Next Article in Journal
Biological and Pharmacological Effects of Synthetic Saponins
Previous Article in Journal
Recent Developments in Transition-Metal Catalyzed Direct C–H Alkenylation, Alkylation, and Alkynylation of Azoles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antimicrobial Activities of Lipopeptides and Polyketides of Bacillus velezensis for Agricultural Applications

by
Muhammad Fazle Rabbee
and
Kwang-Hyun Baek
*
Department of Biotechnology, Yeungnam University, Gyeongsan, Gyeongbuk 38541, Korea
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(21), 4973; https://doi.org/10.3390/molecules25214973
Submission received: 10 September 2020 / Revised: 23 October 2020 / Accepted: 25 October 2020 / Published: 27 October 2020

Abstract

:
Since the discovery of penicillin, bacteria are known to be major sources of secondary metabolites that can function as drugs or pesticides. Scientists worldwide attempted to isolate novel compounds from microorganisms; however, only less than 1% of all existing microorganisms have been successfully identified or characterized till now. Despite the limitations and gaps in knowledge, in recent years, many Bacillus velezensis isolates were identified to harbor a large number of biosynthetic gene clusters encoding gene products for the production of secondary metabolites. These chemically diverse bioactive metabolites could serve as a repository for novel drug discovery. More specifically, current projects on whole-genome sequencing of B. velezensis identified a large number of biosynthetic gene clusters that encode enzymes for the synthesis of numerous antimicrobial compounds, including lipopeptides and polyketides; nevertheless, their biological applications are yet to be identified or established. In this review, we discuss the recent research on synthesis of bioactive compounds by B. velezensis and related Bacillus species, their chemical structures, bioactive gene clusters of interest, as well as their biological applications for effective plant disease management.

1. Introduction

Bacillus velezensis was first identified in 2005 by Ruiz-García et al. by isolating two novel Bacillus species from environmental samples of the Vélez river [1]. This bacterium is known to exert antagonistic effects against plant pathogens via production of diverse antimicrobial compounds [2,3,4]. In 2016, several other Bacillus species previously classified as B. amyloliquefaciens subsp. plantarum, B. methylotrophicus, and B. oryzicola were re-classified as strains of B. velezensis [5]. Phylogenetic analysis based on RNA polymerase beta-subunit gene sequence and core genome, revealed that B. velezensis belongs to a conspecific group consisting of B. velezensis, B. methylotrophicus, and B. amyloliquefaciens subsp. plantarum FZB42 (reclassified as B. velezensis FZB42); however, it is distinct from the closely related species of B. subtilis, B. amyloliquefaciens, and B. siamensis [6].
The plant-associated B. velezensis FZB42 genome was first sequenced in 2007, which revealed the presence of nine giant gene clusters representing approximately 10% of the whole genome. These biosynthetic gene clusters (i.e., srf, bmy, fen, dhb, bac, mln, bae, dfn, and nrs) encode the biosynthetic enzymes for the antimicrobial compounds, namely surfactin, bacillomycin-D, fengycins, bacillibactin, bacilysin, macrolactin, bacillaene, difficidin, and a putative peptide with unknown functions, respectively (Figure 1) [7]. Among the nine gene clusters, five encode the biosynthetic enzymes that are involved in the synthesis of non-ribosomal lipopeptides (LPs), where synthesis takes place on large enzyme complexes of non-ribosomal peptide synthetases (NRPSs). LPs share similar structures consisting of a hydrophilic peptide portion linked to the hydrophobic fatty acid chain, which could be divided into three major sub-families based on the amino acid sequence—surfactins (srf), bacillomycin-D (bmy), and fengycins (fen) or plipastatins (pps) [8]. Moreover, three more polyketide synthase (PKSs) gene clusters were identified that directed the synthesis of polyketides (PKs), e.g., macrolactin (mln), bacillaene (bae), and difficidin (dfn) [9,10]. PKSs and NRPSs function as multi-enzyme complexes that sequentially combine malonyl derivatives and amino acids, respectively. These tailoring enzymes employ different building blocks to synthesize a variety of secondary metabolites with therapeutic potential [11]. The products of bac gene cluster guide the synthesis and export of the antibacterial dipeptide bacilysin [10]. In B. velezensis, all of the three LP and three PK type compounds are biosynthesized via the 4’-phosphopantetheine transferase (Sfp) pathway [12]; however, the production of antibacterial compound bacilysin is independent of this pathway [13]. In addition, two other ribosomally-synthesized bacteriocins classified as amylocyclicin and plantazolicin were identified in B. velezensis, displaying high antibacterial activity against closely related gram-positive bacteria [13,14].
Genome mining of B. velezensis LM2303 revealed 13 biosynthetic gene clusters encoding the enzymes for the production of the secondary metabolites, with biocontrol potentiality against the pathogenic fungus Fusarium graminearum [3]. The production of the metabolites was further confirmed by chemical analysis using ultra-high-performance liquid chromatography-electrospray ionization (ESI)-mass spectrometry (MS) [3]. Among them, three gene clusters encode the enzymes for antifungal metabolites (i.e., surfactin A, iturin A, and fengycin B); eight gene clusters encode the enzymes for antibacterial metabolites (i.e., difficidin, bacilysin, bacillaene, macrolactin, plantazolicin, kijanimicin, butirosin, and surfactin A); and another three gene clusters encode the enzymes for the synthesis of metabolites involved in nutrient uptake (i.e., bacillibactin, teichuronic acid, and molybdenum cofactor) [3]. Under field conditions, LM2303 exhibited strong biocontrol efficacy against F. graminearum, by greatly reducing the incidence of Fusarium head blight, with a control efficiency of approximately 72.3% [3].
Apart from the specific antagonistic activity of B. velezensis against pathogenic microbes (Table 1), this bacterium was also found to contribute to plant protection by competing with harmful microorganisms for vital nutrients like iron, through the secretion of the siderophore bacillibactin (dhb) [15]. Endophytic B. velezensis CCO9 is widely distributed in various parts of the plant body, including cortex, xylem vessel, stems, and leaves, and is known for its protective functions against wheat plant diseases. It was reported that the strain CCO9 stimulates plant resistance and shows 21.64% and 66.67% disease-control efficacy of spot blotch and take-all, respectively [16]. B. velezensis can express induced systemic resistance (ISR) in plants by activating the defense-associated genes of jasmonic acid (JA) and salicylic acid (SA) [17]. B. velezensis PEA1 demonstrated both the antifungal and antiviral activities against Fusarium oxysporum and cucumber mosaic virus (CMV) MN594112 (capable to infect ~1200 plant species around the world), respectively. PEA1 was able to reduce the accumulation of viral coat protein (i.e., CMV-CP) by 2.1 fold, compared to untreated Datura stramonium plant leaves, and it also induces ISR [18]. Most notably, strains of B. velezensis possess genes encoding the enzymes for the production of bioactive compounds related to biocontrol traits acting in the rhizosphere. These genes are activated by exposure to root exudates, following pathogen attacks through the regulation of specific genes, rather than the presence or absence of specific genes [19].
B. velezensis FZB42 is distinguished from the model B. subtilis 168 strain by the ability to suppress the competitive organisms present in the rhizosphere, and helps in plants growth promotion [9]. Despite the high genomic similarity between B. velezensis and B. subtilis, non-plant associated B. subtilis species contribute only 4–5% of genome ability to the synthesis of antimicrobial compounds; however, B. velezensis devotes 10% of its genome to the synthesis of antimicrobial molecules [38]. In recent years, based on phylogenomic analysis of Bacillus genomes, many B. subtilis strains (e.g., B. subtilis 83, B. subtilis BZR 517 etc.) were re-classified as plant-associated B. velezensis species [39,40]. Moreover, several B. subtilis-based commercial biocontrol agents like Serenade® (B. subtilis QST713), Kodiak™ (B. subtilis GB03), Taegro® (B. subtilis var. amyloliquefaciens FZB24) were re-categorized as B. velezensis-based biocontrol agents for agricultural applications (Table 2). These commercial biocontrol agents are widely used to control various pathogenic microorganisms in soil and to protect plants from various foliar bacterial and fungal diseases, during agricultural applications.
In this review, the biosynthesis of antimicrobial compounds from B. velezensis and their antimicrobial activities are described. The antimicrobial compounds can be utilized as biocontrol agents for several agricultural purposes, to eradicate pathogenic microbes. More specifically, we will discuss the past and recent developments in the biosynthesis of LP- and PK-type compounds from B. velezensis and their biological applications, by studying the modes of actions, based on previously published reports.

2. Antimicrobial LPs Synthesized by B. velezensis

LPs produced by B. velezensis are categorized into three distinct families based on the amino acid sequence: surfactins, fengycins, and bacillomycin-D that were originally isolated from B. subtilis [49]. Many microbial LPs are assembled by ribosome-independent pathways through a series of giant enzyme machines known as NRPSs that comprise ~1000 amino acids [49]. NRPSs catalyzes the stimulation of specific amino acids by conversion into corresponding aminoacyl thioesters and the subsequent formation of peptide bonds between activated amino acids [50]. NRPSs are a multi-functional enzyme complex with at least four critical domains essential to direct the non-ribosomal synthesis of peptides. The adenylation (A) domain is the first catalytic domain that activates specific amino acids; the thiolation (T) domain is needed for amino acid tethering; the condensation (C) domain assists in peptide bond formation; and finally, the thioesterase domain (TE) contributes in chain elongation and release of the cyclic peptide [51,52].

2.1. Surfactins

The history of surfactin dates back to 1968, when it was first purified and characterized by Arima et al., as a new bioactive compound in the culture broth of B. subtilis [53]. To date, several surfactin-producing strains are reported from different Bacillus spp., including B. velezensis, B. amyloliquefaciens, B. licheniformis, B. methylotrophicus, and B. thuringiensis [54]. These amphiphilic cyclic LPs comprise a hydrophilic heptapeptide ring structure consisting of the amino acid sequence (Glu-Leu-Leu-Val-Asp-Leu-Leu) attached to a β-hydroxy fatty acid moiety, usually between C-13 and C-16 [55]. There are three distinct forms of surfactins (e.g., surfactin A, B, and C) that are classified, based on variations in the amino acid sequence. The amino acids, namely L-leucine, L-valine, and L-iso-leucine are present in surfactin A, B, and C, respectively, at the position of the amino acid involved in formation of the lactone ring [56]. Surfactins are synthesized by a complex interaction of NRPSs encoded by srfA operon, consisting of four open reading frames (ORFs), namely srfAA, srfAB, srfAC, and srfAD [57]. Among them, srfAA, srfAB, and srfAC ORFs encode the modular enzymes responsible for integrating the seven amino acids into the peptide ring. However, the terminal ORF srfAD, a repair enzyme, encodes a thioesterase/acyltransferase domain that regulates the initiation of surfactin biosynthesis [58].
Isolates of Bacillus spp. produce small amounts of surfactin (<10% of its biomass) that serve as a signaling molecule during inter- or intra-species interactions [59]. Surfactin biosynthesis depends on cell density; however, quorum sensing (QS) [60] prevents the constant production of bacterial cells, thereby, limiting the overall yield of surfactin (Figure 2) [59].
In general, Bacillus cells secrete extracellular signaling factors like ComX pheromones (10-amino-acid modified peptides) continuously into the liquid media. A membrane-anchored histidine kinase receptor, ComP, detects the ComX at a vital concentration and subsequently autophosphorylates its cognate receptor regulator ComA. ComA is a part of the signaling cascade system of ComQXPA that is responsible for QS in several Bacillus spp. Successively, phosphorylated ComA (ComA~P) triggers the transcription of the srfA operon by binding to the promoter site, and initiates surfactin biosynthesis [57]. However, surfactin indirectly interacts with sensor kinase KinC, followed by the phosphorylation of the master response regulator Spo0A. Phosphorylated Spo0A, subsequently, induces the expression of SinI, which antagonizes the repressor SinR that causes the transcription of genes involved in matrix biosynthesis [61]. Thus, surfactin act as a paracrine signaling molecule that triggers other cells to produce the extracellular matrix and inhibit the biosynthesis of surfactins [62]. Paracrine signaling is observed in some bacterial populations, in which ComX indirectly induces the production of extracellular matrix, in a sub-population of cells, but these surfactant-responsive cells can no longer respond to ComX, thus, halting the production of additional surfactin [62].
In addition to the ComX-dependent regulation, several other factors including competence and sporulation-stimulating factor (CSF) and aspartate phosphatase (Rap) proteins, including Rap C, D, F, and H, also regulate the surfactin biosynthesis. CSF is a species-specific extracellular peptide secreted by Bacillus spp. and imported into the cell by oligopeptide permease (Opp; also known as Spo0K) [63]. Subsequently, CSF binds to the Rap proteins, which dephosphorylates ComA~P, thereby, impairing its function. However, the dephosphorylation of ComA~P can be inhibited to promote the transcription of srfA gene and surfactin biosynthesis [64]. These mechanisms would rationally explain why most Bacillus spp. in the liquid culture medium show minimal surfactin biosynthesis (Figure 2).
As a consequence of its amphiphilic structure, surfactin is a powerful and effective bio-surfactant molecule displaying antimicrobial activity against a wide variety of pathogenic microbes (Figure 3), including Ralstonia solanacearum [20], Pseudomonas syringae pv. tomato DC3000 [23], and F. verticillioides [22]. Additionally, surfactin was shown to harbor anti-mycoplasma activity against Mycoplasma hyorhinis [65], and anti-Legionella activity against Legionella pneumophila [66]. In a similar study, surfactins (surfactin B and C) produced by B. velezensis 9D-6, inhibited the growth of P. syringae DC3000 and Clavibacter michiganensis, during an in vitro plate assay. Furthermore, co-cultivation of B. velezensis 9D-6 and P. syringae DC3000, substantially reduced root colonization of DC3000 in A. thaliana seedlings, signifying that 9D-6 employs additional non-antimicrobial mechanisms against phytopathogens [21]. Upon root colonization, the strain B. subtilis 6051 protects Arabidopsis plants from pathogenic bacteria P. syringae DC3000 infection, and reduces plant mortality by 70%, through the combined actions of biofilm formation and surfactin secretion. The level of LPs secreted by B. subtilis 6051 was sufficient to kill the pathogen [23].

2.2. Fengycins

Fengycin or plipastatin, originally discovered from B. subtilis F-29-3 in 1986 is known to exhibit antifungal activity against a broad spectrum of filamentous fungi [67]. The structure of fengycins is composed of cyclic octapeptide containing decapeptides linked to N-terminal β-hydroxy fatty acid chain, usually between C-12 and C-19 [68]. Two isoforms of fengycin, fengycin A and fengycin B differ structurally, due to the presence of Ala/Val dimorphy at the sixth position [69]. Fengycins are synthesized by NRPSs encoded by an operon consisting of five ORFs fenA-E or ppsA-E [49].
Fengycins are assumed to cause cell death of the target organism by interacting with the cell membrane and altering the cell permeability. The findings of scanning electron microscopy (SEM) and transmission electron microscopy (TEM) suggested that treatment of hyphal cells of Magnaporthe grisea with fengycin (20 µg/mL) from B. subtilis BS155, led to the ultrastructural destruction of pathogen hyphae and the loss of cytoplasm, plasma membrane, or cell membrane integrity, which eventually resulted in cell lysis [8].
The antibiotic LP fengycin can be used to treat various plant diseases, e.g., barley head blight disease (F. graminearum) [70], rice blast disease (Magnaporthe grisea) [8], gray mold disease (Botrytis cinerea) [24], maize disease (Rhizomucor variabilis) [25], and cucurbit powdery disease (Podosphaera fusca) [71], etc. Fengycins produced by B. velezensis SQR9, exhibited antagonistic activities against F. oxysporum, F. solani, and Phytophthora parasitica and Verticillium dahliae Kleb [15] Plipastatin A synthesized by B. amyloliquefaciens S76-3 demonstrated superior fungicidal activity against F. graminearum, by inactivating the conidial spores at a minimum inhibitory concentration of 100 μg/mL. Microscopy experiments showed marked morphological changes in conidia and major distortions in the F. graminearum hyphae, with increased vacuolation [72]. However, in contrast to the antifungal activity of this LP, the antibacterial activity of fengycins produced by B. amyloliquefaciens MEP218 against the spot disease-causing Xanthomonas axonopodis pv. vesicatoria in tomato plants were characterized using liquid chromatography ESI-MS/MS [26].

2.3. Bacillomycin-D

Bacillomycin-D belongs to the LPs iturin family, including iturin A, C, D, and E, bacillomycin-F and L, bacillopeptin, and mycosubtilin [73]. This antimicrobial compound is a cyclic heptapeptide bound to the β-amino fatty acid chain between C-15 and C-18. The bmy operon that regulates the biosynthesis of bacillomycin-D comprises four genes (i.e., bmyD, bmyA, bmyB, and bmyC) without orthologues in B. subtilis 168 [74]. Most notably, the bmy gene cluster encoding the enzymes for the synthesis of bacillomycin-D, is separated from fengycin gene cluster, by only 25 kb, within the B. velezensis FZB42 genome, and is positioned exactly at the same location of the iturin-A gene cluster of B. subtilis RB14 [73]. Three pleiotropic regulators (i.e., DegU, DegQ, and ComA) and two sigma factors (i.e., σB and σH) positively regulate the transcriptional activation of the bmy promoter towards the synthesis of bacillomycin-D. Another study demonstrated the role of DegU and ComA in regulating the bacillomycin-D expression. Inactivation of the genes encoding DegU and ComA proteins resulted in an impaired promoter function of the bmy operon. As a consequence, the transcription rate of the bmy operon was three to four fold lower in the mutant derivatives than in parental B. velezensis FZB42 strain [74]. Furthermore, LP bacillomycin-D synthesized by B. velezensis SQR9 acts as a signaling molecule in biofilm formation, due to an increase in the intracellular iron concentration and activation of the KinB-Spo0A-SinI-SinR signal cascade-based synthesis of biofilm matrix components [75].
Bacillomycin-D synthesized by B. velezensis were shown to display antimicrobial activity against different microorganisms, such as X. campestris pv. cucurbitae [29], Aspergillus flavus [76], F. graminearum (Fusarium head blight) [12], F. oxysporum f. sp. cucumerinum (vascular wilt in cucumber plants), etc. [30]. Mutant strains deficient in the production of bacillomycin-D compromised antifungal action, suggesting the role of bacillomycin-D in the antifungal activity of FZB42. SEM and TEM analyses confirmed that bacillomycin-D causes morphological alterations in the cytoplasmic membranes and cell walls of F. graminearum hyphae and conidia. This resulted in the accumulation of reactive oxygen species (ROS), and ultimately triggered the cell lysis of F. graminearum. The 50% effective concentration (EC50) that purified bacillomycin-D and inhibited the activity of F. graminearum was estimated to be about 30 μg/mL [12].

3. Antibacterial PKs Synthesized by B. velezensis

PKs are a natural class of secondary metabolites synthesized by PKSs. To date, more than 10 thousand PK-type compounds are identified from bacteria, fungi, plants, and animals, of which at least 20 were developed as commercial drugs including erythromycin, tetracycline, and lovastatin [77]. The genes encoding PKSs were identified in 1993 during genome sequencing of B. subtilis 168. PKSs catalyzes the decarboxylative Claisen condensation reactions with possible additional alterations through β-reduction, dehydration, or enoyl-reduction reactions that are catalyzed by some PKSs-modifying domains. The multi-enzyme system of PKSs uses acyl carrier proteins that are post-translationally modified with the 4’-phosphopantetheine prosthetic group, to guide the intermediate PK molecule throughout the elongation process [9]. Interestingly, the model strain B. subtilis 168 was shown to contain a large PKS gene cluster designated as pksX; however, this strain was not capable of synthesizing PKs due to mutation in the sfp gene encoding 4’-phosphopantetheine transferase (Sfp) [9].

3.1. Bacillaene

Bacillaene, a novel polyene antibiotic, was discovered from the fermentation broth of B. subtilis that inhibit the prokaryotic protein synthesis, by an unknown mechanism [78]. Among the three giant modular PKSs system in B. velezensis (pks1, pks2, pks3), bacillaene is synthesized by the enzymes encoded by the pks1 (bae) gene cluster, which is an ortholog of the pksX gene cluster of B. subtilis 168 [9]. Despite antibacterial activity of this antibiotic against multi-drug-resistant bacterial isolates, for many years, characterization of bacillaene using the traditional methods based on fractionations was proved challenging, owing to its chemical instability [11]. On exposure to light or room temperature, bacillaene decomposes rapidly, which hindered earlier attempts to identify the biosynthetic pathway of this antibiotic molecule [79]. Antibacterial polyketide bacillaene synthesized by B. velezensis FZB42, exhibited a minor extent of bacteriostatic effect against Erwinia amylovora, a causal agent of fire blight disease [32]. In addition, bacillaene-A synthesized by Bacillus spp. displayed antifungal activity against Termitomyces fungi [33].

3.2. Macrolactin

Macrolactin was originally isolated from the ethyl acetate extract of an unclassified deep-sea bacterium Bacillus spp. Sc026 [80]. In B. velezensis, the pks2 (mlnBCDEFGH) gene cluster encode the enzymes for antibacterial compound macrolactin, which is an inhibitor of the bacterial peptide deformylase [7,81]. The chemical structure of macrolactin is synthesized by the expansion of the acetyl starter unit, by 11 successive Claisen condensation reactions with malonyl-CoA. Currently, approximately 17 different types of macrolactins are identified; however, only four macrolactin forms (e.g., macrolactin-A, macrolactin-D, 7-O-malonyl-macrolactin-A, and 7-O-succinyl-macrolactin) are found in B. velezensis. Of the four, 7-O-malonyl-macrolactin-A was found to have bacteriostatic effects on a variety of gram-positive and multidrug-resistant bacterial pathogens, particularly, methicillin-resistant Staphylococcus aureus, vancomycin-resistant enterococci, and small-colony variant of Burkholderia cepacia [34].

3.3. Difficidin

Difficidin was detected for the first time in the fermentation broth of B. subtilis ATCC-39320 and categorized as an unsaturated macrocyclic polyene lactone phosphate ester in its 22-member family [82]. Difficidin, as well as its oxidized form oxydifficidin, encoded by the enzymes of pks3 (dif) gene cluster, appeared primarily as their alkali ion adducts in the matrix-assisted laser desorption ionization-time of flight mass spectra [9]. Oxydifficidin has a hydroxyl group at the fifth position of the difficidin ring structure [9].
The antibiotic compounds, difficidin and bacilysin, exhibited antibacterial activity against two rice pathogens, X. oryzae pv. oryzae, as well as X. oryzae pv. oryzicola, causing bacterial blight and bacterial leaf streak disease, respectively. In combination, these two compounds affected the cell wall of Xanthomonas, as indicated by SEM and TEM observations. Furthermore, the quantitative real-time PCR results also indicated the downregulation of several X. oryzae genes including rpfF, gumD, glmS, ftsZ, and rrlA, related to the virulence, cell division, and biosynthesis of proteins and cell wall of X. oryzae [35]. In a similar study, a butanolic extract of the B. velezensis DR-08 broth culture containing difficidin and oxydifficidin displayed antibacterial activity against R. solanacearum, a leading causal agent of tomato bacterial wilt with a minimum inhibitory concentration (MIC) value of 12.62 μg/mL. Furthermore, the metabolic extract of this bacterium also inhibited the growth of 14 phytopathogenic bacteria with MIC values ranging from 1.95–500 μg/mL [36].

4. Bacillibactin

Iron is an essential element for all living organisms and serves as a vital cofactor to perform cellular processes including DNA synthesis, respiration, and defense against ROS [83]. Several Bacillus spp. secretes bacillibactin, the catecholic iron siderophore, which is very important in facilitating Fe(III) acquisition, especially when the Bacillus cells experience iron limitation [84]. In B. velezensis, the products of the functional dhb gene cluster was shown to assist in the synthesis of bacillibactin (small molecule iron-chelators). It is a part of a complex transport system that enables the B. velezensis cells to accumulate iron ions and acquire them from their natural environment, under iron-limiting conditions [10]. LPs (i.e., bacillomycin D, fengycins, and surfactins) coupled with bacillibactin synthesized by B. velezensis SQR9 had an antagonistic effect against certain fungal pathogens, including F. oxysporum, F. solani, P. parasitica, where the production of bacillibactin was greatly upregulated. However, mutant strains deficient in LPs and bacillibactin displayed a substantial reduction in antifungal effects, when challenged with these fungal pathogens. These results suggest that bacillibactin plays a passive role in the suppression of microbial pathogens, either by depriving them of essential iron or directly inhibiting the growth [15]. However, there is no experimental evidence of antimicrobial activity of purified bacillibactin, in the absence of known secondary metabolites like LPs or PKs.

5. Bacilysin

Bacilysin is a Trojan horse antibiotic, synthesized by the enzymes of the bacA-E gene cluster (formerly ywfBCDEF) of certain Bacillus spp. [85]. This dipeptide antibiotic [L-alanyl-(2,3-epoxycyclohexanone-4)-L-alanine] was first isolated from the soil bacterium B. subtilis by Foster and Woodruff in 1946 [35], and its structure was established by Walker and Abraham in 1970 [86].
Bacilysin relies on peptide transporters for uptake into the target cells. Once internalized into the susceptible cells, bacilysin is hydrolyzed by cytoplasmic peptidases to non-proteinogenic anticapsin (epoxy-cyclohexanonyl-Ala) and N-terminal L-alanine (Figure 4). The C-terminal epoxy amino acid (anticapsin) of bacilysin is responsible for the antimicrobial activity against pathogenic microorganisms [86]. Anticapsin covalently interacts with the active site of the cell wall biosynthetic enzyme glucosamine synthase, the latter catalyzes the synthesis of glucosamine-6-phosphate from fructose-6-phosphate and glutamine [87]. This covalent binding was caused by the crosslinking between the active thiol of cysteine residue present in the enzyme glucosamine synthase, and the apoxide functional group of anticapsin. Therefore, the bacterial peptidoglycan or fungal mannoprotein biosynthetic pathway was thus blocked, leading to cell protoplasting and lysis [88]. In real-time PCR analysis, it was confirmed that several genes, including glmS, psbA1, mcyB, and ftsZ, which are related to the biosynthesis of peptidoglycan, cell division, and photosynthesis in Microcystis aeruginosa cells, were downregulated in response to bacilysin treatment (4 mg/L) [88].
The antimicrobial action of bacilysin depends on the composition of culture medium and the activity could be reversed by using some antagonists like N-acetylglucosamine, several dipeptides, and amino acids, which might inhibit the transport of this antibiotic into the microbial cells [89]. Bacilysin, synthesized by B. velezensis FZB42, exerted antagonistic activities against S. aureus and C. michiganense subsp. Sepedonicum, which cause ring rot disease in potatoes [37]. In a similar study, bacilysin synthesized by B. velezensis, exhibited strong anti-cyanobacterial activity against M. aeruginosa, which cause harmful algal blooms with a kill rate of 98.78%. However, disruption of a single gene bacB or supplementation of N-acetylglucosamine to the bioassay plates, abolished the inhibitory effect of bacilysin [88]. Analyses using SEM and TEM revealed that exposing X. oryzae pv. oryzae and X. oryzae pv. oryzicola to 50 μg/mL of bacilysin for 12 h, triggered changes in the cell wall structure as well as efflux of intracellular components [35]. In a similar study, TEM revealed the micro- and ultra-structural changes to M. aeruginosa cells treated with 15 mg/L bacilysin for 2 h. The cells were severely damaged and the cytoplasm was condensed, eventually, resulting in plasmolysis of M. aeruginosa cells [88].

6. Conclusions

Over the past few decades, hundreds of antimicrobial drugs were developed from a plethora of microorganisms. These antimicrobial molecules are considered safe for the treatment of various plant diseases, due to their broad-spectrum activity against multiple microbes, reduced toxicity compared to chemical pesticides, environment-friendly nature, and reduced risk of resistance acquisition in pathogenic microbes. Recently characterized bioactive compounds synthesized by B. velezensis demonstrated promising antimicrobial activities suitable for agricultural applications; therefore, the mode of actions of these antimicrobial compounds against various plant pathogens were extensively investigated. Although, some B. subtilis strains are also capable to produce bioactive secondary metabolites, it was reported that the biosynthetic arsenals of B. velezensis is more powerful and diverse than that of B. subtilis. In addition, in recent years, several biocontrol agents that were formulated from B. subtilis strains, were reclassified as B. velezensis-dependent biocontrol agents, based on the availability of genome sequence data. Taken together, B. velezensis could be a versatile and powerful biocontrol agent that can be used as an effective alternative to synthetic agrochemicals, either by using the bacteria itself or by extracting its active compounds. Moreover, the elucidation of the genes responsible for the synthesis of bioactive compounds and strategies to alter these genes using genome-engineering techniques would constitute additional important measures to increase the biosynthesis of metabolites in B. velezensis.

Author Contributions

M.F.R. and K.-H.B. collected the data and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NRF-2019R1F1A1052625.

Acknowledgments

Authors appreciate the research fund provided by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2019R1F1A1052625).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rabbee, M.F.; Ali, M.S.; Choi, J.; Hwang, B.S.; Jeong, S.C.; Baek, K. Bacillus velezensis: A valuable member of bioactive molecules within plant microbiomes. Molecules 2019, 24, 1046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Cao, Y.; Pi, H.; Chandrangsu, P.; Li, Y.; Wang, Y.; Zhou, H.; Xiong, H.; Helmann, J.D.; Cai, Y. Antagonism of two plant-growth promoting Bacillus velezensis isolates against Ralstonia solanacearum and Fusarium oxysporum. Sci. Rep. 2018, 8, 4360. [Google Scholar] [CrossRef] [PubMed]
  3. 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]
  4. 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]
  5. 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 phylogenom. Int. J. Syst. Evol. Microbiol. 2016, 66, 1212–1217. [Google Scholar] [CrossRef]
  6. 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]
  7. Chen, X.H.; Koumoutsi, A.; Scholz, R.; Eisenreich, A.; Schneider, K.; Heinemeyer, I.; Morgenstern, B.; Voss, B.; Hess, W.R.; Reva, O.; et al. Comparative analysis of the complete genome sequence of the plant growth–promoting bacterium Bacillus amyloliquefaciens FZB42. Nat. Biotechnol. 2007, 25, 1007–1014. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. 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]
  10. Chen, X.H.; Koumoutsi, A.; Scholz, R.; Schneider, K.; Vater, J.; Süssmuth, R.; Piel, J.; Borriss, R. Genome analysis of Bacillus amyloliquefaciens FZB42 reveals its potential for biocontrol of plant pathogens. J. Biotechnol. 2009, 140, 27–37. [Google Scholar] [CrossRef]
  11. 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]
  12. 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. Envrion. Microbiol. 2017, 83, e01075-17. [Google Scholar] [CrossRef] [Green Version]
  13. Scholz, R.; Vater, J.; Budiharjo, A.; Wang, Z.; He, Y.; Dietel, K.; Schwecke, T.; Herfort, S.; Lasch, P.; Borriss, R. Amylocyclicin, a novel circular bacteriocin produced by Bacillus amyloliquefaciens FZB42. J. Bacteriol. 2014, 196, 1842–1852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. 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]
  15. Li, B.; Li, Q.; Xu, Z.; Zhang, N.; Shen, Q.; Zhang, R. Responses of beneficial Bacillus amyloliquefaciens SQR9 to different soilborne fungal pathogens through the alteration of antifungal compounds production. Front. Microbiol. 2014, 5, 636. [Google Scholar] [CrossRef] [Green Version]
  16. 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] [PubMed]
  17. Li, C.; Hu, W.C.; Pan, B.; Liu, Y.; Yuan, S.F.; Ding, Y.Y.; Li, R.; Zheng, X.Y.; Shen, Q. Rhizobacterium Bacillus amyloliquefaciens strain SQRT3-mediated induced systemic resistance controls bacterial wilt of tomato. Pedosphere 2017, 27, 1135–1146. [Google Scholar] [CrossRef]
  18. Abdelkhalek, A.; Behiry, S.I.; Al-Askar, A.A. Bacillus velezensis PEA1 inhibits Fusarium oxysporum growth and induces systemic resistance to cucumber mosaic virus. Agronomy 2020, 10, 1312. [Google Scholar] [CrossRef]
  19. Reva, O.N.; Swanevelder, D.Z.H.; Mwita, L.A.; Mwakilili, A.D.; Muzondiwa, D.; Joubert, M.; Chan, W.Y.; Lutz, S.; Ahrens, C.H.; Avdeeva, L.V.; et al. Genetic, epigenetic and phenotypic diversity of four Bacillus velezensis strains used for plant protection or as probiotics. Front. Microbiol. 2019, 10, 2610. [Google Scholar] [CrossRef] [Green Version]
  20. Xiong, H.; Li, Y.; Cai, Y.; Cao, Y.; Wang, Y. Isolation of Bacillus amyloliquefaciens JK6 and identification of its lipopeptides surfactin for suppressing tomato bacterial wilt. RSC Adv. 2015, 5, 82042–82049. [Google Scholar] [CrossRef]
  21. Grady, E.N.; MacDonald, J.; Ho, M.T.; Weselowski, B.; McDowell, T.; Solomon, O.; Renaud, J.; Yuan, Z.C. Characterization and complete genome analysis of the surfactin-producing, plant-protecting bacterium Bacillus velezensis 9D-6. BMC Microbiol. 2019, 19, 5. [Google Scholar] [CrossRef]
  22. Snook, M.E.; Mitchell, T.; Hinton, D.M.; Bacon, C.W. Isolation and characterization of Leu7-surfactin from the endophytic bacterium Bacillus mojavensis RRC 101, a biocontrol agent for Fusarium verticillioides. J. Agric. Food Chem. 2009, 57, 4287–4292. [Google Scholar] [CrossRef]
  23. Bais, H.P.; Fall, R.; Vivanco, J.M. Biocontrol of Bacillus subtilis against infection of Arabidopsis roots by Pseudomonas syringae is facilitated by biofilm formation and surfactin production. Plant Physiol. 2004, 134, 307–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Toral, L.; Rodríguez, M.; Béjar, V.; Sampedro, I. Antifungal activity of lipopeptides from Bacillus XT1 CECT 8661 against Botrytis cinerea. Front. Microbiol. 2018, 9, 1315. [Google Scholar] [CrossRef] [PubMed]
  25. Kulimushi, P.Z.; Arias, A.A.; Franzil, L.; Steels, S.; Ongena, M. Stimulation of fengycin-type antifungal lipopeptides in Bacillus amyloliquefaciens in the presence of the maize fungal pathogen Rhizomucor variabilis. Front. Microbiol. 2017, 8, 850. [Google Scholar] [CrossRef] [PubMed]
  26. Medeot, D.B.; Fernandez, M.; Morales, G.M.; Jofre, E. Fengycins from Bacillus amyloliquefaciens MEP2 18 exhibit antibacterial activity by producing alterations on the cell surface of the pathogens Xanthomonas axonopodis pv. vesicatoria and Pseudomonas aeruginosa PA01. Front. Microbiol. 2020, 10, 3107. [Google Scholar] [CrossRef]
  27. Pajčin, I.; Vlajkov, V.; Frohme, M.; Grebinyk, S.; Grahovac, M.; Mojićević, M.; Grahovac, J. Pepper bacterial spot control by Bacillus velezensis: Bioprocess solution. Microorganisms 2020, 8, 1463. [Google Scholar] [CrossRef]
  28. Adeniji, A.A.; Aremu, O.S.; Babalola, O.O. Selecting lipopeptide-producing, Fusarium-suppressing Bacillus spp.: Metabolomic and genomic probing of Bacillus velezensis NWUMFkBS10.5. MicrobiologyOpen 2019, 8, e00742. [Google Scholar] [CrossRef] [Green Version]
  29. Zeriouh, H.; Romero, D.; García-Gutiérrez, L.; Cazorla, F.M.; De Vicente, A.; Pérez-García, A. The iturin-like lipopeptides are essential components in the biological control arsenal of Bacillus subtilis against bacterial diseases of cucurbits. Mol. Plant Microbe Interact. 2011, 24, 1540–1552. [Google Scholar] [CrossRef] [Green Version]
  30. Xu, Z.; Shao, J.; Li, B.; Yan, X.; Shen, Q.; Zhang, R. Contribution of bacillomycin D in Bacillus amyloliquefaciens SQR9 to antifungal activity and biofilm formation. Appl. Environ. Microbiol. 2013, 79, 808–815. [Google Scholar] [CrossRef] [Green Version]
  31. Luna-Bulbarela, A.; Tinoco-Valencia, R.; Corzo, G.; Kazuma, K.; Konno, K.; Galindo, E.; Serrano-Carreón, L. Effects of bacillomycin D homologues produced by Bacillus amyloliquefaciens 83 on growth and viability of Colletotrichum gloeosporioides at different physiological stages. Biol. Control 2018, 127, 145–154. [Google Scholar] [CrossRef]
  32. Chen, X.H.; Scholz, R.; Borriss, M.; Junge, H.; Mögel, G.; Kunz, S.; Borriss, R. Difficidin and bacilysin produced by plant-associated Bacillus amyloliquefaciens are efficient in controlling fire blight disease. J. Biotechnol. 2009, 140, 38–44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Um, S.; Fraimout, A.; Sapountzis, P.; Oh, D.C.; Poulsen, M. The fungus-growing termite Macrotermes natalensis harbors bacillaene-producing Bacillus sp. that inhibit potentially antagonistic fungi. Sci. Rep. 2013, 3, 3250. [Google Scholar] [CrossRef]
  34. Romero-tabarez, M.; Jansen, R.; Sylla, M.; Lu, H.; Ha, S.; Santosa, D.A.; Timmis, K.N.; Molinari, G. 7-O-malonyl macrolactin A, a new macrolactin antibiotic from Bacillus subtilis active against methicillin-resistant Staphylococcus aureus, vancomycin-resistant Enterococci, and a small-colony variant of Burkholderia cepacia. Antimicrob. Agents Chemother. 2006, 50, 1701–1709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. 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]
  36. 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. 2019, 163, 130–137. [Google Scholar] [CrossRef]
  37. Wu, L.; Wu, H.; Chen, L.; Lin, L.; Borriss, R. Bacilysin overproduction in Bacillus amyloliquefaciens FZB42 markerless derivative strains FZBREP and FZBSPA enhances antibacterial activity. Appl. Microbiol. Biotechnol. 2015, 99, 4255–4263. [Google Scholar] [CrossRef]
  38. Chowdhury, S.P.; Hartmann, A.; Gao, X.W.; Borriss, R. Biocontrol mechanism by root-associated Bacillus amyloliquefaciens FZB42-A review. Front. Microbiol. 2015, 6, 780. [Google Scholar] [CrossRef] [Green Version]
  39. Balderas-Ruíz, K.A.; Bustos, P.; Santamaria, R.I.; González, V.; Cristiano-Fajardo, S.A.; Barrera-Ortíz, S.; Mezo-Villalobos, M.; Aranda-Ocampo, S.; Guevara-García, Á.A.; Galindo, E.; et al. Bacillus velezensis 83 a bacterial strain from mango phyllosphere, useful for biological control and plant growth promotion. AMB Express 2020, 10, 1–9. [Google Scholar] [CrossRef]
  40. Radchenko, V.V.; Vasilyev, I.Y.; Ilnitskaya, E.V.; Garkovenko, A.V.; Asaturova, A.M.; Tomashevich, N.S.; Kozitsyn, A.E.; Milovanov, A.V.; Grigoreva, T.V.; Shternshis, M.V. Draft genome sequence of the plant growth-promoting bacterium Bacillus subtilis strain BZR 517, isolated from winter wheat, now reclassified as Bacillus velezensis strain BZR 517. Microbiol. Resour. Announc. 2020, 9, e00853-20. [Google Scholar] [CrossRef]
  41. Chowdhury, S.P.; Dietel, K.; Rändler, M.; Schmid, M.; Junge, H.; Borriss, R.; Hartmann, A.; Grosch, R. Effects of Bacillus amyloliquefaciens FZB42 on lettuce growth and health under pathogen pressure and its impact on the rhizosphere bacterial community. PLoS ONE 2013, 8, e68818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Romanazzi, G.; Feliziani, E. Botrytis cinerea (Gray Mold). In Postharvest Decay: Control Strategies; Bautista-Banos, Elsevier Inc.: London, UK, 2014; pp. 131–146. [Google Scholar]
  43. Pandin, C.; Le Coq, D.; Deschamps, J.; Védie, R.; Rousseau, T.; Aymerich, S.; Briandet, R. Complete genome sequence of Bacillus velezensis QST713: A biocontrol agent that protects Agaricus bisporus crops against the green mould disease. J. Biotechnol. 2018, 278, 10–19. [Google Scholar] [CrossRef]
  44. Matzen, N.; Heick, T.M.; Jørgensen, L.N. Control of powdery mildew (Blumeria graminis spp.) in cereals by Serenade®ASO (Bacillus amyloliquefaciens (former subtilis) strain QST 713). Biol. Control 2019, 139, 104067. [Google Scholar] [CrossRef]
  45. Brannen, P.M.; Kenney, D.S. Kodiak®-A successful biological-control product for suppression of soil-borne plant pathogens of cotton. J. Ind. Microbiol. Biotechnol. 1997, 19, 169–171. [Google Scholar] [CrossRef]
  46. Borriss, R.; Chen, X.H.; Rueckert, C.; Blom, J.; Becker, A.; Baumgarth, B.; Fan, B.; Pukall, R.; Schumann, P.; Spröer, C.; et al. Relationship of Bacillus amyloliquefaciens clades associated with strains DSM7T and FZB42T: A proposal for Bacillus amyloliquefaciens subsp. amyloliquefaciens subsp. nov. and Bacillus amyloliquefaciens subsp. plantarum subsp. nov. based on complete genome sequence comparisons. Int. J. Syst. Evol. Microbiol. 2011, 61, 1786–1801. [Google Scholar] [PubMed]
  47. Elanchezhiyan, K.; Keerthana, U.; Nagendran, K.; Prabhukarthikeyan, S.R.; Prabakar, K.; Raguchander, T.; Karthikeyan, G. Multifaceted benefits of Bacillus amyloliquefaciens strain FBZ24 in the management of wilt disease in tomato caused by Fusarium oxysporum f. sp. lycopersici. Physiol. Mol. Plant Pathol. 2018, 103, 92–101. [Google Scholar] [CrossRef]
  48. Keerthana, U.; Nagendran, K.; Raguchander, T.; Prabakar, K.; Rajendran, L.; Karthikeyan, G. Deciphering the role of Bacillus subtilis var. amyloliquefaciens in the management of late blight pathogen of potato, Phytophthora infestans. Proc. Natl. Acad. Sci. India Sect. B Biol. Sci. 2018, 88, 1071–1080. [Google Scholar] [CrossRef]
  49. Ongena, M.; Jacques, P. Bacillus lipopeptides: Versatile weapons for plant disease biocontrol. Trends Microbiol. 2008, 16, 115–125. [Google Scholar] [CrossRef]
  50. Corre, C.; Challis, G.L. Exploiting genomics for new natural product discovery in prokaryotes. Chem. Biol. 2010, 2, 429–453. [Google Scholar]
  51. Degen, A.; Mayerthaler, F.; Mootz, H.D.; Di Ventura, B. Context-dependent activity of A domains in the tyrocidine synthetase. Sci. Rep. 2019, 9, 5119. [Google Scholar] [CrossRef] [Green Version]
  52. Bozhüyük, K.A.J.; Linck, A.; Tietze, A.; Kranz, J.; Wesche, F.; Nowak, S.; Fleischhacker, F.; Shi, Y.N.; Grün, P.; Bode, H.B. Modification and de novo design of non-ribosomal peptide synthetases using specific assembly points within condensation domains. Nat. Chem. 2019, 11, 653–661. [Google Scholar] [CrossRef]
  53. 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]
  54. Beltran-Gracia, E.; Macedo-Raygoza, G.; Villafaña-Rojas, J.; Martinez-Rodriguez, A.; Chavez-Castrillon, Y.; Espinosa-Escalante, F.; Di Mascio, P.; Ogura, T.; Beltran-Garcia, M. Production of lipopeptides by fermentation processes: Endophytic bacteria, fermentation strategies and easy methods for bacterial selection. In Fermentation Processes, 1st ed.; Menestrina, G., Serra, M.D., Jozala, A.F., Eds.; Intech Open Science: London, UK, 2017; pp. 260–271. [Google Scholar]
  55. Zhou, D.; Hu, F.; Lin, J.; Wang, W.; Li, S. Genome and transcriptome analysis of Bacillus velezensis BS-37, an efficient surfactin producer from glycerol, in response to D-/L-leucine. MicrobiologyOpen 2019, 8, e79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Singh, P.; Cameotra, S.S. Potential applications of microbial surfactants in biomedical sciences. Trends Biotechnol. 2004, 22, 142–146. [Google Scholar] [CrossRef]
  57. Nakano, M.M.; Magnuson, R.; Myers, A.; Curry, J.; Grossman, A.D.; Zuber, P. srfA is an operon required for surfactin production, competence development, and efficient sporulation in Bacillus subtilis. J. Bacteriol. 1991, 173, 1770–1778. [Google Scholar] [CrossRef] [Green Version]
  58. Steller, S.; Sokoll, A.; Wilde, C.; Bernhard, F.; Franke, P.; Vater, J. Initiation of surfactin biosynthesis and the role of the SrfD-thioesterase protein. Biochemistry 2004, 43, 11331–11343. [Google Scholar] [CrossRef] [PubMed]
  59. Zhi, Y.; Wu, Q.; Xu, Y. Genome and transcriptome analysis of surfactin biosynthesis in Bacillus amyloliquefaciens MT45. Sci. Rep. 2017, 7, 40976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Rutherford, S.T.; Bassler, B.L. Bacterial quorum sensing: Its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2012, 2, a012427. [Google Scholar] [CrossRef] [PubMed]
  61. Bendori, S.O.; Pollak, S.; Hizi, D.; Eldar, A. The RapP-PhrP quorum-sensing system of Bacillus subtilis strain NCIB3610 affects biofilm formation through multiple targets, due to an atypical signal-insensitive allele of RapP. J. Bacteriol. 2015, 197, 592–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. López, D.; Vlamakis, H.; Losick, R.; Kolter, R. Paracrine signaling in a bacterium. Genes Dev. 2009, 23, 1631–1638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Comella, N.; Grossman, A.D. Conservation of genes and processes controlled by the quorum response in bacteria: Characterization of genes controlled by the quorum-sensing transcription factor ComA in Bacillus subtilis. Mol. Microbiol. 2005, 57, 1159–1174. [Google Scholar] [CrossRef]
  64. Hu, F.; Liu, Y.; Li, S. Rational strain improvement for surfactin production: Enhancing the yield and generating novel structures. Microb. Cell Fact. 2019, 18, 42. [Google Scholar] [CrossRef] [Green Version]
  65. Vollenbroich, D.; Pauli, G.; Özel, M.; Vater, J. Antimycoplasma properties and application in cell culture of surfactin, a lipopeptide antibiotic from Bacillus subtilis. Appl. Environ. Microbiol. 1997, 63, 44–49. [Google Scholar] [CrossRef] [Green Version]
  66. Loiseau, C.; Schlusselhuber, M.; Bigot, R. Surfactin from Bacillus subtilis displays an unexpected anti-Legionella activity. Appl. Microbiol. Biotechnol. 2015, 99, 5083–5093. [Google Scholar] [CrossRef]
  67. Vanittanakomt, N.; Loeffler, W. Fengycin—A novel antifungal lipopeptide antibiotic produced by Bacillus subtilis F-29-3. J. Antibiot. 1986, 7, 888–901. [Google Scholar] [CrossRef] [Green Version]
  68. Xu, B.H.; Lu, Y.Q.; Ye, Z.W.; Zheng, Q.W.; Wei, T.; Lin, J.F.; Guo, L.Q. Genomics-guided discovery and structure identification of cyclic lipopeptides from the Bacillus siamensis JFL15. PLoS ONE 2018, 13, e0202893. [Google Scholar] [CrossRef]
  69. Yang, H.; Li, X.; Li, X.; Yu, H.; Shen, Z. Identification of lipopeptide isoforms by MALDI-TOF-MS / MS based on the simultaneous purification of iturin, fengycin, and surfactin by RP-HPLC. Anal. Bioanal. Chem. 2015, 407, 2529–2542. [Google Scholar] [CrossRef]
  70. Kim, K.; Lee, Y.; Ha, A.; Kim, J.I.; Park, A.R.; Yu, N.H.; Son, H.; Choi, G.J.; Park, H.W.; Lee, C.W.; et al. Chemosensitization of Fusarium graminearum to chemical fungicides using cyclic lipopeptides produced by Bacillus amyloliquefaciens strain JCK-12. Front. Plant Sci. 2017, 8, 2010. [Google Scholar] [CrossRef] [Green Version]
  71. Romero, D.; De Vicente, A.; Rakotoaly, R.H.; Dufour, S.E.; Veening, J.; 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]
  72. Gong, A.; Li, H.; Yuan, Q.; Song, X.; Yao, W. 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] [PubMed] [Green Version]
  73. Chen, X.H.; Koumoutsi, A.; Scholz, R.; Borriss, R. More than anticipated-production of antibiotics and other secondary metabolites by Bacillus amyloliquefaciens FZB42. J. Mol. Microbiol. Biotechnol. 2009, 16, 14–24. [Google Scholar] [CrossRef] [PubMed]
  74. Koumoutsi, A.; Chen, X.H.; Vater, J.; Borriss, R. DegU and YczE positively regulate the synthesis of bacillomycin D by Bacillus amyloliquefaciens strain FZB42. Appl. Environ. Microbiol. 2007, 73, 6953–6964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Xu, Z.; Mandic-mulec, I.; Zhang, H.; Zhang, N.; Xu, Z.; Mandic-mulec, I.; Zhang, H.; Liu, Y.; Sun, X.; Feng, H.; et al. 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]
  76. Moyne, A.L.; Shelby, R.; Cleveland, T.E.; Tuzun, S. Bacillomycin D: An iturin with antifungal activity against Aspergillus flavus. J. Appl. Microbiol. 2001, 90, 622–629. [Google Scholar] [CrossRef] [PubMed]
  77. Tan, Z.; Clomburg, J.M.; Cheong, S.; Qian, S.; Gonzalez, R. A polyketoacyl-CoA thiolase-dependent pathway for the synthesis of polyketide backbones. Nat. Catal. 2020, 3, 593–603. [Google Scholar] [CrossRef]
  78. Patel, P.S.; Huangn, S.; Fisher, S.; Pirnik, D.; Aklonis, C.; Dean, L.; Meyers, E.; Fernandes, P.; Mayerlm, 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] [Green Version]
  79. Li, H.; Han, X.; Zhang, J.; Dong, Y.; Xu, S.; Bao, Y.; Chen, C.; Feng, Y.; Cui, Q.; Li, W. An effective strategy for identification of highly unstable bacillaenes. J. Nat. Prod. 2019, 82, 3340–3346. [Google Scholar] [CrossRef]
  80. Jaruchoktaweechai, C.; Suwanborirux, K.; Tanasupawatt, S.; Kittakoop, P.; Menasveta, P. New macrolactins from a marine Bacillus sp. Sc026. J. Nat. Prod. 2000, 63, 984–986. [Google Scholar] [CrossRef]
  81. Schneider, K.; Chen, X.H.; Vater, J.; Franke, P.; Nicholson, G.; Borriss, R.; Sussmuth, R.D. Macrolactin is the polyketide biosynthesis product of the pks2 cluster of Bacillus amyloliquefaciens FZB42. J. Nat. Prod. 2007, 70, 1417–1423. [Google Scholar] [CrossRef]
  82. Wilson, K.E.; Flor, J.E.; Schwartz, R.E.; Joshua, H.; Smith, J.L.; Pelak, B.A.; Liesch, J.M.; Hensens, O.D. Difficidin and oxydifficidin: Novel broad spectrum antibacterial antibiotics produced by Bacillus subtilis. I. Production, taxonomy and antibacterial activity. J. Antibiot. 1987, 40, 1682–1691. [Google Scholar] [CrossRef]
  83. Dunn, L.L.; Rahmanto, Y.S.; Richardson, D.R. Iron uptake and metabolism in the new millennium. Trends Cell Biol. 2007, 17, 93–100. [Google Scholar] [CrossRef]
  84. Fukushima, T.; Allred, B.E.; Sia, A.K.; Nichiporuk, R.; Andersen, U.N.; Raymond, K.N. Gram-positive siderophore-shuttle with iron-exchange from Fe-siderophore to apo-siderophore by Bacillus cereus YxeB. Proc. Natl. Acad. Sci. USA 2013, 110, 13821–13826. [Google Scholar] [CrossRef] [Green Version]
  85. Parker, J.B.; Walsh, C.T. Action and timing of BacC and BacD in the late stages of biosynthesis of the dipeptide antibiotic bacilysin. Biochemistry 2013, 52, 889–901. [Google Scholar] [CrossRef] [Green Version]
  86. Walker, J.E.; Abraham, E.P. The structure of bacilysin and other products of Bacillus subtilis. Biochem. J. 1970, 118, 563–570. [Google Scholar] [CrossRef] [Green Version]
  87. Khan, M.A.; Göpel, Y.; Milewski, S.; Görke, B. Two small RNAs conserved in enterobacteriaceae provide intrinsic resistance to antibiotics targeting the cell wall biosynthesis enzyme glucosamine-6-phosphate synthase. Front. Microbiol. 2016, 7, 908. [Google Scholar] [CrossRef]
  88. Wu, L.; Wu, H.; Chen, L.; Xie, S.; Zang, H.; Borriss, R.; Gao, X. Bacilysin from Bacillus amyloliquefaciens FZB42 has specific bactericidal activity against harmful algal bloom species. Appl. Environ. Microbiol. 2014, 80, 7512–7520. [Google Scholar] [CrossRef] [Green Version]
  89. Kenig, M.; Abraham, E.P. Antimicrobial activities and antagonists of bacilysin and anticapsin. J. Gen. Microbiol. 1976, 94, 37–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Antimicrobial compounds synthesized by B. velezensis. The compounds highlighted in red are synthesized by non-ribosomal peptide synthetases (NRPSs); blue compounds are synthesized by polyketide synthase (PKSs); green color compound bacilysin is synthesized by a ribosome independent pathway.
Figure 1. Antimicrobial compounds synthesized by B. velezensis. The compounds highlighted in red are synthesized by non-ribosomal peptide synthetases (NRPSs); blue compounds are synthesized by polyketide synthase (PKSs); green color compound bacilysin is synthesized by a ribosome independent pathway.
Molecules 25 04973 g001
Figure 2. General pathway that regulates the transcription of srfA operon, which involves extracellular peptide regulated quorum sensing in B. velezensis and B. subtilis. T-bars show the negative regulation of protein interactions; the bent arrow indicates the function of the promoter.
Figure 2. General pathway that regulates the transcription of srfA operon, which involves extracellular peptide regulated quorum sensing in B. velezensis and B. subtilis. T-bars show the negative regulation of protein interactions; the bent arrow indicates the function of the promoter.
Molecules 25 04973 g002
Figure 3. Antimicrobial mechanisms of lipopeptides (LPs) and polyketides (PKs) synthesized by B. velezensis.
Figure 3. Antimicrobial mechanisms of lipopeptides (LPs) and polyketides (PKs) synthesized by B. velezensis.
Molecules 25 04973 g003
Figure 4. Modes of actions of antibacterial activity of bacilysin synthesized by B. velezensis.
Figure 4. Modes of actions of antibacterial activity of bacilysin synthesized by B. velezensis.
Molecules 25 04973 g004
Table 1. Antimicrobial molecules synthesized by B. velezensis to control pathogenic microbes.
Table 1. Antimicrobial molecules synthesized by B. velezensis to control pathogenic microbes.
Compounds Gene ClustersAntimicrobial ActivityReferences
Antibacterial Activity (Diseases)MIC (Pathogens)Antifungal Activity (Diseases)MIC (Pathogens)
SurfactinsSrfA-DCochliobolus carbonum (Leaf spot); P. syringae pv. tomato; R. solanacearum25-100 µg/mL
(P. syringae pv. tomato)
F. verticillioides
(Maiz disease)
-[20,21,22,23]
FengycinsFenA-ER. solanacearum (Tomato wilt), X. euvesicatoria (Pepper spot); X. axonopodis pv. esicatoria-F. oxysporum (banana Fusarium wilt); F. graminearum (Fusarium head blight); B. cinerea (Grey mould); R. variabilis (Maiz disease)20.0 µg/mL
(M. grisea);
100 µg/mL (F. graminearum)
[2,24,25,26,27,28]
Bacillomycin-DbmyA-DX. campestris pv. cucurbitae-Colletotrichum gloeosporioides (Bitter rot); F. graminearum (Fusarium head blight); F. oxysporum f. sp. cucumerinum (Cucumber vascular wilt)30 μg/mL (F. graminearum)[12,29,30,31]
BacillaenebaeBCDE, acpK, baeGHIJLMNRSE. amylovora (Fire blight)-Termitomyces spp.-[32,33]
MacrolactinMlnA-IS. aureus; B. cepacia---[34]
DifficidindfnAYXBCDEFGHIJKLMX. oryzae pv. oryzae (Rice blight) and X. oryzae pv. oryzicola (Rice leaf streak); E. amylovora (Fire blight); R. solanacearum (Tomato wilt)12.62 μg/mL (R. solanacearum);--[32,35,36]
BacilysinbacA-ES. aureus and C. michiganense; X. oryzae pv. oryzae (Rice blight) and X. oryzae pv. oryzicola (Rice leaf streak); E. amylovora (Fire blight)50.0 μg/mL (X. oryzae pv. oryzae and X. oryzae pv. oryzicola)--[32,35,37]
Table 2. Commercial uses of B. velezensis-based biological control in agriculture.
Table 2. Commercial uses of B. velezensis-based biological control in agriculture.
Commercial Name Biocontrol AgentsCurrent Name of Biocontrol Agents (*NCBI Accession Number)Target Pathogens (Disease)ManufacturerReferences
RhizoVital®B. amyloliquefaciens FZB42TB. velezensis FZB42T (CP000560.2)R. solani (Bottom rot in lettuce); E. amylovora (Fire blight disease)ABiTEP, GmbH, Berlin, Germany[20,41]
BotrybelB. velezensisB. velezensisB. cinerea (Gray mold)Agricaldes, Spain[42]
Serenade®B. subtilis QST713B. velezensis QST713 (CP025079.1)Trichoderma aggressivum; Blumeria graminis (Powdery mildew)AgraQuest Inc., California, USA[43,44]
Kodiak™B. subtilis GB03B. velezensis GB03 (AYTJ00000000)F. oxysporum (Fusarium-wilt); R. solani (Cotton disease)Gustafson Inc., Texas, USA[45]
Taegro®B. subtilis var. amyloliquefaciens FZB24B. velezensis FZB24F. oxysporum (Tomato wilt); Phytophthora infestans (Potato late blight)Novozymes, Virginia, USA[46,47,48]
*NCBI: National Center for Biotechnology Information.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fazle Rabbee, M.; Baek, K.-H. Antimicrobial Activities of Lipopeptides and Polyketides of Bacillus velezensis for Agricultural Applications. Molecules 2020, 25, 4973. https://doi.org/10.3390/molecules25214973

AMA Style

Fazle Rabbee M, Baek K-H. Antimicrobial Activities of Lipopeptides and Polyketides of Bacillus velezensis for Agricultural Applications. Molecules. 2020; 25(21):4973. https://doi.org/10.3390/molecules25214973

Chicago/Turabian Style

Fazle Rabbee, Muhammad, and Kwang-Hyun Baek. 2020. "Antimicrobial Activities of Lipopeptides and Polyketides of Bacillus velezensis for Agricultural Applications" Molecules 25, no. 21: 4973. https://doi.org/10.3390/molecules25214973

Article Metrics

Back to TopTop