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 93 毫秒
1.
Y Chen  S Miyata  S Makino    R Moriyama 《Journal of bacteriology》1997,179(10):3181-3187
The exudate of fully germinated spores of Clostridium perfringens S40 in 0.15 M KCI-50 mM potassium phosphate (pH 7.0) was found to contain another spore-lytic enzyme in addition to the germination-specific amidase previously characterized (S. Miyata, R. Moriyama, N. Miyahara, and S. Makino, Microbiology 141:2643-2650, 1995). The lytic enzyme was purified to homogeneity by anion-exchange chromatography and shown to be a muramidase which requires divalent cations (Ca2+, Mg2+, or Mn2+) for its activity. The enzyme was inactivated by sulfhydryl reagents, and sodium thioglycolate reversed the inactivation by Hg2+. The muramidase hydrolyzed isolated spore cortical fragments from a variety of wild-type organisms but had minimal activity on decoated spores and isolated cell walls. However, the enzyme was not capable of digesting isolated cortical fragments from spores of Bacillus subtilis ADD1, which lacks muramic acid delta-lactam in its cortical peptidoglycan. This indicates that the enzyme recognizes the delta-lactam residue peculiar to spore peptidoglycan, suggesting an involvement of the enzyme in spore germination. Immunochemical studies indicated that the muramidase in its mature form is localized on the exterior of the cortex layer in the dormant spore. A gene encoding the muramidase, sleM, was cloned into Escherichia coli, and the nucleotide sequence was determined. The gene encoded a protein of 321 amino acids with a deduced molecular weight of 36,358. The deduced amino acid sequence of the sleM gene indicated that the enzyme is produced in a mature form. It was suggested that the muramidase belongs to a separate group within the lysozyme family typified by the fungus Chalaropsis lysozyme. A possible mechanism for cortex degradation in C. perfringens S40 spores is discussed.  相似文献   

2.
As part of an effort to develop detectors for selected species of bacterial spores, we screened phage display peptide libraries for 7- and 12-mer peptides that bind tightly to spores of Bacillus subtilis. All of the peptides isolated contained the sequence Asn-His-Phe-Leu at the amino terminus and exhibited clear preferences for other amino acids, especially Pro, at positions 5 to 7. We demonstrated that the sequence Asn-His-Phe-Leu-Pro (but not Asn-His-Phe-Leu) was sufficient for tight spore binding. We observed equal 7-mer peptide binding to spores of B. subtilis and its most closely related species, Bacillus amyloliquefaciens, and slightly weaker binding to spores of the closely related species Bacillus globigii. These three species comprise one branch on the Bacillus phylogenetic tree. We did not detect peptide binding to spores of several Bacillus species located on adjacent and nearby branches of the phylogenetic tree nor to vegetative cells of B. subtilis. The sequence Asn-His-Phe-Leu-Pro was used to identify B. subtilis proteins that may employ this peptide for docking to the outer surface of the forespore during spore coat assembly and/or maturation. One such protein, SpsC, appears to be involved in the synthesis of polysaccharide on the spore coat. SpsC contains the Asn-His-Phe-Leu-Pro sequence at positions 6 to 10, and the first five residues of SpsC apparently must be removed to allow spore binding. Finally, we discuss the use of peptide ligands for bacterial detection and the use of short peptide sequences for targeting proteins during spore formation.  相似文献   

3.
Bacillus anthracis spores, the infectious agents of anthrax, are notoriously difficult to remove from contaminated areas because they are resistant to many eradication methods. These resistance properties are due to the spore's dehydration and dormancy and to the multiple protective layers surrounding the spore core, one of which is the cortex. In order for B. anthracis spores to germinate and resume growth, the cortex peptidoglycan must be depolymerized. This study reports on analyses of sleL (yaaH), which encodes a cortex-lytic enzyme. The inactivation of sleL does not affect vegetative growth, spore viability, or the initial stages of germination, including dipicolinic acid release. However, mutant spores exhibit a slight delay in the loss of optical density compared to that of wild-type spores. Mutants also retain more diaminopimelic acid and N-acetylmuramic acid during germination than wild-type spores, suggesting that the cortex peptidoglycan is not being hydrolyzed as rapidly. This finding is supported by high-pressure liquid chromatography analysis of the peptidoglycan structure used to confirm that SleL acts as an N-acetylglucosaminidase. When sleL is inactivated, the cortex peptidoglycan is not depolymerized into small muropeptides but instead is retained within the spore as large fragments. In the absence of the sleL-encoded N-acetylglucosaminidase, other cortex-lytic enzymes break down the cortex peptidoglycan sufficiently to allow rapid germination and outgrowth.  相似文献   

4.
Upon UV irradiation, Bacillus subtilis spore DNA accumulates the novel thymine dimer 5-thyminyl-5,6-dihydrothymine. Spores can repair this "spore photoproduct" (SP) upon germination either by the uvr-mediated general excision repair pathway or by the SP-specific spl pathway, which involves in situ monomerization of SP to two thymines by an enzyme named SP lyase. Mutants lacking both repair pathways produce spores that are extremely sensitive to UV. For cloning DNA that can repair a mutation in the spl pathway called spl-1, a library of EcoRI fragments of chromosomal DNA from B. subtilis 168 was constructed in integrative plasmid pJH101 and introduced by transformation into a mutant B. subtilis strain that carries both the uvrA42 and spl-1 mutations, and transformants whose spores exhibited UV resistance were selected by UV irradiation. With a combination of genetic and physical mapping techniques, the DNA responsible for the restoration of UV resistance was shown to be present on a 2.3-kb EcoRI-HindIII fragment that was mapped to a new locus in the metC-pyrD region of the B. subtilis chromosome immediately downstream from the pstI gene. The spl coding sequence was localized on the cloned fragment by analysis of in vitro-generated deletions and by nucleotide sequencing. The spl nucleotide sequence contains an open reading frame capable of encoding a 40-kDa polypeptide that shows regional amino acid sequence homology to DNA photolyases from a number of bacteria and fungi.  相似文献   

5.
AIMS: To determine the mechanism of the hydrolysis of 4-methylumbelliferyl-beta-D-glucopyranoside (beta-MUG) by germinating and outgrowing spores of Bacillus species. METHODS AND RESULTS: Spores of B. atrophaeus (formerly B. subtilis var. niger, Fritze and Pukall 2001) are used as biological indicators of the efficacy of ethylene oxide sterilization by measurement of beta-MUG hydrolysis during spore germination and outgrowth. It was previously shown that beta-MUG is hydrolysed to 4-methylumbelliferone (MU) during the germination and outgrowth of B. atrophaeus spores (Chandrapati and Woodson 2003), and this was also the case with spores of B. subtilis 168. Germination of spores of either B. atrophaeus or B. subtilis with chloramphenicol reduced beta-MUG hydrolysis by almost 99%, indicating that proteins needed for rapid beta-MUG hydrolysis are synthesized during spore outgrowth. However, the residual beta-MUG hydrolysis during spore germination with chloramphenicol indicated that dormant spores contain low levels of proteins needed for beta-MUG uptake and hydrolysis. With B. subtilis 168 spores that lacked several general proteins of the phosphotransferase system (PTS) for sugar uptake, beta-MUG hydrolysis during spore germination and outgrowth was decreased >99.9%. This indicated that beta-MUG is taken up by the PTS, resulting in the intracellular accumulation of the phosphorylated form of beta-MUG, beta-MUG-6-phosphate (beta-MUG-P). This was further demonstrated by the lack of detectable glucosidase activity on beta-MUG in dormant, germinated and outgrowing spore extracts, while phosphoglucosidase active on beta-MUG-P was readily detected. Dormant B. subtilis 168 spores had low levels of at least four phosphoglucosidases active on beta-MUG-P: BglA, BglH, BglC (originally called YckE) and BglD (originally called YdhP). These enzymes were also detected in spores germinating and outgrowing with beta-MUG, but levels of BglH were the highest, as this enzyme's synthesis was induced ca 100-fold during spore outgrowth in the presence of beta-MUG. Deletion of the genes coding for BglA, BglH, BglC and BglD reduced beta-MUG hydrolysis by germinating and outgrowing spores of B. subtilis 168 at least 99.7%. Assay of glucosidases active on beta-MUG or beta-MUG-P in extracts of dormant and outgrowing spores of B. atrophaeus revealed no enzyme active on beta-MUG and one enzyme that comprised > or =90% of the phosphoglucosidase active on beta-MUG-P. Partial purification and amino-terminal sequence analysis of this phosphoglucosidase identified this enzyme as BglH. CONCLUSIONS: Generation of MU from beta-MUG by germinating and outgrowing spores of B. atrophaeus and B. subtilis is mediated by the PTS-driven uptake and phosphorylation of beta-MUG, followed by phosphoglucosidase action on the intracellular beta-MUG-P. The major phosphoglucosidase catalyzing MU generation from beta-MUG-P in spores of both species is probably BglH. SIGNIFICANCE AND IMPACT OF THE STUDY: This work provides new insight into the mechanism of uptake and hydrolysis of beta-MUG by germinating and outgrowing spores of Bacillus species, in particular B. atrophaeus. The research reported here provides a biological basis for a Rapid Readout Biological Indicator that is used to monitor the efficacy of ethylene oxide sterilization.  相似文献   

6.
目的:利用枯草杆菌芽孢呈递技术制备表达SARS冠状病毒S蛋白受体结合区(RBD)的重组芽孢。方法:将枯草杆菌 CotB 基因构建到基因组整合质粒pDG1664中,再将 RBD 基因连接到 CotB 基因的下游,构建成重组质粒pDG1664-CotB-RBD,通过同源重组整合到PY-79枯草杆菌基因组中;利用红霉素抗性筛选重组菌并进行PCR和DNA测序鉴定,Western印迹鉴定重组菌芽孢表面RBD蛋白的表达情况;用表达RBD的重组芽孢以口服方式免疫小鼠,通过ELISA和流式细胞术检测重组芽孢的免疫原性。结果:制备出枯草杆菌基因组整合了RBD抗原基因的重组菌株RS1931,形成的重组芽孢表达相对分子质量约62×103的CotB-RBD融合蛋白;重组芽孢免疫的小鼠血清RBD抗原特异性IgG抗体滴度在末次免疫后2周可达1∶10880,重组芽孢初免后18周的小鼠脾细胞中IFN-γ+CD4^+、IL-4+CD4^+和IFN-γ+CD8^+T细胞比例上调,表明重组芽孢经口服免疫产生良好的体液免疫和细胞免疫应答。结论:针对SARS冠状病毒S蛋白RBD建立了枯草杆菌芽孢呈递技术方法,制备出在枯草杆菌芽孢表面稳定表达外源RBD蛋白的重组株,获得的重组芽孢具有良好的免疫原性,为开发芽孢呈递型SARS疫苗奠定了基础。  相似文献   

7.
The enzyme CwlJ is involved in the depolymerization of cortex peptidoglycan during germination of spores of Bacillus subtilis. CwlJ with a C-terminal His tag was functional and was extracted from spores by procedures that remove spore coat proteins. However, this CwlJ was not extracted from disrupted spores by dilute buffer, high salt concentrations, Triton X-100, Ca(2+)-dipicolinic acid, dithiothreitol, or peptidoglycan digestion, disappeared during spore germination, and was not present in cotE spores in which the spore coat is aberrant. These findings indicate the following: (i) the reason decoated and cotE spores germinate poorly with dipicolinic acid is the absence of CwlJ from these spores; and (ii) CwlJ is located in the spore coat, presumably tightly associated with one or more other coat proteins.  相似文献   

8.
The gerP1 transposon insertion mutation of Bacillus cereus is responsible for a defect in the germination response of spores to both L-alanine and inosine. The mutant is blocked at an early stage, before loss of heat resistance or release of dipicolinate, and the efficiency of colony formation on nutrient agar from spores is reduced fivefold. The protein profiles of alkaline-extracted spore coats and the spore cortex composition are unchanged in the mutant. Permeabilization of gerP mutant spores by coat extraction procedures removes the block in early stages of germination, although a consequence of the permeabilization procedure in both wild type and mutant is that late germination events are not complete. The complete hexacistronic operon that includes the site of insertion has been cloned and sequenced. Four small proteins encoded by the operon (GerPA, GerPD, GerPB, and GerPF) are related in sequence. A homologous operon (yisH-yisC) can be found in the Bacillus subtilis genome sequence; null mutations in yisD and yisF, constructed by integrational inactivation, result in a mutant phenotype similar to that seen in B. cereus, though somewhat less extreme and equally repairable by spore permeabilization. Normal rates of germination, as estimated by loss of heat resistance, are also restored to a gerP mutant by the introduction of a cotE mutation, which renders the spore coats permeable to lysozyme. The B. subtilis operon is expressed solely during sporulation, and is sigma K-inducible. We hypothesize that the GerP proteins are important as morphogenetic or structural components of the Bacillus spore, with a role in the establishment of normal spore coat structure and/or permeability, and that failure to synthesize these proteins during spore formation limits the opportunity for small hydrophilic organic molecules, like alanine or inosine, to gain access to their normal target, the germination receptor, in the spore.  相似文献   

9.
The GerAA, -AB, and -AC proteins of the Bacillus subtilis spore are required for the germination response to L-alanine as the sole germinant. They are likely to encode the components of the germination apparatus that respond directly to this germinant, mediating the spore's response; multiple homologues of the gerA genes are found in every spore former so far examined. The gerA operon is expressed in the forespore, and the level of expression of the operon appears to be low. The GerA proteins are predicted to be membrane associated. In an attempt to localize GerA proteins, spores of B. subtilis were broken and fractionated to give integument, membrane, and soluble fractions. Using antibodies that detect Ger proteins specifically, as confirmed by the analysis of strains lacking GerA and the related GerB proteins, the GerAA protein and the GerAC+GerBC protein homologues were localized to the membrane fraction of fragmented spores. The spore-specific penicillin-binding protein PBP5*, a marker for the outer forespore membrane, was absent from this fraction. Extraction of spores to remove coat layers did not release the GerAC or AA protein from the spores. Both experimental approaches suggest that GerAA and GerAC proteins are located in the inner spore membrane, which forms a boundary around the cellular compartment of the spore. The results provide support for a model of germination in which, in order to initiate germination, germinant has to permeate the coat and cortex of the spore and bind to a germination receptor located in the inner membrane.  相似文献   

10.
To investigate the outermost structure of the Bacillus subtilis spore, we analyzed the accessibility of antibodies to proteins on spores of B. subtilis. Anti-green fluorescent protein (GFP) antibodies efficiently accessed GFP fused to CgeA or CotZ, which were previously assigned to the outermost layer termed the spore crust. However, anti-GFP antibodies did not bind to spores of strains expressing GFP fused to 14 outer coat, inner coat, or cortex proteins. Anti-CgeA antibodies bound to spores of wild-type and CgeA-GFP strains but not cgeA mutant spores. These results suggest that the spore crust covers the spore coat and is the externally exposed, outermost layer of the B. subtilis spore. We found that CotZ was essential for the spore crust to surround the spore but not for spore coat formation, indicating that CotZ plays a critical role in spore crust formation. In addition, we found that CotY-GFP was exposed on the surface of the spore, suggesting that CotY is an additional component of the spore crust. Moreover, the localization of CotY-GFP around the spore depended on CotZ, and CotY and CotZ depended on each other for spore assembly. Furthermore, a disruption of cotW affected the assembly of CotV-GFP, and a disruption of cotX affected the assembly of both CotV-GFP and CgeA-GFP. These results suggest that cgeA and genes in the cotVWXYZ cluster are involved in spore crust formation.  相似文献   

11.
The outermost layer of spores of the Bacillus cereus family is a loose structure known as the exosporium. Spores of a library of Tn917-LTV1 transposon insertion mutants of B. cereus ATCC 10876 were partitioned into hexadecane; a less hydrophobic mutant that was isolated contained an insertion in the exsA promoter region. ExsA is the equivalent of SafA (YrbA) of Bacillus subtilis, which is also implicated in spore coat assembly; the gene organizations around both are identical, and both proteins contain a very conserved N-terminal cortex-binding domain of ca. 50 residues, although the rest of the sequence is much less conserved. In particular, unlike SafA, the ExsA protein contains multiple tandem oligopeptide repeats and is therefore likely to have an extended structure. The exsA gene is expressed in the mother cell during sporulation. Spores of an exsA mutant are extremely permeable to lysozyme and are blocked in late stages of germination, which require coat-associated functions. Two mutants expressing differently truncated versions of ExsA were constructed, and they showed the same gross defects in the attachment of exosporium and spore coat layers. The protein profile of the residual exosporium harvested from spores of the three mutants--two expressing truncated proteins and the mutant with the original transposon insertion in the promoter region--showed some differences from the wild type and from each other, but the major exosporium glycoproteins were retained. The exsA gene is extremely important for the normal assembly and anchoring of both the spore coat and exosporium layers in spores of B. cereus.  相似文献   

12.
The exosporium is the outermost layer of spores of Bacillus cereus and its close relatives Bacillus anthracis and Bacillus thuringiensis. For these pathogens, it represents the surface layer that makes initial contact with the host. To date, only the BclA glycoprotein has been described as a component of the exosporium; this paper defines 10 more tightly associated proteins from the exosporium of B. cereus ATCC 10876, identified by N-terminal sequencing of proteins from purified, washed exosporium. Likely coding sequences were identified from the incomplete genome sequence of B. anthracis or B. cereus ATCC 14579, and the precise corresponding sequence from B. cereus ATCC 10876 was defined by PCR and sequencing. Eight genes encode likely structural components (exsB, exsC, exsD, exsE, exsF, exsG, exsJ, and cotE). Several proteins of the exosporium are related to morphogenetic and outer spore coat proteins of B. subtilis, but most do not have homologues in B. subtilis. ExsE is processed from a larger precursor, and the CotE homologue appears to have been C-terminally truncated. ExsJ contains a domain of GXX collagen-like repeats, like the BclA exosporium protein of B. anthracis. Although most of the exosporium genes are scattered on the genome, bclA and exsF are clustered in a region flanking the rhamnose biosynthesis operon; rhamnose is part of the sugar moiety of spore glycoproteins. Two enzymes, alanine racemase and nucleoside hydrolase, are tightly adsorbed to the exosporium layer; they could metabolize small molecule germinants and may reduce the sensitivity of spores to these, limiting premature germination.  相似文献   

13.
Bacillus subtilis is an endospore-forming bacterium. There are indications that protein disulfide linkages occur in spores, but the role of thiol-disulfide chemistry in spore synthesis is not understood. Thiol-disulfide oxidoreductases catalyze formation or breakage of disulfide bonds in proteins. CcdA is the only B. subtilis thiol-disulfide oxidoreductase that has previously been shown to play some role in endospore biogenesis. In this work we show that lack of the StoA (YkvV) protein results in spores sensitive to heat, lysozyme, and chloroform. Compared to CcdA deficiency, StoA deficiency results in a 100-fold-stronger negative effect on sporulation efficiency. StoA is a membrane-bound protein with a predicted thioredoxin-like domain probably localized in the intermembrane space of the forespore. Electron microscopy of spores of CcdA- and StoA-deficient strains showed that the spore cortex is absent in both cases. The BdbD protein catalyzes formation of disulfide bonds in proteins on the outer side of the cytoplasmic membrane but is not required for sporulation. Inactivation of bdbD was found to suppress the sporulation defect of a strain deficient in StoA. Our results indicate that StoA is a thiol-disulfide oxidoreductase that is involved in breaking disulfide bonds in cortex components or in proteins important for cortex synthesis.  相似文献   

14.
AIMS: To determine the mechanism of action of inhibitors of the germination of spores of Bacillus species, and where these inhibitors act in the germination process. METHODS AND RESULTS: Spores of various Bacillus species are significant agents of food spoilage and food-borne disease, and inhibition of spore germination is a potential means of reducing such problems. Germination of the following spores was studied: (i) wild-type B. subtilis spores; (ii) B. subtilis spores with a nutrient receptor variant allowing recognition of a novel germinant; (iii) B. subtilis spores with elevated levels of either the variant nutrient receptor or its wild-type allele; (iv) B. subtilis spores lacking all nutrient receptors and (v) wild-type B. megaterium spores. Spores were germinated with a variety of nutrient germinants, Ca2+-dipicolinic acid (DPA) and dodecylamine for B. subtilis spores, and KBr for B. megaterium spores. Compounds tested as inhibitors of germination included alkyl alcohols, a phenol derivative, a fatty acid, ion channel blockers, enzyme inhibitors and several other compounds. Assays used to assess rates of spore germination monitored: (i) the fall in optical density at 600 nm of spore suspensions; (ii) the release of the dormant spore's large depot of DPA; (iii) hydrolysis of the dormant spore's peptidoglycan cortex and (iv) generation of CFU from spores that lacked all nutrient receptors. The results with B. subtilis spores allowed the assignment of inhibitory compounds into two general groups: (i) those that inhibited the action of, or response to, one nutrient receptor and (ii) those that blocked the action of, or response to, several or all of the nutrient receptors. Some of the compounds in groups 1 and 2 also blocked action of at least one cortex lytic enzyme, however, this does not appear to be the primary site of their action in inhibiting spore germination. The inhibitors had rather different effects on germination of B. subtilis spores with nutrients or non-nutrients, consistent with previous work indicating that germination of B. subtilis spores by non-nutrients does not involve the spore's nutrient receptors. In particular, none of the compounds tested inhibited spore germination with dodecylamine, and only three compounds inhibited Ca2+-DPA germination. In contrast, all compounds had very similar effects on the germination of B. megaterium spores with either glucose or KBr. The effects of the inhibitors tested on spores of both Bacillus species were largely reversible. CONCLUSIONS: This work indicates that inhibitors of B. subtilis spore germination fall into two classes: (i) compounds (most alkyl alcohols, N-ethylmaleimide, nifedipine, phenols, potassium sorbate) that inhibit the action of, or response to, primarily one nutrient receptor and (ii) compounds [amiloride, HgCl2, octanoic acid, octanol, phenylmethylsulphonylfluoride (PMSF), quinine, tetracaine, tosyl-l-arginine methyl ester, trifluoperazine] that inhibit the action of, or response to, several nutrient receptors. Action of these inhibitors, is reversible. The similar effects of inhibitors on B. megaterium spore germination by glucose or KBr indicate that inorganic salts likely trigger germination by activating one or more nutrient receptors. The lack of effect of all inhibitors on dodecylamine germination suggests that this compound stimulates germination by creating channels in the spore's inner membrane allowing DPA release. SIGNIFICANCE AND IMPACT OF THE STUDY: This work provides new insight into the steps in spore germination that are inhibited by various chemicals, and the mechanism of action of these inhibitors. The work also provides new insights into the process of spore germination itself.  相似文献   

15.
Three mutant forms of the protease (GPR) that initiates degradation of small, acid-soluble spore proteins (SASP) during germination of spores of Bacillus species have been generated. In one variant (GPR delta), the putative pro sequence removed in conversion of the GPR zymogen (termed P46) to the active enzyme (termed P41) was deleted. GPR delta was expressed in both Escherichia coli and Bacillus subtilis as a polypeptide of 41 kDa (P41) which was active both in vivo and in vitro. The other two variants had changes in the sequence around the site where the pro sequence is removed, making this sequence even more like that recognized and cleaved by GPR in its SASP substrates. One of these variants (GPRS) was synthesized as P46S in both B. subtilis and E. coli, but P46S was processed to P41S earlier in B. subtilis sporulation than was wild-type P46. The second variant (GPREI) was made as P46EI but underwent extremely rapid processing to P41EI in both E. coli and B. subtilis. Expression of elevated (> 100-fold) levels of GPR delta or GPREI blocked sporulation at the time of synthesis of glucose dehydrogenase. Expression of elevated levels of GPRS or low levels (< 20% of the wild-type level) of GPR delta or GPREI did not retard sporulation, but the SASP level in the resultant spores was greatly reduced. Prolonged incubation of P41 delta, P41EI, or wild-type P41, either in vivo or with purified proteins in vitro, resulted in a second self-cleavage event generating a 39-kDa polypeptide termed P39. The sequence in the P(41)-->P(39) cleavage site was also quite similar to that recognized and cleaved by GPR in SASP. Together, these results strongly support a model in which activation of GPR during sporulation by conversion of P(46) to P(41) is a self-processing event triggered by a change in the spore core environment (i.e., dehydration) which precludes attack of the active P(41) on its SASP substrates. However, in the first minutes of spore germination, rapid spore core hydration allows rapid attack of active GPR on SASP.  相似文献   

16.
During endospore formation in Bacillus subtilis, over two dozen polypeptides are assembled into a multilayered structure known as the spore coat, which protects the cortex peptidoglycan (PG) and permits efficient germination. In the initial stages of coat assembly a protein known as CotE forms a ring around the forespore. A second morphogenetic protein, SpoVID, is required for maintenance of the CotE ring during the later stages, when most of proteins are assembled into the coat. Here, we report on a protein that appears to associate with SpoVID during the early stage of coat assembly. This protein, which we call SafA for SpoVID-associated factor A, is encoded by a locus previously known as yrbA. We confirmed the results of a previous study that showed safA mutant spores have defective coats which are missing several proteins. We have extended these studies with the finding that SafA and SpoVID were coimmunoprecipitated by anti-SafA or anti-SpoVID antiserum from whole-cell extracts 3 and 4 h after the onset of sporulation. Therefore, SafA may associate with SpoVID during the early stage of coat assembly. We used immunogold electron microscopy to localize SafA and found it in the cortex, near the interface with the coat in mature spores. SafA appears to have a modular design. The C-terminal region of SafA is similar to those of several inner spore coat proteins. The N-terminal region contains a sequence that is conserved among proteins that associate with the cell wall. This motif in the N-terminal region may target SafA to the PG-containing regions of the developing spore.  相似文献   

17.
Germination of spores of Bacillus subtilis with dodecylamine   总被引:1,自引:0,他引:1  
AIMS: To determine the properties of Bacillus subtilis spores germinated with the alkylamine dodecylamine, and the mechanism of dodecylamine-induced spore germination. METHODS AND RESULTS: Spores of B. subtilis prepared in liquid medium were germinated efficiently by dodecylamine, while spores prepared on solid medium germinated more poorly with this agent. Dodecylamine germination of spores was accompanied by release of almost all spore dipicolinic acid (DPA), degradation of the spore's peptidoglycan cortex, release of the spore's pool of free adenine nucleotides and the killing of the spores. The dodecylamine-germinated spores did not initiate metabolism, did not degrade their pool of small, acid-soluble spore proteins efficiently and had a significantly lower level of core water than did spores germinated by nutrients. As measured by DPA release, dodecylamine readily induced germination of B. subtilis spores that: (a) were decoated, (b) lacked all the receptors for nutrient germinants, (c) lacked both the lytic enzymes either of which is essential for cortex degradation, or (d) had a cortex that could not be attacked by the spore's cortex-lytic enzymes. The DNA in dodecylamine-germinated wild-type spores was readily stained, while the DNA in dodecylamine-germinated spores of strains that were incapable of spore cortex degradation was not. These latter germinated spores also did not release their pool of free adenine nucleotides. CONCLUSIONS: These results indicate that: (a) the spore preparation method is very important in determining the rate of spore germination with dodecylamine, (b) wild-type spores germinated by dodecylamine progress only part way through the germination process, (c) dodecylamine may trigger spore germination by a novel mechanism involving the activation of neither the spore's nutrient germinant receptors nor the cortex-lytic enzymes, and (d) dodecylamine may trigger spore germination by directly or indirectly activating release of DPA from the spore core, through the opening of channels for DPA in the spore's inner membrane. SIGNIFICANCE AND IMPACT OF THE STUDY: These results provide new insight into the mechanism of spore germination with the cationic surfactant dodecylamine, and also into the mechanism of spore germination in general. New knowledge of mechanisms to stimulate spore germination may have applied utility, as germinated spores are much more sensitive to processing treatments than are dormant spores.  相似文献   

18.
Previous work has shown that the degradation of 20% of total protein which occurs early in germination of Bacillus megaterium spores is initiated by an endoprotease. This enzyme is found only in the spore and is active only on the spore proteins degraded during germination. Action of the spore protease in vitro on the three major proteins (Proteins A, B, and C) which are degraded in vivo during germination results in cleavage of one (A and C protein) or two (B protein) peptide bonds. The sequences surrounding the cleavage sites are -Tyr-Glu- Ile-Ala-Ser-Glu-Phe- in the A protein, -Phe-Glu- Ile-Ala-Ser-Glu-Phe- in the C protein, and -Thr-Glu- Phe-Gly-Ser-Glu-Thr-, and -Thr-Glu- Phe-Ala-Ser-Glu-Thr- in the B protein, with cleavage taking place at the glutamyl bond noted by the arrow. The similarity of these four sequences suggests the possibility that the specificity of the spore protease may be due to its requirement for a specific pentapeptide sequence of the type -R-Glu-(Phe or Ile)-(Gly or Ala)-Ser-Glu-R- for recognition and cleavage. However, it is also possible that it is the conformation of the A, B, and C proteins which determines their site of cleavage by the spore protease.  相似文献   

19.
Raman spectroscopy and differential interference contrast (DIC) microscopy were used to monitor the kinetics of nutrient and nonnutrient germination of multiple individual untreated and wet-heat-treated spores of Bacillus cereus and Bacillus megaterium, as well as of several isogenic Bacillus subtilis strains. Major conclusions from this work were as follows. (i) More than 90% of these spores were nonculturable but retained their 1:1 chelate of Ca2+ and dipicolinic acid (CaDPA) when incubated in water at 80 to 95°C for 5 to 30 min. (ii) Wet-heat treatment significantly increased the time, T(lag), at which spores began release of the great majority of their CaDPA during the germination of B. subtilis spores with different nutrient germinants and also increased the variability of T(lag) values. (iii) The time period, ΔT(release), between T(lag) and the time, T(release), at which a spore germinating with nutrients completed the release of the great majority of its CaDPA, was also increased in wet-heat-treated spores. (iv) Wet-heat-treated spores germinating with nutrients had higher values of I(release), the intensity of a spore's DIC image at T(release), than did untreated spores and had much longer time periods, ΔT(lys), for the reduction in I(release) intensities to the basal value due to hydrolysis of the spore's peptidoglycan cortex, probably due at least in part to damage to the cortex-lytic enzyme CwlJ. (v) Increases in T(lag) and ΔT(release) were also observed when wet-heat-treated B. subtilis spores were germinated with the nonnutrient dodecylamine, while the change in I(release) was less significant. (vi) The effects of wet-heat treatment on nutrient germination of B. cereus and B. megaterium spores were generally similar to those on B. subtilis spores. These results indicate that (i) some proteins important in spore germination are damaged by wet-heat treatment, (ii) the cortex-lytic enzyme CwlJ is one germination protein damaged by wet heat, and (iii) the CaDPA release process itself seems likely to be the target of wet-heat damage which has the greatest effect on spore germination.  相似文献   

20.
AIMS: To determine the effect of sporulation temperature on Bacillus subtilis spore resistance and spore composition. METHODS AND RESULTS: Bacillus subtilis spores prepared at temperatures from 22 to 48 degrees C had identical amounts of dipicolinic acid and small, acid-soluble proteins but the core water content was lower in spores prepared at higher temperatures. As expected from this latter finding, spores prepared at higher temperatures were more resistant to wet heat than were spores prepared at lower temperatures. Spores prepared at higher temperatures were also more resistant to hydrogen peroxide, Betadine, formaldehyde, glutaraldehyde and a superoxidized water, Sterilox. However, spores prepared at high and low temperatures exhibited nearly identical resistance to u.v. radiation and dry heat. The cortex peptidoglycan in spores prepared at different temperatures showed very little difference in structure with only a small, albeit significant, increase in the percentage of muramic acid with a crosslink in spores prepared at higher temperatures. In contrast, there were readily detectable differences in the levels of coat proteins in spores prepared at different temperatures and the levels of at least one coat protein, CotA, fell significantly as the sporulation temperature increased. However, this latter change was not due to a reduction in cotA gene expression at higher temperatures. CONCLUSIONS: The temperature of sporulation affects a number of spore properties, including resistance to many different stress factors, and also results in significant alterations in the spore coat and cortex composition. SIGNIFICANCE AND IMPACT OF THE STUDY: The precise conditions for the formation of B. subtilis spores have a large effect on many spore 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