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.
Spores of Clostridium perfringens possess high heat resistance, and when these spores germinate and return to active growth, they can cause gastrointestinal disease. Work with Bacillus subtilis has shown that the spore's dipicolinic acid (DPA) level can markedly influence both spore germination and resistance and that the proteins encoded by the spoVA operon are essential for DPA uptake by the developing spore during sporulation. We now find that proteins encoded by the spoVA operon are also essential for the uptake of Ca(2+) and DPA into the developing spore during C. perfringens sporulation. Spores of a spoVA mutant had little, if any, Ca(2+) and DPA, and their core water content was approximately twofold higher than that of wild-type spores. These DPA-less spores did not germinate spontaneously, as DPA-less B. subtilis spores do. Indeed, wild-type and spoVA C. perfringens spores germinated similarly with a mixture of l-asparagine and KCl (AK), KCl alone, or a 1:1 chelate of Ca(2+) and DPA (Ca-DPA). However, the viability of C. perfringens spoVA spores was 20-fold lower than the viability of wild-type spores. Decoated wild-type and spoVA spores exhibited little, if any, germination with AK, KCl, or exogenous Ca-DPA, and their colony-forming efficiency was 10(3)- to 10(4)-fold lower than that of intact spores. However, lysozyme treatment rescued these decoated spores. Although the levels of DNA-protective alpha/beta-type, small, acid-soluble spore proteins in spoVA spores were similar to those in wild-type spores, spoVA spores exhibited markedly lower resistance to moist heat, formaldehyde, HCl, hydrogen peroxide, nitrous acid, and UV radiation than wild-type spores did. In sum, these results suggest the following. (i) SpoVA proteins are essential for Ca-DPA uptake by developing spores during C. perfringens sporulation. (ii) SpoVA proteins and Ca-DPA release are not required for C. perfringens spore germination. (iii) A low spore core water content is essential for full resistance of C. perfringens spores to moist heat, UV radiation, and chemicals.  相似文献   

2.
AIMS: To determine the reason that spores of Bacillus species, in particular Bacillus subtilis, are able to form colonies with high efficiency on media with very high salt concentrations. METHODS AND RESULTS: Spores of various Bacillus species have a significantly higher plating efficiency on media with high salt concentration (termed osmoresistance) than do log or stationary phase cells. This spore osmoresistance is higher on richer media. Bacillus subtilis spores lacking various small, acid-soluble spore proteins (SASP) were generally significantly less osmoresistant than were wild-type spores, as shown previously (Ruzal et al. 1994). Other results included: (a) spore osmoresistance varied significantly between species; (b) the osmoresistance of spores lacking SASP was not restored well by amino acid osmolytes added to plating media, but was completely restored by glucose; (c) the osmoresistance of spores lacking SASP was restored upon brief germination in the absence of salt in a process that did not require protein synthesis; (d) significant amounts of amino acids generated by SASP degradation were retained within spores upon germination in a medium with high but not low salt; (e) slowing but not abolishing SASP degradation by loss of the SASP-specific germination protease (GPR) did not affect spore osmoresistance; (f) sporulation at higher temperatures produced less osmoresistant spores; and (g) spore osmoresistance was not decreased markedly by the absence of the stress sigma factor for RNA polymerase, sigmaB. CONCLUSIONS: Spore osmoresistance appears as a result of three major factors: (1) specific characteristics of spores and cells of individual species; (2) the precise sporulation conditions that produce the spores; and (3) sufficient energy generation by the germinating and outgrowing spore to allow the spore to adapt to conditions of high osmotic strength; the substrates for this energy generation can come from either the endogenous generation of amino acids by SASP degradation or from the spore's environment, in the form of a readily taken up and metabolized energy source such as glucose. SIGNFICANCE AND IMPACT OF STUDY: These results provide information on the mechanisms of spore osmoresistance, a spore property that can be of major applied significance given the use of high osmotic strength with or without high salt as a means of food preservation.  相似文献   

3.
High-level heat resistance of spores of Bacillus thermoamylovorans poses challenges to the food industry, as industrial sterilization processes may not inactivate such spores, resulting in food spoilage upon germination and outgrowth. In this study, the germination and heat resistance properties of spores of four food-spoiling isolates were determined. Flow cytometry counts of spores were much higher than their counts on rich medium (maximum, 5%). Microscopic analysis revealed inefficient nutrient-induced germination of spores of all four isolates despite the presence of most known germination-related genes, including two operons encoding nutrient germinant receptors (GRs), in their genomes. In contrast, exposure to nonnutrient germinant calcium-dipicolinic acid (Ca-DPA) resulted in efficient (50 to 98%) spore germination. All four strains harbored cwlJ and gerQ genes, which are known to be essential for Ca-DPA-induced germination in Bacillus subtilis. When determining spore survival upon heating, low viable counts can be due to spore inactivation and an inability to germinate. To dissect these two phenomena, the recoveries of spores upon heat treatment were determined on plates with and without preexposure to Ca-DPA. The high-level heat resistance of spores as observed in this study (D120°C, 1.9 ± 0.2 and 1.3 ± 0.1 min; z value, 12.2 ± 1.8°C) is in line with survival of sterilization processes in the food industry. The recovery of B. thermoamylovorans spores can be improved via nonnutrient germination, thereby avoiding gross underestimation of their levels in food ingredients.  相似文献   

4.
Spores of Bacillus subtilis spoVF strains that cannot synthesize dipicolinic acid (DPA) but take it up during sporulation were prepared in medium with various DPA concentrations, and the germination and viability of these spores as well as the DPA content in individual spores were measured. Levels of some other small molecules in DPA-less spores were also measured. These studies have allowed the following conclusions. (i) Spores with no DPA or low DPA levels that lack either the cortex-lytic enzyme (CLE) SleB or the receptors that respond to nutrient germinants could be isolated but were unstable and spontaneously initiated early steps in spore germination. (ii) Spores that lacked SleB and nutrient germinant receptors and also had low DPA levels were more stable. (iii) Spontaneous germination of spores with no DPA or low DPA levels was at least in part via activation of SleB. (iv) The other redundant CLE, CwlJ, was activated only by the release of high levels of DPA from spores. (v) Low levels of DPA were sufficient for the viability of spores that lacked most alpha/beta-type small, acid-soluble spore proteins. (vi) DPA levels accumulated in spores prepared in low-DPA-containing media varied greatly between individual spores, in contrast to the presence of more homogeneous DPA levels in individual spores made in media with high DPA concentrations. (vii) At least the great majority of spores of several spoVF strains that contained no DPA also lacked other major spore small molecules and had gone through some of the early reactions in spore germination.  相似文献   

5.
Aims:  To determine roles of cortex lytic enzymes (CLEs) in Bacillus megaterium spore germination.
Methods and Results:  Genes for B. megaterium CLEs CwlJ and SleB were inactivated and effects of loss of one or both on germination were assessed. Loss of CwlJ or SleB did not prevent completion of germination with agents that activate the spore's germinant receptors, but loss of CwlJ slowed the release of dipicolinic acid (DPA). Loss of both CLEs also did not prevent release of DPA and glutamate during germination with KBr. However, cwlJ sleB spores had decreased viability, and could not complete germination. Loss of CwlJ eliminated spore germination with Ca2+ chelated to DPA (Ca-DPA), but loss of CwlJ and SleB did not affect DPA release in dodecylamine germination.
Conclusions:  CwlJ and SleB play redundant roles in cortex degradation during B. megaterium spore germination, and CwlJ accelerates DPA release and is essential for Ca-DPA germination. The roles of these CLEs are similar in germination of B. megaterium and Bacillus subtilis spores.
Significance and Impact of the Study:  These results indicate that redundant roles of CwlJ and SleB in cortex degradation during germination are similar in spores of Bacillus species; consequently, inhibition of these enzymes will prevent germination of Bacillus spores.  相似文献   

6.
Dipicolinic acid (DPA) comprises approximately 10% of the dry weight of spores of Bacillus species. Although DPA has long been implicated in spore resistance to wet heat and spore stability, definitive evidence on the role of this abundant molecule in spore properties has generally been lacking. Bacillus subtilis strain FB122 (sleB spoVF) produced very stable spores that lacked DPA, and sporulation of this strain with DPA yielded spores with nearly normal DPA levels. DPA-replete and DPA-less FB122 spores had similar levels of the DNA protective alpha/beta-type small acid-soluble spore proteins (SASP), but the DPA-less spores lacked SASP-gamma. The DPA-less FB122 spores exhibited similar UV resistance to the DPA-replete spores but had lower resistance to wet heat, dry heat, hydrogen peroxide, and desiccation. Neither wet heat nor hydrogen peroxide killed the DPA-less spores by DNA damage, but desiccation did. The inability to synthesize both DPA and most alpha/beta-type SASP in strain PS3664 (sspA sspB sleB spoVF) resulted in spores that lost viability during sporulation, at least in part due to DNA damage. DPA-less PS3664 spores were more sensitive to wet heat than either DPA-less FB122 spores or DPA-replete PS3664 spores, and the latter also retained viability during sporulation. These and previous results indicate that, in addition to alpha/beta-type SASP, DPA also is extremely important in spore resistance and stability and, further, that DPA has some specific role(s) in protecting spore DNA from damage. Specific roles for DPA in protecting spore DNA against damage may well have been a major driving force for the spore's accumulation of the high levels of this small molecule.  相似文献   

7.
Superdormant spores of Bacillus subtilis and Bacillus megaterium were isolated in 4 to 12% yields following germination with high nutrient levels that activated one or two germinant receptors. These superdormant spores did not germinate with the initial nutrients or those that stimulated other germinant receptors, and the superdormant spores'' defect was not genetic. The superdormant spores did, however, germinate with Ca2+-dipicolinic acid or dodecylamine. Although these superdormant spores did not germinate with high levels of nutrients that activated one or two nutrient germinant receptors, they germinated with nutrient mixtures that activated more receptors, and using high levels of nutrient mixtures activating more germinant receptors decreased superdormant spore yields. The use of moderate nutrient levels to isolate superdormant spores increased their yields; the resultant spores germinated poorly with the initial moderate nutrient concentrations, but they germinated well with high nutrient concentrations. These findings suggest that the levels of superdormant spores in populations depend on the germination conditions used, with fewer superdormant spores isolated when better germination conditions are used. These findings further suggest that superdormant spores require an increased signal for triggering spore germination compared to most spores in populations. One factor determining whether a spore is superdormant is its level of germinant receptors, since spore populations with higher levels of germinant receptors yielded lower levels of superdormant spores. A second important factor may be heat activation of spore populations, since yields of superdormant spores from non-heat-activated spore populations were higher than those from optimally activated spores.Spores of various Bacillus species are formed in sporulation and are metabolically dormant and very resistant to environmental stress factors (21, 37). While such spores can remain in this dormant, resistant state for long periods, they can return to life rapidly through the process of germination, during which the spore''s dormancy and extreme resistance are lost (36). Spore germination has long been of intrinsic interest, and continues to attract applied interest, because (i) spores of a number of Bacillus species are major agents of food spoilage and food-borne disease and (ii) spores of Bacillus anthracis are a major bioterrorism agent. Since spores are much easier to kill after they have germinated, it would be advantageous to trigger germination of spores in foods or the environment and then readily inactivate the much less resistant germinated spores. However, this simple strategy has been largely nullified because germination of spore populations is heterogeneous, with some spores, often called superdormant spores, germinating extremely slowly and potentially coming back to life long after treatments are applied to inactivate germinated spores (8, 9, 16). The concern over superdormant spores in populations also affects decisions such as how long individuals exposed to B. anthracis spores should continue to take antibiotics, since spores could remain dormant in an individual for long periods and then germinate and cause disease (3, 11).In many species, spore germination can be increased by a prior activation step, generally a sublethal heat treatment, although the changes taking place during heat activation are not known (16). Spore germination in Bacillus species is normally triggered by nutrients such as glucose, amino acids, or purine ribosides (27, 36). These agents bind to germinant receptors located in the spore''s inner membrane that are specific for particular nutrients. In Bacillus subtilis, the GerA receptor responds to l-alanine or l-valine, while the GerB and GerK receptors act cooperatively to respond to a mixture of l-asparagine (or l-alanine), d-glucose, d-fructose and K+ ions (AGFK [or Ala-GFK]) (1, 27, 36). There are even more functional germinant receptors in Bacillus megaterium spores, and these respond to d-glucose, l-proline, l-leucine, l-valine, or even salts, such as KBr (6). Glucose appears to trigger germination of B. megaterium spores through either of two germinant receptors, GerU or GerVB, while l-proline triggers germination through only the GerVB receptor, and KBr germination is greatly decreased by the loss of either GerU or GerVB (6). Nutrient binding to the germinant receptors triggers the release of small molecules from the spore core, most notably the huge depot (∼10% of spore dry weight) of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) present in spores predominantly as a 1:1 diluted chelate with Ca2+ (Ca-DPA) (35, 36). Ca-DPA release then triggers the activation of one of two redundant cortex lytic enzymes (CLEs) that degrade the spore''s peptidoglycan cortex, and cortex degradation completes spore germination and allows progression into outgrowth and then vegetative growth (27, 33, 36).Spore germination can also be triggered by nonnutrient agents, including Ca-DPA and cationic surfactants (27, 33, 36). With B. subtilis spores, Ca-DPA triggers germination by activating one particular CLE, termed CwlJ, and bypasses the spore''s germinant receptors. Germination by the cationic surfactant dodecylamine also bypasses the germinant receptors, and this agent appears to release small molecules including Ca-DPA from the spore core either by opening a normal channel in the spore''s inner membrane for Ca-DPA and other small molecules or by creating such a channel (31, 38, 39).Almost all work on the specifics of the germination of spores of Bacillus species has focused on the majority of spores in populations, and little detailed attention has been paid to that minority of spores that either fail to germinate or germinate extremely slowly. However, it is these latter spores that are most important in unraveling the cause of superdormancy and perhaps suggesting a means to germinate and thus easily inactivate such superdormant spores. Consequently, we have undertaken the task of isolating superdormant spores from spore populations, using buoyant density centrifugation to separate dormant spores from germinated spores. The properties of these purified superdormant spores were then studied, and this information has suggested some reason(s) for spore superdormancy.  相似文献   

8.
Sporulation of a Bacillus subtilis strain (termed alpha(-) beta(-)) lacking the majority of the alpha/beta-type small, acid-soluble spore proteins (SASP) that are synthesized in the developing forespore and saturate spore DNA exhibited a number of differences from that of the wild-type strain, including delayed forespore accumulation of dipicolinic acid, overexpression of forespore-specific genes, and delayed expression of at least one mother cell-specific gene turned on late in sporulation, although genes turned on earlier in the mother cell were expressed normally in alpha(-) beta(-) strains. The sporulation defects in alpha(-) beta(-) strains were corrected by synthesis of chromosome-saturating levels of either of two wild-type, alpha/beta-type SASP but not by a mutant SASP that binds DNA poorly. Spores from alpha(-) beta(-) strains also exhibited less glutaraldehyde resistance and slower outgrowth than did wild-type spores, but at least some of these defects in alpha(-) beta(-) spores were abolished by the synthesis of normal levels of alpha/beta-type SASP. These results indicate that alpha/beta-type SASP may well have global effects on gene expression during sporulation and spore outgrowth.  相似文献   

9.
Aims: To determine the effects of Mn levels in Bacillus megaterium sporulation and spores on spore resistance. Methods and Results: Bacillus megaterium was sporulated with no added MnCl2 and up to 1 mmol l?1 MnCl2. The resultant spores were purified and loosely bound Mn removed, and spore Mn levels were found to vary c. 100‐fold. The Mn level had no effect on spore γ‐radiation resistance, but B. megaterium spores with elevated Mn levels had higher resistance to UVC radiation (as did Bacillus subtilis spores), wet and dry heat and H2O2. However, levels of dipicolinic acid and the DNA‐protective α/β‐type small, acid‐soluble spore proteins were the same in spores with high and low Mn levels. Conclusions: Mn levels either in sporulation or in spores are important factors in determining levels of B. megaterium spore resistance to many agents, with the exception of γ‐radiation. Significance and Impact of the Study: The Mn level in sporulation is an important factor to consider when resistance properties of B. megaterium spores are examined, and will influence the UV resistance of B. subtilis spores, some of which are used as biological dosimeters.  相似文献   

10.
AIMS: To determine the resistance of a variety of Bacillus species spores to a combined high pressure and heat treatment; and to determine the affect of varying sporulation and treatment conditions on the level of inactivation achieved. METHODS AND RESULTS: Spores from eight Bacillus species (40 isolates) were high pressure-heat treated at 600 MPa, 1 min, initial temperature 72 degrees C. The level of inactivation was broad (no inactivation to 6 log10 spores ml(-1) reduction) and it varied within species. Different sporulation agar, high pressure equipment and pressure-transmitting fluid significantly affected the response of some isolates. Varying the initial treatment temperature (75, 85 or 95 degrees C) shifted the relative order of isolate high pressure-heat resistance. CONCLUSIONS: The response of Bacillus spores to combined high pressure-heat treatment is variable and can be attributed to both intrinsic and extrinsic factors. The combined process resulted in a high level of spore inactivation for several Bacillus species and is a potential alternative treatment to traditional heat-only processes. SIGNIFICANCE AND IMPACT OF THE STUDY: Sporulation conditions, processing conditions and treatment temperature all affect the response of Bacillus spores to the combined treatment of high pressure and heat. High levels of spore inactivation can be achieved but the response is variable both within and between species.  相似文献   

11.
This work was undertaken to obtain information on levels of metabolism in dormant spores of Bacillus species incubated for weeks at physiological temperatures. Spores of Bacillus megaterium and Bacillus subtilis strains were harvested shortly after release from sporangia and incubated under various conditions, and dormant spore metabolism was monitored by 31P nuclear magnetic resonance (NMR) analysis of molecules including 3-phosphoglyceric acid (3PGA) and ribonucleotides. Incubation for up to 30 days at 4, 37, or 50°C in water, at 37 or 50°C in buffer to raise the spore core pH from ∼ 6.3 to 7.8, or at 4°C in spent sporulation medium caused no significant changes in ribonucleotide or 3PGA levels. Stage I germinated spores of Bacillus megaterium that had slightly increased core water content and a core pH of 7.8 also did not degrade 3PGA and accumulated no ribonucleotides, including ATP, during incubation for 8 days at 37°C in buffered saline. In contrast, spores incubated for up to 30 days at 37 or 50°C in spent sporulation medium degraded significant amounts of 3PGA and accumulated ribonucleotides, indicative of RNA degradation, and these processes were increased in B. megaterium spores with a core pH of ∼7.8. However, no ATP was accumulated in these spores. These data indicate that spores of Bacillus species stored in water or buffer at low or high temperatures exhibited minimal, if any, metabolism of endogenous compounds, even when the spore core pH was 7.8 and core water content was increased somewhat. However, there was some metabolism in spores stored in spent sporulation medium.  相似文献   

12.
Degradation of small, acid-soluble spore proteins during germination of Bacillus subtilis spores is initiated by a sequence-specific protease called GPR. Western blot (immunoblot) analysis of either Bacillus megaterium or B. subtilis GPR expressed in B. subtilis showed that GPR is synthesized at about the third hour of sporulation in a precursor form and is processed to an approximately 2- to 5-kDa-smaller species 2 to 3 h later, at or slightly before the time of accumulation of dipicolinic acid by the forespore. This was found with both normal levels of expression of B. subtilis and B. megaterium GPR in B. subtilis, as well as when either protein was overexpressed up to 100-fold. The sporulation-specific processing of GPR was blocked in all spoIII, -IV, and -V mutants tested (none of which accumulated dipicolinic acid), but not in a spoVI mutant which accumulated dipicolinic acid. The amino-terminal sequences of the B. megaterium and B. subtilis GPR initially synthesized in sporulation were identical to those predicted from the coding genes' sequences. However, the processed form generated in sporulation lacked 15 (B. megaterium) or 16 (B. subtilis) amino-terminal residues. The amino acid sequence surrounding this proteolytic cleavage site was very homologous to the consensus sequence recognized and cleaved by GPR in its small, acid-soluble spore protein substrates. This observation, plus the efficient processing of overproduced GPR during sporulation, suggests that the GPR precursor may autoproteolyze itself during sporulation. During spore germination, the GPR from either species expressed in B. subtilis was further processed by removal of one additional amino-terminal amino acid (leucine), generating the mature protease which acts during spore germination.  相似文献   

13.
Bacteria of various Bacillus species are able to grow in media with very high osmotic strength in part due to the accumulation of low-molecular-weight osmolytes such as glycine betaine (GB). Cells of Bacillus species grown in rich and minimal media contained low levels of GB, but GB levels were 4- to 60-fold higher in cells grown in media with high salt. GB levels in Bacillus subtilis cells grown in minimal medium were increased approximately 7-fold by GB in the medium and 60-fold by GB plus high salt. GB was present in spores of Bacillus species prepared in media with or without high salt but at lower levels than in comparable growing cells. With spores prepared in media with high salt, GB levels were highest in B. subtilis spores and > or =20-fold lower in B. cereus and B. megaterium spores. Although GB levels in B. subtilis spores were elevated 15- to 30-fold by GB plus high salt in sporulation media, GB levels did not affect spore resistance. GB levels were similar in wild-type B. subtilis spores and spores that lacked major small, acid-soluble spore proteins but were much lower in spores that lacked dipicolinic acid.  相似文献   

14.
Dynamic processes during wet-heat treatment of individual spores of Bacillus cereus, Bacillus megaterium, and Bacillus subtilis at 80 to 90°C were investigated using dual-trap Raman spectroscopy, differential interference contrast (DIC) microscopy, and nucleic acid stain (SYTO 16) fluorescence microscopy. During spore wet-heat treatment, while the spores' 1:1 chelate of Ca(2+) with dipicolinic acid (CaDPA) was released rapidly at a highly variable time T(lag), the levels of spore nucleic acids remained nearly unchanged, and the T(lag) times for individual spores from the same preparation were increased somewhat as spore levels of CaDPA increased. The brightness of the spores' DIC image decreased by ~50% in parallel with CaDPA release, and there was no spore cortex hydrolysis observed. The lateral diameters of the spores' DIC image and SYTO 16 fluorescence image also decreased in parallel with CaDPA release. The SYTO 16 fluorescence intensity began to increase during wet-heat treatment at a time before T(lag) and reached maximum at a time slightly later than T(release). However, the fluorescence intensities of wet-heat-inactivated spores were ~15-fold lower than those of nutrient-germinated spores, and this low SYTO 16 fluorescence intensity may be due in part to the low permeability of the dormant spores' inner membranes to SYTO 16 and in part to nucleic acid denaturation during the wet-heat treatment.  相似文献   

15.
Messenger Ribonucleic Acid of Dormant Spores of Bacillus subtilis   总被引:4,自引:3,他引:1       下载免费PDF全文
Evidence of the presence of messenger ribonucleic acid (mRNA) in dormant spores of Bacillus subtilis has been obtained. The bulk RNA from spores was isolated and labeled in vitro with tritiated dimethyl sulfate. The spore RNA hybridized to 2.4 to 3.2% of the B. subtilis genome. The RNA hybridized to both the complementary heavy and light fractions of deoxyribonucleic acid (DNA). Bulk RNA from log-phase cells competed with virtually all the spore RNA for the heavy DNA fraction and with part of the spore RNA for the light DNA fraction. Bulk RNA from stage IV cells in sporulation also competed with all of the spore RNA for the heavy DNA fraction and with essentially all the spore RNA for the light DNA fraction. These results indicate that dormant spores contain mRNA species present in both log-phase cells and stage IV cells of sporulation. The RNA polymerase in the developing forespore must be able to recognize promotor sites for both log-phase and sporulation genes.  相似文献   

16.
AIMS: To assess the influence of sporulation media on heat resistance, and the use of stress recovery media to measure preservation injury of spores of five representative spoilage bacilli. METHODS AND RESULTS: Bacillus spores prepared on nutrient agar supplemented with Ca2+, Mg2+, Mn2+, Fe2+ and K+ were more heat-resistant than spores obtained from nutrient agar with Mn2+. This increased heat resistance correlated with a decrease in the protoplast water content as determined by buoyant density sedimentation. The degree of preservation injury severity could be assessed on media containing NaCl at moderate pH and organic acids at acid pH. Ca-DPA, K+ or proline were added to the recovery media to demonstrate that heat probably caused injury to both spore germination and the outgrowth system. SIGNIFICANCE AND IMPACT OF THE STUDY: The metal content of sporulation media can strongly effect the validity of preservation resistance studies. The distinctive recovery media developed here can be relevant for assessing and comparing new preservation technologies.  相似文献   

17.
Two major low-molecular weight, acid-soluble proteins (termed A and B proteins) were purified from Bacillus sphaericus spores and had properties similar to those of the analogous proteins from spores of other Bacillus species. These proteins were accumulated late in sporulation, when the developing spores became resistant to UV light, and were degraded during spore germination by a spore protease. A mutant of B. sphaericus unable to make spore cortex because of a block in diaminopimelic acid (DAP) biosynthesis accumulated and maintained levels of the A and B proteins similar to those in the DAP+ parent or the DAP- strain in which cortex formation was restored by growth with DAP. In addition, the DAP- strain grown without DAP acquired a level of UV light resistance identical to that of wild-type spores and at the time of appearance of the A and B proteins. These findings indicate that formation of little, if any, spore cortex is required for acquisition of UV light resistance or maintenance of high levels of A and B proteins. The data provide further support for a role of the A and B proteins in the spore's UV light resistance.  相似文献   

18.
19.
Spores of Bacillus stearothermophilus were exposed to calcium and sodium salts of dipicolinic acid (DPA) in phosphate and Tris acid maleate buffers over the range pH 4.5–10.0. The exposed spores were enumerated using a standard plate counting technique from which the kinetics of colony formation were determined and maximum colony counts were obtained for each condition examined. Exposure of the spores to calcium-DPA (50-40 mmol/l) in Tris acid maleate buffer pH 9.0 maintained at 10°C was found to produce an optimal response. Following this method the total viable population of a spore suspension was enumerated. This was demonstrated statistically using the Wilcoxon rank-sum test for significance. Calcium-DPA was found to produce activation in spores but further germinants and nutrients were required for colony formation. The Ca-DPA treatment was found to be effective in enumerating both naturally dormant spores and heat injured spores.  相似文献   

20.
AIMS: To determine the size distribution of the spores of Bacillus anthracis, and compare its size with other Bacillus species grown and sporulated under similar conditions. METHODS AND RESULTS: Spores from several Bacillus species, including seven strains of B. anthracis and six close neighbours, were prepared and studied using identical media, protocols and instruments. Here, we report the spore length and diameter distributions, as determined by transmission electron microscopy (TEM). We calculated the aspect ratio and volume of each spore. All the studied strains of B. anthracis had similar diameter (mean range between 0.81 +/- 0.08 microm and 0.86 +/- 0.08 microm). The mean lengths of the spores from different B. anthracis strains fell into two significantly different groups: one with mean spore lengths 1.26 +/- 0.13 microm or shorter, and another group of strains with mean spore lengths between 1.49 and 1.67 microm. The strains of B. anthracis that were significantly shorter also sporulated with higher yield at relatively lower temperature. The grouping of B. anthracis strains by size and sporulation temperature did not correlate with their respective virulence. CONCLUSIONS: The spores of Bacillus subtilis and Bacillus atrophaeus (previously named Bacillus globigii), two commonly used simulants of B. anthracis, were considerably smaller in length, diameter and volume than all the B. anthracis spores studied. Although rarely used as simulants, the spores of Bacillus cereus and Bacillus thuringiensis had dimensions similar to those of B. anthracis. SIGNIFICANCE AND IMPACT OF THE STUDY: Spores of nonvirulent Bacillus species are often used as simulants in the development and testing of countermeasures for biodefence against B. anthracis. The data presented here should help in the selection of simulants that better resemble the properties of B. anthracis, and thus, more accurately represent the performance of collectors, detectors and other countermeasures against this threat agent.  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