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 140 毫秒
1.
The kinetics of initiation of germination and inactivation by hydrostatic pressure of phosphate-buffered Bacillus pumilus spores is shown to be a consecutive first-order process at 25 C. The effect of increasing pressure at constant temperature was studied, and rate constants were derived by using the criteria of heat resistance, refractility, and stainability. The calculated volume change of activation (DeltaVdouble dagger) was -139 +/- 6 cm(3)/mole for loss of heat resistance, -158 +/- 8 cm(3)/mole for the loss of refractility, and -153 +/- 4 cm(3)/mole for the change in permeability to dilute stains for the pressure range 800 to 1,010 atm at 25 C. It is suggested that the spore exists as a Donnan phase and that pressure triggers germination by influencing the equilibrium.  相似文献   

2.
The effect of hydrostatic pressures as high as 1,700 atm at 25 C on the heat and radiation resistance of Bacillus pumilus spores was studied. Phosphate-buffered spores were more sensitive to compression than spores suspended in distilled water. Measurements of the turbidity of suspensions, the viability, refractility, stainability, dry weight, and respiratory activity of spores, and calcium and dipicolinic acid release were made for different pressures and times. Initiation of germination occurred at pressures exceeding 500 atm and was the prerequisite for inactivation by compression. The rate of initiation increased with increasing pressure at constant temperature. This result is interpreted as a net decrease in the volume of the system during initiation as a result of increased solvation of the spore components.  相似文献   

3.
Aims:  The aim of this work was to investigate the germination and inactivation of spores of Bacillus species in buffer and milk subjected to high pressure (HP) and nisin.
Methods and Results:  Spores of Bacillus subtilis and Bacillus cereus suspended in milk or buffer were treated at 100 or 500 MPa at 40°C with or without 500 IU ml−1 of nisin. Treatment at 500 MPa resulted in high levels of germination (4 log units) of B. subtilis spores in both milk and buffer; this increased to >6 logs by applying a second cycle of pressure. Viability of B. subtilis spores in milk and buffer was reduced by 2·5 logs by cycled HP, while the addition of nisin (500 IU ml−1) prior to HP treatment resulted in log reductions of 5·7 and 5·9 in phosphate buffered saline and milk, respectively. Physical damage of spores of B. subtilis following HP was apparent using scanning electron microscopy. Treating four strains of B. cereus at 500 MPa for 5 min twice at 40°C in the presence of 500 IU ml−1 nisin proved less effective at inactivating the spores of these isolates compared with B. subtilis and some strain-to-strain variability was observed.
Conclusions:  Although high levels of germination of Bacillus spores could be achieved by combining HP and nisin, complete inactivation was not achieved using the aforementioned treatments.
Significance and Impact of the Study:  Combinations of HP treatment and nisin may be an appealing alternative to heat pasteurization of milk.  相似文献   

4.
AIMS: To elucidate the factors that determine the rate of germination of Bacillus subtilis spores with very high pressure (VHP) and the mechanism of VHP germination. METHODS AND RESULTS: Spores of B. subtilis were germinated rapidly with a VHP of 500 MPa at 50 degrees C. This VHP germination did not require the spore's nutrient-germinant receptors, as found previously, and did not require diacylglycerylation of membrane proteins. However, the spore's pool of dipicolinic acid (DPA) was essential. Either of the two redundant enzymes that degrade the spore's peptidoglycan cortex, and thus allow completion of spore germination, was essential for completion of VHP germination. However, neither of these enzymes was needed for DPA release triggered by VHP treatment. Completion of spore germination as well as DPA release with VHP had an optimum temperature of approx. 60 degrees C, in contrast to an optimum temperature of 40 degrees C for germination with the moderately high pressure of 150 MPa. The rate of spore germination by VHP decreased approx. fourfold when the sporulation temperature increased from 23 degrees C to 44 degrees C, and decreased twofold when 1 mol l(-1) salt was present in sporulation. However, large variations in levels of unsaturated fatty acids in the spore's inner membranes did not affect rates of VHP germination. Complete germination of spores by VHP was not inhibited significantly by killing of spores with several oxidizing agents, and was not inhibited by ethanol, octanol or o-chlorophenol at concentrations that abolish nutrient germination. Completion of spore germination by VHP was also inhibited by Hg(2+), but this ion did not inhibit DPA release caused by VHP. In contrast, dodecylamine, a surfactant that can trigger spore germination, strongly inhibited DPA release caused by VHP treatment. CONCLUSIONS: VHP does not cause spore germination by acting upon the spore's nutrient-germinant receptors, but by directly causing DPA release. This DPA release then leads to subsequent completion of germination. VHP likely acts on the spore's inner membrane to cause DPA release, targeting either a membrane protein or the membrane itself. However, the precise identity of this target is not yet clear. SIGNIFICANCE AND IMPACT OF THE STUDY: There is significant interest in the use of VHP to eliminate or reduce levels of bacterial spores in foods. As at least partial spore germination by pressure is almost certainly essential for subsequent spore killing, knowledge of factors involved and the mechanism of VHP germination are crucial to the understanding of spore killing by VHP. This work provides new insight into factors that can affect the rate of B. subtilis spore germination by VHP, and into the mechanism of VHP germination itself.  相似文献   

5.
The influence of carbon dioxide at 1–55 atm on the germination of Clostridium sporogenes, Clostridium perfringens and Bacillus cereus spores in a complex medium was studied. The germination studies at atmospheric pressure were done in the pH range 5.2–6.7. Controls at the same pH were done in 100% nitrogen. Carbon dioxide at atmospheric pressure (1 atm) inhibited the spore germination of B. cereus spores but strongly enhanced the germination rate of those of the clostridia. Spore germination of Cl. sporogenes and Cl. perfringens was inhibited completely at 10 atm and at 25 atm, respectively. The germination rate in carbon dioxide or nitrogen was generally higher at pH 6.7 than at 5.2–6.0.  相似文献   

6.
The mechanism by which potassium sorbate inhibits Bacillus cereus T and Clostridium botulinum 62A spore germination was investigated. Spores of B. cereus T were germinated at 35 degrees C in 0.08 M sodium-potassium phosphate buffers (pH 5.7 and 6.7) containing various germinants (L-alanine, L-alpha-NH2-n-butyric acid, and inosine) and potassium sorbate. Spores of C. botulinum 62A were germinated in the same buffers but with 10 mM L-lactic acid, 20 mM sodium bicarbonate, L-alanine or L-cysteine, and potassium sorbate. Spore germination was monitored by optical density measurements at 600 nm and phase-contrast microscopy. Inhibition of B. cereus T spore germination was observed when 3,900 micrograms of potassium sorbate per ml was added at various time intervals during the first 2 min of spore exposure to the pH 5.7 germination medium. C. botulinum 62A spore germination was inhibited when 5,200 micrograms of potassium sorbate per ml was added during the first 30 min of spore exposure to the pH 5.7 medium. Potassium sorbate inhibition of germination was reversible for both B. cereus T and C. botulinum 62A spores. Potassium sorbate inhibition of B. cereus T spore germination induced by L-alanine and L-alpha-NH2-n-butyric acid was shown to be competitive in nature. Potassium sorbate was also a competitive inhibitor of L-alanine- and L-cysteine-induced germination of C. botulinum 62A spores.  相似文献   

7.
The causes of Bacillus spore resistance remain unclear. Many structures including a highly compact envelope, low hydration of the protoplast, high concentrations of Ca-chelated dipicolinic acid, and the presence of small acid-soluble spore proteins seem to contribute to resistance. To evaluate the role of internal protoplast composition and hydration, spores of Bacillus subtilis were produced at different osmotic pressures corresponding to water activities of 0.993 (standard), 0.970, and 0.950, using the two depressors (glycerol or NaCl). Sporulation of Bacillus subtilis was slower and reduced in quantity when the water activity was low, taking 4, 10, and 17 days for 0.993, 0.970, and 0.950 water activity, respectively. The spores produced at lower water activity were smaller and could germinate on agar medium at lower water activity than on standard spores. They were also more sensitive to heat (97 degrees C for 5-60 min) than the standard spores but their resistance to high hydrostatic pressure (350 MPa at 40 degrees C for 20 min to 4 h) was not altered. Our results showed that the water activity of the sporulation medium significantly affects spore properties including size, germination capacity, and resistance to heat but has no role in bacterial spore resistance to high hydrostatic pressure.  相似文献   

8.
Here we describe the functional relationship between YabG and transglutaminase (Tgl), enzymes that modify the spore coat proteins of Bacillus subtilis. In wild-type spores at 37 degrees C, Tgl mediates the crosslinking of GerQ into higher molecular mass forms; however, some GerQ multimers are found in tgl mutant spores, indicating that Tgl is not essential. Immunoblotting showed that spores isolated from a yabG mutant after sporulation at 37 degrees C contain only very low levels of GerQ multimers. Heat treatment for 20 min at 60 degrees C, which maximally activates the enzymatic activity of Tgl, caused crosslinking of GerQ in isolated yabG spores but not in tgl/yabG double-mutant spores. In addition, the germination frequency of the tgl/yabG spores in the presence of l-alanine with or without heat activation at 60 degrees C was lower than that of wild-type spores. These findings suggest that Tgl cooperates with YabG to mediate the temperature-dependent modification of the coat proteins, a process associated with spore germination in B. subtilis.  相似文献   

9.
AIMS: To determine the aerobic spore composition and presence of Bacillus sporothermodurans spores in feed concentrate for dairy cattle. METHODS AND RESULTS: Six feed concentrate samples from five different farms were analysed. High levels of spores (up to 10(6) spores g(-1)) were found. Identification of 100 selected isolates was obtained by a combination of fatty acid methyl esters analysis, amplified ribosomal DNA restriction analysis and 16S rDNA sequencing. Ninety-seven isolates could be identified to the species level or assigned to a phylogenetic species group. Most of the isolates obtained after a heat treatment of 10 min at 80 degrees C were identified as members of the B. subtilis group (32 isolates), B. pumilus (25 isolates), B. clausii (eight isolates) and B. licheniformis (eight isolates). The isolates with very heat-resistant spores, obtained after a heat treatment of 30 min at 100 degrees C, were identified as members of the B. subtilis group (five isolates), B. sporothermodurans (three isolates), B. amyloliquefaciens (one isolate), B. oleronius (one isolate) and B. pallidus (one isolate). Bacillus cereus was present in each feed concentrate sample and was isolated using a selective mannitol egg yolk polymyxin agar medium. CONCLUSIONS: Feed concentrate for dairy cattle contains known as well as as yet unknown species of Bacillus and related genera with properties relevant to the dairy sector. SIGNIFICANCE AND IMPACT OF THE STUDY: The results formulate the hypothesis that feed concentrate can be a contamination source of spores, including those of B. sporothermodurans, for raw milk at the farm level.  相似文献   

10.
The effects of moisture and oxygen concentration on germination of Bacillus cereus and B. subtilis var. niger spores were investigated in a simulated Martian environment. Less moisture was required for germination than for vegetative growth of both organisms. A daily freeze-thaw cycle lowered moisture requirements for spore germination and vegetative growth of both organisms, as compared with a constant 35 C environment. Oxygen had a synergistic effect by lowing the moisture requirements for vegetative growth, and possibly germination, of both organisms. Oxygen was not required for spore germination of either organism, but was required for vegetative growth of B. subtilis and for sporulation of both organisms.  相似文献   

11.
Five monoclonal antibodies against bacterial spores of Bacillus cereus T and Clostridium sporogenes PA3679 were developed. Two antibodies (B48 and B183) were selected for their reactivity with B. cereus T spores, two (C33 and C225) were selected for their reactivity with C. sporogenes spores, and one (D89) was selected for its reactivity with both B. cereus and C sporogenes spores. The isotypes of the antibodies were determined to be immunoglobulin G2a (IgG2a) (B48), IgG1 (B183), and IgM (C33, C225, and D89). The antibodies reacted with spores of B. cereus T, Bacillus subtilis subsp. globigii, Bacillus megaterium, Bacillus stearothermophilus, C. sporogenes, Clostridium perfringens, and Desulfotomaculum nigrificans. Antibody D89 also reacted with vegetative cells of B. cereus and C. sporogenes. Analysis of B. cereus spore extracts showed that two of the antigens with which the anti-Bacillus antibodies reacted had molecular masses of 76 kDa and approximately 250 kDa. Immunocytochemical localization indicated that antigens with which B48, B183, and D89 react are on the exosporium of the B. cereus T spore. Antibody D89 reacted with the exosporium and outer cortex of C. sporogenes spores in immunocytochemical localization studies but did not react with extracts of C. sporogenes or B. cereus spores in Western blotting. Some C. sporogenes antigens were not stable during long-term storage at -20 degrees C. Antibodies B48, B183, and D89 should prove to be useful tools for developing immunological methods for the detection of bacterial spores.  相似文献   

12.
The heat activation of bacterial spores was studied by means of differential thermal analysis in the temperature range 30-110 degrees C using the spores of Bacillus cereus. The thermogram showed three endothermic peaks at 56, 95, and 103 degrees C with one exothermic peak at 105 degrees C during the heating process. The spore coat separated from the native spores also showed a peak at 56 degrees C on its heating thermogram. The peak at 56 degrees C was reversible for both native spores and the spore coat. It was suggested that this peak at 56 degrees C might be related to the heat-activation process that takes place in the spore-coat region. It seems that the peak is due to the denaturation or the structural change of the spore-coat protein that might facilitate either the permeation of germination stimulators or the release of some germination inhibitor into or out of the spores.  相似文献   

13.
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   

14.
Several properties of the major proteins degraded during germination of spores of Bacillus cereus, Bacillus megaterium, and Bacillus subtilis have been compared. All of the proteins had low molecular weights (6,000 to 13,000) and lacked cysteine, cystine, and tryptophan. The proteins could be subdivided into two groups: group I (B. megaterium A and C proteins, B. cereus A protein, and B. subtilis alpha and beta proteins) and group II (B. cereus and B. megaterium B proteins and B. subtilis gamma protein). Species in group II had lower levels of (or lacked) the amino acids isoleucine, leucine, methionine, and proline. Similarly, proteins in each group were more closely related immunologically. However, antisera against a B. megaterium group I protein cross-reacted more strongly with the B. megaterium group II protein than with group I proteins from other spore species, whereas antisera against the B. megaterium group II protein cross-reacted most strongly with B. megaterium group I proteins. Analysis of the primary sequences at the amino termini and in the regions of the B. cereus and B. subtilis proteins cleaved by the B. megaterium spore protease revealed that the B. cereus A protein was most similar to the B. megaterium A and C proteins, and the B. cereus B protein and the B. subtilis gamma protein were most similar to the B. megaterium B protein. However, amino terminal sequences within one group of proteins varied considerably, whereas the spore protease cleavage sites were more highly conserved.  相似文献   

15.
The sporicidal activity of chlorhexidine gluconate in aqueous and alcoholic solution against spores of Bacillus subtilis was examined over a broad temperature range. Activity was not observed at 20 degrees C even with concentrations as high as 10% chlorhexidine. Temperatures of 37 degrees-70 degrees C in combination with such high concentrations were required for reductions in spore viability. No viable spores were recoverable after 4 h contact at 55 degrees C with 10% aqueous chlorhexidine and none after 3 h contact with the alcoholic solution. Because of the high concentrations necessary for activity and the possibility of sporostasis occurring from inefficient chlorhexidine inactivation, existing inactivation systems were examined and modified to obtain satisfactory results. The spores of other Bacillus species examined (B. cereus, B. megaterium and B. stearothermophilus) proved to be considerably less resistant than those of B. subtilis. Presence of organic matter had little effect on the activity.  相似文献   

16.
SUMMARY: Well washed spore preparations of Bacillus cereus and B. subtilis were suspended in various nutrient broths, soil extracts, autoclaved soil of various moisture contents, and in two inorganic solutions, phosphate buffer, pH 7·2 and Ringer's solution. These were incubated at 8°, 5°, 1° and 0° for periods up to 270 days. Periodic total and spore counts on plates indicated a progressive decrease in each, associated with germination taking place in all conditions except in the two inorganic media. Enzymic tests indicated secretion and activity of nitratase and gelatinase and, with spores of B. pasteurii , urease, as a result of germination. B. subtilis germinated to a greater extent than B. cereus in each of the nutrient media. Germination of both organisms at 5° was also observed in L - and D -alanine: in the latter it was probably the result of racemization to the L -form.  相似文献   

17.
Fluorescein-conjugated rabbit antibodies to formalized spores of Bacillus anthracis were tested against strains of B. anthracis and other Bacillus species in a subjective immunofluorescence test. The lack of reaction of B. anthracis Vollum spores with conjugated antibody raised against B. anthracis Sterne spores indicated that spores of the Vollum strain lacked a major surface antigen present in most of the other anthrax strains tested, including the non-encapsulated strains Sterne and the Soviet ST1, variants cured of the pX01 plasmid that codes for the toxin, and several virulent strains. Four other antibody preparations, raised against B, anthracis Vollum, New Hampshire, Ames and Strain 15, reacted to an approximately similar degree with spores of all four strains and of Sterne, indicating that Vollum has at least one spore antigen in common with these other strains. The anti-Sterne and anti-Vollum conjugates both displayed cross-reactions with spores of strains of B. cereus, B. coagulans, B. subtilis, B. megaterium, B. polymyxa, B. pumilus and B. thuringiensis. Absorption of the anti-anthrax conjugates with B. cereus NCTC 8035 and NCTC 10320 removed all these cross-reactions, demonstrating the existence of spore antigens specific for anthrax.  相似文献   

18.
Aims:  To determine the germination and inactivation of Bacillus cereus spores lacking various germination proteins using moderately high pressure (MHP) and heat.
Methods:  The inactivation and germination of wild-type B. cereus spores in buffer by MHP (150 MPa) at various temperatures, as well as the MHP inactivation and germination of B. cereus spores lacking individual germinant receptors and monovalent cation antiporters, was determined.
Results:  Loss of individual germinant receptors had no large effects on spore inactivation or germination, although germination of receptor-deficient spores was generally slightly decreased. Loss of the GerN in particular the GerN and GerT antiporters also decreased spore germination by MHP, especially at 40 and 50°C.
Conclusions:  Both inactivation and germination of B. cereus spores by MHP increased with rise of temperature; however, mutant strains lacking individual germinant receptor had similar levels of germination as compared to wild-type spores. To evaluate the role of germinant receptors in MHP, a strain lacking a large number of germinant receptors is needed.
Significance and Impact of the Study:  The results of this work may lead to a better understanding of how MHP causes germination of spores of B. cereus .  相似文献   

19.
Fluorescein-conjugated rabbit antibodies to formalized spores of Bacillus anthracis were tested against strains of B. anthracis and other Bacillus species in a subjective immunofluorescence test. The lack of reaction of B. anthracis Vollum spores with conjugated antibody raised against B. anthracis Sterne spores indicated that spores of the Vollum strain lacked a major surface antigen present in most of the other anthrax strains tested, including the non-encapsulated strains Sterne and the Soviet ST1, variants cured of the pX01 plasmid that codes for the toxin, and several virulent strains. Four other antibody preparations, raised against B. anthracis Vollum, New Hampshire, Ames and Strain 15, reacted to an approximately similar degree with spores of all four strains and of Sterne, indicating that Vollum has at least one spore antigen in common with these other strains. The anti-Sterne and anti-Vollum conjugates both displayed cross-reactions with spores of strains of B. cereus, B. coagulans, B. subtilis, B. megaterium, B. polymyxa, B. pumilus and B. thuringiensis. Absorption of the anti-anthrax conjugates with B. cereus NCTC 8035 and NCTC 10320 removed all these cross-reactions, demonstrating the existence of spore antigens specific for anthrax.  相似文献   

20.
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   

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