首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The rising existence of antimicrobial resistance, confirms the urgent need for new antimicrobial compounds. Lantibiotics are active in a low nanomolar range and represent good compound candidates. The lantibiotic nisin is well studied, thus it is a perfect origin for exploring novel lantibiotics via mutagenesis studies. However, some human pathogens like Streptococcus agalactiae COH1 already express resistance proteins against lantibiotics like nisin.This study presents three nisin variants with mutations in the hinge-region and determine their influence on both the growth inhibition as well as the pore-forming activity. Furthermore, we analyzed the effect of these mutants on the nisin immunity proteins NisI and NisFEG from Lactococcus lactis, as well as the nisin resistance proteins SaNSR and SaNsrFP from Streptococcus agalactiae COH1.We identified the nisin variant 20NMKIV24 with an extended hinge-region, to be an excellent candidate for further studies to eventually overcome the lantibiotic resistance in human pathogens, since these proteins do not recognize this variant well.  相似文献   

2.
Lantibiotics are antimicrobial peptides produced by Gram-positive bacteria and active in the nanomolar range. Nisin is the most intensely studied and used lantibiotic, with applications as food preservative and recognized potential for clinical usage. However, different bacteria that are pathogenic for humans and do not produce nisin, including Streptococcus agalactiae, show an innate resistance that has been related to the nisin resistance protein (NSR), a membrane-associated protease. Here, we report the first-in-class small-molecule inhibitors of SaNSR identified by virtual screening based on a previously derived structural model of the nisin/NSR complex. The inhibitors belong to three different chemotypes, of which the halogenated phenyl-urea derivative NPG9 is the most potent one. Co-administration of NPG9 with nisin yields increased potency compared to nisin alone in SaNSR-expressing bacteria. The binding mode of NPG9, predicted with molecular docking and validated by extensive molecular dynamics simulations, confirms a structure-activity relationship derived from the in vivo data. Saturation transfer difference-NMR experiments demonstrate direct binding of NPG9 to SaNSR and agree with the predicted binding mode. Our results demonstrate the potential to overcome SaNSR-related lantibiotic resistance by small molecules.  相似文献   

3.
Many Gram-positive bacteria produce lantibiotics, genetically encoded and posttranslationally modified peptide antibiotics, which inhibit the growth of other Gram-positive bacteria. To protect themselves against their own lantibiotics these bacteria express a variety of immunity proteins including the LanI lipoproteins. The structural and mechanistic basis for LanI-mediated lantibiotic immunity is not yet understood. Lactococcus lactis produces the lantibiotic nisin, which is widely used as a food preservative. Its LanI protein NisI provides immunity against nisin but not against structurally very similar lantibiotics from other species such as subtilin from Bacillus subtilis. To understand the structural basis for LanI-mediated immunity and their specificity we investigated the structure of NisI. We found that NisI is a two-domain protein. Surprisingly, each of the two NisI domains has the same structure as the LanI protein from B. subtilis, SpaI, despite the lack of significant sequence homology. The two NisI domains and SpaI differ strongly in their surface properties and function. Additionally, SpaI-mediated lantibiotic immunity depends on the presence of a basic unstructured N-terminal region that tethers SpaI to the membrane. Such a region is absent from NisI. Instead, the N-terminal domain of NisI interacts with membranes but not with nisin. In contrast, the C-terminal domain specifically binds nisin and modulates the membrane affinity of the N-terminal domain. Thus, our results reveal an unexpected structural relationship between NisI and SpaI and shed light on the structural basis for LanI mediated lantibiotic immunity.  相似文献   

4.
Nisin, a 3.4 kDa antimicrobial peptide produced by some Lactococcus lactis strains is the most prominent member of the lantibiotic family. Nisin can inhibit cell growth and penetrates the target Gram-positive bacterial membrane by binding to Lipid II, an essential cell wall synthesis precursor. The assembled nisin-Lipid II complex forms pores in the target membrane. To gain immunity against its own-produced nisin, Lactococcus lactis is expressing two immunity protein systems, NisI and NisFEG. Here, we show that the NisI expressing strain displays an IC50 of 73±10 nM, an 8–10-fold increase when compared to the non-expressing sensitive strain. When the nisin concentration is raised above 70 nM, the cells expressing full-length NisI stop growing rather than being killed. NisI is inhibiting nisin mediated pore formation, even at nisin concentrations up to 1 µM. This effect is induced by the C-terminus of NisI that protects Lipid II. Its deletion showed pore formation again. The expression of NisI in combination with externally added nisin mediates an elongation of the chain length of the Lactococcus lactis cocci. While the sensitive strain cell-chains consist mainly of two cells, the NisI expressing cells display a length of up to 20 cells. Both results shed light on the immunity of lantibiotic producer strains, and their survival in high levels of their own lantibiotic in the habitat.  相似文献   

5.
The biosynthesis of the lantibiotics subtilin and nisin is regulated by autoinduction via two-component systems. Although subtilin is structurally closely related to nisin and contains the same lanthionine ring structure, both lantibiotics specifically autoinduce their biosynthesis. Subtilin and also the subtilin-like lantibiotics entianin and ericin autoinduce the two-component system SpaRK of Bacillus subtilis, whereas the biosynthesis of nisin is autoinduced via the two-component system NisRK of Lactococcus lactis. Autoinduction is highly specific for the respective lantibiotic and therefore of major importance for the functional expression of genetically engineered subtilin-like lantibiotics. To identify the structural features required for subtilin autoinduction, subtilin-nisin hybrids and specific point mutations of amino acid position 1 were generated. For subtilin autoinduction, the N-terminal tryptophan is the most important for full SpaK activation. The failure of subtilin to autoinduce the histidine kinase NisK mainly depends on the N-terminal tryptophan, as its single exchange to the aliphatic amino acid residues isoleucine, leucine, and valine provided NisK autoinduction. In addition, the production of subtilin variants which did not autoinduce their own biosynthesis could be rescued upon heterologous coexpression in B. subtilis DSM15029 by the autoinducing subtilin-like lantibiotic entianin.  相似文献   

6.
We analyzed the mode of action of the lantibiotic plantaricin C (PlnC), produced by Lactobacillus plantarum LL441. Compared to the well-characterized type A lantibiotic nisin and type B lantibiotic mersacidin, which are both able to interact with the cell wall precursor lipid II, PlnC displays structural features of both prototypes. In this regard, we found that lipid II plays a key role in the antimicrobial activity of PlnC besides that of pore formation. The pore forming activity of PlnC in whole cells was prevented by shielding lipid II on the cell surface. However, in contrast to nisin, PlnC was not able to permeabilize Lactococcus lactis cells or to form pores in 1,2-dioleoyl-sn-glycero-3-phosphocholine liposomes supplemented with 0.1 mol% purified lipid II. This emphasized the different requirements of these lantibiotics for pore formation. Using cell wall synthesis assays, we identified PlnC as a potent inhibitor of (i) lipid II synthesis and (ii) the FemX reaction, i.e., the addition of the first Gly to the pentapeptide side chain of lipid II. As revealed by thin-layer chromatography, both reactions were clearly blocked by the formation of a PlnC-lipid I and/or PlnC-lipid II complex. On the basis of the in vivo and in vitro activities of PlnC shown in this study and the structural lipid II binding motifs described for other lantibiotics, the specific interaction of PlnC with lipid II is discussed.  相似文献   

7.
While nisin (lantibiotic), lacticin 3147 (lantibiotic) and vancomycin (glycopeptides) are among the best studied lipid II-binding antimicrobials, their relative activities have never been compared. Nisin and lacticin 3147 have been employed/investigated primarily as food preservatives, although they do have potential in terms of veterinary and clinical applications. Vancomycin is used exclusively in clinical therapy. We reveal a higher potency for lacticin 3147 (MIC 0.95?C3.8???g/ml) and vancomycin (MIC 0.78?C1.56???g/ml) relative to that of nisin (MIC 6.28?C25.14???g/ml) against the food-borne pathogen Listeria monocytogenes. A comparison of the activity of the three antimicrobials against nisin resistance mutants of L. monocytogenes also reveals that their susceptibility to vancomycin and lacticin 3147 changed only slightly or not at all. A further assessment of relative activity against a selection of Bacillus cereus, Enterococcus and Staphylococcus aureus targets revealed that vancomycin MICs consistently ranged between 0.78 and 1.56???g/ml against all but one strain. Lacticin 3147 was found to be more effective than nisin against B. cereus (lacticin 3147 MIC 1.9?C3.8???g/ml; nisin MIC 4.1?C16.7???g/ml) and E. faecium and E. faecalis targets (lacticin 3147 MIC from 1.9 to 3.8???g/ml; nisin MIC ??8.3???g/ml). The greater effectiveness of lacticin 3147 is even more impressive when expressed as molar values. However, in agreement with the previous reports, nisin was the more effective of the two lantibiotics against S. aureus strains. This study highlights that in many instances the antimicrobial activity of these leading lantibiotics are comparable with that of vancomycin and emphasizes their particular value with respect to use in situations including foods and veterinary medicine, where the use of vancomycin is not permitted.  相似文献   

8.
In Bacillus subtilis, the extracytoplasmic function (ECF) σ factors σM, σW and σX all contribute to resistance against lantibiotics. Nisin, a model lantibiotic, has a dual mode of action: it inhibits cell wall synthesis by binding lipid II, and this complex also forms pores in the cytoplasmic membrane. These activities can be separated in a nisin hinge‐region variant (N20P M21P) that binds lipid II, but no longer permeabilizes membranes. The major contribution of σM to nisin resistance is expression of ltaSa, encoding a stress‐activated lipoteichoic acid synthase, and σX functions primarily by activation of the dlt operon controlling d ‐alanylation of teichoic acids. Together, σM and σX regulate cell envelope structure to decrease access of nisin to its lipid II target. In contrast, σW is principally involved in protection against membrane permeabilization as it provides little protection against the nisin hinge region variant. σW contributes to nisin resistance by regulation of a signal peptide peptidase (SppA), phage shock proteins (PspA and YvlC, a PspC homologue) and tellurite resistance related proteins (YceGHI). These defensive mechanisms are also effective against other lantibiotics such as mersacidin, gallidermin and subtilin and comprise an important subset of the intrinsic antibiotic resistome of B. subtilis.  相似文献   

9.
Streptococcus mutans, a dental pathogen, secretes different kinds of lantibiotic and nonlantibiotic bacteriocins. For self-protection, a bacteriocin producer strain must possess one or more cognate immunity mechanisms. We report here the identification of one such immunity complex in S. mutans strain GS-5 that confers protection against Smb, a two-component lantibiotic. The immunity complex that we identified is an ABC transporter composed of two proteins: SmbF (the ATPase component) and SmbT (the permease component). Both of the protein-encoding genes are located within the smb locus. We show that GS-5 becomes sensitized to Smb upon deletion of smbT, which makes the ABC transporter nonfunctional. To establish the role SmbFT in providing immunity, we heterologously expressed this ABC transporter complex in four different sensitive streptococcal species and demonstrated that it can confer resistance against Smb. To explore the specificity of SmbFT in conferring resistance, we tested mutacin IV (a nonlantibiotic), nisin (a single peptide lantibiotics), and three peptide antibiotics (bacitracin, polymyxin B, and vancomycin). We found that SmbFT does not recognize these structurally different peptides. We then tested whether SmbFT can confer protection against haloduracin, another two-component lantibiotic that is structurally similar to Smb; SmbFT indeed conferred protection against haloduracin. SmbFT can also confer protection against an uncharacterized but structurally similar lantibiotic produced by Streptococcus gallolyticus. Our data suggest that SmbFT truly displays immunity function and confer protection against Smb and structurally similar lantibiotics.  相似文献   

10.
Many lantibiotics use the membrane bound cell wall precursor Lipid II as a specific target for killing Gram-positive bacteria. Binding of Lipid II usually impedes cell wall biosynthesis, however, some elongated lantibiotics such as nisin, use Lipid II also as a docking molecule for pore formation in bacterial membranes. Although the unique nisin pore formation can be analyzed in Lipid II-doped vesicles, mechanistic details remain elusive. We used optical sectioning microscopy to directly visualize the interaction of fluorescently labeled nisin with membranes of giant unilamellar vesicles containing Lipid II and its various bactoprenol precursors. We quantitatively analyzed the binding and permeation capacity of nisin when applied at nanomolar concentrations. Specific interactions with Lipid I, Lipid II and bactoprenol-diphosphate (C55-PP), but not bactoprenol-phosphate (C55-P), resulted in the formation of large molecular aggregates. For Lipid II, we demonstrated the presence of both nisin and Lipid II in these aggregates. Membrane permeation induced by nisin was observed in the presence of Lipid I and Lipid II, but not in the presence of C55-PP. Notably, the size of the C55-PP–nisin aggregates was significantly smaller than that of the aggregates formed with Lipid I and Lipid II. We conclude that the membrane permeation capacity of nisin is determined by the size of the bactoprenol-containing aggregates in the membrane. Notably, transmitted light images indicated that the formation of large aggregates led to a pinch-off of small vesicles, a mechanism, which probably limits the growth of aggregates and induces membrane leakage.  相似文献   

11.
The biosynthesis of the lantibiotic subtilin is autoinduced in a quorum-sensing mechanism via histidine kinase SpaK. Subtilin-like lantibiotics, such as entianin, ericin S, and subtilin, specifically activated SpaK in a comparable manner, whereas the structurally similar nisin did not provide the signal for SpaK activation at nontoxic concentrations. Surprisingly, nevertheless, nisin if applied together with entianin partly quenched SpaK activation. The N-terminal entianin1–20 fragment (comprising N-terminal amino acids 1 to 20) was sufficient for SpaK activation, although higher concentrations were needed. The N-terminal nisin1–20 fragment also interfered with entianin-mediated activation of SpaK and, remarkably, at extremely high concentrations also activated SpaK. Our data show that the N-terminal entianin1–20 fragment is sufficient for SpaK activation. However, if present, the C-terminal part of the molecule further strongly enhances the activation, possibly by its interference with the cellular membrane. As shown by using lipid II-interfering substances and a lipid II-deficient mutant strain, lipid II is not needed for the sensing mechanism.  相似文献   

12.
Several peptide antibiotics have been described as potent inhibitors of bacterial growth. With respect to their biosynthesis, they can be devided into two classes: (i) those that are synthesized by a non-ribosomal mechanism and (ii) those that are ribosomally synthesized. Subtilin and nisin belong to the ribosomally synthesized peptide antibiotics. They contain the rare amino acids dehydroalanine, dehydrobutyrine, meso-lanthionine, and 3-methyl-lanthionine. They are derived from prepeptides which are post-translationally modiffied and have been termed lantibiotics because of their characteristic lanthionine bridges (Schnell et al. 1988). Nisin is the most prominent lantibiotic and is used as a food preservative due to its high potency against certain gram-positive bacteria (Mattick & Hirsch 1944, 1947; Rayman & Hurst 1984). It is produced by Lactococcus lactis strains belonging to serological group N. The potent bactericidal activities of nisin and other lantibiotics are based on depolarization of energized bacterial cytoplasmic membranes. Breakdown of the membrane potential is initiated by the formation of pores through which molecules of low molecular weight are released. A trans-negative membrane potential of 50 to 100 mV is necessary for pore formation by nisin (Ruhr & Sahl 1985; Sahl et al. 1987). Nisin occurs as a partially amphiphilic molecule (Van de Ven et al. 1991). Apart from the detergent-like effect of nisin on cytoplasmic membranes, an inhibition of murein synthesis has also been discussed as the primary effect (Reisinger et al. 1980). In several countries nisin is used to prevent the growth of clostridia in cheese and canned food. The nisin peptide structure was first described by Gross & Morall (1971), and its structural gene was isolated in 1988 (Buchman et al. 1988; Kaletta & Entian 1989). Nisin has two natural variants, nisin A and nisin Z, which differ in a single amino acid residue at position 27 (histidin in nisin A is replaced by asparagin in nisin Z (Mulders et al. 1991; De Vos et al. 1993). Subtilin is produced by Bacillus subtilis ATCC 6633. Its chemical structure was first unravelled by Gross & Kiltz (1973) and its structural gene was isolated in 1988 (Banerjee & Hansen 1988). Subtilin shares strong similarities to nisin with an identical organization of the lanthionine ring structures (Fig. 1), and both lantibiotics possess similar antibiotic activities. Due to its easy genetic analysis B. subtilis became a very suitable model organism for the identification and characterization of genes and proteins involved in lantibiotic biosynthesis. The pathway by which nisin is produced is very similar to that of subtilin, and the proteins involved share significant homologies over the entire proteins (for review see also De Vos et al. 1995b). The respective genes have been identified adjacent to the structural genes, and are organized in operon-like structures (Fig. 2). These genes are responsible for post-translational modification, transport of the modified prepeptide, proteolytic cleavage, and immunity which prevents toxic effects on the producing bacterium. In addition to this, biosynthesis of subtilin and nisin is strongly regulated by a two-component regulatory system which consists of a histidin kinase and a response regulator protein.  相似文献   

13.
Streptococcus uberis is one of the principal causative agents of bovine mastitis. In this study, we report that S. uberis strain 42 produces a lantibiotic, nisin U, which is 78% identical (82% similar) to nisin A from Lactococcus lactis. The 15.6-kb nisin U locus comprises 11 open reading frames, similar in putative functionality but differing in arrangement from that of the nisin A biosynthetic cluster. The nisin U producer strain exhibits specific resistance (immunity) to nisin U and cross-resistance to nisin A, a finding consistent with the 55% sequence similarity of their respective immunity peptides. Homologues of the nisin U structural gene were identified in several additional S. uberis strains, and in each case cross-protective immunity was expressed to nisin A and to the other producers of nisin U and its variants. To our knowledge, this is the first report both of characterization of a bacteriocin by S. uberis, as well as of a member of the nisin family of peptides in a species other than L. lactis.  相似文献   

14.
While the potential emergence of food spoilage and pathogenic bacteria with resistance to lantibiotics is a concern, the creation of derivatives of starter cultures and adjuncts that can grow in the presence of these antimicrobials may have applications in food fermentations. Here a bank of Lactococcus lactis IL1403 mutants was created and screened, and a number of novel genetic loci involved in lantibiotic resistance were identified.  相似文献   

15.
16.
Lacticin 3147 is a broad-spectrum bacteriocin produced by Lactococcus lactis subsp. lactis DPC3147 which is bactericidal against a range of mastitis-causing streptococci and staphylococci. In this study, both lacticin 3147 and the lantibiotic nisin were separately incorporated into an intramammary teat seal product. The seal containing lacticin 3147 exhibited excellent antimicrobial activity and might form the basis of an improved treatment for the prevention of mastitis in dry cows.  相似文献   

17.
Several Lactococcus lactis strains produce the lantibiotic nisin. The dedicated enzymes NisB and NisC and the transporter NisT modify and secrete the ribosomally synthesized nisin precursor peptide. NisB can function in the absence of the cyclase NisC, yielding the dehydrated prenisin that lacks the thioether rings. A kinetic analysis of nisin production by L. lactis NZ9700 demonstrated that the prenisin was released from the cell into the medium before the processing of the leader sequence occurred. Upon the deletion of nisC, the production of prenisin was reduced by 70%, while in the absence of nisB, the production of prenisin was nearly completely abolished. In cells lacking nisT, no secretion was observed, while the expression of nisABC in these cells resulted in considerable growth rate inhibition caused by the intracellular accumulation of active nisin. Overall, these data indicate that the efficiency of prenisin transport by NisT is markedly enhanced by NisB, suggesting a channeling mechanism of prenisin transfer between the nisin modification enzymes and the transporter.  相似文献   

18.
Streptococcus mutans strain GS-5 produces a two-peptide lantibiotic, Smb, which displays inhibitory activity against a broad spectrum of bacteria, including other streptococci. For inhibition, lantibiotics must recognize specific receptor molecules present on the sensitive bacterial cells. However, so far no such receptor proteins have been identified for any lantibiotics. In this study, using a powerful transposon mutagenesis approach, we have identified in Streptococcus pyogenes a gene that exhibits a receptor-like function for Smb. The protein encoded by that gene, which we named LsrS, is a membrane protein belonging to the CAAX protease family. We also found that nisin, a monopeptide lantibiotic, requires LsrS for its optimum inhibitory activity. However, we found that LsrS is not required for inhibition by haloduracin and galolacticin, both of which are two-peptide lantibiotics closely related to Smb. LsrS appears to be a well-conserved protein that is present in many streptococci, including S. mutans. Inactivation of SMU.662, an LsrS homolog, in S. mutans strains UA159 and V403 rendered the cells refractory to Smb-mediated killing. Furthermore, overexpression of LsrS in S. mutans created cells more susceptible to Smb. Although LsrS and its homolog contain the CAAX protease domain, we demonstrate that inactivation of the putative active sites on the LsrS protein has no effect on its receptor-like function. This is the first report describing a highly conserved membrane protein that displays a receptor-like function for lantibiotics.  相似文献   

19.
The thioether rings in the lantibiotics lacticin 3147 and nisin are posttranslationally introduced by dehydration of serines and threonines, followed by coupling of these dehydrated residues to cysteines. The prepeptides of the two-component lantibiotic lacticin 3147, LtnA1 and LtnA2, are dehydrated and cyclized by two corresponding bifunctional enzymes, LtnM1 and LtnM2, and are subsequently processed and exported via one bifunctional enzyme, LtnT. In the nisin synthetase complex, the enzymes NisB, NisC, NisT, and NisP dehydrate, cyclize, export, and process prenisin, respectively. Here, we demonstrate that the combination of LtnM2 and LtnT can modify, process, and transport peptides entirely different from LtnA2 and that LtnT can process and transport unmodified LtnA2 and unrelated peptides. Furthermore, we demonstrate a higher extent of NisB-mediated dehydration in the absence of thioether rings. Thioether rings apparently inhibited dehydration, which implies alternating actions of NisB and NisC. Furthermore, certain (but not all) NisC-cyclized peptides were exported with higher efficiency as a result of their conformation. Taken together, these data provide further insight into the applicability of Lactococcus lactis strains containing lantibiotic enzymes for the design and production of modified peptides.  相似文献   

20.
Nisin is a pentacyclic peptide antibiotic produced by some Lactococcus lactis strains. Nisin contains dehydroresidues and thioether rings that are posttranslationally introduced by a membrane-associated enzyme complex, composed of a serine and threonine dehydratase NisB and the cyclase NisC. In addition, the transporter NisT is necessary for export of the modified peptide. We studied the potential of L. lactis expressing NisB and NisT to produce peptides whose serines and threonines are dehydrated. L. lactis containing the nisBT genes and a plasmid coding for a specific leader peptide fusion construct efficiently produced peptides with a series of non-naturally occurring multiple flanking dehydrobutyrines. We demonstrated NisB-mediated dehydration of serines and threonines in a C-terminal nisin(1-14) extension of nisin, which implies that also residues more distant from the leader peptide than those occurring in prenisin or any other lantibiotic can be modified. Furthermore, the feasibility and efficiency of generating a library of peptides containing dehydroresidues were demonstrated. In view of the particular shape and reactivity of dehydroamino acids, such a library provides a novel source for screening for peptides with desired biological and physicochemical properties.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号