Review Special Issues

Chemical structure, properties and potential applications of surfactin, as well as advanced strategies for improving its microbial production

  • #Authors contributed equally to this work
  • Surfactin, a cyclic lipopeptide produced by microbes belonging to the genus Bacillus, is one of the most effective biosurfactants available in many industrial fields. However, its low production and high cost have intensively constrained its commercial applications. In this review, we first summarize the molecular structure, biological properties, beneficial roles and potential applications of surfactin in the fields of medical care and food safety, highlighting the great medical and commercial values of making its industrial production into reality. Further, genetic regulation for surfactin biosynthesis and advanced strategies for enhancing its microbial production, including optimizing fermentation conditions, rational genetic engineering and synthetic biology combined with metabolic engineering approaches, are elucidated. Finally, prospects for improving surfactin biosynthesis are discussed, and the establishment of suitable chassis hosts for exogenous production of surfactin might serve as an important strategy in future research.

    Citation: Cheng Zhen, Xian-Feng Ge, Yi-Ting Lu, Wen-Zheng Liu. Chemical structure, properties and potential applications of surfactin, as well as advanced strategies for improving its microbial production[J]. AIMS Microbiology, 2023, 9(2): 195-217. doi: 10.3934/microbiol.2023012

    Related Papers:

    [1] Allison Yaguchi, Dyllan Rives, Mark Blenner . New kids on the block: emerging oleaginous yeast of biotechnological importance. AIMS Microbiology, 2017, 3(2): 227-247. doi: 10.3934/microbiol.2017.2.227
    [2] Yu Wang, Caroline A. Brown, Rachel Chen . Industrial production, application, microbial biosynthesis and degradation of furanic compound, hydroxymethylfurfural (HMF). AIMS Microbiology, 2018, 4(2): 261-273. doi: 10.3934/microbiol.2018.2.261
    [3] Marzieh Sajadi Bami, Payam Khazaeli, Shayan Fakhraei Lahiji, Gholamreza Dehghannoudeh, Ibrahim M. Banat, Mandana Ohadi . Potential of biosurfactant as green pharmaceutical excipients for coating of microneedles: A mini review. AIMS Microbiology, 2024, 10(3): 596-607. doi: 10.3934/microbiol.2024028
    [4] Karina Sałek, Tony Gutierrez . Surface-active biopolymers from marine bacteria for potential biotechnological applications. AIMS Microbiology, 2016, 2(2): 92-107. doi: 10.3934/microbiol.2016.2.92
    [5] Tru Tran, Stephanie N. Dawrs, Grant J. Norton, Ravleen Virdi, Jennifer R. Honda . Brought to you courtesy of the red, white, and blue–pigments of nontuberculous mycobacteria. AIMS Microbiology, 2020, 6(4): 434-450. doi: 10.3934/microbiol.2020026
    [6] Patrick Di Martino . Ways to improve biocides for metalworking fluid. AIMS Microbiology, 2021, 7(1): 13-27. doi: 10.3934/microbiol.2021002
    [7] Zeling Xu, Shuzhen Chen, Weiyan Wu, Yongqi Wen, Huiluo Cao . Type I CRISPR-Cas-mediated microbial gene editing and regulation. AIMS Microbiology, 2023, 9(4): 780-800. doi: 10.3934/microbiol.2023040
    [8] Phu-Tho Nguyen, Tho-Thi Nguyen, Duc-Cuong Bui, Phuoc-Toan Hong, Quoc-Khanh Hoang, Huu-Thanh Nguyen . Exopolysaccharide production by lactic acid bacteria: the manipulation of environmental stresses for industrial applications. AIMS Microbiology, 2020, 6(4): 451-469. doi: 10.3934/microbiol.2020027
    [9] Thomas Bintsis . Lactic acid bacteria as starter cultures: An update in their metabolism and genetics. AIMS Microbiology, 2018, 4(4): 665-684. doi: 10.3934/microbiol.2018.4.665
    [10] Maria Parapouli, Anastasios Vasileiadis, Amalia-Sofia Afendra, Efstathios Hatziloukas . Saccharomyces cerevisiae and its industrial applications. AIMS Microbiology, 2020, 6(1): 1-31. doi: 10.3934/microbiol.2020001
  • Surfactin, a cyclic lipopeptide produced by microbes belonging to the genus Bacillus, is one of the most effective biosurfactants available in many industrial fields. However, its low production and high cost have intensively constrained its commercial applications. In this review, we first summarize the molecular structure, biological properties, beneficial roles and potential applications of surfactin in the fields of medical care and food safety, highlighting the great medical and commercial values of making its industrial production into reality. Further, genetic regulation for surfactin biosynthesis and advanced strategies for enhancing its microbial production, including optimizing fermentation conditions, rational genetic engineering and synthetic biology combined with metabolic engineering approaches, are elucidated. Finally, prospects for improving surfactin biosynthesis are discussed, and the establishment of suitable chassis hosts for exogenous production of surfactin might serve as an important strategy in future research.



    Pathogenic microorganisms, especially pathogenic bacteria, are microbial species that cause a variety of diseases in animals, plants and humans [1], resulting in huge economic losses worldwide every year [2]. Antibiotics serve as one of the main strategies for treating pathogenic bacterial infections, but this frequently allows the adaptive evolution of microbial species to have antibiotic resistance. In addition, many studies have demonstrated that most antibiotics could generate a range of side effects that can be life-threatening [3]. Therefore, it is essential to discover and develop new antimicrobial agents that are not resistant to therapeutic effects and have fewer side effects. Microorganisms have attracted much attention for their potential to produce multiple biologically active metabolites, among which biosurfactants are key targets for research focusing on developing novel antimicrobial agents due to their profound antibacterial and biological activities [4].

    Surfactin is a lipopeptide-based bioactive substance produced by microbial species belonging to the genus Bacillus. As one of the most effective biosurfactants available, the numerous physiological and biochemical activities of surfactin have received considerable attention [5],[6]. According to reports, surfactin has pharmacological effects like antibacterial and antifungal properties [7] and anti-mycoplasma [8], antiviral [9], anti-inflammatory [10] and thrombolytic [9] activities. Surfactin displays an amphiphilic surface activity by the presence of hydrophobic and hydrophilic parts in its molecules, which allow them to aggregate at the interface of two immiscible liquids, thus reducing the interfacial tension of the liquid or different material phases [11]. Due to the good antimicrobial effect and safety in food applications, surfactin could serve as a candidate for inhibiting bacterial growth and maintaining the sensory status of food products during storage. This important property allows surfactin to play a potential role in food preservation [12]. In addition, some studies have pointed out implications of surfactin for improving composition and structure of intestinal microorganisms, indicating its potential application in ameliorating intestinal microbiota dysfunction [13].

    Although surfactin has potential applications in many areas, it is currently not competitive with chemically synthetic surfactants due to the high cost and low yield of its production. Fermentation conditions have been intensively optimized, in order to improve the yield of surfactin. In addition, the use of genetic engineering means that improvement of surfactin production is gaining more and more attention. Moreover, the development of synthetic biology also holds great promises for further improvement of surfactin production [14].

    In this paper, the structure, physiological and biochemical characteristics and antibacterial mechanisms of surfactin and its main applications are reviewed, and the advanced strategies to improve its microbial production are also presented in detail.

    The discovery history of surfactin could date back to 1968, when this biologically active secondary metabolite was first identified in cultures of Bacillus subtilis strains [15]. The structure of surfactin has a cyclic peptide chain with 7 amino acids and a 13–16 carbon atom hydroxy fatty acid chain, which together create a cyclic lactone ring structure. Positions 2, 3, 4, 6 and 7 are occupied by hydrophobic amino acid residues, whereas positions 1 and 5, by glutamyl and aspartyl residues, respectively, add two negative charges to the molecule. In cells, various surfactin isomers often coexist as a combination of various peptide variations with various aliphatic chain lengths [16].

    Surfactin has considerable surface activity because of its amphiphilic nature, which also allows it to reduce surface/interfacial tension and self-assemble in the nanostructure. As a result, it shows physicochemical properties such as foaming [17], emulsification [18], solid surface drying prevention and chelating ability [19],[20]. Recently, it has been suggested that surfactin can also be effective in demulsifying waste crude oil [21]. Its emulsification properties also give it potential for applications in the food and cosmetic sectors, as well as in the pharmaceutical sector [22],[23].

    Surfactin is considered an extremely effective surfactant molecule. At a concentration of 20 µM, it reduces the surface tension of water from 72 mN/m to 27 mN/m, which is approximately 2 log lower than those caused by most other detergents [24],[25]. In terms of aggregation activity, the critical micelle concentration (CMC) value of surfactin decreases continuously as its fatty acid chains become longer [26]. In general, the CMC of biosurfactants is lower than that of chemically synthetic surfactants. The low CMC of biosurfactants, especially surfactin, iturin A and fengycin, makes them very interesting for use in various fields because they are required in smaller quantities compared to petrochemical-derived surfactants [27]. In general, even microbial derived surfactin with low purity has a lower CMC than its synthetic counterpart, which is highly favorable for its application in bioremediation or petroleum recovery.

    The antibacterial activity of surfactin is closely related to its biofilms. Surfactin is an amphiphilic molecule that destabilizes membranes and disrupts their integrity [28]. The mechanisms of action are as follows: insertion into lipid bilayers, chelation of monovalent and divalent cations or modification of membrane permeability by formation of channels [29].

    The arrangement of the hydrocarbon chains and the membrane's thickness are both impacted by surfactin's penetration of the membrane through hydrophobic interactions. The surfactin peptide cycle then exhibits structural changes following this initial collision, which contributes to the interaction process [30]. In vitro, surfactin binding to the membrane causes dehydration of the phospholipid polar head group and severely affects the stability of the bilayer, leading to disturbances in membrane barrier properties. This provides a good explanation for the antimicrobial effect of lipopeptides [31].

    Both monovalent and divalent cations can be propelled by surfactin past organic barriers, with divalent cations doing so more effectively [32]. At the air/water interface, the two acidic residues Glu-1 and Asp-5 are partially neutralized by Na+ and K+, whereas complete neutralization is induced by Ca2+. The cation chelation could result in the inhibition of cyclic phosphodiesterase activity. Furthermore, by neutralizing the surfactin and lipid charges and maintaining the 1:1 surfactin-calcium complex, Ca2+ could enable the surfactin to penetrate the membrane more deeply with the aid of Glu-1 and Asp-5. This effect on the conformation of surfactin promotes deeper insertion of surfactin into the cell membrane and its antimicrobial effect [6].

    Special structures of surfactin give it diverse biological functions that hold great potentials of being widely used in various fields. Potential applications and biological functions of surfactin have been briefly summarized in Table 1. We particularly focus on discussing potential applications and beneficial roles of surfactin in the fields of medical care and food safety in this section.

    Table 1.  Potential applications and biological functions of surfactin.
    Potential applications Biological functions Refs.
    Antibacterial activity Suppress the expression of inflammatory mediators [33]
    Biological control of Arabidopsis root infection [51]
    Antibacterial activity against Brachyspira hyodysenteriae and Clostridium perfringens [45]
    Antiviral activity Decreases the titer of herpes simplex virus (HSV-1) [36]
    Activity on pseudorabies virus in vitro [52]
    Inhibits invasion of epithelial cells by enveloped viruses [53]
    Antimycoplasma activity Complete inactivation of mycoplasma in mammalian monolayer and suspension cell cultures. [37]
    Anticancer activity Growth inhibition of MCF-7 human breast cancer cells [40]
    Inhibits cancer progression through growth inhibition, cell cycle arrest, apoptosis and metastasis arrest [41]
    Induced necrosis-like death of Huh 7.5 hepatocellular carcinoma cells [54]
    Microbial oil recovery Effective in separating oil when used as an emulsifier [18]
    Enhanced oil recovery by 9.2% [55]
    Maintenance of physical and oxidative stability of O/W algal oil DHA emulsions [56]
    Food antistaling agent Reduced the growth of Salmonella enterica in meat [43]
    Helps yogurt maintain its sensory quality and extend its shelf life [44]
    Reduces the growth of mesophilic bacteria and maintains the organoleptic state of the juice [43]
    Extend the shelf lives of fruits, vegetables and grains [45]
    Maintenance of gastrointestinal homeostasis Regulates the intestinal microbiota of broilers [48]
    Improve the intestinal health of tilapia (Oreochromis niloticus) [49]
    Elimination of Staphylococcus aureus colonizing the human intestine by inhibiting population sensing [50]
    Cosmetics field Suitability of surfactin as powerful emulsifier in cosmetics [57]
    Bioremediation Stimulating indigenous microorganisms for enhanced bioremediation of diesel contaminated soil [58]
    Removal of heavy metals from contaminated soil and sediments [59]

     | Show Table
    DownLoad: CSV

    The action of surfactin with the bilayer confers its antibacterial effect. Recent studies have revealed that surfactin has an impact on how endotoxin (lipopolysaccharide, LPS) interacts with eukaryotic cells. This compound, which inhibits endotoxin activity, has the potential to become a new anti-inflammatory agent. Surfactin might lower plasma levels of endotoxin and nitric oxide in septic shock rats as well as suppress the expression of endotoxin-induced inflammatory mediators (IL-1 and iNOS) [33],[34].

    The number of carbon atoms on the fatty acid chain of surfactin has a significant role in the inactivation of viruses, and it has been demonstrated that surfactin exhibits an antiviral impact, particularly against enveloped viruses [35]. Increased fatty acid hydrophobicity in surfactin typically demonstrates potent virus inactivation capabilities. Surfactin attaches to the lipid bilayer during inactivation, causing the viral proteins involved in viral adsorption and invasion to completely detach from the envelope and disintegrate. Surfactin was found to be more effective at inactivating enclosed viruses than unenveloped viruses in in vitro tests on the drug's effects on various viruses, particularly herpes simplex and retroviruses. In medium containing 5% fetal calf serum (FCS), surfactin was active at 25 mM, whereas surfactin at a concentration of 80 mM caused a decrease in herpes simplex virus (HSV-1) titer > 4.4 log10 CCID50/mL within 15 min [36].

    Mycoplasma is the smallest free-living microorganism and is a specialized parasite that causes respiratory inflammation and diseases of the urogenital tract. In in vitro experiments, the proliferation rate and morphology of mycoplasma-contaminated mammalian cells were improved when surfactin was used [37]. In addition, the low toxicity of surfactin in in vivo models and its potential to prevent sexually transmitted diseases caused by genital tract mycoplasma infections are also of great importance.

    Surfactin has been found to exhibit anticancer properties against several cancer cells, including the ability to limit cancer spread and have antiproliferative and apoptotic effects [38]. Surfactin-mediated programmed cell death, which is primarily produced by apoptosis, regulates tissue development and homeostasis in vivo and offers a possible method for treating cancer [39].

    Breast cancer is a malignant tumor that poses a serious health risk to women, and the most extensive research has focused on surfactin's anticancer efficacy against breast cancer cells probably. It was shown that B. subtilis CSY 191-derived surfactin inhibited the growth of human breast cancer cells MCF-7 in a dose-dependent manner with an IC50 of 9.65 µM at 24 hours. Furthermore, compared to the surfactin produced by the CSY 191 strain alone, the high yield obtained by co-fermenting cheonggukjang (a Korean sauce) and CSY 191 strain further boosted the degree of anticancer activity by a factor of two [40].

    In addition to breast cancer, surfactin has shown biological activities against colon, cervical and hepatocellular cancers. Recently, many studies have investigated the mechanism of its anticancer activity. It was shown that surfactin attenuated 12-O-tetradecanoylphorbol-13-acetate (TPA)-induced nuclear translocation and activation of nuclear factor-κB (NF-κB) and activator protein-1 (AP-1), thereby inhibiting cancer cell invasion and metastasis. In addition, surfactin significantly inhibited the expression and activation of matrix metalloproteinase-9 (MMP9), an enzyme that degrades almost all protein components in the ECM (extracellular matrix), disrupting the histological barrier to tumor cell invasion and playing a key role in tumor invasion and metastasis [38]. Moreover, the amphiphilic nature of surfactin makes them readily incorporated into nanopreparations, and the use of nanopreparations offers the advantage of optimized surfactin delivery for improved anticancer therapy [41].

    Due to its small molecular weight, thermal stability, water solubility, non-toxicity and lack of side effects, surfactin is not harmful to natural foods that require long term storage [42]. In a previous study, surfactin and iturin strongly decreased the growth of Salmonella enterica in meat by five orders of magnitude [43]. Studies have shown that antimicrobial peptides (especially surfactin) combined with preservatives can help yogurt maintain sensory quality and extend shelf life [44]. In addition, surfactin has been shown to have a bactericidal effect when added to milk. The lipopeptide produced by B. subtilis fmbJ greatly reduced the growth of mesophilic bacteria, slowed down rancidity and preserved the juice's outstanding organoleptic status in fresh watermelon [43]. Surfactin has also been shown in previous studies to extend the shelf lives of fruits, vegetables and cereals as a biological preservative with minimal toxicity and degradability [45]. These studies illustrate the powerful role of surfactin in food preservation and storage.

    As soil microbes are thought to be transient, unlike the popular probiotic Lactobacillus, representatives of the genus Bacillus were not thought to be a natural component of the human gut microbiota. Despite the fact that some of its members remain in the host for a very long time and create a wide variety of bioactive chemicals, most current investigations of the human intestinal microbiota typically do not take into account the special functional role of the transitory microbiota. Representatives of the genus Bacillus colonize the epithelium as food, water and probiotic preparations enter the gastrointestinal tract (GIT), obfuscating the distinction between the resident and transient microbiota. Although these bacteria represent a small proportion of the microbiome composition, they can have a significant impact on the gut microbiota and the whole body due to the large variety of secreted compounds they produce [46],[47].

    In a recent study, the metabolite surfactin produced by B. subtilis was shown to be effective in improving growth performance, alleviating expression of intestinal inflammatory genes and having a regulatory effect on the intestinal microbiota in broilers [48]. In addition, the addition of antibacterial peptide surfactin to tilapia (Oreochromis niloticus) feed can improve intestinal health by increasing the height of intestinal folds, regulating intestinal flora and increasing intestinal antioxidant capacity [49]. A class of widely used lipopeptides (fengycins) produced by Bacillus has been shown to eliminate Staphylococcus aureus colonizing the human gut by inhibiting population sensing, a process by which bacteria respond to their population density by altering their genetic regulation [50]. These studies highlight the importance of probiotics derived lipopeptides in maintaining intestinal homeostasis and reducing infectious diseases.

    The molecular mechanism behind probiotic efficacy has not yet been fully uncovered or is only partially understood. Future research aiming to decipher the mechanisms that determine the properties of bacterial probiotics will certainly expand the field of scientifically proven probiotics for medical use, which are considered promising alternatives for the prevention of gastrointestinal infections.

    Surfactin is widely biosynthesized not only in B. subtilis but also in Bacillus mojavensis, Bacillus licheniformis, Bacillus circulans, Bacillus nidulans, and Bacillus amylolyticus [60]. Like most cyclic lipopeptides, surfactin is not synthesized by ribosomes but by a special system called non-ribosomal peptide synthase (NRPS) (Figure 1). SrfAA, SrfAB, SrfAC and SrfAD, which comprise a linear array of seven modules, are four modules that make up NRPS, and each is in charge of adding one amino acid [61],[62]. Each module has three catalytic structural domains at a minimum: an adenylation structural domain (A) that selects and activates substrates, a small peptidyl carrier protein (PCP) that transports aminoacyladenosine substrates as enzyme-bound thioesters and a condensation structural domain (C) that forms a peptide bond between acyl-S-PCP intermediates [63]. An extra thioesterase (TE) of the terminating module catalyzes the release of the product by hydrolysis or macrocyclization to produce cyclic or cyclic branched molecules after the epimeric (E) structural domain experiences a stereochemical conversion to produce a d-isomer of some bound residues [64].

    Figure 1.  Non-ribosomal peptide synthetase modules for surfactin biosynthesis. Note: A, adenylation domain, represented by amino acids in red; PCP, peptidyl carrier protein domains, shown in green; C, condensation domain, shown in gray; E, epimerization domain, shown in blue; TE, thioesterase domain, shown in orange.

    The synthesis of surfactin receives control from genetic regulation in addition to being influenced by the NRPS module. Surfactin synthesis is controlled by two genetic motifs, named srf and sfp. SRF is encoded by an inducible manipulator, the srfA motif (25 kb), and composed of four modular open reading frames (ORFs) [65], the function of which has been elucidated above. The second gene critical for surfactant production is sfp, located 4 kb downstream of the srfA manipulator [66]. Sfp encodes an enzyme belonging to the 4′-phosphopantetheinases family, which acts as a non-ribosomal peptide synthesis promoter. SFP enzymes can also convert the inactivated deoxygenated form of surfactin synthase into a holo-form containing cofactors, which can play an active role [67]. SrfA expression is induced at late exponential stages of bacterial growth and regulated by a number of transcriptional regulatory genes. SrfA expression is controlled by ComP-ComA in the quorum-sensing system, a mechanism that senses nutrient stress and regulates the expression of multiple genes [65].

    The biosynthetic system of fatty acids, especially branched-chain fatty acids, also plays an important role in the synthesis of surfactin. The β-ketoacyl-acyl carrier protein synthase III (FabH) of B. subtilis has excellent activity and selectivity for branched-chain fatty acid synthesis precursors and can start the straight-chain and branched-chain fatty acid synthesis cycles by condensing acetyl-CoA, isobutyl-CoA, isopentyl-CoA, or -methylbutyl-CoA with malonyl ACP [68].

    Surfactin has been extensively studied for its antibacterial, antiviral, antitumor and hemolytic effects. However, due to its high economic cost and low synthetic yield, it cannot compete with chemically synthesized surfactants [6]. Although the current market for surfactin is estimated to grow at an annual rate of 2–4%, the high cost of production remains an issue limiting its large-scale synthesis. Moreover, its synthesis requires many steps and expensive mediator components, which also limits its development prospects [60]. Notably, despite certain advancements in the biosurfactant business, poor yields are still produced, and the downstream processes of recycling and purification have not yet achieved a sufficient economic scale [69]. In addition, the low rate of surfactin synthesis may be related to the complex regulation of biosynthesis, which results in production dependent on the cell density. This prevents continuous synthesis and limits the total yield [70].

    In recent years, research on improving surfactin yield by fermentation engineering has gradually increased. Fermentation parameters such pH, temperature, stirring rate, oxygen supply and medium composition are important factors to optimize, as they all affect the yield of surfactin. The surfactin yield of B. subtilis BS5 increased from 1.25 to 1.9 g/L when the pH was increased from 6 to 8 [71].

    Temperature is also an important parameter of the culture medium. In the study of Amani et al., the yield of surfactin increased with increasing temperature, and the highest yield was at 37 °C [72]. The yield of surfactin varied when changing the agitation speed and oxygen supply of the medium, reaching a maximum yield of 6.45 g/L [25].

    However, titers of most wild Bacillus strain-derived surfactin were below 5.0 g/L (Table 2). This is due to the antibiotic and signaling activity of surfactin itself, and this yield is not enough for large-scale production and industrial applications.

    Table 2.  Fermentation conditions for microbial production of surfactin.
    Strains Conditions Modification Surfactin production (g/L) Refs.
    B. subtilis BS5 pH 6.0 1.25 [71]
    B. subtilis BS5 8.0 1.90
    Bacillus amyloliquefaciens 7.0 6.04 [73]
    B. subtilis NLIM 0110 Temperature (°C) 25 1.75 [72]
    B. subtilis NLIM 0110 30 3.20
    B. subtilis NLIM 0110 37 4.00
    B. subtilis NLIM 0110 45 3.50
    B. subtilis ATCC 21332 Agitation speed (rpm) 0 0.44 [25]
    B. subtilis ATCC 21332 200 2.66
    B. subtilis ATCC 21332 250 4.44
    B. subtilis ATCC 21332 300 6.45
    B. subtilis ATCC 21332 350 1.01
    B. subtilis ATCC21332 Oxygen supply (vvm, L min−1) 1.50,3 6.45 [25]
    B. subtilis ATCC21332 1.00,1 3.83
    B. subtilis ATCC21332 0.50,1 2.80
    B. subtilis BS5 Medium compositions MSM plus FeSO4 0.40 [71]
    B. subtilis BS5 MSM plus MnCl2 0.70
    B. subtilis BS5 MSM plus FeCl3 1.00
    B. subtilis BS5 MSM plus ZnSO4 1.75

     | Show Table
    DownLoad: CSV

    In the biosurfactant industry, profitable product application is a major concern. Proper selection of substrates to be used in the production of biosurfactants is a fundamental prerequisite. Water-insoluble substrates are often used in biosurfactant production, due to the fact that microbial communities that produce biosurfactants are often isolated from petroleum and hydrocarbon-contaminated environments [69]. However, increasing evidence showed that water-soluble carbon sources, such as glucose, fructose and sucrose, can also be used as substrates for the synthesis of surface active substances from various microbial communities [74].

    Despite the increasing number of reports on surfactin, the commercial production process remains somewhat difficult, mainly due to the increased cost of chemicals in the growth media used to produce surfactin. Cheap waste biomass is now being used as a substitute for refined carbon and nitrogen sources in order to reduce costs. For example, in a study of surfactin production using low-cost beer waste as a carbon source, B. subtilis produced surfactin at an amount of 210.11 mg/L for 28 hours, with inhibition activity against all tested bacteria, and complete inhibition was achieved against Pseudomonas aeruginosa, indicating that surfactin produced from low-cost substrates has the potential to be a promising bactericide [75]. In addition, a number of studies have been conducted on the production of surfactin from hydrolysate, glutamate, wine lees and other wastes with the aim of reducing its production cost by using cheap carbon sources, and the final yield of surfactin was around 500 mg/L [76],[77].

    Based on a series of problems in surfactin production, the study aims to enhance the different stages of surfactin biosynthesis through advanced techniques to achieve higher surfactin yields by altering transcriptional regulatory genes and promoter substitutions, attenuating competition pathways, increasing precursor supply or enhancing secretion (Figure 2).

    It has been well documented that surfactin production can be significantly increased using genetic engineering approaches (Table 3). Surfactin synthesis is critically regulated by the Psrf promoter, which controls the expression of the srfA operon. In the genetic modification of surfactin, promoter swapping has attracted much attention due to the challenge of heterologous expression of the srfA operon. As previously described, the four genes srfAA, srfAB, srfAC and srfAD are activated by signaling molecules of the quorum-sensing pathway under the control of the Psrf self-inducible promoter responsible for encoding the NRPS of surfactin [78].

    In a previous study [85], four strong promoters, PgroE, Pcdd, PrplK and PsspE, were identified and cloned from the genome of B. subtilis THY-7. The optimal PgroE promoter was screened by single crossover homologous recombination, but the PgroE-containing strain was unable to synthesize surfactin. Subsequent replacement of PsrfA with the sucrose-inducible promoters PsacB and PsacP yielded engineered strains that produced 1.09 and 0.22 g/L of surfactin, respectively. By fusing the PgroE and PsacB ribonucleotide promoters (RAT), an artificial sucrose-inducible Pg1 promoter was generated, and the engineered strain containing the Pg1 substitution produced a surfactin titer of 1.44 g/L. An artificial IPTG-inducible promoter Pg2 was constructed from a PgroE-lacO fusion, which then replaced the PsrfA locus on the chromosome to yield THY-7/Pg2-srfA, whose surfactin titer increased to 5.98 g/L. On top of this, a new promoter Pg3 was generated by adding two point mutations in the -35 and -10 regions, which produced up to 9.74 g/L of surfactin, 15.6-fold higher than the original strain. In addition, a library of PsrfA derivatives was optimized by shortening the sequence of PsrfA and changing the nucleotides in the conserved regions of -35, -15 and -10 regions, using green fluorescent protein (GFP) as the reporter protein. The strongest promoter, P10, which was screened, had 150% higher GFP expression intensity than PsrfA and was able to efficiently heterologously express the exogenous protein [89]. As a tailored SrfA expression promoter for each strain may be more appropriate, it is crucial to investigate efficient promoters to increase surfactant production.

    Figure 2.  Genetic engineering strategies for enhancing microbial production of surfactin. Note: (A) Attenuating competition pathway module. The red arrow and the black cross indicate the deletion of pathways. (B) Enhancing precursor supply module. Overexpressed genes and important precursors are highlighted by green and red, respectively. (C) Transcriptional regulation of srf operon module. T-bar indicates the negative effects on DNA binding or protein interactions. “P” in the circle represents the phosphoryl group. A bent arrow represents the promoter. The red arrows and the black crosses indicate the deletion of negative transcriptional regulators. (D) Increasing efflux and resistance module. Specific roles of the proteins are listed following black arrows.
    Table 3.  Genetic manipulations for improving surfactin yield.
    Strains Modification Surfactin production (g/l) Refs.
    Branched chain fatty acids biosynthesis
    B. subtilis 168S23 Promoter substitution of key genes for precursor material 8.5 [79]
    B. subtilis TS1726 Overexpression of key genes 13.37 [80]
    Branched chain amino acids biosynthesis
    B. subtilis H1 Inhibition of negative transcriptional regulators 0.75 [81]
    B. amyloliquefaciens MT45 Upregulation of synthetic genes in L-glu / L-asp 2.93 [70]
    B. subtilis TS1726ΔspoIVB Overexpression of key genes and medium optimization 16.7 [82]
    Surfactin efflux
    B. subtilis 168S12 Overexpression of transporter proteins 3.8 [83]
    B. subtilis TS662 Overexpression of transporter genes 1.67 [84]
    Promoter engineering
    B. subtilis THY-7/Pg3-srfA Replacement of PsrfA with Pg3 9.74 [85]
    B. subtilis fmbR-1 Replacement of PsrfA with Pspac 3.86 [86]
    Regulation of transcription factors
    B. subtilis TS1726ΔspoIVB Deleting regulator of sporulation 9.6 [82]
    B. subtilis (pHT43-comXphrC) Overexpression of positive regulators 0.135 [87]
    B. subtilis 168S35 Overexpression of positive regulators and deletion of negative regulators 12.8 [83]
    Genome reduction
    B. amyloliquefaciens GR167 4.18% reduction of genome 9.7% increased [88]

     | Show Table
    DownLoad: CSV

    The expression of the srfA operon is controlled by a number of transcriptional regulatory genes in addition to the promoter. ComX and CSF are two typical peptides that control quorum-sensing in B. subtilis. By phosphorylating the transcription factor ComA, ComX and CSF each use a different mechanism to start the srfA operon's transcription (ComA-P) [90],[91]. In previous studies, the expression of srfA was successfully reduced by repressing the quorum-sensing system ComQXP transcription in B. subtilis. Overexpression of the two signal factors encoded by comX and phrC in B. subtilis could stimulate transcription of the srfA operon, thereby increasing the production of surfactin [87],[92].

    Multiple global regulators and proteins also control the expression of the srfA operon. CodY can restrict srfA transcription by binding directly to the srfA promoter area, which can be activated by high amino acid concentrations. Inactivation of codY could increase surfactin production in B. subtilis 168 by roughly 10-fold [93]. In addition, other negative regulators of srfA have been studied in detail and reported, such as sinI, spX, raP and perR [70],[83],[94]. Moreover, the synthesis of surfactin was increased after knocking out spore synthesis-related genes [82]. This may be related to the utilization of nutrients by spore synthesis. At present, quorum-sensing directed regulatory network needs to be further investigated, and more signal peptides and transcriptional regulators of surfactin synthesis need to be explored.

    Different metabolic engineering techniques, such as (i) increased supply of branched-ketoacyl-CoA, (ii) increased synthesis of malonyl-ACP and (iii) overexpression of the whole fatty acid synthase complex, were used to augment surfactin production in terms of branched fatty acid supply.

    The catalytic interaction of branched-chain α-keto-CoA with malonyl-ACP by FabHB (β-ketoacyl-carrier protein synthase III) initiates the synthesis of branched-chain fatty acids. In order to increase the intracellular content of branched α-keto CoA to increase the supply of branched fatty acid biosynthetic precursors, the branched α-keto acid dehydrogenase complex (BKD) is overexpressed. BKD is a complex manipulator of branched keto acid biosynthesis, because the dehydrogenase activity of BKD requires lipidation. Overexpression of BKD in B. subtilis may compete with other lipoic acid-dependent complexes for lipidation of proteins and inhibits cell proliferation and surfactin production. It has been shown that lipoic acid biosynthesis in B. subtilis is dependent on three enzymes, LIPA, LIPL and LIPM. Overexpression of lipALM to eliminate the competitive lipoylation process between BKD and other lipoic acid-dependent complexes led to a surfactin yield of 4.6 g/L, an increase of 21% compared to the parental strain [83]. Another targeting pathway for modification is the synthesis of malonyl-ACP. In the fatty acid biosynthetic pathway, the acetyl CoA carboxylase complex (ACCD, ACCA, ACCB, ACCC) catalyzes the conversion of acetyl-coA to malonyl-coA. FabD (malonyl-coA: acyl carrier protein transacylase) then converts the malonyl-coA to malonyl-ACP, starting the synthesis and expansion of the fatty acid chain. According to these findings, overexpression of the endogenous accDABC and fabD increased surfactin production by 14% [83]. FBF (β-ketoacyl-acyl carrier protein synthase II), FBG (β-ketoacyl-acyl carrier protein reductase), FBZ (β-hydroxyacyl-acyl carrier protein dehydratase) and FBI made up the fatty acid synthase complex (cyclic enoyl-acyl carrier protein reductase). After overexpressing the fatty acid synthase complex with increasing supplies of both branched-ketoacyl-coA and malonyl-coA, the surfactin production was further raised to 8.5 g/L. However, if these interventions are combined with transcriptional enhancement of the srf gene cluster, surfactin production can be further increased to 12.8 g/L, reaching 42% of the theoretical production [83].

    In addition to fatty acids, amino acids are also important precursors for the biosynthesis of surfactin. The genes yrpC, racE or murC for L-glutamate synthesis and bkdAA and bkdAB for L-leucine and L-valine synthesis were inhibited by CRISPRi technology, reducing the metabolic flux of competing amino biosynthetic pathways, which increased the production of surfactin and the proportion of C14 subtypes [81]. After overexpression of the gene for the leucine pathway in the non-spore producing strain TS1726, more leucine (5 g/L) was added, and the surfactin titer reached 16.7 g/L [82].

    In B. subtilis 168, although biofilm synthesis is not possible, the gene clusters epsA-O and tasA-sipW-yqxM, which are associated with biofilm synthesis, still have high transcriptional activity. EpsA-O and tasA-sipW-yqxM are responsible for polysaccharide and amyloid synthesis, respectively, in B. subtilis. Xu et al. knocked out the eps gene operon and the tasA-sipW-yqxM operon, respectively, and surfactin production increased 1.8 and 1.3-fold, respectively. When both operons were knocked out simultaneously, the surfactin yield reached 1.4 g/L, which was 2.5-fold higher than the original. On top of this, further knockdown of the gene clusters pps, dhb and pks, encoding nonribosomal peptide fengycin, the siderophore bacillibactin and an unknown polyketide, increased the surfactin yield by about 3.3-fold [83]. This is probably because the biosynthesis of these different proteins, lipopeptides and polyketides will inevitably compete with surfactin production for energy, NADH and direct precursors.

    With the development of systems biology, homologous recombination and genome reduction techniques have become popular research topics for the construction of various chassis cells. In a previous study, genome reduction in B. amyloliquefaciens LL3 strain deleted about 4.18% of non-essential genes, reducing the competition for surfactin precursor material and energy, leading to a significant increase in surfactin production [88].

    Tsuge et al. determined the correlation between the expression level of the yerP gene and the resistance of B. subtilis to surfactin, and they hypothesized that yerP was associated with surfactin efflux. Although there was no direct evidence that YerP can cause surfactin efflux, the inherent metabolite surfactin of B. subtilis severely inhibited the growth of YerP-deficient strains, suggesting that the enhanced susceptibility of mutant strains to surfactin was closely linked to the dysfunction of YerP [95]. Li et al. showed that overexpression of three lipopeptide transporters (YcxA, KrsE and YerP), the functioning of which depended on proton motive force, resulted in 89%, 52% and 145% increases in surfactin export, respectively [84]. Although the mechanism of self-resistance in surfactin-producing strains was not well known, the production of surfactin was increased to 3.8 g/L after overexpression of resistance proteins SwrC, AcrB and LiaIHGFSR [83]. These studies hold promise for a better understanding of surfactin efflux.

    With the development of synthetic biology, various genes have been heterologously expressed, leading to successfully heterologous biosynthesis of various molecular compounds [96]. Studies on heterologous expression of surfactin in non-Bacillus hosts have not been reported, as cloning of large DNA fragments and stable expression of heterologous biosynthetic genes remain challenging [97]. Several synthetic biology tools have been developed to clone entire biosynthetic gene clusters directly from complex genomes, such as tools for yeast transformation-associated recombination (TARs) [98], yTREX systems [97] and multiplex CRISPR-TAR [99] for promoter engineering. These tools have been widely used to clone and modify biosynthetic genes, including some genes related to surfactin synthesis. In the future, synthetic biology tools could be used to engineer more stable genes to ensure efficient expression of biosynthetic gene clusters and to replace traditional surfactin-producing hosts.

    In the future, a new research direction is to use synthetic biotechnology and metabolic engineering to improve the ability of strains to obtain surfactin from xylose as the sole carbon source. When xylose was used as the sole carbon source, B. subtilis 168 reduced organic acid by-products and was able to efficiently produce surfactin (up to 2.074 g/L). However, the utilization of xylose is limited by the production characteristics of the strain and enzyme and the inhibition of AraE by the AraR protein [100]. Hu et al. found that B. subtilis 168 could effectively use xylose to eliminate the inhibition of AraE by AraR when combined with organic nitrogen sources. Recombinant strain BSFX022 increased the supply of precursor fatty acyl CoA by overexpressing sfp and bte genes and yhfl gene, thereby overexpressing 4′-phosphopanthenol transferase, medium-chain acyl-acyl carrier protein (ACP) thioesterase and fatty acyl CoA ligase, ultimately producing 2.203 g/L of surfactin [101].

    Nevertheless, B. subtilis using xylose as the sole carbon source for surfactin synthesis displayed some shortcomings, such as the complex and rigorous mechanism of CCR (carbon catabolite repression) and insufficient metabolic regulation of xylose utilization. A new research direction is to use synthetic biotechnology combined with metabolic engineering to gradually improve xylose metabolism in the host and enhance the ability of the strain to produce surfactin. An “optimization” approach for this pathway was established by applying synthetic biology design principles (learn, design, build and test) (Figure 3) and transcriptome optimization of the target bacterial gene cluster, combined with optimization of promoters, RBS (ribosome binding sites), terminators and other components in the context of metabolic engineering, promising further improvements in surfactin production [102].

    Figure 3.  Synthetic biology and metabolic engineering of xylose utilization as exemplified by Crispr technology.

    Caulobacter crescentus has been found to be able to metabolize xylose efficiently, and many studies have been conducted to establish a new pathway for xylose utilization by heterologously expressing its related enzymes in other microorganisms [103]. Related approaches include evolutionary engineering, which identifies mutations that favor the growth of B. subtilis on xylose. These mutations would be designed to derepress AraR and produce recombinant B. subtilis that could utilize xylose for rapid growth. The strains were designed to effectively improve xylose conversion efficiency. Screening for xylose-specific transporter proteins by enhancing the expression of the transporter protein AraE and designing homologous or heterologous transporter proteins can also greatly improve xylose utilization. In conclusion, the continued development of synthetic biology will advance research on the production of surfactin from xylose, thus providing another option for inexpensive industrial production.

    Surfactin has stimulated a lot of interest because of its specific structure and biological activity, including its activity in microbial oil recovery and emulsification properties. This paper reviews the potential applications and beneficial effects of surfactin, such as antibacterial effects, antiviral effects, anticancer effects, application in food preservation and maintenance of gastrointestinal homeostasis. However, although surfactin has great medical and commercial value, achieving its industrial production has been a challenge because of the high cost and low yield of its synthesis. At present, the highest titers of surfactin by microbial production could reach up to 16.7 g/L according to the most updated report [82], which offers a robust strategy combining genetic engineering and fermentation optimization for re-modeling B. subtilis to further improve its fermentation efficiency and industrial application for surfactin production. Here, we describe the non-ribosomal peptide synthesis mode of surfactin synthesis and the transcriptional regulators that regulate its synthesis. We also highlight some of the latest strategies that could be used to improve the yield of surfactin, including fermentation engineering approaches, rational genetic engineering and the combination of synthetic biology and metabolic engineering.

    Notably, the establishment of suitable hosts for exogenous expression of surfactin biosynthetic genes to provide chassis strains for achieving efficient surfactin production should be an important direction for future research. Meanwhile, with the development of systematic biology, modeling of metabolic networks will help us to explore the interactions between metabolites and key metabolic modules and identify more information that may be useful for surfactin production. In addition, it has been found that the abundance of surfactin in the medium decreases as fermentation proceeds to the late stages of incubation (nutrient depletion), which may be related to an increase in protease activity [104]. The degradation process of surfactin has been little studied, and understanding its mechanism might also be important to further enhance microbial biosynthesis of surfactin for industrial production.


    Acknowledgments



    This work was supported by the National Natural Science Foundation of China (grant number 32202121).

    Conflict of interest



    The authors declare no competing interests.

    [1] Kaper JB, Nataro JP, Mobley H (2004) Pathogenic Escherichia coli. Nat Rev Microbiol 2: 123-140. https://doi.org/10.1038/nrmicro818
    [2] Wang YX, Ye ZZ, Si CY, et al. (2012) Application of aptamer based biosensors for detection of pathogenic microorganisms. Chinese J Anal Chem 40: 634-642. https://doi.org/10.1016/S1872-2040(11)60542-2
    [3] Cunha BA (2001) Antibiotic side effects. Med Clin N Am 85: 149. https://doi.org/10.1016/S0025-7125(05)70309-6
    [4] Rangarajan V, Clarke KG (2015) Process development and intensification for enhanced production of Bacillus lipopeptides. Biotechnol Genet Eng Revs 31: 46-68. https://doi.org/10.1080/02648725.2016.1166335
    [5] Wei YH, Wang LF, Chang JS, et al. (2003) Identification of induced acidification in iron-enriched cultures of Bacillus subtilis during biosurfactant fermentation. J Biosci Bioeng 96: 174-178. https://doi.org/10.1263/jbb.96.174
    [6] Seydlova G, Svobodova J (2008) Review of surfactin chemical properties and the potential biomedical applications. Cent Eur J Med 3: 123-133. https://doi.org/10.2478/s11536-008-0002-5
    [7] Das P, Mukherjee S, Sen R (2008) Antimicrobial potential of a lipopeptide biosurfactant derived from a marine Bacillus circulans. J Appl Microbiol 104: 1675-1684. https://doi.org/10.1111/j.1365-2672.2007.03701.x
    [8] Boettcher C, Kell H, Holzwarth JF, et al. (2010) Flexible loops of thread-like micelles are formed upon interaction of L-alpha-dimyristoyl-phosphatidylcholine with the biosurfactant surfactin as revealed by cryo-electron tomography. Biophys Chem 149: 22-27. https://doi.org/10.1016/j.bpc.2010.03.006
    [9] Sachdev DP, Cameotra SS (2013) Biosurfactants in agriculture. Appl Microbiol Biot 97: 1005-1016. https://doi.org/10.1007/s00253-012-4641-8
    [10] Zhang YY, Liu C, Dong B, et al. (2015) Anti-inflammatory activity and mechanism of surfactin in lipopolysaccharide-activated macrophages. Inflammation 38: 756-764. https://doi.org/10.1007/s10753-014-9986-y
    [11] Gang HZ, Liu JF, Mu BZ (2011) Molecular dynamics study of surfactin monolayer at the air/water interface. J Phys Chem B 115: 12770-12777. https://doi.org/10.1021/jp206350j
    [12] Barros F, de Quadros CP, Marstica MR, et al. (2007) Surfactin: Chemical, technological and functional properties for food applications. Quim Nova 30: 409-414. https://doi.org/10.1590/S0100-40422007000200031
    [13] Zhai SW, Chen XH, Wang MH (2017) Effects of different levels of dietary surfactin supplementation on intestinal morphology, and intestinal microflora of growth retarded marbled eel juveniles (Anguilla marmaorata). Isr J Aquacult-Bamid 69: 1433.
    [14] Xia L, Wen JP (2022) Available strategies for improving the biosynthesis of surfactin: A review. Crit Rev Biotechnol . https://doi.org/10.1080/07388551.2022.2095252
    [15] Arima K, Kakinuma A, Tamura G (1968) Surfactin, a crystalline peptidelipid surfactant produced by Bacillus subtilis: isolation, characterization and its inhibition of fibrin clot formation. Biochem Bioph Res Co 31: 488-494. https://doi.org/10.1016/0006-291X(68)90503-2
    [16] Hue N, Serani L, Laprevote O (2001) Structural investigation of cyclic peptidolipids from Bacillus subtilis by high-energy tandem mass spectrometry. Rapid Commun Mass Sp 15: 203-209. https://doi.org/10.1002/1097-0231(20010215)15:3<203::AID-RCM212>3.0.CO;2-6
    [17] Fei D, Zhou GW, Yu ZQ, et al. (2020) Low-toxic and nonirritant biosurfactant surfactin and its performances in detergent formulations. J Surfactants Deterg 23: 109-118. https://doi.org/10.1002/jsde.12356
    [18] Long XW, He N, He YK, et al. (2017) Biosurfactant surfactin with pH-regulated emulsification activity for efficient oil separation when used as emulsifier. Bioresource Technol 241: 200-206. https://doi.org/10.1016/j.biortech.2017.05.120
    [19] Shakerifard P, Gancel F, Jacques P, et al. (2009) Effect of different Bacillus subtilis lipopeptides on surface hydrophobicity and adhesion of Bacillus cereus 98/4 spores to stainless steel and Teflon. Biofouling 25: 533-541. https://doi.org/10.1080/08927010902977943
    [20] Marcelino L, Puppin-Rontani J, Coutte F, et al. (2019) Surfactin application for a short period (10/20 s) increases the surface wettability of sound dentin. Amino Acids 51: 1233-1240. https://doi.org/10.1007/s00726-019-02750-
    [21] Yang ZY, Zu YQ, Zhu JS, et al. (2020) Application of biosurfactant surfactin as a pH-switchable biodemulsifier for efficient oil recovery from waste crude oil. Chemosphere 240: 124946. https://doi.org/10.1016/j.chemosphere.2019.124946
    [22] Zouari R, Besbes S, Ellouze-Chaabouni S, et al. (2016) Cookies from composite wheat-sesame peels flours: Dough quality and effect of Bacillus subtilis SPB1 biosurfactant addition. Food Chem 194: 758-769. https://doi.org/10.1016/j.foodchem.2015.08.064
    [23] Ohadi M, Shahravan A, Dehghannoudeh N, et al. (2020) Potential use of microbial surfactant in microemulsion drug delivery system: A systematic review. Drug Des Dev Ther 14: 541-550. https://doi.org/10.2147/DDDT.S232325
    [24] Heerklotz H, Seelig J (2001) Detergent-like action of the antibiotic peptide surfactin on lipid membranes. Biophys J 81: 1547-1554. https://doi.org/10.1016/S0006-3495(01)75808-0
    [25] Yeh MS, Wei YH, Chang JS (2005) Enhanced production of surfactin from Bacillus subtilis by addition of solid carriers. Biotechnol Progr 21: 1329-1334. https://doi.org/10.1021/bp050040c
    [26] Liu Q, Lin JZ, Wang WD, et al. (2015) Production of surfactin isoforms by Bacillus subtilis BS-37 and its applicability to enhanced oil recovery under laboratory conditions. Biochem Eng J 93: 31-37. https://doi.org/10.1016/j.bej.2014.08.023
    [27] Das P, Yang XP, Ma L (2014) Analysis of biosurfactants from industrially viable Pseudomonas strain isolated from crude oil suggests how rhamnolipids congeners affect emulsification property and antimicrobial activity. Front Microbiol 5: 696. https://doi.org/10.3389/fmicb.2014.00696
    [28] Bernheimer AW, Avigad LS (1970) Nature and properties of a cytolytic agent produced by Bacillus subtilis. J Gen Appl Microbiol 61: 361-369. https://doi.org/10.1099/00221287-61-3-361
    [29] Bouffioux O, Berquand A, Eeman M, et al. (2007) Molecular organization of surfactin-phospholipid monolayers: Effect of phospholipid chain length and polar head. Bba-Biomembranes 1768: 1758-1768. https://doi.org/10.1016/j.bbamem.2007.04.015
    [30] Maget-Dana R, Ptak M (1995) Interactions of surfactin with membrane models. Biophys J 68: 1937-1943. https://doi.org/10.1016/S0006-3495(95)80370-X
    [31] Carrillo C, Teruel JA, Aranda FJ, et al. (2003) Molecular mechanism of membrane permeabilization by the peptide antibiotic surfactin. Bba-Biomembranes 1611: 91-97. https://doi.org/10.1016/S0005-2736(03)00029-4
    [32] Thimon L, Peypoux F, Maget-Dana R, et al. (1992) Interactions of bioactive lipopeptides, iturin A and surfactin from Bacillus subtilis. Biotechnol Appl Bioc 16: 144-151.
    [33] Hwang YH, Park BK, Lim JH, et al. (2007) Lipopolysaccharide-binding and neutralizing activities of surfactin C in experimental models of septic shock. Eur J Pharmacol 556: 166-171. https://doi.org/10.1016/j.ejphar.2006.10.031
    [34] Hwang MH, Lim JH, Yun HI, et al. (2005) Surfactin C inhibits the lipopolysaccharide-induced transcription of interleukin-1 beta and inducible nitric oxide synthase and nitric oxide production in murine RAW 264.7 cells. Biotechnol Lett 27: 1605-1608. https://doi.org/10.1007/s10529-005-2515-1
    [35] Huang X, Wei Z, Gao X, et al. (2008) Study on antiviral activity of surfactin on Porcine parvovirus in vitro. Acta Agriculturae Zhejiangensis 20: 349-352.
    [36] Vollenbroich D, Ozel M, Vater J, et al. (1997) Mechanism of inactivation of enveloped viruses by the biosurfactant surfactin from Bacillus subtilis. Biologicals 25: 289-297. https://doi.org/10.1006/biol.1997.0099
    [37] Vollenbroich D, Pauli G, Ozel M, et al. (1997) Antimycoplasma properties and application in cell culture of surfactin, a lipopeptide antibiotic from Bacillus subtilis. Appl Environ Microb 63: 44-49. https://doi.org/10.1128/AEM.63.1.44-49.1997
    [38] Park SY, Kim JH, Lee YJ, et al. (2013) Surfactin suppresses TPA-induced breast cancer cell invasion through the inhibition of MMP-9 expression. Int J Oncol 42: 287-296. https://doi.org/10.3892/ijo.2012.1695
    [39] Peterlik M, Grant WB, Cross HS (2009) Calcium, vitamin D and cancer. Anticancer Res 29: 3687-3698.
    [40] Lee JH, Nam SH, Seo WT, et al. (2012) The production of surfactin during the fermentation of cheonggukjang by potential probiotic Bacillus subtilis CSY191 and the resultant growth suppression of MCF-7 human breast cancer cells. Food Chem 131: 1347-1354. https://doi.org/10.1016/j.foodchem.2011.09.133
    [41] Wu YS, Ngai SC, Goh BH, et al. (2017) Anticancer activities of surfactin and potential application of nanotechnology assisted surfactin delivery. Front Pharmacol 8: 761. https://doi.org/10.3389/fphar.2017.00761
    [42] Chen XY, Lu YJ, Shan MY, et al. (2022) A mini-review: mechanism of antimicrobial action and application of surfactin. World J Microb Biot 38: 143. https://doi.org/10.1007/s11274-022-03323-3
    [43] Huang XQ, Gao XP, Zheng LY, et al. (2009) Optimization of sterilization of salmonella enteritidis in meat by surfactin and iturin using a response surface method. Int J Pept Res Ther 15: 61-67. https://doi.org/10.1007/s10989-008-9164-x
    [44] Wang Y, Tian JH, Shi FF, et al. (2021) Protective effect of surfactin on copper sulfate-induced inflammation, oxidative stress, and hepatic injury in zebrafish. Microbiol Immunol 65: 410-421. https://doi.org/10.1111/1348-0421.12924
    [45] Horng YB, Yu YH, Dybus A, et al. (2019) Antibacterial activity of Bacillus species-derived surfactin on Brachyspira hyodysenteriae and Clostridium perfringens. Amb Express 9: 188. https://doi.org/10.1186/s13568-019-0914-2
    [46] Rowland I, Gibson G, Heinken A, et al. (2018) Gut microbiota functions: metabolism of nutrients and other food components. Eur J Nutr 57: 1-24. https://doi.org/10.1007/s00394-017-1445-8
    [47] Jandhyala SM, Talukdar R, Subramanyam C, et al. (2015) Role of the normal gut microbiota. World J Gastroentero 21: 8787-8803. https://doi.org/10.3748/wjg.v21.i29.8787
    [48] Yu YH, Wu CM, Chen WJ, et al. (2021) Effectiveness of Bacillus licheniformis-fermented products and their derived antimicrobial lipopeptides in controlling coccidiosis in broilers. Animals-Basel 11: 3576. https://doi.org/10.3390/ani11123576
    [49] Zhai SW, Shi QC, Chen XH (2016) Effects of dietary surfactin supplementation on growth performance, intestinal digestive enzymes activities and some serum biochemical parameters of tilapia (Oreochromis niloticus) fingerlings. Ital J Anim Sci 15: 318-324. https://doi.org/10.1080/1828051X.2016.1175325
    [50] Piewngam P, Zheng Y, Nguyen TH, et al. (2018) Pathogen elimination by probiotic Bacillus via signalling interference. Nature 562: 532. https://doi.org/10.1038/s41586-018-0616-y
    [51] Bais HP, Fall R, Vivanco JM (2004) Biocontrol of Bacillus subtilis against infection of Arabidopsis roots by Pseudomonas syringae is facilitated by biofilm formation and surfactin production. Plant Physiol 134: 307-319. https://doi.org/10.1104/pp.103.028712
    [52] Huang X, Wei Z, Gao X, et al. (2009) Study of antiviral activity of surfactin on pseudorabies virus in vitro. J Biol 26: 41-43.
    [53] Yuan LF, Zhang SA, Wang YH, et al. (2018) Surfactin inhibits membrane fusion during invasion of epithelial cells by enveloped viruses. J Virol 92. https://doi.org/10.1128/JVI.00809-18
    [54] Zhou SN, Liu G, Wu SM (2020) Marine bacterial surfactin CS30-2 induced necrosis-like cell death in Huh7.5 liver cancer cells. J Oceanol Limnol 38: 826-833. https://doi.org/10.1007/s00343-019-9129-2
    [55] Wang D, Liu Y, Yang Z, et al. (2008) Application of surfactin in microbial enhanced oil recovery. Acta Petrolei Sinica 29: 111-115.
    [56] He Z, Zhao H, Lu Z (2017) Effect of surfactin as surfactant on physical and oxidation stability of O/W DHA-rich algae oil emulsion. Food Sci 38: 146-151.
    [57] Ganesan NG, Rangarajan V (2021) A kinetics study on surfactin production from Bacillus subtilis MTCC 2415 for application in green cosmetics. Biocatal Agr Biotech 33. https://doi.org/10.1016/j.bcab.2021.102001
    [58] Whang LM, Liu P, Ma CC, et al. (2008) Application of biosurfactants, rhamnolipid, and surfactin, for enhanced biodegradation of diesel-contaminated water and soil. J Hazard Mater 151: 155-163. https://doi.org/10.1016/j.jhazmat.2007.05.063
    [59] Mulligan CN (2005) Environmental applications for biosurfactants. Environ Pollut 133: 183-198. https://doi.org/10.1016/j.envpol.2004.06.009
    [60] Chen WC, Juang RS, Wei YH (2015) Applications of a lipopeptide biosurfactant, surfactin, produced by microorganisms. Biochem Eng J 103: 158-169. https://doi.org/10.1016/j.bej.2015.07.009
    [61] Kaneda T (1977) Fatty acids of the genus Bacillus: an example of branched-chain preference. Bacteriol Rev 41: 391-418. https://doi.org/10.1128/MMBR.41.2.391-418.1977
    [62] Youssef NH, Duncan KE, McInerney MJ (2005) Importance of 3-hydroxy fatty acid composition of lipopeptides for biosurfactant activity. Appl Environ Microb 71: 7690-7695. https://doi.org/10.1128/AEM.71.12.7690-7695.2005
    [63] Ongena M, Jacques P (2008) Bacillus lipopeptides: versatile weapons for plant disease biocontrol. Trends Microbiol 16: 115-125. https://doi.org/10.1016/j.tim.2007.12.009
    [64] Kopp F, Marahiel MA (2007) Macrocyclization strategies in polyketide and nonribosomal peptide biosynthesis. Nat Prod Rep 24: 735-749. https://doi.org/10.1039/b613652b
    [65] Hamoen LW, Venema G, Kuipers OP (2003) Controlling competence in Bacillus subtilis: shared use of regulators. Microbiol-Sgm 149: 9-17. https://doi.org/10.1099/mic.0.26003-0
    [66] Nakano MM, Corbell N, Besson J, et al. (1992) Isolation and characterization of sfp: a gene that functions in the production of the lipopeptide biosurfactant, surfactin, in Bacillus subtilis. Mol Gen Genet 232: 313-321. https://doi.org/10.1007/BF00280011
    [67] Lambalot RH, Gehring AM, Flugel RS, et al. (1996) A new enzyme superfamily-the phosphopantetheinyl transferases. Chem Biol 3: 923-936. https://doi.org/10.1016/S1074-5521(96)90181-7
    [68] Hu FX, Liu YY, Li S (2019) Rational strain improvement for surfactin production: enhancing the yield and generating novel structures. Microb Cell Fact 18: 42. https://doi.org/10.1186/s12934-019-1089-x
    [69] Banat IM, Satpute SK, Cameotra SS, et al. (2014) Cost effective technologies and renewable substrates for biosurfactants' production. Front Microbiol 5: 697. https://doi.org/10.3389/fmicb.2014.00697
    [70] Zhi Y, Wu Q, Xu Y (2017) Genome and transcriptome analysis of surfactin biosynthesis in Bacillus amyloliquefaciens MT45. Sci Rep-Uk 7: 40976. https://doi.org/10.1038/srep40976
    [71] Abdel-Mawgoud AM, Aboulwafa MM, Hassouna N (2008) Optimization of surfactin production by Bacillus subtilis isolate BS5. Appl Biochem Biotech 150: 305-325. https://doi.org/10.1007/s12010-008-8155-x
    [72] Amani H, Haghighi M, Keshtkar MJ (2013) Production and optimization of microbial surfactin by Bacillus subtilis for ex situ enhanced oil recovery. Petrol Sci Technol 31: 1249-1258. https://doi.org/10.1080/10916466.2010.542416
    [73] Yang N, Wu Q, Xu Y (2020) Fermentation optimization for the production of surfactin by Bacillus amyloliquefaciens. Chinese Biotechnol 40: 51-58.
    [74] Satpute SK, Banat IM, Dhakephalkar PK, et al. (2010) Biosurfactants, bioemulsifiers and exopolysaccharides from marine microorganisms. Biotechnol Adv 28: 436-450. https://doi.org/10.1016/j.biotechadv.2010.02.006
    [75] Nazareth TC, Zanutto CP, Tripathi L, et al. (2020) The use of low-cost brewery waste product for the production of surfactin as a natural microbial biocide. Biotechnol Rep 28: e537. https://doi.org/10.1016/j.btre.2020.e00537
    [76] Ramirez IM, Vaz DA, Banat IM, et al. (2016) Hydrolysis of olive mill waste to enhance rhamnolipids and surfactin production. Bioresource Technol 205: 1-6. https://doi.org/10.1016/j.biortech.2016.01.016
    [77] Zhi Y, Wu Q, Xu Y (2017) Production of surfactin from waste distillers' grains by co-culture fermentation of two Bacillus amyloliquefaciens strains. Bioresource Technol 235: 96-103. https://doi.org/10.1016/j.biortech.2017.03.090
    [78] Willenbacher J, Mohr T, Henkel M, et al. (2016) Substitution of the native srfA promoter by constitutive P-veg in two B. subtilis strains and evaluation of the effect on Surfactin production. J Biotechnol 224: 14-17. https://doi.org/10.1016/j.jbiotec.2016.03.002
    [79] Dhali D, Coutte F, Arias AA, et al. (2017) Genetic engineering of the branched fatty acid metabolic pathway of Bacillus subtilis for the overproduction of surfactin C-14 isoform. Biotechnol J 12: 1600574. https://doi.org/10.1002/biot.201600574
    [80] Wang MM, Yu HM, Shen ZY (2019) Antisense RNA-based strategy for enhancing surfactin production in Bacillus subtilis TS1726 via overexpression of the unconventional biotin carboxylase II to enhance ACCase activity. Acs Synth Biol 8: 251-256. https://doi.org/10.1021/acssynbio.8b00459
    [81] Wang CY, Cao YX, Wang YP, et al. (2019) Enhancing surfactin production by using systematic CRISPRi repression to screen amino acid biosynthesis genes in Bacillus subtilis. Microb Cell Fact 18: 90. https://doi.org/10.1186/s12934-019-1139-4
    [82] Wang MM, Yu HM, Li X, et al. (2020) Single-gene regulated non-spore-forming Bacillus subtilis: Construction, transcriptome responses, and applications for producing enzymes and surfactin. Metab Eng 62: 235-248. https://doi.org/10.1016/j.ymben.2020.08.008
    [83] Wu Q, Zhi Y, Xu Y (2019) Systematically engineering the biosynthesis of a green biosurfactant surfactin by Bacillus subtilis 168. Metab Eng 52: 87-97. https://doi.org/10.1016/j.ymben.2018.11.004
    [84] Li X, Yang H, Zhang DL, et al. (2015) Overexpression of specific proton motive force-dependent transporters facilitate the export of surfactin in Bacillus subtilis. J Ind Microbiol Biot 42: 93-103. https://doi.org/10.1007/s10295-014-1527-z
    [85] Jiao S, Li X, Yu HM, et al. (2017) In situ enhancement of surfactin biosynthesis in Bacillus subtilis using novel artificial inducible promoters. Biotechnol Bioeng 114: 832-842. https://doi.org/10.1002/bit.26197
    [86] Atwa NA, El-Shatoury E, Elazzazy A, et al. (2013) Enhancement of surfactin production by Bacillus velezensis NRC-1 strain using a modified bench-top bioreactor. J Food Agric Environ 11: 169-174.
    [87] Jung J, Yu KO, Ramzi AB, et al. (2012) Improvement of surfactin production in Bacillus subtilis using synthetic wastewater by overexpression of specific extracellular signaling peptides, comX and phrC. Biotechnol Bioeng 109: 2349-2356. https://doi.org/10.1002/bit.24524
    [88] Zhang F, Huo KY, Song XY, et al. (2020) Engineering of a genome-reduced strain Bacillus amyloliquefaciens for enhancing surfactin production. Microb Cell Fact 19: 223. https://doi.org/10.1186/s12934-020-01485-z
    [89] Cheng JT, Guan CR, Cui WJ, et al. (2016) Enhancement of a high efficient autoinducible expression system in Bacillus subtilis by promoter engineering. Protein Expres Purif 127: 81-87. https://doi.org/10.1016/j.pep.2016.07.008
    [90] Pottathil M, Jung A, Lazazzera BA (2008) CSF, a species-specific extracellular signaling peptide for communication among strains of Bacillus subtilis and Bacillus mojavensis. J Bacteriol 190: 4095-4099. https://doi.org/10.1128/JB.00187-08
    [91] Shank EA, Kolter R (2011) Extracellular signaling and multicellularity in Bacillus subtilis. Curr Opin Microbiol 14: 741-747. https://doi.org/10.1016/j.mib.2011.09.016
    [92] Ohsawa T, Tsukahara K, Sato T, et al. (2006) Superoxide stress decreases expression of srfA through inhibition of transcription of the comQXP quorum-sensing locus in Bacillus subtilis. J Biochem 139: 203-211. https://doi.org/10.1093/jb/mvj023
    [93] Coutte F, Niehren J, Dhali D, et al. (2015) Modeling leucine's metabolic pathway and knockout prediction improving the production of surfactin, a biosurfactant from Bacillus subtilis. Biotechnol J 10: 1216-1234. https://doi.org/10.1002/biot.201400541
    [94] Hayashi K, Ohsawa T, Kobayashi K, et al. (2005) The H2O2 stress-responsive regulator PerR positively regulates srfA expression in Bacillus subtilis. J Bacteriol 187: 6659-6667. https://doi.org/10.1128/JB.187.19.6659-6667.2005
    [95] Tsuge K, Ohata Y, Shoda M (2001) Gene yerP, involved in surfactin self-resistance in Bacillus subtilis. Antimicrob Agents Ch 45: 3566-3573. https://doi.org/10.1128/AAC.45.12.3566-3573.2001
    [96] Nah HJ, Pyeon HR, Kang SH, et al. (2017) Cloning and heterologous expression of a large-sized natural product biosynthetic gene cluster in Streptomyces species. Front Microbiol 8: 394. https://doi.org/10.3389/fmicb.2017.00394
    [97] Weihmann R, Domrose A, Drepper T, et al. (2020) Protocols for yTREX/Tn5-based gene cluster expression in Pseudomonas putida. Microb Biotechnol 13: 250-262. https://doi.org/10.1111/1751-7915.13402
    [98] Yamanaka K, Reynolds KA, Kersten RD, et al. (2014) Direct cloning and refactoring of a silent lipopeptide biosynthetic gene cluster yields the antibiotic taromycin A. P Natl Acad Sci Usa 111: 1957-1962. https://doi.org/10.1073/pnas.1319584111
    [99] Lee N, Larionov V, Kouprina N (2015) Highly efficient CRISPR/Cas9-mediated TAR cloning of genes and chromosomal loci from complex genomes in yeast. Nucleic Acids Res 43: e55. https://doi.org/10.1093/nar/gkv112
    [100] Krispin O, Allmansberger R (1998) The Bacillus subtilis AraE protein displays a broad substrate specificity for several different sugars. J Bacteriol 180: 3250-3252. https://doi.org/10.1128/JB.180.12.3250-3252.1998
    [101] Hu FX, Liu YY, Lin JZ, et al. (2020) Efficient production of surfactin from xylose-rich corncob hydrolysate using genetically modified Bacillus subtilis 168. Appl Microbiol Biot 104: 4017-4026. https://doi.org/10.1007/s00253-020-10528-9
    [102] Naseri G, Koffas M (2020) Application of combinatorial optimization strategies in synthetic biology. Nat Commun 11: 2446. https://doi.org/10.1038/s41467-020-16175-y
    [103] Stephens C, Christen B, Fuchs T, et al. (2007) Genetic analysis of a novel pathway for D-xylose metabolism in Caulobacter crescentus. J Bacteriol 189: 2181-2185. https://doi.org/10.1128/JB.01438-06
    [104] Nitschke M, Pastore GM (2004) Biosurfactant production by Bacillus subtilis using cassava-processing effluent. Appl Biochem Biotech 112: 163-172. https://doi.org/10.1385/ABAB:112:3:163
  • This article has been cited by:

    1. Mohadeseh Hassanisaadi, Surfactin as a multifaceted biometabolite for sustainable plant defense: a review, 2024, 2239-7264, 10.1007/s42161-024-01645-9
    2. Sara E. Badawey, Lamia Heikal, Mohamed Teleb, Marwa Abu-Serie, Basant A. Bakr, Sherine N. Khattab, Labiba El-Khordagui, Biosurfactant-amphiphilized hyaluronic acid: A dual self-assembly anticancer nanoconjugate and drug vector for synergistic chemotherapy, 2024, 271, 01418130, 132545, 10.1016/j.ijbiomac.2024.132545
    3. Heidi Schalchli, Claudio Lamilla, Olga Rubilar, Gabriela Briceño, Felipe Gallardo, Nelson Durán, Andrés Huenchupan, María Cristina Diez, Production and characterization of a biosurfactant produced by Bacillus amyloliquefaciens C11 for enhancing the solubility of pesticides, 2023, 11, 22133437, 111572, 10.1016/j.jece.2023.111572
    4. P.L. Fernandes, E.M. Rodrigues, M.J. McInerney, M.R. Tótola, Microbial enhanced oil recovery: Use of metabolic products of Bacillus subtilis RI4914 in non-consolidated porous media and influence of environmental parameters, 2024, 697, 09277757, 134431, 10.1016/j.colsurfa.2024.134431
    5. Xiaohua Qi, Wei Liu, Xin He, Chunmei Du, A review on surfactin: molecular regulation of biosynthesis, 2023, 205, 0302-8933, 10.1007/s00203-023-03652-3
    6. Fumihiro Ishikawa, Chiharu Uchida, Genzoh Tanabe, Proteolytic Regulation in the Biosynthesis of Natural Product Via a ClpP Protease System, 2024, 19, 1554-8929, 1794, 10.1021/acschembio.4c00304
    7. Hung-Yu Shu, Chien-Chi Chen, Hsin-Tzu Ku, Chun-Lin Wang, Keh-Ming Wu, Hui-Ying Weng, Shih-Tung Liu, Chyi-Liang Chen, Cheng-Hsun Chiu, David Rasko, Complete genome sequence of Bacillus halotolerans F29-3, a fengycin-producing strain , 2024, 13, 2576-098X, 10.1128/mra.01246-23
    8. Toshiaki Taira, Ryodai Moriyama, Kenichi Sakai, Hideki Sakai, Tomohiro Imura, Solid-phase synthesis of amphiphilic cyclic peptides as a surfactin mimetic, 2024, 702, 09277757, 134989, 10.1016/j.colsurfa.2024.134989
    9. Andrea Castaldi, Bich Ngan Truong, Quyen Thi Vu, Thi Hong Minh Le, Arul Marie, Gaël Le Pennec, Florent Rouvier, Jean-Michel Brunel, Arlette Longeon, Van Cuong Pham, Thi Mai Huong Doan, Marie-Lise Bourguet-Kondracki, Computational Methods Reveal a Series of Cyclic and Linear Lichenysins and Surfactins from the Vietnamese Marine Sediment-Derived Streptomyces Strain G222, 2024, 29, 1420-3049, 1458, 10.3390/molecules29071458
    10. Karen C. L. Fung, Henrique S. Dornelles, Maria B. A. Varesche, Tony Gutierrez, From Wastewater Treatment Plants to the Oceans: A Review on Synthetic Chemical Surfactants (SCSs) and Perspectives on Marine-Safe Biosurfactants, 2023, 15, 2071-1050, 11436, 10.3390/su151411436
    11. Noelia Hernández-Ortiz, Pedro A. Sánchez-Murcia, Celia Gil-Campillo, Mirian Domenech, Daniel Lucena-Agell, Rafael Hortigüela, Sonsoles Velázquez, María José Camarasa, Noemí Bustamante, Sonia de Castro, Margarita Menéndez, Design, synthesis and structure-activity relationship (SAR) studies of an unusual class of non-cationic fatty amine-tripeptide conjugates as novel synthetic antimicrobial agents, 2024, 15, 1663-9812, 10.3389/fphar.2024.1428409
    12. Lynne Itelson, Mayan Merav, Shai Haymi, Shmuel Carmeli, Micha Ilan, Diversity and Activity of Bacteria Cultured from a Cup—The Sponge Calyx nicaeensis, 2024, 22, 1660-3397, 440, 10.3390/md22100440
    13. Salome Dini, Alaa El-Din A. Bekhit, Shahin Roohinejad, Jim M. Vale, Dominic Agyei, The Physicochemical and Functional Properties of Biosurfactants: A Review, 2024, 29, 1420-3049, 2544, 10.3390/molecules29112544
    14. Irene Russo Krauss, Rodolfo Esposito, Luigi Paduano, Gerardino D'Errico, From composite molecular structures to a multiplicity of supramolecular aggregates: The role of intermolecular interactions in biosurfactant self-assembly, 2024, 70, 13590294, 101792, 10.1016/j.cocis.2024.101792
    15. Ricardo Valencia Albornoz, Diego Oyarzún, Karl Burgess, Optimisation of surfactin yield in Bacillus using data-efficient active learning and high-throughput mass spectrometry, 2024, 23, 20010370, 1226, 10.1016/j.csbj.2024.02.012
    16. Madalena Lourenço, Noélia Duarte, Isabel A. C. Ribeiro, Exploring Biosurfactants as Antimicrobial Approaches, 2024, 17, 1424-8247, 1239, 10.3390/ph17091239
    17. Hyun Soo Kim, Jeong Won Ahn, Kongara Damodar, Jung Youl Park, Yeong-Min Yoo, Seong Soo Joo, Identification and characterization of a surfactin from Pseudomonas gessardii: A symbiotic bacterium with potent anticancer activity, 2024, 739, 0006291X, 150989, 10.1016/j.bbrc.2024.150989
    18. Isha Abhyankar, Swarali Hirlekar, Asmita Prabhune, Anuya Nisal, Bridging the gap: An investigation of biosurfactants-polymer systems, 2024, 72, 13590294, 101806, 10.1016/j.cocis.2024.101806
    19. Anmol Gupta, Fahad Khan, Pratibha Pandey, Manikant Tripathi, Neelam Pathak, A comprehensive review on the role of biosurfactants in remediation of heavy metals from contaminated environment, 2024, 1088-9868, 1, 10.1080/10889868.2024.2427076
    20. Anna Zdziennicka, Bronisław Jańczuk, Adsorption and wetting properties of biosurfactants, trtions and their mixtures in aqueous and water-ethanol environment, 2024, 00018686, 103379, 10.1016/j.cis.2024.103379
    21. Rita de Cássia Barreto Silva-Portela, Carolina Fonseca Minnicelli, Júlia Firme Freitas, Marbella Maria Bernardes Fonseca, Douglas Felipe de Lima Silva, Kamila Karla Silva-Barbalho, Raul Maia Falcão, Thiago Bruce, João Vitor Ferreira Cavalcante, Rodrigo Juliani Siqueira Dalmolin, Lucymara Fassarella Agnez-Lima, Unlocking the transcriptional profiles of an oily waste-degrading bacterial consortium, 2025, 485, 03043894, 136866, 10.1016/j.jhazmat.2024.136866
    22. Alexey Mikhailovich Neurov, Anna Andreevna Zaikina, Evgeniya Valer’evna Prazdnova, Ranjan Anuj, Dmitriy Vladimirovich Rudoy, Modulation of Stress-Related Protein in the African Catfish (Clarias gariepinus) Using Bacillus-Based Non-Ribosomal Peptides, 2024, 15, 2036-7481, 2743, 10.3390/microbiolres15040182
    23. Lívia Vieira Araujo de Castilho, Carolina Reis Guimarães, Lucy Seldin, Márcia Nitschke, Denise Maria Guimarães Freire, Pseudomonas fluorescens and Listeria monocytogenes Planktonic Cells and Biofilms Are Inhibited by Surfactin from Bacillus velezensis H2O-1, 2024, 13, 2227-9717, 18, 10.3390/pr13010018
    24. Jiaqi Zhang, Xinmiao Yang, Jiajia Qiu, Wen Zhang, Jie Yang, Jinzhi Han, Li Ni, The Characterization, Biological Activities, and Potential Applications of the Antimicrobial Peptides Derived from Bacillus spp.: A Comprehensive Review, 2024, 1867-1306, 10.1007/s12602-024-10447-5
    25. Nadin Darwiche, Christelle Dufresne, Agnès Chartier, Bérengère Claude, Cyril Colas, Laëtitia Fougère, Muriel Sebban, Marie-Elisabeth Lucchesi, Stéphane Le Floch, Reine Nehmé, Glycolipid and Lipopeptide Biosurfactants: Structural Classes and Characterization—Rhamnolipids as a Model, 2024, 1040-8347, 1, 10.1080/10408347.2024.2441428
    26. Natalia Markelova, Anastasia Chumak, Antimicrobial Activity of Bacillus Cyclic Lipopeptides and Their Role in the Host Adaptive Response to Changes in Environmental Conditions, 2025, 26, 1422-0067, 336, 10.3390/ijms26010336
    27. Vincenza Casella, Gerardo Della Sala, Silvia Scarpato, Carmine Buonocore, Costanza Ragozzino, Pietro Tedesco, Daniela Coppola, Giovanni Andrea Vitale, Donatella de Pascale, Fortunato Palma Esposito, Novel Insights into the Nobilamide Family from a Deep-Sea Bacillus: Chemical Diversity, Biosynthesis and Antimicrobial Activity Towards Multidrug-Resistant Bacteria, 2025, 23, 1660-3397, 41, 10.3390/md23010041
    28. Caja Dinesen, Manca Vertot, Scott A Jarmusch, Carlos N Lozano-Andrade, Aaron J C Andersen, Ákos T Kovács, Subtilosin A production is influenced by surfactin levels in Bacillus subtilis , 2025, 6, 2633-6693, 10.1093/femsml/uqae029
    29. Zhengjun Pang, Wenshuo Zhang, Bo Zhang, Shiza Nawaz, Fenghuan Wang, Yonghong Liao, Biosynthesis and modification strategies of novel cyclic lipopeptide secreted by Bacillus spp.: research progress, 2025, 13595113, 10.1016/j.procbio.2025.01.023
    30. Taise B. Martins, Janaina Debon, Willibaldo Schmidell, Hugo M. Soares, Effect of oxygen mass transfer coefficient (kLa) on surfactin production by Bacillus subtilis ATCC 21332, 2025, 1678-4383, 10.1007/s43153-024-00528-x
    31. Alexander Hermann, Eric Hiller, Philipp Hubel, Lennart Biermann, Elvio Henrique Benatto Perino, Oscar Paul Kuipers, Rudolf Hausmann, Lars Lilge, Genetic Code Expansion for Controlled Surfactin Production in a High Cell-Density Bacillus subtilis Strain, 2025, 13, 2076-2607, 353, 10.3390/microorganisms13020353
    32. Gloria Romero Vega, Paola Gallo Stampino, Bio-Based Surfactants and Biosurfactants: An Overview and Main Characteristics, 2025, 30, 1420-3049, 863, 10.3390/molecules30040863
    33. Syeda Amna Farooq, Shazia Khaliq, Saeed Ahmad, Neelma Ashraf, Muhammad Afzal Ghauri, Munir Ahmad Anwar, Kalsoom Akhtar, Jean Francois Picimbon, Application of Combined Irradiation Mutagenesis Technique for Hyperproduction of Surfactin in Bacillus velezensis Strain AF_3B, 2025, 2025, 1687-918X, 10.1155/ijm/5570585
    34. Brianne R. Zbylicki, Sierra Cochran, David S. Weiss, Craig D. Ellermeier, Jimmy D. Ballard, Identification of two glycosyltransferases required for synthesis of membrane glycolipids in Clostridioides difficile , 2025, 2150-7511, 10.1128/mbio.03512-24
    35. Barbara B. Gerbelli, Pedro T. Sodré, Pedro L.O. Filho, Mauricio D. Coutinho-Neto, Ian W. Hamley, Jani Seitsonen, Wendel A. Alves, Enhancing pesticide detection: The role of serine in lipopeptide nanostructures and their self-assembly dynamics, 2025, 00219797, 137271, 10.1016/j.jcis.2025.137271
    36. Hongwei Yu, Guangdong Zhou, Wenlong Zhang, Bonaventure Aman Omondi, Alberto Cenci, Kunhua Liu, Juhua Liu, Huacai Fan, Shu Li, Mathieu Rouard, Si-Jun Zheng, Bacillus velezensis YN2111 reveals the potential role of Amine Oxidase in Fusarium Wilt response in Cavendish bananas, 2025, 08855765, 102653, 10.1016/j.pmpp.2025.102653
    37. Afsana Habib Jui, Md Nur Hossain, Sadia Afrin, Banasree Bhowmik, Chaminda Senaka Ranadheera, Md. Habibur Rahman Bhuiyan, Microbial Biosurfactants: Prospect and Challenges for Application in Food Industry, 2025, 8755-9129, 1, 10.1080/87559129.2025.2478199
    38. Eduarda Guimarães Sousa, Gabriela Munis Campos, Ludmila Silva Quaresma, Thaís Fernandes Mendonça Mota, Nédia de Castilhos Ghisi, Gabriel Camargos Gomes, Rhayane Cristina Viegas Santos, Beatriz Gamarano Rocha de Souza, Éric Guédon, Siomar de Castro Soares, Joyce da Cruz Ferraz Dutra, Vasco Ariston de Carvalho Azevedo, Exploring Bacillus velezensis in a biomedical context: a systematic review, 2025, 2, 3064-9765, 10.20935/AcadMolBioGen7598
    39. Deepak A. Yaraguppi, Prasanth DSNBK, Sangameshwar G. Halkavatagi, A Review on Lipopeptide Production, Structural Characterization, and Its Applications, 2025, 1550-9087, 10.1089/ind.2024.0055
    40. V. S. Trefilov, E. Yu. Lindin, M. V. Monakhova, O. V. Kisil, M. B. Viryasov, T. S. Oretskaya, E. A. Kubareva, Instrumental Approaches to the Detection and Quantification of Surfactin, 2025, 51, 1068-1620, 465, 10.1134/S1068162024606517
    41. George Michael T. Nicolas, Secondary Metabolites from Bacillus spp. Probiotics as Potential Treatments for Multidrug-Resistant Pathogens: A Comprehensive Review, 2025, 26665174, 100392, 10.1016/j.crmicr.2025.100392
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4692) PDF downloads(361) Cited by(41)

Figures and Tables

Figures(3)  /  Tables(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog