Next Article in Journal
Antimicrobial Stewardship in College and University Health Settings: A Public Health Opportunity
Next Article in Special Issue
Curcumin: Biological Activities and Modern Pharmaceutical Forms
Previous Article in Journal
The Effect of Antibiotics on Planktonic Cells and Biofilm Formation Ability of Collected Arcobacter-like Strains and Strains Isolated within the Czech Republic
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antimicrobial Bacillus: Metabolites and Their Mode of Action

1
Griffith Institute for Drug Discovery (GRIDD), Griffith University, Nathan, QLD 4111, Australia
2
School of Environment and Science, Griffith University, Nathan, QLD 4111, Australia
3
Bioproton Pty Ltd., Acacia Ridge, QLD 4110, Australia
*
Author to whom correspondence should be addressed.
Antibiotics 2022, 11(1), 88; https://doi.org/10.3390/antibiotics11010088
Submission received: 14 December 2021 / Revised: 7 January 2022 / Accepted: 8 January 2022 / Published: 12 January 2022
(This article belongs to the Special Issue Antimicrobial Natural Products)

Abstract

:
The agricultural industry utilizes antibiotic growth promoters to promote livestock growth and health. However, the World Health Organization has raised concerns over the ongoing spread of antibiotic resistance transmission in the populace, leading to its subsequent ban in several countries, especially in the European Union. These restrictions have translated into an increase in pathogenic outbreaks in the agricultural industry, highlighting the need for an economically viable, non-toxic, and renewable alternative to antibiotics in livestock. Probiotics inhibit pathogen growth, promote a beneficial microbiota, regulate the immune response of its host, enhance feed conversion to nutrients, and form biofilms that block further infection. Commonly used lactic acid bacteria probiotics are vulnerable to the harsh conditions of the upper gastrointestinal system, leading to novel research using spore-forming bacteria from the genus Bacillus. However, the exact mechanisms behind Bacillus probiotics remain unexplored. This review tackles this issue, by reporting antimicrobial compounds produced from Bacillus strains, their proposed mechanisms of action, and any gaps in the mechanism studies of these compounds. Lastly, this paper explores omics approaches to clarify the mechanisms behind Bacillus probiotics.

1. Introduction

Probiotics are live microorganisms that can be consumed by its host to confer a range of health benefits. These benefits include the production of antimicrobial metabolites, restoration of the host microbiota, modulation of the immune system, and the release of digestive enzymes to improve nutrient uptake [1]. For example, Bacillus subtilis MA139 restored microbiota diversity in finishing pigs, improved their resistance to pathogenic illnesses, and promoted animal health and growth [2]. This increase in animal production makes probiotics a suitable alternative to antibiotic use in animals, due to the WHO advocating for its restricted use and its subsequent ban by the EU in 2006 [3].
Probiotics are commonly used in animal feed production, which do not contribute to antibiotic resistance and may even reduce it [4]. Selective probiotic bacteria have been used to treat antibiotic-associated diarrhea (AAD), a common side-effect of antibiotic use. Antibiotics elevate the risk of AAD by disrupting the diversity of the gut biota, allowing the proliferation of opportunistic pathogens such as Clostridium difficile [5]. This issue can be tackled through the use of probiotics, which inhibit pathogen growth and restabilize the intestinal microbiota back to normal levels [6]. Furthermore, probiotics can bind to the intestinal walls of its host and competitively exclude competing pathogens. Additionally, these probiotics produce a plethora of antimicrobial compounds that target pathogenic bacteria, which has driven the search for a potent probiotic strain for industrial use.
The issue lies in the presence of antibiotic resistance genes, with the commonly used Lactobacillus showing frequent resistance to vancomycin, ciprofloxacin, and aminoglycosides [7]. This development has driven the research into other probiotic genera not yet explored such as Bacillus. Bacillus probiotics are pore-forming bacteria that can survive the harsh conditions needed for pelletizing and can tolerate aerobic conditions for industrial use, unlike Lactobacillus and Bifidobacterium [8].
Several review papers have been published in the literature summarizing Bacillus metabolites, structural classes, and their antimicrobial activities [9,10,11]. However, no literature is available investigating the mechanisms of action of the antimicrobial metabolites from Bacillus. In this review, we summarized 47 antimicrobial compounds based on their molecular targets in the cell wall, plasma membrane, intracellular processes, and other emerging targets.

2. A Glance of Bioactive Bacillus and Their Antimicrobial Metabolites

To gain a good understanding of antimicrobial Bacillus sp., and hence their potential as a probiotic supplement, we conducted a literature review on antimicrobial Bacillus. Google-Scholar, PubMed, Scopus, and Science-Direct electronic databases were used to identify original scientific research papers. The terms ‘antimicrobial Bacillus’ and ‘mechanism of action’ were used as filters, with the earliest possible time range. Our literature search revealed that 1389 Bacillus strains have been reported for antimicrobial activity, composed of 27 different species (Figure 1). The most commonly reported species included subtilis (n = 348), amyloliquefaciens (n = 214), licheniformis (n = 114), circulans (n = 89), thuringiensis (n = 73), pumilus (n = 61), velezensis (n = 60), megaterium (n = 17), and mojavensis (n = 17) (Figure 1). The literature review also suggested that a substantial number of Bacillus species were not identified (n = 293). From the antimicrobial Bacillus sp., 47 metabolites have been identified and their mechanisms of actions reported [12]. We herein report the chemical structures of the metabolites, their antimicrobial activity, and mechanism of action. Details regarding these compounds, including source strain, anti-microbial activity, molecular target, and references are provided in Supplementary Table S1.

3. Antimicrobial Metabolites and Their Mechanism of Action

3.1. Metabolites Targeting the Cell Wall

The cell wall is a selectively permeable layer that has a distinct layer of polysaccharides, peptidoglycans, and fungi-specific chitins and glucans [13]. This structure is located outside the plasma membrane and acts as a permeable barrier, which regulates the entry of metabolites into the cell and protects it against external stresses (Figure 2a). The cell wall is a promising target for drug development due to its absence in mammalian cells, and several Bacillus strains have been shown to target this structure by releasing enzymes (amylase, cellulase, chitinase, chitosanase, glucanase, and protease) and antimicrobial metabolites. From the reported 47 compounds with clearly defined mechanisms, 9 compounds target the cell wall (Figure 2a).
The peptidoglycan layer provides integrity and protection to the cell. This layer is comprised of linear glycan strands, which alternate between N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc) residues linked by β-1-4 bonds [14]. Bacitracin, an antibiotic first isolated from B. licheniformis, primarily acts on gram-positive bacteria such as Streptococcus mutans (MIC = 78.12 μg/mL) [15,16]. This antibiotic is comprised of a mixture of compounds, which include bacitracin A (1), B and C. Bacitracin A (Figure 3) prevents the dephosphorylation of undecaprenyl pyrophosphate (C55-PP) to undecaprenyl phosphate (C55-P), which prevents the formation of lipid I/II and the eventual disruption of the peptidoglycan layer [17]. Additionally, recent scanning-electron microscopy (SEM) analysis has shown that bacitracin inhibits the formation of biofilm by Streptococcus mutans by downregulating several genes related to cell division and biofilm [16].
Glucosamine-6-phosphate synthetase (G6PS) is an enzyme that catalyzes the production of uridine diphosphate N-acetylglucosamine (UDP-GlcNAc), which is a precursor for peptidoglycan synthesis [18]. Bacilysin (2) and its chlorinated derivative chlorotetaine (3) (Figure 3) were first isolated from B. subtilis A14 and B. amyloliquefaciens ZJU-2011, respectively [19,20]. Both compounds are active against a broad range of bacteria, with bacilysin inhibiting E. coli at MIC = 0.001 µg/mL and chlorotetaine inhibiting Candidas spp. and Aspergillus niger at an MIC value of 1.8–7.8 μg/mL [20,21]. Bacilysin first enters the cell by binding to a transmembrane transport protein and is subsequently hydrolyzed to anticapsin, a G6PS inhibitor [22]. Kanosamine (4) (Figure 3) produced from B. cereus UW85 inhibits a wide array of plant-related pathogens (i.e., Phytophthora medicaginis M2913 with an MIC = 25 µg/mL) [23]. Kanosmine inhibits Candida albicans by utilizing the glucose transport system to transport itself into the cell, where it is subsequently phosphorylated to kanosamine-6-phosphate [24]. Kanosmine-6-phosphate inhibits G6PS, leading to the in septum deformation and cell agglutination of C. albicans.
Lipid II is a peptidoglycan intermediate, which is formed when the glycosyltransferase MurG catalyzes the addition of N-acetylglucosamine (GlcNAc) to lipid I [25]. Lipid II subsequently translocates across the plasma membrane, where it transfers MurNaC and GlcNAc to the peptidoglycan layer [26] (Figure 2a). Lipid II is generally conserved throughout microbes and has been studied as a target for various antimicrobial compounds, especially lantibiotics [27]. Lantibiotics are a class of large ribosomal compounds, typically around 3000kDa, and contain unique lanthionine and β-methyllanthionine residues [28]. These lantibiotics are often further divided based on the enzymes involved in their biosynthesis, which includes class I (5, 6) and class II (7, 8, 9) lantibiotics (Figure 4). Subtilin (5) is a class I lantibiotic isolated from B. subtilis 6633 [29]. This metabolite inhibits gram-positive bacteria, with MIC of 0.05 µg/mL (Micrococcus luteus NDCO8166) [29]. Binding studies show that subtilin binds to lipid II and pyrophosphate-containing intermediates. These pyrophosphate intermediates coat the outer cell membranes, and subtilin attaches to these intermediates, forming membrane pores [30]. These pores release essential metabolites, which eventually lead to cell death. Clausin (6), a class I lantibiotic produced by B. clausii O/C, inhibits gram-positive microbes (e.g., Micrococcus luteus, MRSA with MICs = 16 mg/L and 128 mg/L respectively) [31,32]. Clausin interacts with both lipid I/II and GlcNAc, forming stable complexes, which obstruct its role in peptidoglycan biosynthesis and hindering microbial growth [31].
A class II lantibiotic, mersacidin (7), was first isolated from Bacillus sp. HIL Y-85,54728 and shows activity against a range of gram-positive bacteria including Staphylococcus aureus SG511 with an MIC = 1 μg/mL [33,34]. Mersacidin associates with lipid II, which interferes with peptidoglycan biosynthesis and obstructs the growth of the microbe [35]. The class II lantibiotic amylolysin A (8), produced by B. amyloliquefaciens GA1, targets gram-positive bacteria such as Enterococcus faecium RFB128 with a MIC = 0.3µg/mL [36]. Amylolysin A exerts its antimicrobial effect by two separate mechanisms [37]. First, amylolysin A interacts with lipid II to hinder the biosynthesis of peptidoglycan. Secondly, amylolysin A induces the formation of membrane pores, leading to cell lysis. Haloduracin (9), a class II lantibiotic isolated from B. halodurans C-125, targets gram-positive bacteria such as Lactococcus lactis HP ATCC 11602 (MIC = 0.4 μg/mL) [38]. Structural analysis has highlighted that haloduracin is comprised of two parts, Halα and Halβ. Halα binds to lipid II in a 2:1 stoichiometry, preventing peptidoglycan biosynthesis. Halβ (2330 Da), however, binds to the anionic lipids of the cell membrane, resulting in pore formation [39].

3.2. Metabolites Targeting Plasma Membrane

The plasma membrane is composed of a phospholipid bilayer, which separates the intracellular compartment from the extracellular environment and may selectively transport metabolites across the membrane [40]. From the reviewed 47 Bacillus metabolites, 23 were identified to target different processes of the cell membrane (Figure 2b).
The lipid bilayer controls the permeability and shape of the plasma membrane and is affected by the negative-charged outer phospholipid layer [41]. Any changes to this membrane, whether by altering its lipid composition or the phospholipid layer, may distort its function as a barrier to the extracellular environment, releasing essential ions from the cell, eventually leading to cell death. ε-Poly-L-lysine (10) (Figure 5) is a homopolymer produced from B. subtilis SDNS, which exerts antimicrobial activity against gram-positive and gram-negative bacteria, as well as fungi (e.g., 600 μg/mL for Ralstonia solanacearum) [42]. ε-Poly-L-lysine electrostatically attaches to the phospholipid layer of the plasma membrane, which disturbs the membrane permeability to eventually lead to cell death [43,44]. Plantazolicin (11) (Figure 5), a product of B. velezensis FZB42, has been identified as a bacteriocin of interest, due to its restrictive spectrum against clinically relevant pathogens, such as B. anthracis, with an MIC value of 1–16 μg/mL [45]. This is highly relevant due to the very serious nature of anthrax. Further mechanism studies revealed that plantazolicin induces higher membrane fluidity and increases the proportion of cardiolipin, a cholesterol associated with higher osmotic stress [45]. Octapeptins are a class of lipooctapeptide antibiotics that were first isolated from B. circulans and that primarily inhibit gram-negative bacteria, with weaker activity on gram-positive bacteria and fungi [46]. Membrane microscopy studies show that octapeptin B (12) (E. coli SC 9251 MIC = 0.3 µg/mL) (Figure 5), produced from B. circulans ATCC 21656, disrupts the ion permeability of the membrane, which reduces the membrane proton gradient [47]. This translates into extensive membrane damage, the efflux of charged metabolites, and cell lysis. The aurantinins B-D (13–15) (Figure 5), a class of metabolites isolated from B. subtilis FMB60, exhibit similar MIC value for certain clinically relevant strains (i.e., Clostridium sporogenes CICC 10385 with a MIC ≤ 0.78 µg/mL, methicillin-resistant Staphylococcus aureus (MRSA) with an MIC = 6.25 µg/mL) [48]. SEM and transmission electron microscopy (TEM) studies show that the aurantinins cause plasma membrane lysis, leading to the efflux of metabolites from the cytoplasm [48]. However, these compounds require further structural elucidation to determine their precise stereochemistry. Myriocin (16) (Figure 5), produced from B. amyloliquefaciens LZN01, exerts antifungal activities against Candidas albicans (MIC = 1.0 µg/mL) [49]. SEM and TEM microscopy studies have indicated that myriocin binds to serine palmitoyl transferase and disrupts the plasma membrane, causing leakage and eventual pore formation [50]. Further omics analysis has revealed that myriocin alters the expression changes related to sphingolipid metabolism, glycerophospholipid metabolism, steroid biosynthesis, ABC transporters, and protein processing [51]. These genes are all relevant to the plasma membrane, suggesting that myriocin may target the expression of DNA. Gramcidins are a class of antibiotic decapeptides synthesized by Aneurinibacillus migulanus (formerly B. brevis) and consist of linear gramicidin A, B, C, and the circular gramicidin S. Gramicidin A (17) (Figure 5), a 15 amino-acid peptide, destroys gram-positive bacteria (Streptococcus pyogenes with a MIC = 33 nM) [52]. Unlike other antimicrobial metabolites, gramicidin A forms a single ion channel, which distorts the membrane and allows the passage of cations across the membrane [53]. Once inside, gramicidin A can also induce the formation of reactive oxygen species (ROS), which damages the intracellular DNA, mitochondria and triggers necrosis [54]. The gram-positive bacteria Aneurinibacillus migulanus (formerly B. brevis natto) inhibits several gram-positive, gram-negative, and fungi microbials (e.g., Staphylococcus aureus with a MIC value of 3.9 µg/mL) by producing gramicidin S (18) (Figure 5) [55]. Gramicidin S interacts with the plasma membrane by forming oligomeric β-barrel pores, which destroys the barrier properties of the membrane [56,57]. Further in vivo studies have shown that gramicidin S binds to the DNA and inhibits transcription and cell growth [58].
Pore-formation metabolites act in a concentration-dependent manner, by forming ion-like channels that release vital ions from the cell, leading to cell death. At low concentrations, these metabolites form unilamellar vesicles on the outer lipid membrane, distorting the shape of the cell, and eventually, lead to apoptosis [59,60,61]. At higher concentrations, these metabolites aggregate to form pores at the plasma membrane, causing the leakage of nucleic acids, essential ions, and ATP from the cell to cause necrosis [59,62,63,64]. Bacillus metabolites that typically utilize this mechanism includes the class of compounds known as lipopeptides. Lipopeptides are composed of a cyclic oligopeptide, attached to a flexible lipid tail, and consist of several groups including the surfactins, fengycins, and iturins [65]. Surfactins were first isolated from a culture broth of B. subtilis and include the compounds surfactin A (19), B (20), C (21), and lichenysin (22) (Figure 6) [66]. Surfactins exert their antibacterial activities by acting on the plasma membrane through the pore-forming mechanism [67]. Additionally, surfactins (21-22) can breakdown bacterial biofilms by decreasing the percentage of alkali-soluble polysaccharides and downregulating the expression of genes involved in biofilm formation such as icaA and icaD [68]. Lastly, surfactins can also induce the grapevine immune system in response to infection [69]. Fengycins (23-26) (Figure 6) are antifungal lipopeptides first isolated from B. subtilis F-29-3 (e.g., Rolani stolonifera with a MIC = 400 µg/mL) [70]. These fengycin molecules are often reported as membrane disruptors, either by deforming membrane shape or by causing pores, leading to cell death [71]. More recent studies have additional antimicrobial mechanisms of action for fengycin A (23) and fengycin B (24). Fengycin A can alter the gene expression related to cell wall synthesis, which alters cell components and increases hydrophobicity [72]. Furthermore, fengycin B155, a mixture of fengycin A (23) and fengycin B (24), is able to disrupt multiple intracellular components of the cell [73]. These processes include the inhibition of the mitochondria membrane potential, the condensation of chromatin involved in replication, the cleavage of DNA repair protein (poly (ARP-ribose) polymerase), and the accumulation of ROS [73]. Lastly, fengycins have been shown to inhibit quorum sensing, due to their structural similarity to S. aureus accessory gene regulator (Agr) [74]. Agr is a virulence factor that mediates the cell-to-cell communication between cells, and its inhibition prevents the aggregation and biofilm formation needed to promote survival [75]. Plipastatin A (26) is a lipopeptide commonly associated with the fengycin family due to its structural similarity and antifungal properties (Fusarium oxysporum with a MIC = 16 μg/mL) [76]. TEM analysis demonstrated that plipastatins disrupt the cell wall, membrane, and cytoskeleton of Fusarium oxysporum, causing intracellular leakage and eventual cell death.
Iturins (27–30) (Figure 7) are cyclic lipopeptides that includes iturin A (27), bacillomycin D (28), bacillomycin L (29), and mycosubtilin (30) [77]. These peptides primarily inhibit fungi by binding to the cell membrane with its fatty acid tail to form ion-conducting or phospholipid–lipopeptide sterol complexes [78]. Optical and fluorescence microscopy studies have revealed that iturin A (27) severely damages the plasma membranes of Fusarium graminearum at a MIC = 5 µg/mL by forming a large pore and inhibiting hyphae growth [79]. Iturin A can stimulate oxidative stress, leading to mitochondria damage and the eventual destruction of the cell [80]. Lastly, iturin A increases the transcription of immune defense genes in several plants [81]. Bacillomycin D (28) exerts antifungal properties against Colletotrichum gloeosporioides with an MIC of 2.2 µg/mL [82,83]. SEM and TEM analysis confirmed bacillomycin D’s ability to target both cell wall and plasma membrane, leading to the leakage of intracellular organelles [82]. Bacillomycin D can disrupt the cell membrane by upregulating the expression of genes involved in ergosterol synthesis and oxidative stress [84]. These sterols adjoin to the membrane, distorting its shape and eventually releasing vital intracellular components to the environment [84]. Additionally, bacillomycin D can increase the expression of specific genes to produce ROS molecules and cellular antioxidant enzymes including deoxyivalentol, glutathione reductase, and thioredoxin [85]. Bacillomycin D has also been reported to act as a biofilm activator by binding to the matrix complex KinB-Spo0A-SinI-SinR, which triggers the production of biofilm [86]. Lastly, bacillomycin D stimulates the expression of genes involved in mediated defense responses and enzymatic proteins that can be released to target competing growth [86]. B. amyloliquefaciens K103 produces the potent antifungal metabolite bacillomycin L (29) (Saccaromyces cerevisiae with a MIC = 30 µg/mL) [78,87]. Like other iturins, bacillomycin L primarily acts on the plasma membrane, forming pores that releases its intracellular components outside the cell [88]. Studies have shown that bacillomycin L binds to sterols on the membrane, destroying the membrane and killing the cell [89]. Bacillomycin L can also alter the expression of 39 different genes in Rhizoctonia solani related to cellular stress, such as calcium homeostasis, energy metabolism, protein degradation, RNA processing, and carbohydrate metabolism [90]. Mycosubtilin (30), an antibiotic from the iturin group, inhibits the growth of fungal Saccharomyces cerevisiae with a MIC of 10 µg/mL [78]. Increased concentrations of mycosubtilin causes the lysis of the phospholipid layer, either by the aggregation of lipopeptides or clustering of mycosubtilin [91]. This binding increases membrane permeability, leading to metabolite release and the eventual lysis of the cell [92]. Mycosubtilin can also activate the salicylic acid and jasmonic acid signaling pathways involved in the immune response to pathogenic microbes [69].
Mycobacillin (31) (Figure 8), an antifungal polypeptide sourced from B. subtilis B3, is active against Aspergillus niger at 20 µg/mL [93,94]. Mycobacillin has been reported to bind to ATP transporter on the plasma membrane, leading to the excessive release of ATP and the subsequent starvation of the cells [94,95]. Subtilosin A (32) (Figure 8) is a sactipeptide produced by B. subtilis 168 that processes antibacterial activity against both gram-positive and gram-negative pathogens (i.e., Gardnerella vaginalis MIC = 7.2 µg/mL) [96,97]. Its specific mechanism of action involves subtilosin A anchoring to a membrane receptor, whilst electrostatically binding to the plasma membrane [98]. This electrostatic binding dissipates the transmembrane pH gradient, causing an efflux of intracellular ATP that starves the cell and eventually leads to its death. Subtilosin A has also been shown to inhibit biofilm formation, presumably by blocking quorum sensing between cells [99].

3.3. Metabolites Targeting Intracellular Processes

Bacillus metabolites may cross the plasma membrane and bind to several intracellular targets essential for cell survival. These intracellular processes include DNA transcription, RNA translation, and protein metabolism needed for energy production. Transcription is the first step in gene expression, in which information from a gene is used to construct a functional product such as a protein. For a protein-coding gene, the RNA copy, or transcript, carries the information needed to build a protein. From the 47 compounds reviewed in this paper, 11 compounds primarily target the intracellular processes.
Zwittermicin A (33) (Figure 9), an aminopolyl antibiotic produced by B. cereus UW85, inhibits gram-positive and gram-negative bacteria, as well as fungi (i.e., Erwinia herbicola L S005 with a MIC = 60 µg/mL) [100]. Zwittermicin A disrupts cellular growth by targeting either DNA transcription and replication via inhibition of two enzymes, gyrase and topoisomerase [101]. Difficidin (34) (Figure 9), a highly unsaturated macrolide phosphate first isolated from B. subtilis ATCC 39320, can inhibit both gram-positive and negative strains such as Rolani solanacearum with a MIC value of 12.62 µg/mL of [102,103]. Microscopy analysis has revealed that difficidin downregulates the genes related to cell wall synthesis, protein production, and DNA replication [104]. Sublancin (35) (Figure 9), a glycosylated peptide produced by B. subtilis 168, displays antibacterial activities (i.e., methicillin-resistant Staphylococcus aureus ATCC43300 with a MIC = 15 μM) [105]. Mechanism investigations suggest that sublancin enters the cytoplasm and reduces DNA transcription and translation [106].
The amicoumacins are a class of dihydroisocoumarin compounds, produced by B. pumilus, that exert antibacterial, antifungal, and anti-inflammatory properties. In particular, amicoumacin A (36) (Figure 10), produced by B. pumilus BN-103, inhibits B. subtilis 1779 with an MIC = 20.0 µg/mL. Further studies have shown that amicoumacin A inhibits the protein synthesis of methicillin-resistant Staphylococcus aureus by stabilizing the mRNA at the terminal E site on the ribosome during protein synthesis [107]. This disruption results in the perturbation of the membrane, leading to energy dissipation and eventual cell death [107,108]. Prumycin (37) (Figure 10), isolated from a culture broth of B. amyloliquefaciens SD-32, exerts bactericidal and fungicidal effects, such as on S. sclerotiorum, with an MIC value of 1.56 µg/mL [109,110,111]. Prumycin inhibits the protein synthesis of Sacrina lutea, preventing the activation of amino acids needed for protein synthesis and the transfer of amino acids to RNA [110].
Thiocillin (38) (Figure 10), produced by B. cereus ATCC 14579, has been previously reported to only target gram-positive bacteria but has recently been shown to also target gram-negative bacteria [112]. Its mechanism on gram-positive bacteria works by targeting the 50S ribosome and inhibiting its role in protein synthesis [113]. In contrast, thiocillin targets the gram-negative bacterium Pseudomonas aeruginosa by binding to ferrioxamine receptor FoxA, which disrupts the proton motive force to inhibit translation [113]. Hetiamacin E and F (39–40) (Figure 10) produced from B. subtilis PJS display antibacterial activity against methicillin-resistant Staphylococcus aureus, with MIC values of 8–16 µg/mL and 32 µg/mL, respectively [114]. Hetiamacin E and F inhibit protein biosynthesis, resulting in the disruption of mRNA translation, leading to cell death [114]. Rhizocticin A (41) (Figure 10) is a potent antifungal first produced from B. subtilis 6633. Its bioactivity data shows that it is active against a range of budding and filamentous fungi (bioactivity not avaliable) [115]. Mutant analysis suggests that rhizocticin utilizes the peptide transport system to enter the cytoplasm, where it forms the fungitoxic L-2-amino-5-phosphono-3-cis-pentenioc acid (L-APPA). L-APPA interferes with threonine metabolism, which inhibits cell growth [116].
Macrolactin N (42) (Figure 10), a novel macrolactin produced by B. subtilis A29, is shown to inhibit Staphylococcus aureus peptide deformylase (PDF), with an MIC of 100 µM [117]. PDFs are essential bacterial specific metalloenzymes, which removes formyl groups during polypeptide elongation [117]. The inhibition of these PDFs leave bacteria unable to hydrolyze these polypeptides and hinder its ability to synthesize proteins [117]. Azoxybacilin (43) (Figure 10), first isolated from B. cereus NR2991 and B. cereus Frankland, is active against a broad spectrum of mycelial fungi, such as Candida albicans (IC50 = 1.2 mg/mL) [118,119]. Its mechanism involves the interruption of the sulfur fixation pathway, an essential support system for microbial growth, by decreasing the expression of sulfate assimilation genes including MET10 and MET4 [118]. MET10 regulates the expression of sulfite reductase, and MET4 is the transactivator of MET10. The reduction of the gene expression in the sulfur-fixation pathway disrupts this support system and eventually leads to cell growth inhibition.

3.4. Metabolites Interacting with Other Emerging Targets

Quorum sensing, also known as cell-to-cell communication, is the regulation of a microbial gene expression in response to its cell density [120]. This mechanism relies on small chemical indicators and has been linked to pathogen virulence, due to its effect on cell reproduction, mobility, and biofilm formation [121]. Biofilms are extracellular adhesive structures produced by various strains of bacteria that assist in their tolerance to UV, acidity conditions, and vulnerability to antimicrobial metabolites [122]. Several key groups of Bacillus metabolites have been shown to interfere with this process [123]. Nonetheless, Bacillus metabolites such as stigmatellin Y (44) (Figure 11) have been identified as a biofilm inhibitor [124]. Stigmatellin Y is shown to inhibit Pseudomonas aeruginosa biofilm formation, presumably by acting as a competitive inhibitor to the quorum sensing mediator PqsR [124]. Bacillaene (45) (Figure 11) has been identified as a biofilm inhibitor produced by numerous B. subtilis strains [125]. Analysis of mutant strains revealed that bacillaene inhibits the biofilm of Campylobacter jejuni, preventing the formation of microcolonies and eventually disrupting their microbial growth.
Siderophores are small molecules secreted by microorganisms that are involved in iron (Fe2+) uptake from the environment [126]. Iron is an essential metabolite for microbial growth and strategies have been developed to starve pathogenic microorganisms using these siderophores. Siderophores produced by Bacillus strains include bacillibactin (46) and schizokinen (47) (Figure 11), which were first isolated from B. subtilis and B. megaterium ATCC 19213, respectively [127,128]. These metabolites facilitate the uptake of ferric ions (Fe3+) from the environment to the bacterial cell using specific membrane receptors to enter the host cell [129]. Once inside, these ions are reduced to ferrous (Fe2+) ions for use in microbial growth [130].

4. Conclusions Remarks and Future Directions

This paper reviews the current literature on antimicrobial compounds from Bacillus sp. and their mechanism of action. Further analysis on the source of antimicrobial compounds and their mechanism of action revealed some interesting trends. In terms of number of strains that produce antimicrobial metabolites, the most prolific is subtilis (n = 73), followed by amyloliquefaciens (n = 52) and velezensis (n = 22) (Figure 12a). B. subtilis is a common bacterium in soil and one of the most-studied Bacillus sp. Research has shown that these species are strongly related to each other, with several papers suggesting that amyloliquefaciens be renamed as velezensis due to its similarity in conserved genomic sequence [131,132]. The least reported of these Bacillus sp. is B. thuringiensis, with only two strains producing antimicrobial compounds in the literature. This highlights the lack of studies for this species and may warrant further investigation.
Further analysis on mechanism of action (Figure 12b) reveals that the cell membrane is the most popular target of different species of Bacillus and their metabolites (n = 122), followed by quorum sensing (n = 79), intracellular processes (n = 73), and the cell wall (n = 57). Quorum sensing is an interesting emerging target, as more species and metabolites (n = 79) hinder the process and hence, inhibit cell-to-cell communication. Further analysis also notes that many Bacillus species and their metabolites exert their antimicrobial activity through not only one but multiple mechanisms.
Several publications noted the geographic location of Bacillus, as well as the source of the bacteria. Further analysis based the information provided in the literature reveals that the majority of identified strains are from Asia (n = 37), followed by South America (n = 8) and the Middle East (n = 4). This observation may indicate that these strains share genomic similarities or properties, however, it may also stem from the research laboratories located in these sites and could be a byproduct of a focus on probiotic research at these locations. Additionally, the top three sources that these strains were isolated are from soil, local produce, and waterways. These findings reinforce the use of soil-based screening as a rich source of microorganisms. It also highlights the recent trend in investigating food produce as a source of Bacillus isolates. This is either guided by historical evidence of their antimicrobial properties or the anecdotal knowledge of their safe use and consumption.
The advancements of omics technologies are essential for the rapid screening of future probiotics. The characterization of the genome and biochemical properties allows the selection of particular strains with properties suitable for industrial use. A number of omics techniques have been developed to provide valuable information on the characteristics, optimization, and metabolic pathways behind antimicrobial activity [133]. One example uses omics to a rapid screen of selected Bacillus strains for specific gene markers known for antimicrobial activity [133]. For example, the genomic screening of B. velezensis CC09 revealed the loci for iturin A previously not identified in its initial screening [134].
In-depth analysis of these pathways and the precursors may reveal optimal conditions needed to produce these metabolites [135]. Wiegand utilized metabolomics and genome mining to provide insight into the expression of DNA under various fermentation conditions. These conditions includes pH levels, temperatures, and oxygen levels, which result in the discovery of optimal conditions needed to express the antimicrobial gene of interest and maximizing their yield [136]. This technique, alongside computational modelling systems, may reveal other conditions unexplored such as the ratio of carbon to nitrogen in fermentation media and the presence of small metabolites and co-culturing in order to further maximize the production of antimicrobial metabolites. As production is required, especially when optimizing for commercial purposes, these techniques can open up the field in the use of bacteria as a source of antimicrobial compounds to tackle the declining rate of antimicrobial compounds being discovered.

Supplementary Materials

Author Contributions

Conceptualization, X.C., I.E.C., and Y.F.; data curation, C.T.; writing—original draft preparation, C.T.; writing—review and editing, X.C., I.E.C. and Y.F.; supervision, X.C., I.E.C., and Y.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was supported, in part, by Bioproton Pty. Ltd. (Ph.D. scholarship for C.T.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hill, C.; Guarner, F.; Reid, G.; Gibson, G.R.; Merenstein, D.J.; Pot, B.; Morelli, L.; Canani, R.B.; Flint, H.J.; Salminen, S.; et al. The International Scientific Association for Probiotics and Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat. Rev. Gastroenterol. Hepatol. 2014, 11, 506–514. [Google Scholar] [CrossRef] [Green Version]
  2. Liu, P.; Zhao, J.; Guo, P.; Lu, W.; Geng, Z.; Levesque, C.L.; Johnston, L.J.; Wang, C.; Liu, L.; Zhang, J.; et al. Dietary Corn Bran Fermented by Bacillus subtilis MA139 Decreased Gut Cellulolytic Bacteria and Microbiota Diversity in Finishing Pigs. Front. Cell. Infect. Microbiol. 2017, 7, 526. [Google Scholar] [CrossRef]
  3. Kahn, L.H. Antimicrobial resistance: A One Health perspective. Trans. R. Soc. Trop. Med. Hyg. 2017, 111, 255–260. [Google Scholar] [CrossRef] [PubMed]
  4. Lim, J.M.; Duong, M.C.; Hsu, L.Y.; Tam, C.C. Determinants influencing antibiotic use in Singapore’s small-scale aquaculture sectors: A qualitative study. PLoS ONE 2020, 15, e0228701. [Google Scholar] [CrossRef]
  5. Carlet, J. The gut is the epicentre of antibiotic resistance. Antimicrob. Resist. Infect. Control 2012, 1, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Zhang, W.; Xin, H.; Jiang, N.; Lv, Z.; Shu, J.; Shi, H. Bacillus Amyloliquefaciens-9 as an Alternative Approach to Cure Diarrhea in Saanen Kids. Animals 2021, 11, 592. [Google Scholar] [CrossRef]
  7. Anisimova, E.A.; Yarullina, D.R. Antibiotic Resistance of LACTOBACILLUS Strains. Curr. Microbiol. 2019, 76, 1407–1416. [Google Scholar] [CrossRef]
  8. Abel-Santos, E. Chapter 9—Endospores, Sporulation and Germination. In Molecular Medical Microbiology, 2nd ed.; Tang, Y.-W., Sussman, M., Liu, D., Poxton, I., Schwartzman, J., Eds.; Academic Press: Boston, MA, USA, 2015; pp. 163–178. [Google Scholar] [CrossRef]
  9. Adeniji, A.A.L.D.; Babalola, O.O. Bacillus velezensis: Phylogeny, useful applications, and avenues for exploitation. Appl. Microbiol. Biotechnol. 2019, 103, 3669–3682. [Google Scholar] [CrossRef]
  10. 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] [Green Version]
  11. Ortiz, A.; Sansinenea, E. Chemical Compounds Produced by Bacillus sp. Factories and Their Role in Nature. Mini Rev. Med. Chem. 2019, 19, 373–380. [Google Scholar] [CrossRef] [PubMed]
  12. Zhao, X.; Kuipers, O.P. Identification and classification of known and putative antimicrobial compounds produced by a wide variety of Bacillales species. BMC Genom. 2016, 17, 882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Nishino, K.; Hsu, F.-F.; Turk, J.; Cromie, M.J.; Wösten, M.M.S.M.; Groisman, E.A. Identification of the lipopolysaccharide modifications controlled by the Salmonella PmrA/PmrB system mediating resistance to Fe(III) and Al(III). Mol. Microbiol. 2006, 61, 645–654. [Google Scholar] [CrossRef] [Green Version]
  14. Vollmer, W.; Blanot, D.; De Pedro, M.A. Peptidoglycan structure and architecture. FEMS Microbiol. Rev. 2008, 32, 149–167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Johnson, B.A.; Anker, H.; Meleney, F.L. Bacitracin: A new antibiotic produced by a member of the B. subtilis group. Science 1945, 102, 376–377. [Google Scholar] [CrossRef] [PubMed]
  16. Zaidi, S.; Singh, S.L.; Khan, A.U. Exploring antibiofilm potential of bacitracin against streptococcus mutans. Microb. Pathog. 2020, 149, 104279. [Google Scholar] [CrossRef]
  17. Siewert, G.; Strominger, J.L. Bacitracin: An inhibitor of the dephosphorylation of lipid pyrophosphate, an intermediate in the biosynthesis of the peptidoglycan of bacterial cell walls. Proc. Natl. Acad. Sci. USA 1967, 57, 767–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Mahlstedt, S.A.; Walsh, C.T. Investigation of Anticapsin Biosynthesis Reveals a Four-Enzyme Pathway to Tetrahydrotyrosine in Bacillus subtilis. Biochemistry 2010, 49, 912–923. [Google Scholar] [CrossRef] [Green Version]
  19. Newton, G.G. Antibiotics from a strain of B. subtilis; bacilipin A and B and bacilysin. Br. J. Exp. Pathol. 1949, 30, 306–319. [Google Scholar]
  20. Wang, T.; Wu, M.B.; Chen, Z.J.; Lin, J.P.; Yang, L.R. Separation, determination and antifungal activity test of the products from a new Bacillus amyloliquefaciens. Nat. Prod. Res. 2016, 30, 1215–1218. [Google Scholar] [CrossRef]
  21. Kenig, M.; Abraham, E.P. Antimicrobial activities and antagonists of bacilysin and anticapsin. J. Gen. Microbiol. 1976, 94, 37–45. [Google Scholar] [CrossRef] [Green Version]
  22. Kenig, M.; Vandamme, E.; Abraham, E.P. The mode of action of bacilysin and anticapsin and biochemical properties of bacilysin-resistant mutants. J. Gen. Microbiol. 1976, 94, 46–54. [Google Scholar] [CrossRef] [Green Version]
  23. Milner, J.L.; Silo-Suh, L.; Lee, J.C.; He, H.; Clardy, J.; Handelsman, J. Production of kanosamine by Bacillus cereus UW85. Appl. Env. Microbiol. 1996, 62, 3061–3065. [Google Scholar] [CrossRef] [Green Version]
  24. Janiak, A.M.; Milewski, S. Mechanism of antifungal action of kanosamine. Med. Mycol. 2001, 39, 401–408. [Google Scholar] [CrossRef] [PubMed]
  25. van Heijenoort, J. Lipid intermediates in the biosynthesis of bacterial peptidoglycan. Microbiol. Mol. Biol. Rev. 2007, 71, 620–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Derouaux, A.; Turk, S.; Olrichs, N.K.; Gobec, S.; Breukink, E.; Amoroso, A.; Offant, J.; Bostock, J.; Mariner, K.; Chopra, I.; et al. Small molecule inhibitors of peptidoglycan synthesis targeting the lipid II precursor. Biochem. Pharm. 2011, 81, 1098–1105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Chugunov, A.; Pyrkova, D.; Nolde, D.; Polyansky, A.; Pentkovsky, V.; Efremov, R. Lipid-II forms potential “landing terrain” for lantibiotics in simulated bacterial membrane. Sci. Rep. 2013, 3, 1678. [Google Scholar] [CrossRef] [Green Version]
  28. Ross, A.C.; Vederas, J.C. Fundamental functionality: Recent developments in understanding the structure–activity relationships of lantibiotic peptides. J. Antibiot. 2011, 64, 27–34. [Google Scholar] [CrossRef] [Green Version]
  29. Chan, W.C.; Bycroft, B.W.; Leyland, M.L.; Lian, L.Y.; Roberts, G.C. A novel post-translational modification of the peptide antibiotic subtilin: Isolation and characterization of a natural variant from Bacillus subtilis A.T.C.C. 6633. Biochem. J. 1993, 291 Pt 1, 23–27. [Google Scholar] [CrossRef] [Green Version]
  30. Parisot, J.; Carey, S.; Breukink, E.; Chan, W.C.; Narbad, A.; Bonev, B. Molecular mechanism of target recognition by subtilin, a class I lanthionine antibiotic. Antimicrob. Agents Chemother. 2008, 52, 612–618. [Google Scholar] [CrossRef] [Green Version]
  31. Bouhss, A.; Al-Dabbagh, B.; Vincent, M.; Odaert, B.; Aumont-Nicaise, M.; Bressolier, P.; Desmadril, M.; Mengin-Lecreulx, D.; Urdaci, M.C.; Gallay, J. Specific interactions of clausin, a new lantibiotic, with lipid precursors of the bacterial cell wall. Biophys. J. 2009, 97, 1390–1397. [Google Scholar] [CrossRef] [Green Version]
  32. Ahire, J.J.; Kashikar, M.S.; Lakshmi, S.G.; Madempudi, R. Identification and characterization of antimicrobial peptide produced by indigenously isolated Bacillus paralicheniformis UBBLi30 strain. 3 Biotech 2020, 10, 112. [Google Scholar] [CrossRef] [PubMed]
  33. Brötz, H.; Bierbaum, G.; Leopold, K.; Reynolds, P.E.; Sahl, H.G. The lantibiotic mersacidin inhibits peptidoglycan synthesis by targeting lipid II. Antimicrob. Agents Chemother. 1998, 42, 154–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Sass, P.; Jansen, A.; Szekat, C.; Sass, V.; Sahl, H.G.; Bierbaum, G. The lantibiotic mersacidin is a strong inducer of the cell wall stress response of Staphylococcus aureus. BMC Microbiol. 2008, 8, 186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Brötz, H.; Bierbaum, G.; Reynolds, P.E.; Sahl, H.G. The lantibiotic mersacidin inhibits peptidoglycan biosynthesis at the level of transglycosylation. Eur. J. Biochem. 1997, 246, 193–199. [Google Scholar] [CrossRef]
  36. Arguelles Arias, A.; Joris, B.; Fickers, P. Dual mode of action of amylolysin: A type-B lantibiotic produced by Bacillus amyloliquefaciens GA1. Protein Pept. Lett. 2014, 21, 336–340. [Google Scholar] [CrossRef]
  37. Arguelles Arias, A.; Ongena, M.; Devreese, B.; Terrak, M.; Joris, B.; Fickers, P. Characterization of amylolysin, a novel lantibiotic from Bacillus amyloliquefaciens GA1. PLoS ONE 2013, 8, e83037. [Google Scholar] [CrossRef] [Green Version]
  38. Oman, T.J.; van der Donk, W.A. Insights into the mode of action of the two-peptide lantibiotic haloduracin. ACS Chem. Biol. 2009, 4, 865–874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Oman, T.J.; Lupoli, T.J.; Wang, T.S.; Kahne, D.; Walker, S.; van der Donk, W.A. Haloduracin α binds the peptidoglycan precursor lipid II with 2:1 stoichiometry. J. Am. Chem. Soc. 2011, 133, 17544–17547. [Google Scholar] [CrossRef]
  40. Ferrante, T.; Viola, F.; Balliano, G.; Oliaro-Bosso, S. Difference in the late ergosterol biosynthesis between yeast spheroplasts and intact cells. Acta Biochim. Pol. 2016, 63, 371–375. [Google Scholar] [CrossRef]
  41. Casares, D.; Escribá, P.V.; Rosselló, C.A. Membrane Lipid Composition: Effect on Membrane and Organelle Structure, Function and Compartmentalization and Therapeutic Avenues. Int. J. Mol. Sci. 2019, 20, 2167. [Google Scholar] [CrossRef] [Green Version]
  42. Rodrigues, B.; Morais, T.P.; Zaini, P.A.; Campos, C.S.; Almeida-Souza, H.O.; Dandekar, A.M.; Nascimento, R.; Goulart, L.R. Antimicrobial activity of Epsilon-Poly-l-lysine against phytopathogenic bacteria. Sci. Rep. 2020, 10, 11324. [Google Scholar] [CrossRef]
  43. El-Sersy, N.A.; Abdelwahab, A.E.; Abouelkhiir, S.S.; Abou-Zeid, D.M.; Sabry, S.A. Antibacterial and anticancer activity of ε-poly-L-lysine (ε-PL) produced by a marine Bacillus subtilis sp. J. Basic Microbiol. 2012, 52, 513–522. [Google Scholar] [CrossRef]
  44. Hyldgaard, M.; Mygind, T.; Vad, B.S.; Stenvang, M.; Otzen, D.E.; Meyer, R.L. The antimicrobial mechanism of action of epsilon-poly-l-lysine. Appl. Environ. Microbiol. 2014, 80, 7758–7770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Molohon, K.J.; Saint-Vincent, P.M.B.; Park, S.; Doroghazi, J.R.; Maxson, T.; Hershfield, J.R.; Flatt, K.M.; Schroeder, N.E.; Ha, T.; Mitchell, D.A. Plantazolicin is an ultra-narrow spectrum antibiotic that targets the Bacillus anthracis membrane. ACS Infect. Dis. 2016, 2, 207–220. [Google Scholar] [CrossRef] [Green Version]
  46. Velkov, T.; Gallardo-Godoy, A.; Swarbrick, J.D.; Blaskovich, M.A.T.; Elliott, A.G.; Han, M.; Thompson, P.E.; Roberts, K.D.; Huang, J.X.; Becker, B.; et al. Structure, Function, and Biosynthetic Origin of Octapeptin Antibiotics Active against Extensively Drug-Resistant Gram-Negative Bacteria. Cell Chem. Biol. 2018, 25, 380–391.e385. [Google Scholar] [CrossRef] [Green Version]
  47. Rosenthal, K.S.; Ferguson, R.A.; Storm, D.R. Mechanism of action of EM 49, membrane-active peptide antibiotic. Antimicrob. Agents Chemother. 1977, 12, 665–672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Yang, J.; Zhu, X.; Cao, M.; Wang, C.; Zhang, C.; Lu, Z.; Lu, F. Genomics-Inspired Discovery of Three Antibacterial Active Metabolites, Aurantinins B, C, and D from Compost-Associated Bacillus subtilis fmb60. J. Agric. Food Chem. 2016, 64, 8811–8820. [Google Scholar] [CrossRef]
  49. Rollin-Pinheiro, R.; Bayona-Pacheco, B.; Domingos, L.T.S.; da Rocha Curvelo, J.A.; de Castro, G.M.M.; Barreto-Bergter, E.; Ferreira-Pereira, A. Sphingolipid Inhibitors as an Alternative to Treat Candidiasis Caused by Fluconazole-Resistant Strains. Pathogens 2021, 10, 856. [Google Scholar] [CrossRef] [PubMed]
  50. Wadsworth, J.M.; Clarke, D.J.; McMahon, S.A.; Lowther, J.P.; Beattie, A.E.; Langridge-Smith, P.R.; Broughton, H.B.; Dunn, T.M.; Naismith, J.H.; Campopiano, D.J. The chemical basis of serine palmitoyltransferase inhibition by myriocin. J. Am. Chem. Soc. 2013, 135, 14276–14285. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, H.; Wang, Z.; Xu, W.; Wang, K. Comprehensive transcriptomic and proteomic analyses identify intracellular targets for myriocin to induce Fusarium oxysporum f. sp. niveum cell death. Microb. Cell Fact. 2021, 20, 69. [Google Scholar] [CrossRef]
  52. Takada, Y.; Itoh, H.; Paudel, A.; Panthee, S.; Hamamoto, H.; Sekimizu, K.; Inoue, M. Discovery of gramicidin A analogues with altered activities by multidimensional screening of a one-bead-one-compound library. Nat. Commun. 2020, 11, 4935. [Google Scholar] [CrossRef]
  53. Hladky, S.B.; Haydon, D.A. Ion transfer across lipid membranes in the presence of gramicidin A. I. Studies of the unit conductance channel. Biochim. Biophys. Acta 1972, 274, 294–312. [Google Scholar] [CrossRef]
  54. Liou, J.W.; Hung, Y.J.; Yang, C.H.; Chen, Y.C. The antimicrobial activity of gramicidin A is associated with hydroxyl radical formation. PLoS ONE 2015, 10, e0117065. [Google Scholar] [CrossRef] [Green Version]
  55. Swierstra, J.; Kapoerchan, V.; Knijnenburg, A.; van Belkum, A.; Overhand, M. Structure, toxicity and antibiotic activity of gramicidin S and derivatives. Eur. J. Clin. Microbiol. Infect. Dis. 2016, 35, 763–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Afonin, S.; Dürr, U.H.; Wadhwani, P.; Salgado, J.; Ulrich, A.S. Solid State NMR Structure Analysis of the Antimicrobial Peptide Gramicidin S in Lipid Membranes: Concentration-Dependent Re-alignment and Self-Assembly as a β-Barrel. Top Curr. Chem. 2008, 273, 139–154. [Google Scholar] [CrossRef]
  57. Kaprel’iants, A.S.; Nikiforov, V.V.; Miroshnikov, A.I.; Snezhkova, L.G.; Eremin, V.A.; Ostrovskiĭ, D.N. Membranes of bacteria and mechanism of action of the antibiotic gramicidin S. Biokhimiia 1977, 42, 329–337. [Google Scholar] [PubMed]
  58. Frangou-Lazaridis, M.; Seddon, B. Effect of gramicidin S on the transcription system of the producer Bacillus brevis Nagano. J. Gen. Microbiol. 1985, 131, 437–449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Grau, A.; Ortiz, A.; de Godos, A.; Gómez-Fernández, J.C. A biophysical study of the interaction of the lipopeptide antibiotic iturin A with aqueous phospholipid bilayers. Arch. Biochem. Biophys. 2000, 377, 315–323. [Google Scholar] [CrossRef]
  60. Inès, M.; Dhouha, G. Lipopeptide surfactants: Production, recovery and pore forming capacity. Peptides 2015, 71, 100–112. [Google Scholar] [CrossRef]
  61. 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]
  62. Ambroggio, E.E.; Separovic, F.; Bowie, J.H.; Fidelio, G.D.; Bagatolli, L.A. Direct visualization of membrane leakage induced by the antibiotic peptides: Maculatin, citropin, and aurein. Biophys. J. 2005, 89, 1874–1881. [Google Scholar] [CrossRef] [Green Version]
  63. Lei, S.; Zhao, H.; Pang, B.; Qu, R.; Lian, Z.; Jiang, C.; Shao, D.; Huang, Q.; Jin, M.; Shi, J. Capability of iturin from Bacillus subtilis to inhibit Candida albicans in vitro and in vivo. Appl. Microbiol. Biotechnol. 2019, 103, 4377–4392. [Google Scholar] [CrossRef] [PubMed]
  64. Peypoux, F.; Besson, F.; Michel, G.; Delcambe, L. Preparation and antibacterial activity upon Micrococcus luteus of derivatives of iturin A, mycosubtilin and bacillomycin L, antibiotics from Bacillus subtilis. J. Antibiot. 1979, 32, 136–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Romero, D.; de Vicente, A.; Rakotoaly, R.H.; Dufour, S.E.; Veening, J.W.; Arrebola, E.; Cazorla, F.M.; Kuipers, O.P.; Paquot, M.; Pérez-García, A. The iturin and fengycin families of lipopeptides are key factors in antagonism of Bacillus subtilis toward Podosphaera fusca. Mol. Plant Microbe Interact. 2007, 20, 430–440. [Google Scholar] [CrossRef] [Green Version]
  66. Arima, K.; Kakinuma, A.; Tamura, G. Surfactin, a crystalline peptidelipid surfactant produced by Bacillus subtilis: Isolation, characterization and its inhibition of fibrin clot formation. Biochem. Biophys. Res. Commun. 1968, 31, 488–494. [Google Scholar] [CrossRef]
  67. Sheppard, J.D.; Jumarie, C.; Cooper, D.G.; Laprade, R. Ionic channels induced by surfactin in planar lipid bilayer membranes. Biochim. Biophys. Acta 1991, 1064, 13–23. [Google Scholar] [CrossRef]
  68. Liu, J.; Li, W.; Zhu, X.; Zhao, H.; Lu, Y.; Zhang, C.; Lu, Z. Surfactin effectively inhibits Staphylococcus aureus adhesion and biofilm formation on surfaces. Appl. Microbiol. Biotechnol. 2019, 103, 4565–4574. [Google Scholar] [CrossRef]
  69. Farace, G.; Fernandez, O.; Jacquens, L.; Coutte, F.; Krier, F.; Jacques, P.; Clément, C.; Barka, E.A.; Jacquard, C.; Dorey, S. Cyclic lipopeptides from Bacillus subtilis activate distinct patterns of defence responses in grapevine. Mol. Plant Pathol. 2015, 16, 177–187. [Google Scholar] [CrossRef]
  70. Tao, Y.; Bie, X.M.; Lv, F.X.; Zhao, H.Z.; Lu, Z.X. Antifungal activity and mechanism of fengycin in the presence and absence of commercial surfactin against Rhizopus stolonifer. J. Microbiol. 2011, 49, 146–150. [Google Scholar] [CrossRef]
  71. Liu, Y.; Zhang, J.; Wang, S.; Guo, Y.; He, T.; Zhou, R. A Novel Adjuvant “Sublancin” Enhances Immune Response in Specific Pathogen-Free Broiler Chickens Inoculated with Newcastle Disease Vaccine. J. Immunol. Res. 2019, 2019, 1016567. [Google Scholar] [CrossRef] [Green Version]
  72. Zhang, L.; Sun, C. Fengycins, Cyclic Lipopeptides from Marine Bacillus subtilis Strains, Kill the Plant-Pathogenic Fungus Magnaporthe grisea by Inducing Reactive Oxygen Species Production and Chromatin Condensation. Appl. Environ. Microbiol. 2018, 84, e00445-18. [Google Scholar] [CrossRef] [Green Version]
  73. Johnson, V.L.; Ko, S.C.; Holmstrom, T.H.; Eriksson, J.E.; Chow, S.C. Effector caspases are dispensable for the early nuclear morphological changes during chemical-induced apoptosis. J. Cell Sci. 2000, 113 Pt 17, 2941–2953. [Google Scholar] [CrossRef] [PubMed]
  74. Piewngam, P.; Zheng, Y.; Nguyen, T.H.; Dickey, S.W.; Joo, H.S.; Villaruz, A.E.; Glose, K.A.; Fisher, E.L.; Hunt, R.L.; Li, B.; et al. Pathogen elimination by probiotic Bacillus via signalling interference. Nature 2018, 562, 532–537. [Google Scholar] [CrossRef]
  75. Chung, L.K.; Raffatellu, M. Probiotic fengycins dis(Agr)ee with Staphylococcus aureus colonization. Cell Res. 2019, 29, 93–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Gao, L.; Han, J.; Liu, H.; Qu, X.; Lu, Z.; Bie, X. Plipastatin and surfactin coproduction by Bacillus subtilis pB2-L and their effects on microorganisms. Antonie Van Leeuwenhoek 2017, 110, 1007–1018. [Google Scholar] [CrossRef] [PubMed]
  77. Cawoy, H.; Debois, D.; Franzil, L.; De Pauw, E.; Thonart, P.; Ongena, M. Lipopeptides as main ingredients for inhibition of fungal phytopathogens by Bacillus subtilis/amyloliquefaciens. Microb. Biotechnol. 2015, 8, 281–295. [Google Scholar] [CrossRef]
  78. Besson, F.; Peypoux, F.; Michel, G.; Delcambe, L. Antifungal activity upon Saccharomyces cerevisiae of iturin A, mycosubtilin, bacillomycin L and of their derivatives; inhibition of this antifungal activity by lipid antagonists. J. Antibiot. 1979, 32, 828–833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Gong, A.D.; Li, H.P.; Yuan, Q.S.; Song, X.S.; Yao, W.; He, W.J.; Zhang, J.B.; Liao, Y.C. Antagonistic mechanism of iturin A and plipastatin A from Bacillus amyloliquefaciens S76-3 from wheat spikes against Fusarium graminearum. PLoS ONE 2015, 10, e0116871. [Google Scholar] [CrossRef] [Green Version]
  80. Wang, Y.; Zhang, C.; Liang, J.; Wu, L.; Gao, W.; Jiang, J. Iturin A Extracted from Bacillus subtilis WL-2 Affects Phytophthora infestans via Cell Structure Disruption, Oxidative Stress, and Energy Supply Dysfunction. Front. Microbiol. 2020, 11, 2205. [Google Scholar] [CrossRef]
  81. Tunsagool, P.; Leelasuphakul, W.; Jaresitthikunchai, J.; Phaonakrop, N.; Roytrakul, S.; Jutidamrongphan, W. Targeted transcriptional and proteomic studies explicate specific roles of Bacillus subtilis iturin A, fengycin, and surfactin on elicitation of defensive systems in mandarin fruit during stress. PLoS ONE 2019, 14, e0217202. [Google Scholar] [CrossRef] [Green Version]
  82. Jin, P.; Wang, H.; Tan, Z.; Xuan, Z.; Dahar, G.Y.; Li, Q.X.; Miao, W.; Liu, W. Antifungal mechanism of bacillomycin D from Bacillus velezensis HN-2 against Colletotrichum gloeosporioides Penz. Pestic. Biochem. Physiol. 2020, 163, 102–107. [Google Scholar] [CrossRef]
  83. Peypoux, F.; Besson, F.; Michel, G.; Lenzen, C.; Dierickx, L.; Delcambe, L. Characterization of a new antibiotic of iturin group: Bacillomycin D. J. Antibiot. 1980, 33, 1146–1149. [Google Scholar] [CrossRef] [Green Version]
  84. Wu, T.; Chen, M.; Zhou, L.; Lu, F.; Bie, X.; Lu, Z. Bacillomycin D effectively controls growth of Malassezia globosa by disrupting the cell membrane. Appl. Microbiol. Biotechnol. 2020, 104, 3529–3540. [Google Scholar] [CrossRef]
  85. Gu, Q.; Yang, Y.; Yuan, Q.; Shi, G.; Wu, L.; Lou, Z.; Huo, R.; Wu, H.; Borriss, R.; Gao, X. Bacillomycin D Produced by Bacillus amyloliquefaciens Is Involved in the Antagonistic Interaction with the Plant-Pathogenic Fungus Fusarium graminearum. Appl. Environ. Microbiol. 2017, 83, e01075-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Xu, Z.; Mandic-Mulec, I.; Zhang, H.; Liu, Y.; Sun, X.; Feng, H.; Xun, W.; Zhang, N.; Shen, Q.; Zhang, R. Antibiotic Bacillomycin D Affects Iron Acquisition and Biofilm Formation in Bacillus velezensis through a Btr-Mediated FeuABC-Dependent Pathway. Cell Rep. 2019, 29, 1192–1202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Zhang, B.; Dong, C.; Shang, Q.; Cong, Y.; Kong, W.; Li, P. Purification and partial characterization of bacillomycin L produced by Bacillus amyloliquefaciens K103 from lemon. Appl. Biochem. Biotechnol. 2013, 171, 2262–2272. [Google Scholar] [CrossRef]
  88. Quentin, M.J.; Besson, F.; Peypoux, F.; Michel, G. Action of peptidolipidic antibiotics of the iturin group on erythrocytes. Effect of some lipids on hemolysis. Biochim. Biophys. Acta 1982, 684, 207–211. [Google Scholar] [CrossRef]
  89. Besson, F.; Michel, G. Action of the antibiotics of the iturin group on artificial membranes. J. Antibiot. 1984, 37, 646–651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Zhang, B.; Qin, Y.; Han, Y.; Dong, C.; Li, P.; Shang, Q. Comparative proteomic analysis reveals intracellular targets for bacillomycin L to induce Rhizoctonia solani Kühn hyphal cell death. Biochim. Biophys. Acta 2016, 1864, 1152–1159. [Google Scholar] [CrossRef]
  91. Loison, C.; Nasir, M.N.; Benichou, E.; Besson, F.; Brevet, P.-F. Multi-scale modeling of mycosubtilin lipopeptides at the air/water interface: Structure and optical second harmonic generation. Phys. Chem. Chem. Phys. 2014, 16, 2136–2148. [Google Scholar] [CrossRef]
  92. Besson, F.; Michel, G. Action of mycosubtilin, an antifungal antibiotic of Bacillus subtilis, on the cell membrane of Saccharomyces cerevisiae. Microbios 1989, 59, 113–121. [Google Scholar]
  93. Bhattacharyya, P.; Bose, S.K. Effect of mycobacillin, an antifungal polypeptide antibiotic, on the producer Bacillus subtilis B3. Indian J. Med. Res. 1967, 55, 1025–1029. [Google Scholar] [PubMed]
  94. Chowdhury, B.; Das, S.K.; Bose, S.K. Use of resistant mutants to characterize the target of mycobacillin in Aspergillus niger membranes. Microbiology 1998, 144, 1123–1130. [Google Scholar] [CrossRef] [Green Version]
  95. Das, S.K.; Mukherjee, S.; Majumdar, S.; Basu, S.; Bose, S.K. Physico-chemical interaction of mycobacillin with Aspergillus niger protoplast membrane, the site of its action. J. Antibiot. 1987, 40, 1036–1043. [Google Scholar] [CrossRef] [Green Version]
  96. Babasaki, K.; Takao, T.; Shimonishi, Y.; Kurahashi, K. Subtilosin A, a new antibiotic peptide produced by Bacillus subtilis 168: Isolation, structural analysis, and biogenesis. J. Biochem. 1985, 98, 585–603. [Google Scholar] [CrossRef]
  97. Turovskiy, Y.; Cheryian, T.; Algburi, A.; Wirawan, R.E.; Takhistov, P.; Sinko, P.J.; Chikindas, M.L. Susceptibility of Gardnerella vaginalis biofilms to natural antimicrobials subtilosin, ε-poly-L-lysine, and lauramide arginine ethyl ester. Infect. Dis. Obs. Gynecol. 2012, 2012, 284762. [Google Scholar] [CrossRef]
  98. Noll, K.S.; Sinko, P.J.; Chikindas, M.L. Elucidation of the Molecular Mechanisms of Action of the Natural Antimicrobial Peptide Subtilosin against the Bacterial Vaginosis-associated Pathogen Gardnerella vaginalis. Probiotics Antimicrob. Proteins 2011, 3, 41–47. [Google Scholar] [CrossRef] [Green Version]
  99. Algburi, A.; Zehm, S.; Netrebov, V.; Bren, A.B.; Chistyakov, V.; Chikindas, M.L. Subtilosin Prevents Biofilm Formation by Inhibiting Bacterial Quorum Sensing. Probiotics Antimicrob. Proteins 2017, 9, 81–90. [Google Scholar] [CrossRef] [PubMed]
  100. Hao, Z.; Yan, L.; Liu, J.; Song, F.; Zhang, J.; Li, X. Extraction of antibiotic zwittermicin A from Bacillus thuringiensis by macroporous resin and silica gel column chromatography. Biotechnol. Appl. Biochem. 2015, 62, 369–374. [Google Scholar] [CrossRef] [PubMed]
  101. Stabb, E.V.; Handelsman, J. Genetic analysis of zwittermicin A resistance in Escherichia coli: Effects on membrane potential and RNA polymerase. Mol. Microbiol. 1998, 27, 311–322. [Google Scholar] [CrossRef] [Green Version]
  102. Im, S.M.; Yu, N.H.; Joen, H.W.; Kim, S.O.; Park, H.W.; Park, A.R.; Kim, J.-C. Biological control of tomato bacterial wilt by oxydifficidin and difficidin-producing Bacillus methylotrophicus DR-08. Pestic. Biochem. Physiol. 2020, 163, 130–137. [Google Scholar] [CrossRef]
  103. Zweerink, M.M.; Edison, A. Difficidin and oxydifficidin: Novel broad spectrum antibacterial antibiotics produced by Bacillus subtilis. III. Mode of action of difficidin. J. Antibiot. 1987, 40, 1692–1697. [Google Scholar] [CrossRef] [Green Version]
  104. Wu, L.; Wu, H.; Chen, L.; Yu, X.; Borriss, R.; Gao, X. Difficidin and bacilysin from Bacillus amyloliquefaciens FZB42 have antibacterial activity against Xanthomonas oryzae rice pathogens. Sci. Rep. 2015, 5, 12975. [Google Scholar] [CrossRef]
  105. Wang, S.; Wang, Q.; Zeng, X.; Ye, Q.; Huang, S.; Yu, H.; Yang, T.; Qiao, S. Use of the Antimicrobial Peptide Sublancin with Combined Antibacterial and Immunomodulatory Activities To Protect against Methicillin-Resistant Staphylococcus aureus Infection in Mice. J. Agric. Food Chem. 2017, 65, 8595–8605. [Google Scholar] [CrossRef] [PubMed]
  106. Wu, C.; Biswas, S.; Garcia De Gonzalo, C.V.; van der Donk, W.A. Investigations into the Mechanism of Action of Sublancin. ACS Infect. Dis 2019, 5, 454–459. [Google Scholar] [CrossRef]
  107. Maksimova, E.M.; Vinogradova, D.S.; Osterman, I.A.; Kasatsky, P.S.; Nikonov, O.S.; Milón, P.; Dontsova, O.A.; Sergiev, P.V.; Paleskava, A.; Konevega, A.L. Multifaceted Mechanism of Amicoumacin A Inhibition of Bacterial Translation. Front. Microbiol. 2021, 12, 172. [Google Scholar] [CrossRef] [PubMed]
  108. Lama, A.; Pané-Farré, J.; Chon, T.; Wiersma, A.M.; Sit, C.S.; Vederas, J.C.; Hecker, M.; Nakano, M.M. Response of methicillin-resistant Staphylococcus aureus to amicoumacin A. PLoS ONE 2012, 7, e34037. [Google Scholar] [CrossRef] [Green Version]
  109. Tanaka, K.; Fukuda, M.; Amaki, Y.; Sakaguchi, T.; Inai, K.; Ishihara, A.; Nakajima, H. Importance of prumycin produced by Bacillus amyloliquefaciens SD-32 in biocontrol against cucumber powdery mildew disease. Pest. Manag. Sci. 2017, 73, 2419–2428. [Google Scholar] [CrossRef] [PubMed]
  110. Schwartz, J.L.; Katagiri, M.; Omura, S.; Tishler, M. The mechanism of prumycin action. J. Antibiot. 1974, 27, 379–385. [Google Scholar] [CrossRef] [Green Version]
  111. Hata, T.; Omura, S.; Katagiri, M.; Atsumi, K.; Awaya, J. A new antifungal antibiotic, prumycin. J. Antibiot. 1971, 24, 900–901. [Google Scholar] [CrossRef] [Green Version]
  112. Chan, D.C.K.; Burrows, L.L. Thiocillin and micrococcin exploit the ferrioxamine receptor of Pseudomonas aeruginosa for uptake. J. Antimicrob. Chemother. 2021, 76, 2029–2039. [Google Scholar] [CrossRef]
  113. Bleich, R.; Watrous, J.D.; Dorrestein, P.C.; Bowers, A.A.; Shank, E.A. Thiopeptide antibiotics stimulate biofilm formation in Bacillus subtilis. Proc. Natl. Acad. Sci. USA 2015, 112, 3086–3091. [Google Scholar] [CrossRef] [Green Version]
  114. Wang, T.; Lu, Q.; Sun, C.; Lukianov, D.; Osterman, I.A.; Sergiev, P.V.; Dontsova, O.A.; Hu, X.; You, X.; Liu, S.; et al. Hetiamacin E and F, New Amicoumacin Antibiotics from Bacillus subtilis PJS Using MS/MS-Based Molecular Networking. Molecules 2020, 25, 4446. [Google Scholar] [CrossRef]
  115. Gahungu, M.; Arguelles-Arias, A.; Fickers, P.; Zervosen, A.; Joris, B.; Damblon, C.; Luxen, A. Synthesis and biological evaluation of potential threonine synthase inhibitors: Rhizocticin A and Plumbemycin A. Bioorg. Med. Chem. 2013, 21, 4958–4967. [Google Scholar] [CrossRef] [PubMed]
  116. Kugler, M.; Loeffler, W.; Rapp, C.; Kern, A.; Jung, G. Rhizocticin A, an antifungal phosphono-oligopeptide of Bacillus subtilis ATCC 6633: Biological properties. Arch. Microbiol. 1990, 153, 276–281. [Google Scholar] [CrossRef] [PubMed]
  117. Yoo, J.S.; Zheng, C.J.; Lee, S.; Kwak, J.H.; Kim, W.G. Macrolactin N, a new peptide deformylase inhibitor produced by Bacillus subtilis. Bioorg. Med. Chem. Lett. 2006, 16, 4889–4892. [Google Scholar] [CrossRef]
  118. Aoki, Y.; Yamamoto, M.; Hosseini-Mazinani, S.M.; Koshikawa, N.; Sugimoto, K.; Arisawa, M. Antifungal azoxybacilin exhibits activity by inhibiting gene expression of sulfite reductase. Antimicrob. Agents Chemother. 1996, 40, 127–132. [Google Scholar] [CrossRef] [Green Version]
  119. Fujiu, M.; Sawairi, S.; Shimada, H.; Takaya, H.; Aoki, Y.; Okuda, T.; Yokose, K. Azoxybacilin, a novel antifungal agent produced by Bacillus cereus NR2991. Production, isolation and structure elucidation. J. Antibiot. 1994, 47, 833–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Zeriouh, H.; de Vicente, A.; Pérez-García, A.; Romero, D. Surfactin triggers biofilm formation of Bacillus subtilis in melon phylloplane and contributes to the biocontrol activity. Environ. Microbiol. 2014, 16, 2196–2211. [Google Scholar] [CrossRef]
  121. Caro-Astorga, J.; Frenzel, E.; Perkins, J.R.; Álvarez-Mena, A.; de Vicente, A.; Ranea, J.A.G.; Kuipers, O.P.; Romero, D. Biofilm formation displays intrinsic offensive and defensive features of Bacillus cereus. NPJ Biofilms Microbiomes 2020, 6, 3. [Google Scholar] [CrossRef] [Green Version]
  122. Karunakaran, E.; Biggs, C.A. Mechanisms of Bacillus cereus biofilm formation: An investigation of the physicochemical characteristics of cell surfaces and extracellular proteins. Appl. Microbiol. Biotechnol. 2011, 89, 1161–1175. [Google Scholar] [CrossRef] [PubMed]
  123. Singh, A.A.; Singh, A.K.; Nerurkar, A. Disrupting the quorum sensing mediated virulence in soft rot causing Pectobacterium carotovorum by marine sponge associated Bacillus sp. OA10. World J. Microbiol. Biotechnol. 2021, 37, 5. [Google Scholar] [CrossRef]
  124. Boopathi, S.; Vashisth, R.; Manoharan, P.; Kandasamy, R.; Sivakumar, N. Stigmatellin Y—An anti-biofilm compound from Bacillus subtilis BR4 possibly interferes in PQS-PqsR mediated quorum sensing system in Pseudomonas aeruginosa. Bioorg. Med. Chem. Lett. 2017, 27, 2113–2118. [Google Scholar] [CrossRef] [PubMed]
  125. Li, H.; Han, X.; Dong, Y.; Xu, S.; Chen, C.; Feng, Y.; Cui, Q.; Li, W. Bacillaenes: Decomposition Trigger Point and Biofilm Enhancement in Bacillus. ACS Omega 2021, 6, 1093–1098. [Google Scholar] [CrossRef] [PubMed]
  126. Tonziello, G.; Caraffa, E.; Pinchera, B.; Granata, G.; Petrosillo, N. Present and future of siderophore-based therapeutic and diagnostic approaches in infectious diseases. Infect. Dis. Rep. 2019, 11, 8208. [Google Scholar] [CrossRef] [Green Version]
  127. Dertz, E.A.; Xu, J.; Stintzi, A.; Raymond, K.N. Bacillibactin-mediated iron transport in Bacillus subtilis. J. Am. Chem. Soc. 2006, 128, 22–23. [Google Scholar] [CrossRef]
  128. Hu, X.; Boyer, G.L. Siderophore-Mediated Aluminum Uptake by Bacillus megaterium ATCC 19213. Appl. Environ. Microbiol. 1996, 62, 4044–4048. [Google Scholar] [CrossRef] [Green Version]
  129. Segond, D.; Abi Khalil, E.; Buisson, C.; Daou, N.; Kallassy, M.; Lereclus, D.; Arosio, P.; Bou-Abdallah, F.; Nielsen Le Roux, C. Iron acquisition in Bacillus cereus: The roles of IlsA and bacillibactin in exogenous ferritin iron mobilization. PLoS Pathog. 2014, 10, e1003935. [Google Scholar] [CrossRef] [Green Version]
  130. Chen, X.H.; Vater, J.; Piel, J.; Franke, P.; Scholz, R.; Schneider, K.; Koumoutsi, A.; Hitzeroth, G.; Grammel, N.; Strittmatter, A.W.; et al. Structural and functional characterization of three polyketide synthase gene clusters in Bacillus amyloliquefaciens FZB 42. J. Bacteriol. 2006, 188, 4024–4036. [Google Scholar] [CrossRef] [Green Version]
  131. Dunlap, C.A.; Kim, S.J.; Kwon, S.W.; Rooney, A.P. Bacillus velezensis is not a later heterotypic synonym of Bacillus amyloliquefaciens; Bacillus methylotrophicus, Bacillus amyloliquefaciens subsp. plantarum and ‘Bacillus oryzicola’ are later heterotypic synonyms of Bacillus velezensis based on phylogenomics. Int. J. Syst. Evol. Microbiol. 2016, 66, 1212–1217. [Google Scholar] [CrossRef]
  132. Fan, B.; Blom, J.; Klenk, H.P.; Borriss, R. Bacillus amyloliquefaciens, Bacillus velezensis, and Bacillus siamensis Form an “Operational Group B. amyloliquefaciens” within the B. subtilis Species Complex. Front. Microbiol. 2017, 8, 22. [Google Scholar] [CrossRef] [Green Version]
  133. Cohen, A.; Bont, L.; Engelhard, D.; Moore, E.; Fernández, D.; Kreisberg-Greenblatt, R.; Oved, K.; Eden, E.; Hays, J.P. A multifaceted ‘omics’ approach for addressing the challenge of antimicrobial resistance. Future Microbiol. 2015, 10, 365–376. [Google Scholar] [CrossRef] [Green Version]
  134. Kang, X.; Zhang, W.; Cai, X.; Zhu, T.; Xue, Y.; Liu, C. Bacillus velezensis CC09: A Potential ‘Vaccine’ for Controlling Wheat Diseases. Mol. Plant-Microbe Interact. 2018, 31, 623–632. [Google Scholar] [CrossRef]
  135. Alkema, W.; Boekhorst, J.; Wels, M.; van Hijum, S.A. Microbial bioinformatics for food safety and production. Brief. Bioinform. 2016, 17, 283–292. [Google Scholar] [CrossRef] [Green Version]
  136. Wiegand, S.; Voigt, B.; Albrecht, D.; Bongaerts, J.; Evers, S.; Hecker, M.; Daniel, R.; Liesegang, H. Fermentation stage-dependent adaptations of Bacillus licheniformis during enzyme production. Microb. Cell Factories 2013, 12, 120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Amin, A.; Khan, M.A.; Ehsanullah, M.; Haroon, U.; Azam, S.M.; Hameed, A. Production of peptide antibiotics by Bacillus sp. GU 057 indigenously isolated from saline soil. Braz. J. Microbiol. 2012, 43, 1340–1346. [Google Scholar] [CrossRef] [Green Version]
  138. Zhu, J.; Li, L.; Wu, F.; Wu, Y.; Wang, Z.; Chen, X.; Li, J.; Cai, D.; Chen, S. Metabolic Engineering of Aspartic Acid Supply Modules for Enhanced Production of Bacitracin in Bacillus licheniformis. ACS Synth. Biol. 2021, 10, 2243–2251. [Google Scholar] [CrossRef]
  139. Haavik, H.I.; Vessia, B. Bacitracin production by the high-yielding mutant bacillus licheniformis strain al: Stimulatory effect of l-leucine. Acta Pathol. Microbiol. Scand. Sect. B Microbiol. 1978, 86B, 67–70. [Google Scholar] [CrossRef]
  140. Ganchev, K.; Kozhukharova, L. Bacitracin biosynthesis by Bacillus licheniformis 16. Acta Microbiol. Bulg. 1984, 15, 38–42. [Google Scholar] [PubMed]
  141. Egorov, N.S.; Loriia, Z.K.; Vybornykh, S.N.; Khamrun, R. Effect of bacitracin on the sporulation of Bacillus licheniformis 28 KA. Nauchnye Dokl. vysshei shkoly. Biol. Nauk. 1985, 6, 89–91. [Google Scholar]
  142. Vitković, L.; Sadoff, H.L. In vitro production of bacitracin by proteolysis of vegetative Bacillus licheniformis cell protein. J. Bacteriol. 1977, 131, 897–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Jamil, B.; Hasan, F.; Hameed, A.; Ahmed, S. Isolation of Bacillus subtilis MH-4 from soil and its potential of polypeptidic antibiotic production. Pak. J. Pharm. Sci. 2007, 20, 26–31. [Google Scholar] [PubMed]
  144. Amin, A.; Khan, M.A.; Ahmad, T. Optimized antimicrobial peptide (Bacitracin) production by immobilized and free cells and of Bacillus Spp. GU215 using Wood chips and silicon polymer beads. Pak. J. Pharm. Sci. 2013, 26, 1077–1082. [Google Scholar] [PubMed]
  145. Lu, J.Y.; Zhou, K.; Huang, W.T.; Zhou, P.; Yang, S.; Zhao, X.; Xie, J.; Xia, L.; Ding, X. A comprehensive genomic and growth proteomic analysis of antitumor lipopeptide bacillomycin Lb biosynthesis in Bacillus amyloliquefaciens X030. Appl. Microbiol. Biotechnol. 2019, 103, 7647–7662. [Google Scholar] [CrossRef]
  146. Rogers, H.J.; Newton, G.G.F.; Abraham, E.P. Production and purification of bacilysin. Biochem. J. 1965, 97, 573–578. [Google Scholar] [CrossRef] [Green Version]
  147. Ozcengiz, G.; Alaeddinoglu, N.G.; Demain, A.L. Regulation of biosynthesis of bacilysin by Bacillus subtilis. J. Ind. Microbiol. Biotechnol. 1990, 6, 91–100. [Google Scholar] [CrossRef]
  148. Chung, S.; Kong, H.; Buyer, J.S.; Lakshman, D.K.; Lydon, J.; Kim, S.-D.; Roberts, D.P. Isolation and partial characterization of Bacillus subtilis ME488 for suppression of soilborne pathogens of cucumber and pepper. Appl. Microbiol. Biotechnol. 2008, 80, 115–123. [Google Scholar] [CrossRef]
  149. Phister, T.G.; O’Sullivan, D.J.; McKay, L.L. Identification of bacilysin, chlorotetaine, and iturin a produced by Bacillus sp. strain CS93 isolated from pozol, a Mexican fermented maize dough. Appl. Environ. Microbiol. 2004, 70, 631–634. [Google Scholar] [CrossRef] [Green Version]
  150. Vairagkar, U.; Mirza, Y. Antagonistic Activity of Antimicrobial Metabolites Produced from Seaweed-Associated Bacillus amyloliquefaciens MTCC 10456 Against Malassezia spp. Probiotics Antimicrob. Proteins 2021, 13, 1228–1237. [Google Scholar] [CrossRef]
  151. Panneerselvam, P.; Senapati, A.; Kumar, U.; Sharma, L.; Lepcha, P.; Prabhukarthikeyan, S.R.; Jahan, A.; Parameshwaran, C.; Govindharaj, G.P.P.; Lenka, S.; et al. Antagonistic and plant-growth promoting novel Bacillus species from long-term organic farming soils from Sikkim, India. 3 Biotech 2019, 9, 416. [Google Scholar] [CrossRef]
  152. AlGburi, A.; Alazzawi, S.A.; Al-Ezzy, A.I.A.; Weeks, R.; Chistyakov, V.; Chikindas, M.L. Potential Probiotics Bacillus subtilis KATMIRA1933 and Bacillus amyloliquefaciens B-1895 Co-Aggregate with Clinical Isolates of Proteus mirabilis and Prevent Biofilm Formation. Probiotics Antimicrob. Proteins 2020, 12, 1471–1483. [Google Scholar] [CrossRef] [PubMed]
  153. Ahire, J.J.; Kashikar, M.S.; Madempudi, R.S. Survival and Germination of Bacillus clausii UBBC07 Spores in in vitro Human Gastrointestinal Tract Simulation Model and Evaluation of Clausin Production. Front. Microbiol. 2020, 11, 1010. [Google Scholar] [CrossRef]
  154. Brötz, H.; Bierbaum, G.; Markus, A.; Molitor, E.; Sahl, H.G. Mode of action of the lantibiotic mersacidin: Inhibition of peptidoglycan biosynthesis via a novel mechanism? Antimicrob. Agents Chemother. 1995, 39, 714–719. [Google Scholar] [CrossRef] [Green Version]
  155. Appleyard, A.N.; Choi, S.; Read, D.M.; Lightfoot, A.; Boakes, S.; Hoffmann, A.; Chopra, I.; Bierbaum, G.; Rudd, B.A.; Dawson, M.J.; et al. Dissecting Structural and Functional Diversity of the Lantibiotic Mersacidin. Chem. Biol. 2009, 16, 490–498. [Google Scholar] [CrossRef] [Green Version]
  156. He, P.; Hao, K.; Blom, J.; Rückert, C.; Vater, J.; Mao, Z.; Wu, Y.; Hou, M.; He, P.; He, Y.; et al. Genome sequence of the plant growth promoting strain Bacillus amyloliquefaciens subsp. plantarum B9601-Y2 and expression of mersacidin and other secondary metabolites. J. Biotechnol. 2013, 164, 281–291. [Google Scholar] [CrossRef] [PubMed]
  157. Molinatto, G.; Puopolo, G.; Sonego, P.; Moretto, M.; Engelen, K.; Viti, C.; Ongena, M.; Pertot, I. Complete genome sequence of Bacillus amyloliquefaciens subsp. plantarum S499, a rhizobacterium that triggers plant defences and inhibits fungal phytopathogens. J. Biotechnol. 2016, 238, 56–59. [Google Scholar] [CrossRef] [PubMed]
  158. Scholz, R.; Molohon, K.J.; Nachtigall, J.; Vater, J.; Markley, A.L.; Süssmuth, R.D.; Mitchell, D.A.; Borriss, R. Plantazolicin, a novel microcin B17/streptolysin S-like natural product from Bacillus amyloliquefaciens FZB42. J. Bacteriol. 2011, 193, 215–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Chen, L.; Heng, J.; Qin, S.; Bian, K. A comprehensive understanding of the biocontrol potential of Bacillus velezensis LM2303 against Fusarium head blight. PLoS ONE 2018, 13, e0198560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Meyers, E.; Brown, W.E.; Principe, P.A.; Rathnum, M.L.; Parker, W.L. EM49, a new peptide antibiotic. I. Fermentation, isolation, and preliminary characterization. J. Antibiot. 1973, 26, 444–448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Wang, H.; Wang, Z.; Liu, Z.; Wang, K.; Xu, W. Membrane disruption of Fusarium oxysporum f. sp. niveum induced by myriocin from Bacillus amyloliquefaciens LZN01. Microb. Biotechnol. 2021, 14, 517–534. [Google Scholar] [CrossRef]
  162. Marahiel, M.A.; Lurz, R.; Kleinkauf, H. Characterization of chromosomal and membrane associated plasmid in Bacillus brevis ATCC 9999. J. Antibiot. 1981, 34, 323–330. [Google Scholar] [CrossRef] [Green Version]
  163. Udalova, T.P.; Gus’Kova, T.M.; Silaev, A.B. Growth of Bacillus brevis var. G.B. and formation of gramicidin S in relation to the intensity of aeration. Mikrobiologiia 1972, 41, 280–286. [Google Scholar]
  164. Jiang, J.; Gao, L.; Bie, X.; Lu, Z.; Liu, H.; Zhang, C.; Lu, F.; Zhao, H. Identification of novel surfactin derivatives from NRPS modification of Bacillus subtilis and its antifungal activity against Fusarium moniliforme. BMC Microbiol. 2016, 16, 31. [Google Scholar] [CrossRef] [Green Version]
  165. Kim, P.I.; Ryu, J.; Kim, Y.H.; Chi, Y.-T. Production of Biosurfactant Lipopeptides Iturin A, Fengycin and Surfactin A from Bacillus subtilis CMB32 for Control of Colletotrichum gloeosporioides. J. Microbiol. Biotechnol. 2010, 20, 138–145. [Google Scholar] [CrossRef] [Green Version]
  166. Chen, M.; Wang, J.; Zhu, Y.; Liu, B.; Yang, W.; Ruan, C. Antibacterial activity against Ralstonia solanacearum of the lipopeptides secreted from the Bacillus amyloliquefaciens strain FJAT-2349. J. Appl. Microbiol. 2019, 126, 1519–1529. [Google Scholar] [CrossRef]
  167. 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. (Orig. Res.) 2017, 8, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Chen, L.; Wu, Y.; Chong, X.; Xin, Q.; Wang, D.; Bian, K. Seed-borne endophytic Bacillus velezensis LHSB1 mediate the biocontrol of peanut stem rot caused by Sclerotium rolfsii. J. Appl. Microbiol. 2020, 128, 803–813. [Google Scholar] [CrossRef] [PubMed]
  169. Cao, L.; Pan, L.; Gong, L.; Yang, Y.; He, H.; Li, Y.; Peng, Y.; Li, N.; Yan, L.; Ding, X.; et al. Interaction of a novel Bacillus velezensis (BvL03) against Aeromonas hydrophila in vitro and in vivo in grass carp. Appl. Microbiol. Biotechnol. 2019, 103, 8987–8999. [Google Scholar] [CrossRef]
  170. Zidour, M.; Belguesmia, Y.; Cudennec, B.; Grard, T.; Flahaut, C.; Souissi, S.; Drider, D. Genome Sequencing and Analysis of Bacillus pumilus ICVB403 Isolated from Acartia tonsa Copepod Eggs Revealed Surfactin and Bacteriocin Production: Insights on Anti-Staphylococcus Activity. Probiotics Antimicrob. Proteins 2019, 11, 990–998. [Google Scholar] [CrossRef] [PubMed]
  171. Sarwar, A.; Hassan, M.N.; Imran, M.; Iqbal, M.; Majeed, S.; Brader, G.; Sessitsch, A.; Hafeez, F.Y. Biocontrol activity of surfactin A purified from Bacillus NH-100 and NH-217 against rice bakanae disease. Microbiol. Res. 2018, 209, 1–13. [Google Scholar] [CrossRef]
  172. Zhang, B.; Li, Y.; Zhang, Y.; Qiao, H.; He, J.; Yuan, Q.; Chen, X.; Fan, J. High-cell-density culture enhances the antimicrobial and freshness effects of Bacillus subtilis S1702 on table grapes (Vitis vinifera cv. Kyoho). Food Chem. 2019, 286, 541–549. [Google Scholar] [CrossRef] [PubMed]
  173. Huang, J.; Wei, Z.; Tan, S.; Mei, X.; Shen, Q.; Xu, Y. Suppression of Bacterial Wilt of Tomato by Bioorganic Fertilizer Made from the Antibacterial Compound Producing Strain Bacillus amyloliquefaciens HR62. J. Agric. Food Chem. 2014, 62, 10708–10716. [Google Scholar] [CrossRef] [PubMed]
  174. Aleti, G.; Lehner, S.; Bacher, M.; Compant, S.; Nikolic, B.; Plesko, M.; Schuhmacher, R.; Sessitsch, A.; Brader, G. Surfactin variants mediate species-specific biofilm formation and root colonization in Bacillus. Environ. Microbiol. 2016, 18, 2634–2645. [Google Scholar] [CrossRef] [PubMed]
  175. Pueyo, M.T.; Bloch, C.; Carmona-Ribeiro, A.M.; di Mascio, P. Lipopeptides Produced by a Soil Bacillus Megaterium Strain. Microb. Ecol. 2009, 57, 367–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Zhao, P.; Quan, C.; Wang, Y.; Wang, J.; Fan, S. Bacillus amyloliquefaciens Q-426 as a potential biocontrol agent against Fusarium oxysporum f. sp. spinaciae. J. Basic Microbiol. 2014, 54, 448–456. [Google Scholar] [CrossRef]
  177. Malfanova, N.; Kamilova, F.; Validov, S.; Shcherbakov, A.; Chebotar, V.; Tikhonovich, I.; Lugtenberg, B. Characterization of Bacillus subtilis HC8, a novel plant-beneficial endophytic strain from giant hogweed. Microb. Biotechnol. 2011, 4, 523–532. [Google Scholar] [CrossRef] [Green Version]
  178. Vanittanakom, N.; Loeffler, W.; Koch, U.; Jung, G. Fengycin—A novel antifungal lipopeptide antibiotic produced by Bacillus subtilis F-29-3. J. Antibiot. 1986, 39, 888–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Ben Abdallah, D.; Frikha-Gargouri, O.; Tounsi, S. Bacillus amyloliquefaciens strain 32a as a source of lipopeptides for biocontrol of Agrobacterium tumefaciens strains. J. Appl. Microbiol. 2015, 119, 196–207. [Google Scholar] [CrossRef]
  180. Pecci, Y.; Rivardo, F.; Martinotti, M.G.; Allegrone, G. LC/ESI-MS/MS characterisation of lipopeptide biosurfactants produced by the Bacillus licheniformis V9T14 strain. Biol. Mass Spectrom. 2010, 45, 772–778. [Google Scholar] [CrossRef]
  181. Li, X.-Y.; Mao, Z.-C.; Wang, Y.-H.; Wu, Y.-X.; He, Y.-Q.; Long, C.-L. ESI LC-MS and MS/MS Characterization of Antifungal Cyclic Lipopeptides Produced by Bacillus subtilis XF-1. J. Mol. Microbiol. Biotechnol. 2012, 22, 83–93. [Google Scholar] [CrossRef]
  182. Ma, Z.; Wang, N.; Hu, J.; Wang, S. Isolation and characterization of a new iturinic lipopeptide, mojavensin A produced by a marine-derived bacterium Bacillus mojavensis B0621A. J. Antibiot. 2012, 65, 317–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Ding, L.; Guo, W.; Chen, X. Exogenous addition of alkanoic acids enhanced production of antifungal lipopeptides in Bacillus amyloliquefaciens Pc3. Appl. Microbiol. Biotechnol. 2019, 103, 5367–5377. [Google Scholar] [CrossRef] [PubMed]
  184. Nastro, R.A.; Arguelles-Arias, A.; Ongena, M.; Di Costanzo, A.; Trifuoggi, M.; Guida, M.; Fickers, P. Antimicrobial Activity of Bacillus amyloliquefaciens ANT1 Toward Pathogenic Bacteria and Mold: Effects on Biofilm Formation. Probiotics Antimicrob. Proteins 2013, 5, 252–258. [Google Scholar] [CrossRef]
  185. Malfanova, N.; Franzil, L.; Lugtenberg, B.; Chebotar, V.; Ongena, M. Cyclic lipopeptide profile of the plant-beneficial endophytic bacterium Bacillus subtilis HC8. Arch. Microbiol. 2012, 194, 893–899. [Google Scholar] [CrossRef] [Green Version]
  186. Hernández-Morales, A.; Martínez-Peniche, R.A.; Arvizu-Gómez, J.L.; Arvizu-Medrano, S.M.; Rodríguez-Ontiveros, A.; Ramos-López, M.A.; Pacheco-Aguilar, J.R. Production of a Mixture of Fengycins with Surfactant and Antifungal Activities by Bacillus sp. MA04, a Versatile PGPR. Indian J. Microbiol. 2018, 58, 208–213. [Google Scholar] [CrossRef]
  187. Escobar, V.V.; Ceballos, I.; Mira, J.J.; Argel, L.E.; Peralta, S.O.; Romero-Tabarez, M. Fengycin C Produced by Bacillus subtilis EA-CB0015. J. Nat. Prod. 2013, 76, 503–509. [Google Scholar] [CrossRef]
  188. Klich, M.A.; Lax, A.R.; Bland, J. Inhibition of some mycotoxigenic fungi by iturin A, a peptidolipid produced by Bacillus subtilis. Mycopathologia 1991, 116, 77–80. [Google Scholar] [CrossRef] [PubMed]
  189. Dang, Y.; Zhao, F.; Liu, X.; Fan, X.; Huang, R.; Gao, W.; Wang, S.; Yang, C. Enhanced production of antifungal lipopeptide iturin A by Bacillus amyloliquefaciens LL3 through metabolic engineering and culture conditions optimization. Microb. Cell Factories 2019, 18, 68. [Google Scholar] [CrossRef]
  190. Cho, S.J.; Lee, S.K.; Cha, B.J.; Kim, Y.H.; Shin, K.S. Detection and characterization of the Gloeosporium gloeosporioides growth inhibitory compound iturin A from Bacillus subtilis strain KS03. FEMS Microbiol. Lett. 2003, 223, 47–51. [Google Scholar] [CrossRef] [Green Version]
  191. Mizumoto, S.; Hirai, M.; Shoda, M. Production of lipopeptide antibiotic iturin A using soybean curd residue cultivated with Bacillus subtilis in solid-state fermentation. Appl. Microbiol. Biotechnol. 2006, 72, 869–875. [Google Scholar] [CrossRef]
  192. Chen, H.; Yuan, C.; Cai, K.; Zheng, Z.; Yu, Z. Purification and identification of iturin A from Bacillus subtilis JA by electrospray ionization mass spectrometry. Acta Microbiol. Sin. 2008, 48, 116–120. [Google Scholar]
  193. Thasana, N.; Prapagdee, B.; Rangkadilok, N.; Sallabhan, R.; Aye, S.L.; Ruchirawat, S.; Loprasert, S. Bacillus subtilis SSE4 produces subtulene A, a new lipopeptide antibiotic possessing an unusual C15 unsaturated β-amino acid. FEBS Lett. 2010, 584, 3209–3214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Arrebola, E.; Jacobs, R.; Korsten, L. Iturin A is the principal inhibitor in the biocontrol activity of Bacillus amyloliquefaciens PPCB004 against postharvest fungal pathogens. J. Appl. Microbiol. 2010, 108, 386–395. [Google Scholar] [CrossRef]
  195. Jin, H.; Zhang, X.; Li, K.; Niu, Y.; Guo, M.; Hu, C.; Wan, X.; Gong, Y.; Huang, F. Direct Bio-Utilization of Untreated Rapeseed Meal for Effective Iturin A Production by Bacillus subtilis in Submerged Fermentation. PLoS ONE 2014, 9, e111171. [Google Scholar] [CrossRef]
  196. Yamamoto, S.; Shiraishi, S.; Suzuki, S. Are cyclic lipopeptides produced by Bacillus amyloliquefaciens S13-3 responsible for the plant defence response in strawberry against Colletotrichum gloeosporioides? Lett. Appl. Microbiol. 2015, 60, 379–386. [Google Scholar] [CrossRef] [PubMed]
  197. Chen, D.; Liu, X.; Li, C.; Tian, W.; Shen, Q.; Shen, B. Isolation of Bacillus amyloliquefaciens S20 and its application in control of eggplant bacterial wilt. J. Environ. Manag. 2014, 137, 120–127. [Google Scholar] [CrossRef]
  198. Saechow, S.; Thammasittirong, A.; Kittakoop, P.; Prachya, S.; Thammasittirong, S.N.-R. Antagonistic Activity against Dirty Panicle Rice Fungal Pathogens and Plant Growth-Promoting Activity of Bacillus amyloliquefaciens BAS23. J. Microbiol. Biotechnol. 2018, 28, 1527–1535. [Google Scholar] [CrossRef]
  199. Han, Y.; Zhang, B.; Shen, Q.; You, C.; Yu, Y.; Li, P.; Shang, Q. Purification and Identification of Two Antifungal Cyclic Peptides Produced by Bacillus amyloliquefaciens L-H15. Appl. Biochem. Biotechnol. 2015, 176, 2202–2212. [Google Scholar] [CrossRef] [PubMed]
  200. 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] [Green Version]
  201. Zhang, Z.; Ding, Z.; Zhong, J.; Zhou, J.; Shu, D.; Luo, D.; Yang, J.; Tan, H. Improvement of iturin A production in Bacillus subtilis ZK0 by overexpression of the comA and sigA genes. Lett. Appl. Microbiol. 2017, 64, 452–458. [Google Scholar] [CrossRef]
  202. Athukorala, S.N.; Fernando, W.G.; Rashid, K.Y. Identification of antifungal antibiotics of Bacillus species isolated from different microhabitats using polymerase chain reaction and MALDI-TOF mass spectrometry. Can. J. Microbiol. 2009, 55, 1021–1032. [Google Scholar] [CrossRef]
  203. Shi, J.; Zhu, X.; Lu, Y.; Zhao, H.; Lu, F.; Lu, Z. Improving Iturin A Production of Bacillus amyloliquefaciens by Genome Shuffling and Its Inhibition Against Saccharomyces cerevisiae in Orange Juice. Front. Microbiol. 2018, 9, 2683. [Google Scholar] [CrossRef]
  204. Kaushal, M.; Kumar, A.; Kaushal, R. Bacillus pumilus strain YSPMK11 as plant growth promoter and bicontrol agent against Sclerotinia sclerotiorum. 3 Biotech 2017, 7, 90. [Google Scholar] [CrossRef] [Green Version]
  205. Moryl, M.; Spętana, M.; Dziubek, K.; Paraszkiewicz, K.; Różalska, S.; Płaza, G.A.; Rozalski, A. Antimicrobial, antiadhesive and antibiofilm potential of lipopeptides synthesised by Bacillus subtilis, on uropathogenic bacteria. Acta Biochim. Pol. 2015, 62, 725–732. [Google Scholar] [CrossRef] [PubMed]
  206. Habe, H.; Taira, T.; Imura, T. Screening of a Bacillus subtilis Strain Producing Multiple Types of Cyclic Lipopeptides and Evaluation of Their Surface-tension-lowering Activities. J. Oleo Sci. 2017, 66, 785–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Xu, B.-H.; Ye, Z.-W.; Zheng, Q.-W.; Wei, T.; Lin, J.-F.; Guo, L.-Q. Isolation and characterization of cyclic lipopeptides with broad-spectrum antimicrobial activity from Bacillus siamensis JFL15. 3 Biotech 2018, 8, 444. [Google Scholar] [CrossRef]
  208. Ambrico, A.; Trupo, M.; Magarelli, R.A. Influence of Phenotypic Dissociation in Bacillus subtilis Strain ET-1 on Iturin A Production. Curr. Microbiol. 2019, 76, 1487–1494. [Google Scholar] [CrossRef] [PubMed]
  209. Lee, T.; Park, D.; Kim, K.; Lim, S.M.; Yu, N.H.; Kim, S.; Kim, H.-Y.; Jung, K.S.; Jang, J.Y.; Park, J.-C.; et al. Characterization of Bacillus amyloliquefaciens DA12 Showing Potent Antifungal Activity against Mycotoxigenic Fusarium Species. Plant Pathol. J. 2017, 33, 499–507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Chen, W.; Li, X.; Ma, X.; Chen, S.; Kang, Y.; Yang, M.; Huang, F.; Wan, X. Simultaneous hydrolysis with lipase and fermentation of rapeseed cake for iturin A production by Bacillus amyloliquefaciens CX-20. BMC Biotechnol. 2019, 19, 98. [Google Scholar] [CrossRef] [Green Version]
  211. Lin, C.; Tsai, C.-H.; Chen, P.-Y.; Wu, C.-Y.; Chang, Y.-L.; Yang, Y.-L.; Chen, Y.-L. Biological control of potato common scab by Bacillus amyloliquefaciens Ba01. PLoS ONE 2018, 13, e0196520. [Google Scholar] [CrossRef] [Green Version]
  212. Waewthongrak, W.; Leelasuphakul, W.; Mccollum, G. Cyclic Lipopeptides from Bacillus subtilis ABS–S14 Elicit Defense-Related Gene Expression in Citrus Fruit. PLoS ONE 2014, 9, e109386. [Google Scholar] [CrossRef]
  213. Wu, G.; Liu, Y.; Xu, Y.; Zhang, G.; Shen, Q.-R.; Zhang, R. Exploring Elicitors of the Beneficial Rhizobacterium Bacillus amyloliquefaciens SQR9 to Induce Plant Systemic Resistance and Their Interactions with Plant Signaling Pathways. Mol. Plant-Microbe Interactions 2018, 31, 560–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Zhao, Z.; Wang, Q.; Wang, K.; Brian, K.; Liu, C.; Gu, Y. Study of the antifungal activity of Bacillus vallismortis ZZ185 in vitro and identification of its antifungal components. Bioresour. Technol. 2010, 101, 292–297. [Google Scholar] [CrossRef]
  215. Mácha, H.; Marešová, H.; Juříková, T.; Švecová, M.; Benada, O.; Škríba, A.; Baránek, M.; Novotný, Č.; Palyzová, A. Killing Effect of Bacillus velezensis FZB42 on a Xanthomonas campestris pv. Campestris (Xcc) Strain Newly Isolated from Cabbage Brassica oleracea Convar. Capitata (L.): A Metabolomic Study. Microorganisms 2021, 9, 1410. [Google Scholar] [CrossRef]
  216. Elkahoui, S.; Djébali, N.; Karkouch, I.; Ibrahim, A.H.; Kalai, L.; Bachkovel, S.; Tabbene, O.; Limam, F. Mass spectrometry identification of antifungal lipopeptides from Bacillus sp. BCLRB2 against Rhizoctonia solani and Sclerotinia sclerotiorum. Microb. Pathog. 2014, 50, 184–188. [Google Scholar] [CrossRef]
  217. Nam, J.; Alam, S.T.; Kang, K.; Choi, J.; Seo, M.-H. Anti-staphylococcal activity of a cyclic lipopeptide, C15-bacillomycin D, produced by Bacillus velezensis NST. J. Appl. Microbiol. 2021, 131, 93–104. [Google Scholar] [CrossRef] [PubMed]
  218. Soussi, S.; Essid, R.; Hardouin, J.; Gharbi, D.; Elkahoui, S.; Tabbene, O.; Cosette, P.; Jouenne, T.; Limam, F. Utilization of Grape Seed Flour for Antimicrobial Lipopeptide Production by Bacillus amyloliquefaciens C5 Strain. Appl. Biochem. Biotechnol. 2019, 187, 1460–1474. [Google Scholar] [CrossRef] [PubMed]
  219. Radovanovic, N.; Milutinović, M.; Mihajlovski, K.; Jović, J.; Nastasijević, B.; Rajilic-Stojanovic, M.; Dimitrijević-Branković, S. Biocontrol and plant stimulating potential of novel strain Bacillus sp. PPM3 isolated from marine sediment. Microb. Pathog. 2018, 120, 71–78. [Google Scholar] [CrossRef] [PubMed]
  220. Walton, R.B.; Woodruff, H.B. A crystalline antifungal agent, mycosubtilin, isolated from subtilin broth. J. Clin. Investig. 1949, 28, 924–926. [Google Scholar] [CrossRef]
  221. Fickers, P.; Guez, J.-S.; Damblon, C.; Leclère, V.; Béchet, M.; Jacques, P.; Joris, B. High-Level Biosynthesis of the Anteiso-C 17 Isoform of the Antibiotic Mycosubtilin in Bacillus subtilis and Characterization of Its Candidacidal Activity. Appl. Environ. Microbiol. 2009, 75, 4636–4640. [Google Scholar] [CrossRef] [Green Version]
  222. Chevanet, C.; Besson, F.; Michel, G. Effect of various growth conditions on spore formation and Bacillomycin L production in Bacillus subtilis. Can. J. Microbiol. 1986, 32, 254–258. [Google Scholar] [CrossRef] [PubMed]
  223. Li, X.; Zhang, Y.; Wei, Z.; Guan, Z.; Cai, Y.; Liao, X. Antifungal Activity of Isolated Bacillus amyloliquefaciens SYBC H47 for the Biocontrol of Peach Gummosis. PLoS ONE 2016, 11, e0162125. [Google Scholar] [CrossRef] [Green Version]
  224. Sengupta, S.; Bose, S. Properties and localisation of mycobacillin-synthesising enzyme system in Bacillus subtilis B3. Biochim. et Biophys. Acta (BBA)-Gen. Subj. 1971, 237, 120–122. [Google Scholar] [CrossRef]
  225. Velho, R.V.; Caldas, D.G.G.; Medina, L.F.C.; Tsai, S.M.; Brandelli, A. Real-time PCR investigation on the expression of sboA and ituD genes in Bacillus spp. Lett. Appl. Microbiol. 2011, 52, 660–666. [Google Scholar] [CrossRef] [PubMed]
  226. Liu, X.; Lee, J.Y.; Jeong, S.-J.; Cho, K.M.; Kim, G.M.; Shin, J.-H.; Kim, J.-S.; Kim, J.H. Properties of a Bacteriocin Produced by Bacillus subtilis EMD4 Isolated from Ganjang (Soy Sauce). J. Microbiol. Biotechnol. 2015, 25, 1493–1501. [Google Scholar] [CrossRef] [Green Version]
  227. Sutyak, K.; Wirawan, R.; Aroutcheva, A.; Chikindas, M. Isolation of the Bacillus subtilis antimicrobial peptide subtilosin from the dairy product-derived Bacillus amyloliquefaciens. J. Appl. Microbiol. 2008, 104, 1067–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Liu, Q.; Gao, G.; Xu, H.; Qiao, M. Identification of the bacteriocin subtilosin A and loss of purL results in its high-level production in Bacillus amyloliquefaciens. Res. Microbiol. 2012, 163, 470–478. [Google Scholar] [CrossRef]
  229. Parveen Rani, R.; Anandharaj, M.; Hema, S.; Deepika, R.; David Ravindran, A. Purification of Antilisterial Peptide (Subtilosin A) from Novel Bacillus tequilensis FR9 and Demonstrate Their Pathogen Invasion Protection Ability Using Human Carcinoma Cell Line. Front Microbiol. 2016, 7, 1910. [Google Scholar] [CrossRef] [Green Version]
  230. Liu, H.; Yin, S.; An, L.; Zhang, G.; Cheng, H.; Xi, Y.; Cui, G.; Zhang, F.; Zhang, L. Complete genome sequence of Bacillus subtilis BSD-2, a microbial germicide isolated from cultivated cotton. J. Biotechnol. 2016, 230, 26–27. [Google Scholar] [CrossRef]
  231. Li, Q.; Liao, S.; Wei, J.; Xing, D.; Xiao, Y.; Yang, Q. Isolation of Bacillus subtilis strain SEM-2 from silkworm excrement and characterisation of its antagonistic effect against Fusarium spp. Can. J. Microbiol. 2020, 66, 401–412. [Google Scholar] [CrossRef] [Green Version]
  232. Chen, L.; Shi, H.; Heng, J.; Wang, D.; Bian, K. Antimicrobial, plant growth-promoting and genomic properties of the peanut endophyte Bacillus velezensis LDO2. Microbiol. Res. 2019, 218, 41–48. [Google Scholar] [CrossRef] [PubMed]
  233. Gao, X.-Y.; Liu, Y.; Miao, L.-L.; Li, E.-W.; Sun, G.-X.; Liu, Z.-P. Characterization and mechanism of anti-Aeromonas salmonicida activity of a marine probiotic strain, Bacillus velezensis V4. Appl. Microbiol. Biotechnol. 2017, 101, 3759–3768. [Google Scholar] [CrossRef]
  234. Sharma, D.; Singh, S.S.; Baindara, P.; Sharma, S.; Khatri, N.; Grover, V.; Patil, P.B.; Korpole, S. Surfactin Like Broad Spectrum Antimicrobial Lipopeptide Co-produced With Sublancin From Bacillus subtilis Strain A52: Dual Reservoir of Bioactives. Front. Microbiol. 2020, 11, 1167. [Google Scholar] [CrossRef] [PubMed]
  235. Li, Y.; Xu, Y.; Liu, L.; Han, Z.; Lai, P.Y.; Guo, X.; Zhang, X.; Lin, W.; Qian, P.-Y. Five New Amicoumacins Isolated from a Marine-Derived Bacterium Bacillus subtilis. Mar. Drugs 2012, 10, 319–328. [Google Scholar] [CrossRef] [Green Version]
  236. Park, H.B.; Perez, C.E.; Perry, E.; Crawford, J.M. Activating and Attenuating the Amicoumacin Antibiotics. Molecules 2016, 21, 824. [Google Scholar] [CrossRef] [PubMed]
  237. Itoh, J.; Shomura, T.; Omoto, S.; Miyado, S.; Yuda, Y.; Shibata, U.; Inouye, S. Isolation, Physicochemical Properties and Biological Activities of Amicoumacins Produced by Bacillus pumilus. Agric. Biol. Chem. 1982, 46, 1255–1259. [Google Scholar] [CrossRef]
  238. Wang, D.; Li, J.; Zhu, G.; Zhao, K.; Jiang, W.; Li, H.; Wang, W.; Kumar, V.; Dong, S.; Zhu, W.; et al. Mechanism of the Potential Therapeutic Candidate Bacillus subtilis BSXE-1601 Against Shrimp Pathogenic Vibrios and Multifunctional Metabolites Biosynthetic Capability of the Strain as Predicted by Genome Analysis. Front. Microbiol. 2020, 11, 581802. [Google Scholar] [CrossRef] [PubMed]
  239. Liao, R.; Duan, L.; Lei, C.; Pan, H.; Ding, Y.; Zhang, Q.; Chen, D.; Shen, B.; Yu, Y.; Liu, W. Thiopeptide Biosynthesis Featuring Ribosomally Synthesized Precursor Peptides and Conserved Posttranslational Modifications. Chem. Biol. 2009, 16, 141–147. [Google Scholar] [CrossRef] [Green Version]
  240. Akasapu, S.; Hinds, A.B.; Powell, W.C.; Walczak, M.A. Total synthesis of micrococcin P1 and thiocillin I enabled by Mo(vi) catalyst. Chem. Sci. 2019, 10, 1971–1975. [Google Scholar] [CrossRef] [Green Version]
  241. Villa-Rodriguez, E.; Moreno-Ulloa, A.; Castro-Longoria, E.; Parra-Cota, F.I.; Santos-Villalobos, S.D.L. Integrated omics approaches for deciphering antifungal metabolites produced by a novel Bacillus species, B. cabrialesii TE3T, against the spot blotch disease of wheat (Triticum turgidum L. subsp. durum). Microbiol. Res. 2021, 251, 126826. [Google Scholar] [CrossRef]
  242. Kino, K.; Kotanaka, Y.; Arai, T.; Yagasaki, M. A novel L-amino acid ligase from Bacillus subtilis NBRC3134, a microorganism producing peptide-antibiotic rhizocticin. Biosci. Biotechnol. Biochem. 2009, 73, 901–907. [Google Scholar] [CrossRef] [Green Version]
  243. Patel, P.S.; Huang, S.; Fisher, S.; Pirnik, D.; Aklonis, C.; Dean, L.; Meyers, E.; Fernandes, P.; Mayerl, F. Bacillaene, a Novel Inhibitor of Procaryotic Protein Synthesis Produced by Bacillus subtilis: Production, Taxonomy, Isolation, Physico-chemical Characterization and Biological Activity. J. Antibiot. 1995, 48, 997–1003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Lv, J.; Da, R.; Cheng, Y.; Tuo, X.; Wei, J.; Jiang, K.; Monisayo, A.O.; Han, B. Mechanism of Antibacterial Activity of Bacillus amyloliquefaciens C-1 Lipopeptide toward Anaerobic Clostridium difficile. BioMed Res. Int. 2020, 2020, 3104613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Chen, K.; Tian, Z.; Luo, Y.; Cheng, Y.; Long, C.-A. Antagonistic Activity and the Mechanism of Bacillus amyloliquefaciens DH-4 Against Citrus Green Mold. Phytopathology 2018, 108, 1253–1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Nonejuie, P.; Trial, R.M.; Newton, G.L.; Lamsa, A.; Perera, V.R.; Aguilar, J.; Liu, W.-T.; Dorrestein, P.C.; Pogliano, J.; Pogliano, K. Application of bacterial cytological profiling to crude natural product extracts reveals the antibacterial arsenal of Bacillus subtilis. J. Antibiot. 2016, 69, 353–361. [Google Scholar] [CrossRef] [Green Version]
  247. Daas, M.S.; Acedo, J.; Rosana, A.R.; Orata, F.; Reiz, B.; Zheng, J.; Nateche, F.; Case, R.J.; Kebbouche-Gana, S.; Vederas, J.C. Bacillus amyloliquefaciens ssp. plantarum F11 isolated from Algerian salty lake as a source of biosurfactants and bioactive lipopeptides. FEMS Microbiol. Lett. 2018, 365, fnx248. [Google Scholar] [CrossRef] [PubMed]
  248. Butcher, R.A.; Schroeder, F.C.; Fischbach, M.A.; Straight, P.D.; Kolter, R.; Walsh, C.T.; Clardy, J. The identification of bacillaene, the product of the PksX megacomplex in Bacillus subtilis. Proc. Natl. Acad. Sci. USA 2007, 104, 1506–1509. [Google Scholar] [CrossRef] [Green Version]
  249. Dimopoulou, A.; Theologidis, I.; Benaki, D.; Koukounia, M.; Zervakou, A.; Tzima, A.; Diallinas, G.; Hatzinikolaou, D.G.; Skandalis, N. Direct Antibiotic Activity of Bacillibactin Broadens the Biocontrol Range of Bacillus amyloliquefaciens MBI600. mSphere 2021, 6, e0037621. [Google Scholar] [CrossRef]
  250. Peters, W.J.; Warren, R.A.J. Itoic Acid Synthesis in Bacillus subtilis. J. Bacteriol. 1968, 95, 360–366. [Google Scholar] [CrossRef] [Green Version]
  251. Ollinger, J.; Song, K.-B.; Antelmann, H.; Hecker, M.; Helmann, J.D. Role of the Fur Regulon in Iron Transport in Bacillus subtilis. J. Bacteriol. 2006, 188, 3664–3673. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The number of Bacillus strains reported for each species.
Figure 1. The number of Bacillus strains reported for each species.
Antibiotics 11 00088 g001
Figure 2. Metabolites targeting (a) cell wall and (b) plasma membrane.
Figure 2. Metabolites targeting (a) cell wall and (b) plasma membrane.
Antibiotics 11 00088 g002
Figure 3. Chemical structures of bacitracin A (1), bacilysin (2), chlorotetaine (3), and kanosamine (4).
Figure 3. Chemical structures of bacitracin A (1), bacilysin (2), chlorotetaine (3), and kanosamine (4).
Antibiotics 11 00088 g003
Figure 4. Chemical structures of subtilin (5), clausin (6), mersacidin (7), amylolysin A (8), and haloduracin (9).
Figure 4. Chemical structures of subtilin (5), clausin (6), mersacidin (7), amylolysin A (8), and haloduracin (9).
Antibiotics 11 00088 g004
Figure 5. Chemical structures of ε-poly-L-Lysine (10), plantazolicin (11), octapeptin B (12), aurantinin B (13), aurantinin C (14), aurantinin (D) (15), myriocin (16), gramicidin A (17), and gramicidin S (18).
Figure 5. Chemical structures of ε-poly-L-Lysine (10), plantazolicin (11), octapeptin B (12), aurantinin B (13), aurantinin C (14), aurantinin (D) (15), myriocin (16), gramicidin A (17), and gramicidin S (18).
Antibiotics 11 00088 g005
Figure 6. Chemical structures of surfactin A–C (1921), lichenysin (22), and fengycin A–D (2326).
Figure 6. Chemical structures of surfactin A–C (1921), lichenysin (22), and fengycin A–D (2326).
Antibiotics 11 00088 g006
Figure 7. Chemical structures of iturin A (27), bacillomycin D (28), bacillomycin L (29), and mycosubtilin (30).
Figure 7. Chemical structures of iturin A (27), bacillomycin D (28), bacillomycin L (29), and mycosubtilin (30).
Antibiotics 11 00088 g007
Figure 8. Chemical structures of mycobacillin (31) and subtilosin A (32).
Figure 8. Chemical structures of mycobacillin (31) and subtilosin A (32).
Antibiotics 11 00088 g008
Figure 9. Chemical structures of zwittermicin A (33), difficidin (34), and sublancin (35).
Figure 9. Chemical structures of zwittermicin A (33), difficidin (34), and sublancin (35).
Antibiotics 11 00088 g009
Figure 10. Chemical structures of amicoumacin A (36), prumycin (37), thiocillin (38), hetiamacin E (39), hetiamacin F (40), rhizocticin A (41), macrolactin N (42), and azoxybacilin (43).
Figure 10. Chemical structures of amicoumacin A (36), prumycin (37), thiocillin (38), hetiamacin E (39), hetiamacin F (40), rhizocticin A (41), macrolactin N (42), and azoxybacilin (43).
Antibiotics 11 00088 g010
Figure 11. Chemical structures of stigmatellin Y (44), bacillaene (45), bacillibactin (46), and schizokinen (47).
Figure 11. Chemical structures of stigmatellin Y (44), bacillaene (45), bacillibactin (46), and schizokinen (47).
Antibiotics 11 00088 g011
Figure 12. The analysis of (a) the number of strains in each species and the (b) mechanism of action targeted by each strain.
Figure 12. The analysis of (a) the number of strains in each species and the (b) mechanism of action targeted by each strain.
Antibiotics 11 00088 g012
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tran, C.; Cock, I.E.; Chen, X.; Feng, Y. Antimicrobial Bacillus: Metabolites and Their Mode of Action. Antibiotics 2022, 11, 88. https://doi.org/10.3390/antibiotics11010088

AMA Style

Tran C, Cock IE, Chen X, Feng Y. Antimicrobial Bacillus: Metabolites and Their Mode of Action. Antibiotics. 2022; 11(1):88. https://doi.org/10.3390/antibiotics11010088

Chicago/Turabian Style

Tran, Charlie, Ian E. Cock, Xiaojing Chen, and Yunjiang Feng. 2022. "Antimicrobial Bacillus: Metabolites and Their Mode of Action" Antibiotics 11, no. 1: 88. https://doi.org/10.3390/antibiotics11010088

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