.V.; Crispim, S.M.; Nicoli, J.R.; Brandelli, A.; et al. Bacillus spp. isolated from puba as a source of biosurfactants and antimicrobial lipopeptides. Front. Microbiol. 2017, 8, 61–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Kourmentza, K.; Gromada, X.; Michael, N.; Degraeve, C.; Vanier, G.; Ravallec, R.; Coutte, F.; Karatzas, K.A.; Jauregi, P. Antimicrobial activity of lipopeptide biosurfactants against foodborne pathogen and food spoilage microorganisms and their cytotoxicity. Front. Microbiol. 2021, 11, 3398–3412. [Google Scholar] [CrossRef] [PubMed]
  58. Isa, M.H.M.; Shannaq, M.A.-H.F.; Mohamed, N.; Hassan, A.R.; Al-Shorgani, N.K.N.; Hamid, A.A. Antibacterial activity of surfactin produced by Bacillus subtilis MSH1. Trans. Sci. Tech. 2017, 4, 402–407. [Google Scholar]
  59. Calvo, H.; Mendiara, I.; Arias, E.; Blanco, D.; Venturini, M. The role of iturin A from B. amyloliquefaciens BUZ-14 in the inhibition of the most common postharvest fruit rots. Food Microbiol. 2019, 82, 62–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Sur, S.; Romo, T.D.; Grossfield, A. Selectivity and mechanism of fengycin, an antimicrobial lipopeptide, from molecular dynamics. J. Phys. Chem. B 2018, 122, 2219–2226. [Google Scholar] [CrossRef]
  61. Meena, K.; Sharma, A.; Kanwar, S. Microbial lipopeptides and their medical applications. Ann. Pharmacol. Pharm. 2017, 2, 1126. [Google Scholar]
  62. Qi, G.; Zhu, F.; Du, P.; Yang, X.; Qiu, D.; Yu, Z.; Chen, J.; Zhao, X. Lipopeptide induces apoptosis in fungal cells by a mitochondria-dependent pathway. Peptides 2010, 31, 1978–1986. [Google Scholar] [CrossRef] [PubMed]
  63. Inès, M.; Dhouha, G. Lipopeptide surfactants: Production, recovery and pore forming capacity. Peptides 2015, 71, 100–112. [Google Scholar] [CrossRef]
  64. Straus, S.K.; Hancock, R.E. Mode of action of the new antibiotic for Gram-positive pathogens daptomycin: Comparison with cationic antimicrobial peptides and lipopeptides. Biochim. Biophys. Acta Biomembr. BBA-Biomembr. 2006, 1758, 1215–1223. [Google Scholar] [CrossRef] [Green Version]
  65. Ma, Z.; Zhang, S.; Sun, K.; Hu, J. Identification and characterization of a cyclic lipopeptide iturin A from a marine-derived Bacillus velezensis 11-5 as a fungicidal agent to Magnaporthe oryzae in rice. J. Plant. Dis. Prot. 2020, 127, 15–24. [Google Scholar] [CrossRef]
  66. Touré, Y.; Ongena, M.; Jacques, P.; Guiro, A.; Thonart, P. Role of lipopeptides produced by Bacillus subtilis GA1 in the reduction of grey mould disease caused by Botrytis cinerea on apple. J. Appl. Microbiol. 2004, 96, 1151–1160. [Google Scholar] [CrossRef] [PubMed]
  67. Solecka, J.; Zajko, J.; Postek, M.; Rajnisz, A. Biologically active secondary metabolites from Actinomycetes. Open Life Sci. 2012, 7, 373–390. [Google Scholar] [CrossRef]
  68. Dholakiya, R.N.; Kumar, R.; Mishra, A.; Mody, K.H.; Jha, B. Antibacterial and antioxidant activities of novel actinobacteria strain isolated from Gulf of Khambhat, Gujarat. Front. Microbiol. 2017, 8, 2420. [Google Scholar] [CrossRef]
  69. Siddharth, S.; Vittal, R.R. Evaluation of antimicrobial, enzyme inhibitory, antioxidant and cytotoxic activities of partially purified volatile metabolites of marine Streptomyces sp. S2A. Microorganisms 2018, 6, 72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Dhakal, D.; Sohng, J.K.; Pandey, R.P. Engineering actinomycetes for biosynthesis of macrolactone polyketides. Micro. Cell Fact. 2019, 18, 137. [Google Scholar] [CrossRef] [PubMed]
  71. Pacios-Michelena, S.; Aguilar González, C.N.; Alvarez-Perez, O.B.; Rodriguez-Herrera, R.; Chávez-González, M.; Arredondo Valdés, R.; Ascacio Valdés, J.A.; Govea Salas, M.; Ilyina, A. Application of Streptomyces antimicrobial compounds for the control of phytopathogens. Front. Sustain. Food Syst. 2021, 5, 310–324. [Google Scholar] [CrossRef]
  72. Widowati, I.; Zainuri, M.; Kusumaningrum, H.P.; Susilowati, R.; Hardivillier, Y.; Leignel, V.; Bourgougnon, N.; Mouget, J.L. Antioxidant Activity of Three Microalgae Dunaliella Salina, Tetraselmis chuii and Isochrysis galbana Clone Tahiti. In Proceedings of the IOP Conference Series: Earth and Environmental Science, Bali, Indonesia, 19–22 October 2017; IOP Publishing: Bristol, UK, 2017; p. 012067. [Google Scholar]
  73. McArthur, K.A.; Mitchell, S.S.; Tsueng, G.; Rheingold, A.; White, D.J.; Grodberg, J.; Lam, K.S.; Potts, B.C. Lynamicins a−e, chlorinated bisindole pyrrole antibiotics from a novel marine actinomycete. J. Nat. Prod. 2008, 71, 1732–1737. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, J.; Wang, J.-D.; Liu, C.-X.; Yuan, J.-H.; Wang, X.-J.; Xiang, W.-S. A new prenylated indole derivative from endophytic actinobacteria Streptomyces sp. neau-D50. Nat. Prod. Res. 2014, 28, 431–437. [Google Scholar] [CrossRef] [PubMed]
  75. Djinni, I.; Defant, A.; Kecha, M.; Mancini, I. Metabolite profile of marine-derived endophytic Streptomyces sundarbansensis WR 1 L 1 S 8 by liquid chromatography–mass spectrometry and evaluation of culture conditions on antibacterial activity and mycelial growth. J. Appl. Microbiol. 2014, 116, 39–50. [Google Scholar] [CrossRef]
  76. Sebak, M.; Saafan, A.E.; Abdelghani, S.; Bakeer, W.; Moawad, A.S.; El-Gendy, A.O. Isolation and optimized production of putative antimicrobial compounds from Egyptian soil isolate Streptomyces sp. MS. 10. Beni-Suef Univ. J. Basic Appl. Sci. 2021, 10, 8. [Google Scholar] [CrossRef]
  77. Qureshi, K.A.; Bholay, A.D.; Rai, P.K.; Mohammed, H.A.; Khan, R.A.; Azam, F.; Jaremko, M.; Emwas, A.-H.; Stefanowicz, P.; Waliczek, M.; et al. Isolation, characterization, anti-MRSA evaluation, and in-silico multi-target anti-microbial validations of actinomycin X2 and actinomycin D produced by novel Streptomyces smyrnaeus UKAQ_23. Sci. Rep. 2021, 11, 14539. [Google Scholar] [CrossRef]
  78. Yang, X.; Peng, T.; Yang, Y.; Li, W.; Xiong, J.; Zhao, L.; Ding, Z. Antimicrobial and antioxidant activities of a new benzamide from endophytic Streptomyces sp. YIM 67086. Natl. Prod. Res. 2015, 29, 331–335. [Google Scholar] [CrossRef]
  79. Ding, W.-J.; Zhang, S.-Q.; Wang, J.-H.; Lin, Y.-X.; Liang, Q.-X.; Zhao, W.-J.; Li, C.-Y. A new di-O-prenylated flavone from an actinomycete Streptomyces sp. MA-12. J. Asian Nat. Prod. Res. 2013, 15, 209–214. [Google Scholar] [CrossRef]
  80. Lee, L.-H.; Zainal, N.; Azman, A.-S.; Eng, S.-K.; Goh, B.-H.; Yin, W.-F.; Ab Mutalib, N.-S.; Chan, K.-G. Diversity and antimicrobial activities of actinobacteria isolated from tropical mangrove sediments in Malaysia. Sci. World J. 2014, 2014, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Jaroszewicz, W.; Bielańska, P.; Lubomska, D.; Kosznik-Kwaśnicka, K.; Golec, P.; Grabowski, Ł.; Wieczerzak, E.; Dróżdż, W.; Gaffke, L.; Pierzynowska, K. Antibacterial, antifungal and anticancer activities of compounds produced by newly isolated Streptomyces strains from the Szczelina Chochołowska cave (Tatra Mountains, Poland). Antibiotics 2021, 10, 1212. [Google Scholar] [CrossRef] [PubMed]
  82. Prado-Alonso, L.; Pérez-Victoria, I.; Malmierca, M.G.; Montero, I.; Rioja-Blanco, E.; Martín, J.; Reyes, F.; Méndez, C.; Salas, J.A.; Olano, C. Colibrimycins, novel halogenated hybrid PKS-NRPS compounds produced by Streptomyces sp. CS147. Appl. Environ. Microbiol. 2021. [Google Scholar] [CrossRef]
  83. Veilumuthu, P.; Siva, R.; Godwin Christopher, J. Extraction and Characterization of Antibiotic Compounds produced by Streptomyces spp. VITGV01 against Selected Human Pathogens. 2021. Available online: https://www.researchsquare.com/article/rs-321054/v1 (accessed on 20 April 2021).
  84. Kang, H.-K.; Hyun, C.G. Anti-inflammatory effect of d-(+)-cycloserine through inhibition of NF-κB and MAPK signaling pathways in LPS-induced RAW 264.7 macrophages. Nat. Prod. Commun. 2020, 15, 1–11. [Google Scholar] [CrossRef]
  85. Robertsen, H.L.; Musiol-Kroll, E.M. Actinomycete-derived polyketides as a source of antibiotics and lead structures for the development of new antimicrobial drugs. Antibiotics 2019, 8, 157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Hug, J.J.; Bader, C.D.; Remškar, M.; Cirnski, K.; Müller, R. Concepts and methods to access novel antibiotics from actinomycetes. Antibiotics 2018, 7, 44. [Google Scholar] [CrossRef] [Green Version]
  87. Yang, X.; Yang, Y.; Peng, T.; Yang, F.; Zhou, H.; Zhao, L.; Xu, L.; Ding, Z. A new cyclopeptide from endophytic Streptomyces sp. YIM 64018. Nat. Prod. Commun. 2013, 8, 1753–1754. [Google Scholar] [CrossRef] [Green Version]
  88. Yu, Z.; Zhao, L.-X.; Jiang, C.-L.; Duan, Y.; Wong, L.; Carver, K.C.; Schuler, L.A.; Shen, B. Bafilomycins produced by an endophytic actinomycete Streptomyces sp. YIM56209. J. Antibiot. 2011, 64, 159–162. [Google Scholar] [CrossRef]
  89. Li, Y.; Han, L.; Rong, H.; Li, L.; Zhao, L.; Wu, L.; Xu, L.; Jiang, Y.; Huang, X. Diastaphenazine, a new dimeric phenazine from an endophytic Streptomyces diastaticus subsp. ardesiacus. J. Antibiot. 2015, 68, 210–212. [Google Scholar] [CrossRef] [PubMed]
  90. Savi, D.C.; Shaaban, K.A.; Vargas, N.; Ponomareva, L.V.; Possiede, Y.M.; Thorson, J.S.; Glienke, C.; Rohr, J. Microbispora sp. LGMB259 endophytic actinomycete isolated from Vochysia divergens (Pantanal, Brazil) producing β-carbolines and indoles with biological activity. Curr. Microbiol. 2015, 70, 345–354. [Google Scholar] [CrossRef]
  91. Zhou, H.; Yang, Y.; Peng, T.; Li, W.; Zhao, L.; Xu, L.; Ding, Z. Metabolites of Streptomyces sp., an endophytic actinomycete from Alpinia oxyphylla. Nat. Prod. Res. 2014, 28, 265–267. [Google Scholar] [CrossRef]
  92. Yan, L.-L.; Han, N.-N.; Zhang, Y.-Q.; Yu, L.-Y.; Chen, J.; Wei, Y.-Z.; Li, Q.-P.; Tao, L.; Zheng, G.-H.; Yang, S.-E. Antimycin A 18 produced by an endophytic Streptomyces albidoflavus isolated from a mangrove plant. J. Antibiot. 2010, 63, 259–261. [Google Scholar] [CrossRef]
  93. Manimaran, M.; Gopal, J.V.; Kannabiran, K. Antibacterial activity of Streptomyces sp. VITMK1 isolated from mangrove Soil of pichavaram, Tamil Nadu, India. Proc. Natl. Acad. Sci. USA India Sect. B Biol. Sci. 2017, 87, 499–506. [Google Scholar] [CrossRef]
  94. Wang, P.; Xi, L.; Liu, P.; Wang, Y.; Wang, W.; Huang, Y.; Zhu, W. Diketopiperazine derivatives from the marine-derived actinomycete Streptomyces sp. FXJ7.328. Mar. Drugs 2013, 11, 1035–1049. [Google Scholar] [CrossRef] [Green Version]
  95. Bulut, N.; Kocyigit, U.M.; Gecibesler, I.H.; Dastan, T.; Karci, H.; Taslimi, P.; Durna Dastan, S.; Gulcin, I.; Cetin, A. Synthesis of some novel pyridine compounds containing bis-1, 2, 4-triazole/thiosemicarbazide moiety and investigation of their antioxidant properties, carbonic anhydrase, and acetylcholinesterase enzymes inhibition profiles. J. Biochem. Mol. Toxicol. 2018, 32, e22006. [Google Scholar] [CrossRef] [PubMed]
  96. Gulcin, İ. Antioxidants and antioxidant methods: An updated overview. Arch. Toxicol. 2020, 94, 651–715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Rodriguez-Valera, F.; Juez, G.; Kushner, D. Halocins: Salt-dependent bacteriocins produced by extremely halophilic rods. Can. J. Microbiol. 1982, 28, 151–154. [Google Scholar] [CrossRef]
  98. Gaglione, R.; Pirone, L.; Farina, B.; Fusco, S.; Smaldone, G.; Aulitto, M.; Dell’Olmo, E.; Roscetto, E.; Del Gatto, A.; Fattorusso, R.; et al. Insights into the anticancer properties of the first antimicrobial peptide from Archaea. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 2155–2164. [Google Scholar] [CrossRef] [Green Version]
  99. Besse, A.; Peduzzi, J.; Rebuffat, S.; Carré-Mlouka, A. Antimicrobial peptides and proteins in the face of extremes: Lessons from archaeocins. Biochimie 2015, 118, 344–355. [Google Scholar] [CrossRef]
  100. Platas, G.; Meseguer, I.; Amils, R. Purification and biological characterization of halocin H1 from Haloferax mediterranei M2a. Int. Microbiol. 2002, 5, 15–19. [Google Scholar] [CrossRef]
  101. Meseguer, I.; Rodriguez-Valera, F. Production and purification of halocin H4. FEMS Microbiol. Lett. 1985, 28, 177–182. [Google Scholar] [CrossRef]
  102. Torreblanca, M.; Meseguer, I.; Rodríguez-Valera, F. Halocin H6, a bacteriocin from Haloferax gibbonsii. Microbiology 1989, 135, 2655–2661. [Google Scholar] [CrossRef] [Green Version]
  103. Li, Y.; Xiang, H.; Liu, J.; Zhou, M.; Tan, H. Purification and biological characterization of halocin C8, a novel peptide antibiotic from Halobacterium strain AS7092. Extremophiles 2003, 7, 401–407. [Google Scholar] [CrossRef] [PubMed]
  104. Price, L.B.; Shand, R.F. Halocin S8: A 36-amino-acid microhalocin from the haloarchaeal strain S8a. J. Bacteriol. 2000, 182, 4951–4958. [Google Scholar] [CrossRef] [Green Version]
  105. Ebert, K.; Goebel, W.; Rdest, U.; Surek, B. Genes and genome structures in the archaebacteria. Syst. Appl. Microbiol. 1986, 7, 30–35. [Google Scholar] [CrossRef]
  106. Pašić, L.; Velikonja, B.H.; Ulrih, N.P. Optimization of the culture conditions for the production of a bacteriocin from halophilic archaeon Sech7a. Prep. Biochem. Biotechnol. 2008, 38, 229–245. [Google Scholar] [CrossRef] [PubMed]
  107. Shand, R.F.; Leyva, K.J. Peptide and protein antibiotics from the domain Archaea: Halocins and sulfolobicins. In Bacteriocins: Ecology and Evolution; Riley, M.A., Chavan, M.A., Eds.; Springer: Berlin/Heidelberg, Germany, 2007; pp. 93–109. [Google Scholar]
  108. Karthikeyan, P.; Bhat, S.G.; Chandrasekaran, M. Halocin SH10 production by an extreme haloarchaeon Natrinema sp. BTSH10 isolated from salt pans of South India. Saudi J. Biol. Sci. 2013, 20, 205–212. [Google Scholar] [CrossRef] [Green Version]
  109. Wali, A.F.; Al Dhaheri, Y.; Ramakrishna Pillai, J.; Mushtaq, A.; Rao, P.G.; Rabbani, S.A.; Firdous, A.; Elshikh, M.S.; Farraj, D.A.A. Lc-ms phytochemical screening, in vitro antioxidant, antimicrobial and anticancer activity of microalgae Nannochloropsis oculata extract. Separations 2020, 7, 54. [Google Scholar] [CrossRef]
  110. Tan, L.T.-H.; Chan, K.-G.; Khan, T.M.; Bukhari, S.I.; Saokaew, S.; Duangjai, A.; Pusparajah, P.; Lee, L.-H.; Goh, B.-H. Streptomyces sp. MUM212 as a source of antioxidants with radical scavenging and metal chelating properties. Front. Pharmacol. 2017, 8, 276–293. [Google Scholar] [CrossRef]
  111. de Castro, I.; Mendo, S.; Caetano, T. Antibiotics from Haloarchaea: What can we learn from comparative genomics? Mar. Biotechnol. 2020, 22, 308–316. [Google Scholar] [CrossRef]
  112. Sahli, K.; Gomri, M.A.; Esclapez, J.; Gómez-Villegas, P.; Ghennai, O.; Bonete, M.J.; León, R.; Kharroub, K. Bioprospecting and characterization of pigmented halophilic archaeal strains from Algerian hypersaline environments with analysis of carotenoids produced by Halorubrum sp. BS2. J. Basic Microbiol. 2020, 60, 624–638. [Google Scholar] [CrossRef]
  113. Ghanmi, F.; Carré-Mlouka, A.; Vandervennet, M.; Boujelben, I.; Frikha, D.; Ayadi, H.; Peduzzi, J.; Rebuffat, S.; Maalej, S. Antagonistic interactions and production of halocin antimicrobial peptides among extremely halophilic prokaryotes isolated from the solar saltern of Sfax, Tunisia. Extremophiles 2016, 20, 363–374. [Google Scholar] [CrossRef] [PubMed]
  114. Ghanmi, F.; Carré-Mlouka, A.; Zarai, Z.; Mejdoub, H.; Peduzzi, J.; Maalej, S.; Rebuffat, S. The extremely halophilic archaeon Halobacterium salinarum ETD5 from the solar saltern of Sfax (Tunisia) produces multiple halocins. Res. Microbiol. 2020, 171, 80–90. [Google Scholar] [CrossRef] [PubMed]
  115. Besse, A.; Vandervennet, M.; Goulard, C.; Peduzzi, J.; Isaac, S.; Rebuffat, S.; Carré-Mlouka, A. Halocin C8: An antimicrobial peptide distributed among four halophilic archaeal genera: Natrinema, Haloterrigena, Haloferax, and Halobacterium. Extremophiles 2017, 21, 623–638. [Google Scholar] [CrossRef] [PubMed]
  116. Quehenberger, J.; Shen, L.; Albers, S.-V.; Siebers, B.; Spadiut, O. Sulfolobus—A potential key organism in future biotechnology. Front. Microbiol. 2017, 8, 2474–2485. [Google Scholar] [CrossRef]
  117. Roscetto, E.; Contursi, P.; Vollaro, A.; Fusco, S.; Notomista, E.; Catania, M.R. Antifungal and anti-biofilm activity of the first cryptic antimicrobial peptide from an archaeal protein against Candida spp. clinical isolates. Sci. Rep. 2018, 8, 17570. [Google Scholar] [CrossRef] [PubMed]
  118. Kumar, V.A.; Tiwari, S.K. Halocin HA1: An archaeocin produced by the haloarchaeon Haloferax larsenii HA1. Process Biochem. 2017, 61, 202–208. [Google Scholar] [CrossRef]
  119. Quadri, I.; Hassani, I.I.; l’Haridon, S.; Chalopin, M.; Hacène, H.; Jebbar, M. Characterization and antimicrobial potential of extremely halophilic archaea isolated from hypersaline environments of the Algerian Sahara. Microbiol. Res. 2016, 186–187, 119–131. [Google Scholar] [CrossRef]
  120. Photolo, M.M.; Mavumengwana, V.; Sitole, L.; Tlou, M.G. Antimicrobial and antioxidant properties of a bacterial endophyte, Methylobacterium radiotolerans MAMP 4754, isolated from Combretum erythrophyllum seeds. Int. J. Microbiol. 2020, 2020, 9483670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Ayyanna, R.; Ankaiah, D.; Arul, V. Anti-inflammatory and antioxidant properties of probiotic bacterium Lactobacillus mucosae AN1 and Lactobacillus fermentum SNR1 in Wistar Albino rats. Front. Microbiol. 2018, 9, 3063–3075. [Google Scholar] [CrossRef]
  122. Revilla, G.; Ramos, F.; López-Nieto, M.; Alvarez, E.; Martin, J. Glucose represses formation of delta-(L-alpha-aminoadipyl)-L-cysteinyl-D-valine and isopenicillin N synthase but not penicillin acyltransferase in Penicillium chrysogenum. J. Bacteriol. 1986, 168, 947–952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Zanca, D.M.; Martin, J.F. Carbon catabolite regulation of the conversion of penicillin N into cephalosporin C. J. Antibiot. 1983, 36, 700–708. [Google Scholar] [CrossRef] [Green Version]
  124. Malve, H. Exploring the ocean for new drug developments: Marine pharmacology. J. Pharm. Bioallied. Sci. 2016, 8, 83. [Google Scholar] [CrossRef] [PubMed]
  125. Li, Y.; Lan, N.; Xu, L.; Yue, Q. Biosynthesis of pneumocandin lipopeptides and perspectives for its production and related echinocandins. Appl. Microbiol. Biotechnol. 2018, 102, 9881–9891. [Google Scholar] [CrossRef]
  126. Huang, Z.; Cai, X.; Shao, C.; She, Z.; Xia, X.; Chen, Y.; Yang, J.; Zhou, S.; Lin, Y. Chemistry and weak antimicrobial activities of phomopsins produced by mangrove endophytic fungus Phomopsis sp. ZSU-H76. Phytochemistry 2008, 69, 1604–1608. [Google Scholar] [CrossRef] [PubMed]
  127. Strobel, G.; Daisy, B. Bioprospecting for microbial endophytes and their natural products. Microbiol. Mol. Biol. Rev. 2003, 67, 491–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Liu, H.; Yan, C.; Li, C.; You, T.; She, Z. Naphthoquinone derivatives with anti-inflammatory activity from mangrove-derived endophytic fungus Talaromyces sp. SK-S009. Molecules 2020, 25, 576. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, C.-L.; Zheng, B.-Q.; Lao, J.-P.; Mao, L.-J.; Chen, S.-Y.; Kubicek, C.P.; Lin, F.-C. Clavatol and patulin formation as the antagonistic principle of Aspergillus clavatonanicus, an endophytic fungus of Taxus mairei. Appl. Microbiol. Biotechnol. 2008, 78, 833–840. [Google Scholar] [CrossRef]
  130. Wu, S.-H.; Chen, Y.-W.; Shao, S.-C.; Wang, L.-D.; Li, Z.-Y.; Yang, L.-Y.; Li, S.-L.; Huang, R. Ten-membered lactones from Phomopsis sp., an endophytic fungus of Azadirachta indica. J. Nat. Prod. 2008, 71, 731–734. [Google Scholar] [CrossRef] [PubMed]
  131. Toghueo, R.M.K.; Sahal, D.; Boyom, F.F. Recent advances in inducing endophytic fungal specialized metabolites using small molecule elicitors including epigenetic modifiers. Phytochemistry 2020, 174, 112338. [Google Scholar] [CrossRef] [PubMed]
  132. 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] [PubMed] [Green Version]
  133. 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]
  134. Wu, B.; Wiese, J.; Labes, A.; Kramer, A.; Schmaljohann, R.; Imhoff, J.F. Lindgomycin, an unusual antibiotic polyketide from a marine fungus of the Lindgomycetaceae. Mar. Drugs 2015, 13, 4617–4632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Gao, S.-S.; Li, X.-M.; Du, F.-Y.; Li, C.-S.; Proksch, P.; Wang, B.-G. Secondary metabolites from a marine-derived endophytic fungus Penicillium chrysogenum QEN-24S. Mar. Drugs 2011, 9, 59–70. [Google Scholar] [CrossRef] [Green Version]
  136. Wang, C.; Tang, S.; Cao, S. Antimicrobial compounds from marine fungi. Phytochem. Rev. 2020, 20, 1–33. [Google Scholar] [CrossRef]
  137. Li, X.; Li, X.-M.; Zhang, P.; Wang, B.-G. A new phenolic enamide and a new meroterpenoid from marine alga-derived endophytic fungus Penicillium oxalicum EN-290. J. Asian Nat. Prod. Res. 2015, 17, 1204–1212. [Google Scholar] [CrossRef]
  138. 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]
  139. Gao, S.-S.; Li, X.-M.; Zhang, Y.; Li, C.-S.; Cui, C.-M.; Wang, B.-G. Comazaphilones A-F, azaphilone derivatives from the marine sediment-derived fungus Penicillium commune QSD-17. J. Nat. Prod. 2011, 74, 256–261. [Google Scholar] [CrossRef] [PubMed]
  140. 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]
  141. Fukuda, T.; Kurihara, Y.; Kanamoto, A.; Tomoda, H. Terretonin G, a new sesterterpenoid antibiotic from marine-derived Aspergillus sp. OPMF00272. J. Antibiot. 2014, 67, 593–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Prompanya, C.; Dethoup, T.; Bessa, L.J.; Pinto, M.M.; Gales, L.; Costa, P.M.; Silva, A.; Kijjoa, A. New isocoumarin derivatives and meroterpenoids from the marine sponge-associated fungus Aspergillus similanensis sp. nov. KUFA 0013. Mar. Drugs 2014, 12, 5160–5173. [Google Scholar] [CrossRef]
  143. Ding, L.; Li, T.; Liao, X.; He, S.; Xu, S. Asperitaconic acids A–C, antibacterial itaconic acid derivatives produced by a marine-derived fungus of the genus Aspergillus. J. Antibiot. 2018, 71, 902–904. [Google Scholar] [CrossRef]
  144. Peng, X.; Wang, Y.; Zhu, T.; Zhu, W. Pyrazinone derivatives from the coral-derived Aspergillus ochraceus LCJ11-102 under high iodide salt. Arch. Pharm. Res. 2018, 41, 184–191. [Google Scholar] [CrossRef]
  145. Wang, R.; Guo, Z.K.; Li, X.M.; Chen, F.X.; Zhan, X.F.; Shen, M.H. Spiculisporic acid analogues of the marine-derived fungus, Aspergillus candidus strain HDf2, and their antibacterial activity. Antonie Van Leeuwenhoek 2015, 108, 215–219. [Google Scholar] [CrossRef]
  146. Zhu, F.; Chen, G.; Chen, X.; Huang, M.; Wan, X. Aspergicin, a new antibacterial alkaloid produced by mixed fermentation of two marine-derived mangrove epiphytic fungi. Chem. Nat. Compd. 2011, 47, 767–769. [Google Scholar] [CrossRef]
  147. Zhang, Y.; Wang, S.; Li, X.-M.; Cui, C.-M.; Feng, C.; Wang, B.-G. New sphingolipids with a previously unreported 9-methyl-C 20-sphingosine moiety from a marine algous endophytic fungus Aspergillus niger EN-13. Lipids 2007, 42, 759–764. [Google Scholar] [CrossRef]
  148. Yang, G.; Sandjo, L.; Yun, K.; Leutou, A.S.; Kim, G.-D.; Choi, H.D.; Kang, J.S.; Hong, J.; Son, B.W. Flavusides A and B, antibacterial cerebrosides from the marine-derived fungus Aspergillus flavus. Chem. Pharm. Bull. 2011, 59, 1174–1177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Hawas, U.W.; El-Beih, A.A.; El-Halawany, A.M. Bioactive anthraquinones from endophytic fungus Aspergillus versicolor isolated from red sea algae. Arch. Pharm. Res. 2012, 35, 1749–1756. [Google Scholar] [CrossRef] [PubMed]
  150. Li, H.-L.; Li, X.-M.; Yang, S.-Q.; Cao, J.; Li, Y.-H.; Wang, B.-G. Induced terreins production from marine red algal-derived endophytic fungus Aspergillus terreus EN-539 co-cultured with symbiotic fungus Paecilomyces lilacinus EN-531. J. Antibiot. 2020, 73, 108. [Google Scholar] [CrossRef] [PubMed]
  151. Zhu, A.; Yang, M.-Y.; Zhang, Y.-H.; Shao, C.-L.; Wang, C.-Y.; Hu, L.-D.; Cao, F.; Zhu, H.-J. Absolute configurations of 14, 15-hydroxylated prenylxanthones from a marine-derived Aspergillus sp. fungus by chiroptical methods. Sci. Rep. 2018, 8, 1–10. [Google Scholar]
  152. Liu, Z.; Dong, Z.; Qiu, P.; Wang, Q.; Yan, J.; Lu, Y.; Wasu, P.-a.; Hong, K.; She, Z. Two new bioactive steroids from a mangrove-derived fungus Aspergillus sp. Steroids 2018, 140, 32–38. [Google Scholar] [CrossRef]
  153. Pruksakorn, P.; Arai, M.; Kotoku, N.; Vilchèze, C.; Baughn, A.D.; Moodley, P.; Jacobs Jr, W.R.; Kobayashi, M. Trichoderins, novel aminolipopeptides from a marine sponge-derived Trichoderma sp., are active against dormant mycobacteria. Bioorg. Med. Chem. Lett. 2010, 20, 3658–3663. [Google Scholar] [CrossRef]
  154. Pang, X.; Lin, X.; Yang, J.; Zhou, X.; Yang, B.; Wang, J.; Liu, Y. Spiro-phthalides and isocoumarins isolated from the marine-sponge-derived fungus Setosphaeria sp. SCSIO41009. J. Nat. Prod. 2018, 81, 1860–1868. [Google Scholar] [CrossRef]
  155. Song, T.; Chen, M.; Chai, W.; Zhang, Z.; Lian, X.-Y. New bioactive pyrrospirones C− I from a marine-derived fungus Penicillium sp. ZZ380. Tetrahedron 2018, 74, 884–891. [Google Scholar] [CrossRef]
  156. Song, T.; Chen, M.; Ge, Z.-W.; Chai, W.; Li, X.-C.; Zhang, Z.; Lian, X.-Y. Bioactive penicipyrrodiether A, an adduct of GKK1032 analogue and phenol A derivative, from a marine-sourced fungus Penicillium sp. ZZ380. J. Org. Chem. 2018, 83, 13395–13401. [Google Scholar] [CrossRef]
  157. Shen, H.-S.; Shao, S.; Chen, J.-C.; Zhou, T. Antimicrobials from mushrooms for assuring food safety. Compr. Rev. Food Sci. Food Saf. 2017, 16, 316–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Thu, Z.M.; Myo, K.K.; Aung, H.T.; Clericuzio, M.; Armijos, C.; Vidari, G. Bioactive phytochemical constituents of wild edible mushrooms from Southeast Asia. Molecules 2020, 25, 1972. [Google Scholar] [CrossRef]
  159. Mygind, P.H.; Fischer, R.L.; Schnorr, K.M.; Hansen, M.T.; Sönksen, C.P.; Ludvigsen, S.; Raventós, D.; Buskov, S.; Christensen, B.; De Maria, L.; et al. Plectasin is a peptide antibiotic with therapeutic potential from a saprophytic fungus. Nature 2005, 437, 975–980. [Google Scholar] [CrossRef] [PubMed]
  160. Wong, J.H.; Ng, T.B.; Wang, H.; Sze, S.C.W.; Zhang, K.Y.; Li, Q.; Lu, X. Cordymin, an antifungal peptide from the medicinal fungus Cordyceps militaris. Phytomedicine 2011, 18, 387–392. [Google Scholar] [CrossRef] [PubMed]
  161. Gebreyohannes, G.; Nyerere, A.; Bii, C.; Sbhatu, D.B. Investigation of antioxidant and antimicrobial activities of different extracts of Auricularia and Termitomyces species of mushrooms. Sci. World J. 2019, 2019, 1–10. [Google Scholar] [CrossRef]
  162. Poompouang, S.; Suksomtip, M. Isolation and characterization of an antifungal peptide from fruiting bodies of edible mushroom Lentinus squarrosulus Mont. Malays. J. Microbiol. 2016, 12, 43–49. [Google Scholar] [CrossRef]
  163. Irshad, A.; Sarwar, N.; Sadia, H.; Malik, K.; Javed, I.; Irshad, A.; Afzal, M.; Abbas, M.; Rizvi, H. Comprehensive facts on dynamic antimicrobial properties of polysaccharides and biomolecules-silver nanoparticle conjugate. Int. J. Biol. Macromol. 2020, 145, 189–196. [Google Scholar] [CrossRef]
  164. Hamamoto, E.; Kimura, N.; Nishino, S.; Ishihara, A.; Otani, H.; Osaki-Oka, K. Antimicrobial activity of the volatile compound 3,5-dichloro-4-methoxybenzaldehyde, produced by the mushroom Porostereum spadiceum, against plant-pathogenic bacteria and fungi. J. App. Microbiol. 2021, 131, 1431–1439. [Google Scholar] [CrossRef]
  165. Subrata, G.; Gunjan, B.; Prakash, P.; Mandal, S.C.; Krishnendu, A. Antimicrobial activities of basidiocarps of wild edible mushrooms of West Bengal, India. Int. J. Pharmtech. Res. 2012, 4, 1554–1560. [Google Scholar]
  166. Sevindi, M. Phenolic content, antioxidant and antimicrobial potential of Melanoleuca melaleuca edible mushroom. J. Anim. Plant Sci. 2021, 31, 824–830. [Google Scholar]
  167. Endo, H.; Niioka, M.; Kobayashi, N.; Tanaka, M.; Watanabe, T. Butyrate-producing probiotics reduce nonalcoholic fatty liver disease progression in rats: New insight into the probiotics for the gut-liver axis. PLoS ONE 2013, 8, e63388. [Google Scholar]
  168. Fakruddin, M.; Hossain, M.N.; Ahmed, M.M. Antimicrobial and antioxidant activities of Saccharomyces cerevisiae IFST062013, a potential probiotic. BMC Complement. Altern. Med. 2017, 17, 64. [Google Scholar] [CrossRef] [Green Version]
  169. Breinig, F.; Sendzik, T.; Eisfeld, K.; Schmitt, M.J. Dissecting toxin immunity in virus-infected killer yeast uncovers an intrinsic strategy of self-protection. Proc. Nat. Acad. Sci. USA 2006, 103, 3810–3815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Muccilli, S.; Wemhoff, S.; Restuccia, C.; Meinhardt, F. Exoglucanase-encoding genes from three Wickerhamomyces anomalus killer strains isolated from olive brine. Yeast 2013, 30, 33–43. [Google Scholar] [CrossRef] [PubMed]
  171. Gao, D.; Gao, Z.; Zhu, G. Antioxidant effects of Lactobacillus plantarum via activation of transcription factor Nrf2. Food Funct. 2013, 4, 982–989. [Google Scholar] [CrossRef]
  172. Diao, Y.; Xin, Y.; Zhou, Y.; Li, N.; Pan, X.; Qi, S.; Qi, Z.; Xu, Y.; Luo, L.; Wan, H. Extracellular polysaccharide from Bacillus sp. strain LBP32 prevents LPS-induced inflammation in RAW 264.7 macrophages by inhibiting NF-κB and MAPKs activation and ROS production. Int. Immunopharmacol. 2014, 18, 12–19. [Google Scholar] [CrossRef] [PubMed]
  173. Riccio, G.; Lauritano, C. Microalgae with immunomodulatory activities. Mar. Drugs 2020, 18, 2. [Google Scholar] [CrossRef] [Green Version]
  174. Moreno-Garcia, L.; Adjallé, K.; Barnabé, S.; Raghavan, G. Microalgae biomass production for a biorefinery system: Recent advances and the way towards sustainability. Renew. Sustain. Energy Rev. 2017, 76, 493–506. [Google Scholar] [CrossRef]
  175. Desbois, A.P.; Mearns-Spragg, A.; Smith, V.J. A fatty acid from the diatom Phaeodactylum tricornutum is antibacterial against diverse bacteria including multi-resistant Staphylococcus aureus (MRSA). Mar. Biotechnol. 2009, 11, 45–52. [Google Scholar] [CrossRef] [PubMed]
  176. Arguelles, E. Proximate analysis, antibacterial activity, total phenolic content and antioxidant capacity of a green microalga Scenedesmus quadricauda (Turpin) Brébisson. Asian J. Microbiol. Biotechnol. Environ. Sci. 2018, 20, 150–158. [Google Scholar]
  177. Maadane, A.; Merghoub, N.; Mernissi, N.E.; Ainane, T.; Amzazi, S.; El Arroussi, H.; Benhima, R.; Amzazi, S.; Bakri, Y.; Wahby, I. Antioxidant activity of some Moroccan marine microalgae: Pufa profiles, carotenoids and phenolic content. J. Biotechnol. 2015, 215, 13–19. [Google Scholar] [CrossRef]
  178. Zielinski, D.; Fraczyk, J.; Debowski, M.; Zielinski, M.; Kaminski, Z.J.; Kregiel, D.; Jacob, C.; Kolesinska, B. Biological activity of hydrophilic extract of Chlorella vulgaris grown on post-fermentation leachate from a biogas plant supplied with stillage and maize silage. Molecules 2020, 25, 1790. [Google Scholar] [CrossRef] [Green Version]
  179. Alsenani, F.; Tupally, K.R.; Chua, E.T.; Eltanahy, E.; Alsufyani, H.; Parekh, H.S.; Schenk, P.M. Evaluation of microalgae and cyanobacteria as potential sources of antimicrobial compounds. Saudi Pharm. J. 2020, 28, 1834–1841. [Google Scholar] [CrossRef] [PubMed]
  180. Maadane, A.; Merghoub, N.; Mernissi, N.E.; Ainane, T.; Amzazi, S.; Bakri, I.W. Antimicrobial activity of marine microalgae isolated from Moroccan coastlines. J. Microbiol. Biotechnol. Food Sci. 2021, 2021, 1257–1260. [Google Scholar] [CrossRef] [Green Version]
  181. Washida, K.; Koyama, T.; Yamada, K.; Kita, M.; Uemura, D. Karatungiols A and B, two novel antimicrobial polyol compounds, from the symbiotic marine dinoflagellate Amphidinium sp. Tetrahedron Lett. 2006, 47, 2521–2525. [Google Scholar] [CrossRef]
  182. Ghaidaa, H.; Neihaya, H.; Nada, Z.M.; Amna, M. The Biofilm Inhibitory Potential of Compound Produced from Chlamydomonas reinhardtii Against Pathogenic Microorganisms. Baghdad Sci. J. 2020, 17, 34–41. [Google Scholar]
  183. Pratt, R.; Daniels, T.; Eiler, J.J.; Gunnison, J.; Kumler, W.; Oneto, J.F.; Strait, L.A.; Spoehr, H.; Hardin, G.; Milner, H. Chlorellin, an antibacterial substance from Chlorella. Science 1944, 99, 351–352. [Google Scholar] [CrossRef]
  184. Bancalari, E.; Martelli, F.; Bernini, V.; Neviani, E.; Gatti, M. Bacteriostatic or bactericidal? Impedometric measurements to test the antimicrobial activity of Arthrospira platensis extract. Food Control 2020, 118, 107380. [Google Scholar] [CrossRef]
  185. Guzmán, F.; Wong, G.; Román, T.; Cárdenas, C.; Alvárez, C.; Schmitt, P.; Albericio, F.; Rojas, V. Identification of antimicrobial peptides from the microalgae Tetraselmis suecica (Kylin) Butcher and bactericidal activity improvement. Mar. Drugs 2019, 17, 453. [Google Scholar] [CrossRef] [Green Version]
  186. Mendiola, J.A.; Torres, C.F.; Toré, A.; Martín-Álvarez, P.J.; Santoyo, S.; Arredondo, B.O.; Señoráns, F.J.; Cifuentes, A.; Ibáñez, E. Use of supercritical CO2 to obtain extracts with antimicrobial activity from Chaetoceros muelleri microalga. A correlation with their lipidic content. Eur. Food Res. Technol. 2007, 224, 505–510. [Google Scholar] [CrossRef]
  187. Kokou, F.; Makridis, P.; Kentouri, M.; Divanach, P. Antibacterial activity in microalgae cultures. Aquac. Res. 2012, 43, 1520–1527. [Google Scholar] [CrossRef]
  188. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2015, 32, 116–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Jones, E.; Salin, V.; Williams, G. Nisin and the Market for Commericial Bacteriocins; Texas Agribusiness Market Research Center (TAMRC) Consumer and Product Research Report No; Texas A&M University: College Station, TX, USA, 2005. [Google Scholar]
  190. Soltani, S.; Hammami, R.; Cotter, P.D.; Rebuffat, S.; Said, L.B.; Gaudreau, H.; Bédard, F.; Biron, E.; Drider, D.; Fliss, I. Bacteriocins as a new generation of antimicrobials: Toxicity aspects and regulations. FEMS Microbiol. Rev. 2021, 45, 1–24. [Google Scholar] [CrossRef] [PubMed]
  191. Xing, L.; Westphal, A. Inhibition of Fusarium virgulifomre by prokaryotes in vitro. Subtrop. Plant. Sci. 2007, 59, 24–29. [Google Scholar]
  192. Sauermann, R.; Rothenburger, M.; Graninger, W.; Joukhadar, C. Daptomycin: A review 4 years after first approval. Pharmacology 2008, 81, 79–91. [Google Scholar] [CrossRef]
  193. Fazle Rabbee, M.; Baek, K.-H. Antimicrobial activities of lipopeptides and polyketides of Bacillus velezensis for agricultural applications. Molecules 2020, 25, 4973. [Google Scholar] [CrossRef]
  194. Karim, A.; Brown, E.A. Isolation and identification of novel sulfur-containing metabolites of spironolactone (Aldactone®). Steroids 1972, 20, 41–62. [Google Scholar] [CrossRef]
Table 5. Antimicrobial compounds extracted from endophytic fungi.
Table 5. Antimicrobial compounds extracted from endophytic fungi.
CompoundChemical StructureProducerActive AgainstHostReferences
e 1, 4-naphthoquinone derivatives-Talaromyces sp.
SK-S009
Pseudomonas sp.Kandelia obovata[128]
Clavatol Biomolecules 11 01860 i009Aspergillus clavatonanicus,
Aspergillus elegans KUFA0015
Botrytiscinerea, Didymella bryoniae, Fusarium oxysporum f. sp. cucumerinum, Rhizoctonia solani, and Pythium ultimumTaxus mairei,
Monanchora unguiculate (Marine sponge)
[129]
Lactones Biomolecules 11 01860 i010Phomopsis sp. YM 311483A. niger, Botrytis cinere, and FusariumAzadirachta indica[130]
Jesterone Biomolecules 11 01860 i011Pestalotiopsis jesteriPythium ultimum, Phytophthora citrophthora, Rhizoctonia solani and Sclerotinia sclerotiorumFragraea bodenii[131]
Peniciadametizine A Biomolecules 11 01860 i012Penicillium adametzioides AS-53, Penicillium janthinellum strain HDN13-309Alternaria brassicaSponge collected at the Hainan Island of China, roots of Sonneratia caseolaris[132]
Table 6. Antimicrobial compounds extracted from marine fungi.
Table 6. Antimicrobial compounds extracted from marine fungi.
CompoundsStructureProducerActive AgainstEnvironment SourceReferences
Penicisteroid A Biomolecules 11 01860 i013Penicillium chrysogenum
QEN-24S
A. niger and Alternaria brassicaeMarine algae associated Penicillium sp.[135]
Arisugacin K Biomolecules 11 01860 i014P. echinulatumE. coliMarine alga Chondrus ocellatus.[136]
Methyl (Z)-3-(3, 4-dihydroxyphenyl)-2-
Formamidoacrylate
Biomolecules 11 01860 i015P. oxalicum EN-
290
S. aureusMarine algae associated Penicillium [137]
Chermesins A and B Biomolecules 11 01860 i016P. chermesinum EN-480,C. albicans,
E. coli, M. luteus, and V. alginolyticus
Marine algae associated Penicillium [138]
Comazaphilones C (C–E) Biomolecules 11 01860 i017P. commune QSD-17AntibacterialPenicillium sp. from marine sediments[139]
Penicibilaenes A-P. bilaiae MA-267Colletotrichum gloeosporioidesRhizospheric soil of Lumnitzera racemosa[140]
Xylarinonericin D and E Biomolecules 11 01860 i018Penicillium sp. H1Fusarium oxysporum f. sp. Cubense (antifungal)Beibu Gulf nearby Guangxi[136]
Terretonin G Biomolecules 11 01860 i019Aspergillus sp. OPMF00272S. aureus FDA209P, Bacillus
subtillis PCI219 and Micrococus
luteus (ATCC9341)
Ishigaki island[141]
Schevalone E -A. similanensis sp. nov.MRSASponge Rhabdermia sp. from the coral reef of the Similan Island[142]
Asperitaconic
acids A–C
-A. niger LS11S. aureusSponges-associated Aspergillus sp.[143]
Ochramide B Biomolecules 11 01860 i020A. ochraceus LCJ11-102Enterobacter
aerogenes
Marine sponge Dichotella gemmacea [144]
Spiculisporic
acids F and G
-A. candidus HDf2P. solanacearum and S. aureusMarine animals associated Aspergillus sp.[145]
Aspergicin Biomolecules 11 01860 i021Aspergillus sp. FSY-01 and Aspergillus sp.
FSW-02
S. aureus, S. epidermidis, B. subtilis, B. dysenteriae, B.
proteus and E. coli,
Mangrove Avicennia marina in Guangdong.[146]
Asperamide Biomolecules 11 01860 i022A. niger EN-13C. albicansMarine algae associated Aspergillus sp.[147]
Flavusides A and B Biomolecules 11 01860 i023A. flavusS. aureus
and MRSA
Marine algae associated Aspergillus sp.[148]
Isorhodoptilometrin-
1-methyl ether
Biomolecules 11 01860 i024A. versicolorB. subtilis, B. cereus and
S. aureus
Marine algae associated Aspergillus sp.[149]
Asperterrein Biomolecules 11 01860 i025Paecilomyces lilacinus EN-531 and A. terreus EN-539Alternaria brassicae, E. coli, Edwardsiella
tarda, Physalospora piricola, and
S. aureus
Marine algae associated Aspergillus sp.[150]
Speradine A Biomolecules 11 01860 i026A. tamarii M143Mycrococcus luteusdriftwood in Okinawa[136]
Versiperol A Biomolecules 11 01860 i027A. versicolor MCCC 3A00080S. aureusseawater-associated Aspergillus sp.[151]
Ergosterdiacids A and B Biomolecules 11 01860 i028Aspergillus sp.M. tuberculosisMarine sediments
associated Aspergillus sp.
[152]
Heptapeptide RHM1-Acremonium sp. HM1S. epidermidisMarine sponges-
associated fungi
Trichoderins A Biomolecules 11 01860 i029Trichoderma sp. 05FI48M. smegmatisMarine sponges-
associated fungi
[153]
Botryorhodines I and J -Setosphaeria sp. SCSIO
41009
Colletotrichum
asianum
Marine sponge
Callyspongia sp.
[154]
Table 7. Selected antimicrobial extracts from microalgae.
Table 7. Selected antimicrobial extracts from microalgae.
MicroalgaeTarget MicroorganismActive ExtractReferences
Scenedesmus quadricaudaS. aureus and P. aeruginosaMethanolic extract[176]
Tetraselmis sp.E. coli, P. aeruginosa, and S. aureusEthanolic extract[177]
Phaeodactylum tricornutumListonella anguillarum, Lactococcus garvieae, Vibrio spp. and MRSAEicosapentaenoic acid[175]
C. vulgarisSteinernema feltiaeHydrophilic extracts[178]
Skeletonema costatumListeria monocytogenesExtra-metabolites
S. costatumVibrio spp., Pseudomonas sp. and Listeria monocytogenesUnsaturated, saturated
long-chain fatty acids
[179]
Haematococcus pluvialisE. coli, S. aureus, Candida albicansShort-chain fatty acids
(butanoic acid and
methyl lactate), Astaxanthin
[180]
Amphidinium sp.A. niger, Trichomonas foetusKaratungiols[181]
Chlamydomonas reinhardtiiA. niger, A. fumigatus, C. albicans, S. aureus and E. coliMethanolic extracts[182]
Table 8. List of selected antimicrobial compounds with commercial trade name and uses.
Table 8. List of selected antimicrobial compounds with commercial trade name and uses.
CompoundsBrand NameCompany NameCountryUsesReferences
NisinNisaplin®DaniscoDenmarkUsed as a food preservative[189]
NisinNovasinDaniscoDenmarkUsed as a food preservative[189]
NisinDelvo®NisDSMNetherlandsUsed as a food preservative[190]
NisinChrisin®Chris HansenDenmarkUsed as a food preservative[190]
Nisin-Duke Thomson’s InternationalIndiaUsed as a food preservative[191]
Nisin-Ecobio Biotech Co. Ltd.ChinaUsed as a food preservative[191]
Delvocid-Natamycin-Duke Thomson’s InternationalIndiaUsed as a food preservative[191]
NatamycinDelvocidDSMNetherlandsUsed as a food preservative[191]
DaptomycinCubicinNovartis India Ltd.IndiaUsed to treat bacterial infections[192]
LipopeptidesRhizoVital®ABiTEP, GmbHGermanyBiological control in agriculture[191]
LipopeptidesKodiakGustafson Inc.USABiological control in agriculture[191]
LipopeptidesTaegro®NovozymesUSABiological control in agriculture[191]
LipopeptidesSerenade®AgraQuest Inc.USABiological control in agriculture[191]
LipopeptidesBotrybelAgricaldesSpainBiological control in agriculture[193]
SpironolactoneAldactone®Pfizer MedicalUSATo treat various diseases[194]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rani, A.; Saini, K.C.; Bast, F.; Varjani, S.; Mehariya, S.; Bhatia, S.K.; Sharma, N.; Funk, C. A Review on Microbial Products and Their Perspective Application as Antimicrobial Agents. Biomolecules 2021, 11, 1860. https://doi.org/10.3390/biom11121860

AMA Style

Rani A, Saini KC, Bast F, Varjani S, Mehariya S, Bhatia SK, Sharma N, Funk C. A Review on Microbial Products and Their Perspective Application as Antimicrobial Agents. Biomolecules. 2021; 11(12):1860. https://doi.org/10.3390/biom11121860

Chicago/Turabian Style

Rani, Alka, Khem Chand Saini, Felix Bast, Sunita Varjani, Sanjeet Mehariya, Shashi Kant Bhatia, Neeta Sharma, and Christiane Funk. 2021. "A Review on Microbial Products and Their Perspective Application as Antimicrobial Agents" Biomolecules 11, no. 12: 1860. https://doi.org/10.3390/biom11121860

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