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 31 毫秒
1.
Short exposures of Bacillus spores to nutrient germinants can commit spores to germinate when germinants are removed or their binding to the spores'' nutrient germinant receptors (GRs) is inhibited. Bacillus subtilis spores were exposed to germinants for various periods, followed by germinant removal to prevent further commitment. Release of spore dipicolinic acid (DPA) was then measured by differential interference contrast microscopy to monitor germination of multiple individual spores, and spores did not release DPA after 1 to 2 min of germinant exposure until ∼7 min after germinant removal. With longer germinant exposures, percentages of committed spores with times for completion of DPA release (Trelease) greater than the time of germinant removal (Tb) increased, while the time TlagTb, where Tlag represents the time when rapid DPA release began, was decreased but rapid DPA release times (ΔTrelease = TreleaseTlag) were increased; Factors affecting average Trelease values and the percentages of committed spores were germinant exposure time, germinant concentration, sporulation conditions, and spore heat activation, as previously shown for commitment of spore populations. Surprisingly, germination of spores given a 2nd short germinant exposure 30 to 45 min after a 1st exposure of the same duration was significantly higher than after the 1st exposure, but the number of spores that germinated in the 2nd germinant exposure decreased as the interval between germinant exposures increased up to 12 h. The latter results indicate that spores have some memory, albeit transient, of their previous exposure to nutrient germinants.  相似文献   

2.
Germination of Bacillus spores with a high pressure (HP) of ∼150 MPa is via activation of spores'' germinant receptors (GRs). The HP germination of multiple individual Bacillus subtilis spores in a diamond anvil cell (DAC) was monitored with phase-contrast microscopy. Major conclusions were that (i) >95% of wild-type spores germinated in 40 min in a DAC at ∼150 MPa and 37°C but individual spores'' germination kinetics were heterogeneous; (ii) individual spores'' HP germination kinetic parameters were similar to those of nutrient-triggered germination with a variable lag time (Tlag) prior to a period of the rapid release (ΔTrelease) of the spores'' dipicolinic acid in a 1:1 chelate with Ca2+ (CaDPA); (iii) spore germination at 50 MPa had longer average Tlag values than that at ∼150 MPa, but the ΔTrelease values at the two pressures were identical and HPs of <10 MPa did not induce germination; (iv) B. subtilis spores that lacked the cortex-lytic enzyme CwlJ and that were germinated with an HP of 150 MPa exhibited average ΔTrelease values ∼15-fold longer than those for wild-type spores, but the two types of spores exhibited similar average Tlag values; and (v) the germination of wild-type spores given a ≥30-s 140-MPa HP pulse followed by a constant pressure of 1 MPa was the same as that of spores exposed to a constant pressure of 140 MPa that was continued for ≥35 min; (vi) however, after short 150-MPa HP pulses and incubation at 0.1 MPa (ambient pressure), spore germination stopped 5 to 10 min after the HP was released. These results suggest that an HP of ∼150 MPa for ≤30 s is sufficient to fully activate spores'' GRs, which remain activated at 1 MPa but can deactivate at ambient pressure.  相似文献   

3.
The simultaneous nutrient germination of hundreds of individual wild-type spores of three Bacillus species and a number of Bacillus subtilis strains has been measured by two new methods, and rates of release of the great majority of the large pool of dipicolinic acid (DPA) from individual spores of B. subtilis strains has been measured by Raman spectroscopy with laser tweezers. The results from these analyses and published data have allowed a number of significant conclusions about the germination of spores of Bacillus species as follows. (i) The time needed for release of the great majority of a Bacillus spore''s DPA once rapid DPA release had begun (ΔTrelease) during nutrient germination was independent of the concentration of nutrient germinant used, the level of the germinant receptors (GRs) that recognize nutrient germinants used and heat activation prior to germination. Values for ΔTrelease were generally 0.5 to 3 min at 25 to 37°C for individual wild-type spores. (ii) Despite the conclusion above, germination of individual spores in populations was very heterogeneous, with some spores in wild-type populations completing germination ≥15-fold slower than others. (iii) The major factor in the heterogeneity in germination of individual spores in populations was the highly variable lag time, Tlag, between mixing spores with nutrient germinants and the beginning of ΔTrelease. (iv) A number of factors decrease spores'' Tlag values including heat activation, increased levels of GRs/spore, and higher levels of nutrient germinants. These latter factors appear to affect the level of activated GRs/spore during nutrient germination. (v) The conclusions above lead to the simple prediction that a major factor causing heterogeneity in Bacillus spore germination is the number of functional GRs in individual spores, a number that presumably varies significantly between spores in populations.Spores of various Bacillus species are metabolically dormant and can survive for years in this state (30). However, spores constantly sense their environment, and if appropriate small molecules termed germinants are present, spores can rapidly return to life in the process of germination followed by outgrowth (25, 29, 30). The germinants that most likely trigger spore germination in the environment are low-molecular-weight nutrient molecules, the identities of which are strain and species specific, including amino acids, sugars, and purine nucleosides. Metabolism of these nutrient germinants is not needed for the triggering of spore germination. Rather, these germinants are recognized by germinant receptors (GRs) located in the spore''s inner membrane that recognize their cognate germinants in a stereospecific manner (17, 24, 25, 29). Spores have a number of such GRs, with three functional GRs in Bacillus subtilis spores and even more in Bacillus anthracis, Bacillus cereus, and Bacillus megaterium spores (6, 29, 30). Binding of nutrient germinants to some single GRs is sufficient to trigger spore germination, for example the triggering of B. subtilis spore germination by binding of l-alanine or l-valine to the GerA GR. However, many GRs cooperate such that binding of germinants by ≥2 different GRs is needed to trigger germination (2, 29): for example, the triggering of B. subtilis spore germination by the binding of components of a mixture of l-asparagine, d-glucose, d-fructose, and K+ ions (AGFK) to the GerB and GerK GRs. The binding of nutrient germinants to GRs triggers subsequent events in germination, although how this is accomplished is not known.The first readily measured biochemical event after addition of nutrient germinants to Bacillus spores is the rapid release of the spore''s large depot (∼10% of spore dry weight) of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) plus its chelated divalent cations, predominantly Ca2+ (Ca-DPA), from the spore core (25, 29). Ca-DPA release then results in the activation of two redundant cortex-lytic enzymes (CLEs), CwlJ and SleB, which hydrolyze the spore''s peptidoglycan cortex layer (16, 22, 27, 29). CwlJ is activated by Ca-DPA as it is released from the spore while SleB is activated only after most DPA is released (17, 20, 22, 26, 27). Cortex hydrolysis ultimately allows the spore core to expand and take up more water, raising the core water content from the 35 to 45% of wet weight in the dormant spore to the 80% of wet weight characteristic of growing cells. Full hydration of the spore core then allows enzyme action, metabolism, and macromolecular synthesis to resume in the now fully germinated spore.Germination of spores in populations is very heterogeneous, with some spores germinating rapidly and some extremely slowly (4, 5, 9, 11, 13-15, 19, 26, 31, 32). Where it has been studied, the reason for this heterogeneity has been suggested to be due to a variable lag period (Tlag) between the time of mixing spores with a germinant and the time at which rapid DPA release begins, since once rapid DPA release begins, the time required for release of almost all DPA as well as for subsequent cortex hydrolysis is generally rather short compared to Tlag values in individual spores (5, 11, 13-15, 19, 26, 31, 32). The times required for DPA release and cortex hydrolysis are also similar in wild-type spores with both very short and long Tlag values (5, 15, 19, 27). The reasons for the variability in Tlag times between individual spores in populations are not known, although there are reports that both activation of spores for germination by a sublethal heat treatment (heat activation) as well as increasing concentrations of nutrient germinants can shorten Tlag values (12, 14, 15, 18, 32). However, there has been no detailed study of the causes of the variability in Tlag values between very large numbers of individual spores in populations.In order to study the heterogeneity in spore germination thoroughly, methods are needed to follow the germination of hundreds of individual spores over several hours. Initial studies of the germination of individual spores examined a single spore in a phase-contrast microscope and followed the germination of this spore by changes in the core''s refractive index due to DPA release and core swelling (14, 15, 32, 34). However, this method is labor-intensive for gathering data with hundreds of individual spores. More recently, confocal microscopy and then surface adsorption and optical tweezers have been used to capture single spores, and germination events have been followed by methods such as Raman spectroscopy to directly measure DPA release, as well as phase-contrast microscopy and elastic light scattering (3, 5, 9, 10, 19, 26). While the latter recent advances have allowed accumulation of much information about germination, collection of this type of data for large numbers of individual spores is still labor-intensive, although use of dual optical traps (35) and perhaps multiple traps in the future may alleviate this problem. However, phase-contrast microscopy plus appropriate computer software has recently allowed the monitoring of many hundreds of individual spores for several hours, with automated assessment of various changes in the cells during the period of observation (19). In the present work, we have used both phase-contrast and differential interference contrast (DIC) microscopy to monitor the germination of many hundreds of individual spores of three Bacillus species adhered on either an agarose pad or a glass coverslip for 1 to 2 h. This work, as well as examination of times needed for release of most DPA once rapid DPA release has begun during germination of individual spores under a variety of conditions, has allowed detailed examination of the effects of heat activation, nutrient germinant concentration, GR numbers per spore, and individual CLEs on spore germination heterogeneity and on values of Tlag for individual spores.  相似文献   

4.
Aim: To analyse the dynamic germination of hundreds of individual superdormant (SD) Bacillus subtilis spores. Methods and Results: Germination of hundreds of individual SD B. subtilis spores with various germinants and under different conditions was followed by multifocus Raman microspectroscopy and differential interference contrast microscopy for 12 h and with temporal resolutions of ≤30 s. SD spores germinated poorly with the nutrient germinant used to isolate them and with alternate germinants targeting the germinant receptor (GR) used originally. The mean times following mixing of spores and nutrient germinants to initiate and complete fast release of Ca‐dipicolinic acid (CaDPA) (Tlag and Trelease times, respectively) of SD spores were much longer than those of dormant spores. However, the ΔTrelease times (Trelease?Tlag) of SD spores were essentially identical to those of dormant spores. SD spores germinated almost as well as dormant spores with nutrient germinants targeting GRs different from the one used to isolate the SD spores and with CaDPA that does not trigger spore germination via GRs. Conclusions: Since (i) ΔTrelease times were essentially identical in GR‐dependent germination of SD and dormant spores; (ii) rates of GR‐independent germination of SD and dormant spores were identical; (iii) large increases in Tlag times were the major difference in the GR‐dependent germination of SD as compared with spores; and (iv) higher GR levels are correlated with shorter Tlag times, these results are consistent with the hypothesis that low levels of a GR are the major reason that some spores in a population are SD with germinants targeting this same GR. Significance and Impact of the Study: This study provides information on the dynamic germination of individual SD spores and improves the understanding of spore superdormancy.  相似文献   

5.
High-precision measurements of size changes of individual bacterial spores based on ellipse fitting to bright-field images recorded with a digital camera were employed to monitor the germination of Bacillus spores with a precision of ∼5 nm. To characterize the germination of individual spores, we recorded bright-field and phase-contrast images and found that the timing of changes in their normalized intensities coincided, so the bright-field images can be used to characterize spore size and refractility changes during germination. The major conclusions from this work were as follows. (i) The sizes of germinating B. cereus spores were nearly unchanged until Trelease, the time of the completion of CaDPA (a 1:1 chelate of Ca2+ and dipicolinic acid [DPA]) release after addition of nutrient germinants. (ii) The minor axis of germinating B. cereus spores rapidly increased by ∼50 nm in a few seconds right after Trelease, while the major axis was slightly decreased or unchanged. Both the minor and major axes remained unchanged for a further 30 to 45 s and then increased by 100 to 200 nm by Tlys, the time of completion of cortex lysis. (iii) Individual spores in a population showed significant heterogeneity in the timing of germination events, such as Trelease and Tlys, but also variation in size changes during germination. (iv) Bacillus subtilis wild-type spores, B. subtilis spores lacking the cortex-lytic enzyme CwlJ, and wild-type Bacillus megaterium spores showed similar kinetics of size changes during nutrient germination. The size increases in germinating spores probably result from uptake of water and cortex lysis after completion of CaDPA release.  相似文献   

6.
This study is aimed at the development and application of a convenient and rapid optical assay to monitor the wet-heat resistance of bacterial endospores occurring in food samples. We tested the feasibility of measuring the release of the abundant spore component dipicolinic acid (DPA) as a probe for heat inactivation. Spores were isolated from the laboratory type strain Bacillus subtilis 168 and from two food product isolates, Bacillus subtilis A163 and Bacillus sporothermodurans IC4. Spores from the lab strain appeared much less heat resistant than those from the two food product isolates. The decimal reduction times (D values) for spores from strains 168, A163, and IC4 recovered on Trypticase soy agar were 1.4, 0.7, and 0.3 min at 105°C, 120°C, and 131°C, respectively. The estimated Z values were 6.3°C, 6.1°C, and 9.7°C, respectively. The extent of DPA release from the three spore crops was monitored as a function of incubation time and temperature. DPA concentrations were determined by measuring the emission at 545 nm of the fluorescent terbium-DPA complex in a microtiter plate fluorometer. We defined spore heat resistance as the critical DPA release temperature (Tc), the temperature at which half the DPA content has been released within a fixed incubation time. We found Tc values for spores from Bacillus strains 168, A163, and IC4 of 108°C, 121°C, and 131°C, respectively. On the basis of these observations, we developed a quantitative model that describes the time and temperature dependence of the experimentally determined extent of DPA release and spore inactivation. The model predicts a DPA release rate profile for each inactivated spore. In addition, it uncovers remarkable differences in the values for the temperature dependence parameters for the rate of spore inactivation, DPA release duration, and DPA release delay.  相似文献   

7.
As previously reported, gerP Bacillus subtilis spores were defective in nutrient germination triggered via various germinant receptors (GRs), and the defect was eliminated by severe spore coat defects. The gerP spores'' GR-dependent germination had a longer lag time between addition of germinants and initiation of rapid release of spores'' dipicolinic acid (DPA), but times for release of >90% of DPA from individual spores were identical for wild-type and gerP spores. The gerP spores were also defective in GR-independent germination by DPA with its associated Ca2+ divalent cation (CaDPA) but germinated better than wild-type spores with the GR-independent germinant dodecylamine. The gerP spores exhibited no increased sensitivity to hypochlorite, suggesting that these spores have no significant coat defect. Overexpression of GRs in gerP spores did lead to faster germination via the overexpressed GR, but this was still slower than germination of comparable gerP+ spores. Unlike wild-type spores, for which maximal nutrient germinant concentrations were between 500 μM and 2 mM for l-alanine and ≤10 mM for l-valine, rates of gerP spore germination increased up to between 200 mM and 1 M l-alanine and 100 mM l-valine, and at 1 M l-alanine, the rates of germination of wild-type and gerP spores with or without all alanine racemases were almost identical. A high pressure of 150 MPa that triggers spore germination by activating GRs also triggered germination of wild-type and gerP spores identically. All these results support the suggestion that GerP proteins facilitate access of nutrient germinants to their cognate GRs in spores'' inner membrane.  相似文献   

8.
Germination of Bacillus anthracis spores occurs when nutrients such as amino acids or purine nucleosides stimulate specific germinant receptors located in the spore inner membrane. The gerPABCDEF operon has been suggested to play a role in facilitating the interaction between germinants and their receptors in spores of Bacillus subtilis and Bacillus cereus. B. anthracis mutants containing deletions in each of the six genes belonging to the orthologue of the gerPABCDEF operon, or deletion of the entire operon, were tested for their ability to germinate. Deletion of the entire gerP operon resulted in a significant delay in germination in response to nutrient germinants. These spores eventually germinated to levels equivalent to wild-type, suggesting that an additional entry point for nutrient germinants may exist. Deletions of each individual gene resulted in a similar phenotype, with the exception of ΔgerPF, which showed no obvious defect. The removal of two additional gerPF-like orthologues was necessary to achieve the germination defect observed for the other mutants. Upon physical removal of the spore coat, the mutant lacking the full gerP operon no longer exhibited a germination defect, suggesting that the GerP proteins play a role in spore coat permeability. Additionally, each of the gerP mutants exhibited a severe defect in calcium-dipicolinic acid (Ca-DPA)–dependent germination, suggesting a role for the GerP proteins in this process. Collectively, these data implicate all GerP proteins in the early stages of spore germination.  相似文献   

9.
Pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) constitutes approximately 10% of Bacillus subtilis spore dry weight and has been shown to play a significant role in the survival of B. subtilis spores exposed to wet heat and to 254-nm UV radiation in the laboratory. However, to date, no work has addressed the importance of DPA in the survival of spores exposed to environmentally relevant solar UV radiation. Air-dried films of spores containing DPA or lacking DPA due to a null mutation in the DPA synthetase operon dpaAB were assayed for their resistance to UV-C (254 nm), UV-B (290 to 320 nm), full-spectrum sunlight (290 to 400 nm), and sunlight from which the UV-B portion was filtered (325 to 400 nm). In all cases, air-dried DPA-less spores were significantly more UV sensitive than their isogenic DPA-containing counterparts. However, the degree of difference in UV resistance between the two strains was wavelength dependent, being greatest in response to radiation in the UV-B portion of the spectrum. In addition, the inactivation responses of DPA-containing and DPA-less spores also depended strongly upon whether spores were exposed to UV as air-dried films or in aqueous suspension. Spores lacking the gerA, gerB, and gerK nutrient germination pathways, and which therefore rely on chemical triggering of germination by the calcium chelate of DPA (Ca-DPA), were also more UV sensitive than wild-type spores to all wavelengths tested, suggesting that the Ca-DPA-mediated spore germination pathway may consist of a UV-sensitive component or components.  相似文献   

10.
A major event in the nutrient germination of spores of Bacillus species is release of the spores'' large depot of dipicolinic acid (DPA). This event is preceded by both commitment, in which spores continue through germination even if germinants are removed, and loss of spore heat resistance. The latter event is puzzling, since spore heat resistance is due largely to core water content, which does not change until DPA is released during germination. We now find that for spores of two Bacillus species, the early loss in heat resistance during germination is most likely due to release of committed spores'' DPA at temperatures not lethal for dormant spores. Loss in spore acid resistance during germination also paralleled commitment and was also associated with the release of DPA from committed spores at acid concentrations not lethal for dormant spores. These observations plus previous findings that DPA release during germination is preceded by a significant release of spore core cations suggest that there is a significant change in spore inner membrane permeability at commitment. Presumably, this altered membrane cannot retain DPA during heat or acid treatments innocuous for dormant spores, resulting in DPA-less spores that are rapidly killed.  相似文献   

11.
Spores of Bacillus species are said to be committed when they continue through nutrient germination even when germinants are removed or their binding to spores'' nutrient germinant receptors (GRs) is both reversed and inhibited. Measurement of commitment and the subsequent release of dipicolinic acid (DPA) during nutrient germination of spores of Bacillus cereus and Bacillus subtilis showed that heat activation, increased nutrient germinant concentrations, and higher average levels of GRs/spore significantly decreased the times needed for commitment, as well as lag times between commitment and DPA release. These lag times were also decreased dramatically by the action of one of the spores'' two redundant cortex lytic enzymes (CLEs), CwlJ, but not by the other CLE, SleB, and CwlJ action did not affect the timing of commitment. The timing of commitment and the lag time between commitment and DPA release were also dependent on the specific GR activated to cause spore germination. For spore populations, the lag times between commitment and DPA release were increased significantly in spores that germinated late compared to those that germinated early, and individual spores that germinated late may have had lower appropriate GR levels/spore than spores that germinated early. These findings together provide new insight into the commitment step in spore germination and suggest several factors that may contribute to the large heterogeneity among the timings of various events in the germination of individual spores in spore populations.Spores of Bacillus species can remain dormant for long times and are extremely resistant to a variety of environmental stresses (26). However, under appropriate conditions, normally upon the binding of specific nutrients to spores'' nutrient germinant receptors (GRs), spores can come back to active growth through a process called germination followed by outgrowth (19, 20, 25, 26). Germination of Bacillus subtilis spores can be triggered by l-alanine or l-valine or a combination of l-asparagine, d-glucose, d-fructose, and K+ (AGFK). These nutrient germinants trigger germination by binding to and interacting with GRs that have been localized to the spore''s inner membrane (12, 20). l-Alanine and l-valine bind to the GerA GR, while the AGFK mixture triggers germination by interacting with both the GerB and GerK GRs (25). Normally, l-asparagine alone does not trigger B. subtilis spore germination. However, a mutant form of the GerB GR, termed GerB*, displays altered germinant specificity such that l-asparagine alone will trigger the germination of gerB* mutant spores (1, 18).A number of events occur in a defined sequence during spore germination. Initially, exposure of spores to nutrient germinants causes a reaction that commits spores to germinate, even if the germinant is removed or displaced from its cognate GR (7, 10, 21, 27, 28). This commitment step is followed by release of monovalent cations, as well as the spore core''s large pool of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) along with divalent cations, predominantly Ca2+, that are chelated with DPA (Ca-DPA). In Bacillus spores, the release of Ca-DPA triggers the hydrolysis of spores'' peptidoglycan cortex by either of two cortex lytic enzymes (CLEs), CwlJ and SleB (11, 16, 23). CwlJ is activated during germination by Ca-DPA as it is being released from individual spores, while SleB activation requires that most Ca-DPA be released (14, 16, 17). Cortex hydrolysis, in turn, allows the spore core to expand and fully hydrate, which leads to activation of enzymes and initiation of metabolism in the spore core (21, 25).As noted above, commitment is the first event that can be assessed during spore germination, although the precise mechanism of commitment is not known. Since much has been learned about proteins important in spore germination in the many years since commitment was last studied (25, 26), it seemed worth reexamining commitment, with the goal of determining those factors that influence this step in the germination process. Knowledge of factors important in determining kinetics of commitment could then lead to an understanding of what is involved in this reaction.Kinetic analysis of spore germination, as well as commitment, has mostly been based on the decrease in optical density at 600 nm (OD600) of spore suspensions, which monitors a combination of events that occur well after commitment, including DPA release, cortex hydrolysis, and core swelling (25-27). In the current work, we have used a germination assay that measures DPA release, an early event in spore germination, and have automated this assay to allow routine measurement of commitment, as well as DPA release from large numbers of spore samples simultaneously. This assay has allowed comparison of the kinetics of DPA release and commitment during germination and study of the effects of heat activation, germinant concentration, GR levels, and CLEs on commitment.  相似文献   

12.
The Bacillus cereus spore surface layers consist of a coat surrounded by an exosporium. We investigated the interplay between the sporulation temperature and the CotE morphogenetic protein in the assembly of the surface layers of B. cereus ATCC 14579 spores and on the resulting spore properties. The cotE deletion affects the coat and exosporium composition of the spores formed both at the suboptimal temperature of 20°C and at the optimal growth temperature of 37°C. Transmission electron microscopy revealed that ΔcotE spores had a fragmented and detached exosporium when formed at 37°C. However, when produced at 20°C, ΔcotE spores showed defects in both coat and exosporium attachment and were susceptible to lysozyme and mutanolysin. Thus, CotE has a role in the assembly of both the coat and exosporium, which is more important during sporulation at 20°C. CotE was more represented in extracts from spores formed at 20°C than at 37°C, suggesting that increased synthesis of the protein is required to maintain proper assembly of spore surface layers at the former temperature. ΔcotE spores formed at either sporulation temperature were impaired in inosine-triggered germination and resistance to UV-C and H2O2 and were less hydrophobic than wild-type (WT) spores but had a higher resistance to wet heat. While underscoring the role of CotE in the assembly of B. cereus spore surface layers, our study also suggests a contribution of the protein to functional properties of additional spore structures. Moreover, it also suggests a complex relationship between the function of a spore morphogenetic protein and environmental factors such as the temperature during spore formation.  相似文献   

13.
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   

14.
We have studied pressure-induced germination of Bacillus subtilis spores at moderate (100 MPa) and high (500 to 600 MPa) pressures. Although we found comparable germination efficiencies under both conditions by using heat sensitivity as a criterion for germination, the sensitivity of pressure-germinated spores to some other agents was found to depend on the pressure used. Spores germinated at 100 MPa were more sensitive to pressure (>200 MPa), UV light, and hydrogen peroxide than were those germinated at 600 MPa. Since small, acid-soluble proteins (SASPs) and dipicolinic acid (DPA) are known to be involved in spore resistance to UV light and hydrogen peroxide, we studied the fate of these compounds during pressure germination. DPA was released upon both low- and high-pressure germination, but SASP degradation, which normally accompanies nutrient-induced germination, occurred upon low-pressure germination but not upon high-pressure germination. These results adequately explain the UV and hydrogen peroxide resistance of spores germinated at 600 MPa. The resistance to pressure inactivation of 600-MPa-germinated spores could also, at least partly, be attributed to α/β-type SASPs, since mutants deficient in α/β-type SASPs were more sensitive to inactivation at 600 MPa. Further, germination at 100 MPa resulted in rapid ATP generation, as is the case in nutrient-induced germination, but no ATP was formed during germination at 600 MPa. These results suggest that spore germination can be initiated by low- and high-pressure treatments but is arrested at an early stage in the latter case. The implications for the use of high pressure as a preservation treatment are discussed.  相似文献   

15.
Germination of dormant Bacillus subtilis spores with specific nutrient germinants is dependent on a number of inner membrane (IM) proteins, including (i) the GerA, GerB, and GerK germinant receptors (GRs) that respond to nutrient germinants; (ii) the GerD protein, essential for optimal GR function; and (iii) SpoVA proteins, essential for the release of the spore-specific molecule dipicolinic acid (DPA) during spore germination. Levels of GR A and C subunit proteins, GerD, and SpoVAD in wild-type spores were determined by Western blot analysis of spore fractions or total disrupted spores by comparison with known amounts of purified proteins. Surprisingly, after disruption of decoated B. subtilis spores with lysozyme and fractionation, ∼90% of IM fatty acids and GR subunits remained with the spores'' insoluble integument fraction, indicating that yields of purified IM are low. The total lysate from disrupted wild-type spores contained ∼2,500 total GRs/spore: GerAA and GerAC subunits each at ∼1,100 molecules/spore and GerBC and GerKA subunits each at ∼700 molecules/spore. Levels of the GerBA subunit determined previously were also predicted to be ∼700 molecules/spore. These results indicate that the A/C subunit stoichiometry in GRs is most likely 1:1, with GerA being the most abundant GR. GerD and SpoVAD levels were ∼3,500 and ∼6,500 molecules/spore, respectively. These values will be helpful in formulating mathematic models of spore germination kinetics as well as setting lower limits on the size of the GR-GerD complex in the spores'' IM, termed the germinosome.  相似文献   

16.
Populations of Bacillus subtilis spores in which 90 to 99.9% of the spores had been killed by moist heat gave only two fractions on equilibrium density gradient centrifugation: a fraction comprised of less dense spores that had lost their dipicolinic acid (DPA), undergone significant protein denaturation, and were all dead and a fraction with the same higher density as that of unheated spores. The latter fraction from heat-killed spore populations retained all of its DPA, but ≥98% of the spores could be dead. The dead spores that retained DPA germinated relatively normally with nutrient and nonnutrient germinants, but the outgrowth of these germinated spores was significantly compromised, perhaps because they had suffered damage to some proteins such that metabolic activity during outgrowth was greatly decreased. These results indicate that DPA release takes place well after spore killing by moist heat and that DPA release during moist-heat treatment is an all-or-nothing phenomenon; these findings also suggest that damage to one or more key spore proteins causes spore killing by moist heat.  相似文献   

17.
Aims: To analyse the germination and its heterogeneity of individual spores of Clostridium perfringens. Methods and Results: Germination of individual wild‐type Cl. perfringens spores was followed by monitoring Ca‐dipicolinic acid (CaDPA) release and by differential interference contrast (DIC) microscopy. Following the addition of KCl that acts via germinant receptors (GRs), there was a long variable lag period (Tlag) with slow release of c. 25% of CaDPA, then rapid release of remaining CaDPA in c. 2 min (ΔTrelease) and a parallel decrease in DIC image intensity, and a final decrease of c. 25% in DIC image intensity during spore cortex hydrolysis. Spores lacking the essential cortex‐lytic enzyme (CLE) (sleC spores) exhibited the same features during GR‐dependent germination, but with longer average Tlag values, and no decrease in DIC image intensity because of cortex hydrolysis after full CaDPA release. The Tlag of wild‐type spores in KCl germination was increased significantly by lower germinant concentrations and suboptimal heat activation. Wild‐type and sleC spores had identical average Tlag and ΔTrelease values in dodecylamine germination that does not utilize GRs. Conclusions: Most of these results were essentially identical to those reported for the germination of individual spores of Bacillus species. However, individual sleC Cl. perfringens spores germinated inefficiently with either KCl or exogenous CaDPA, in contrast to CLE‐deficient Bacillus spores, indicating that germination of these species’ spores is not completely identical. Significance and Impact of the Study: This work provides information on the kinetic germination and its heterogeneity of individual spores of Cl. perfringens.  相似文献   

18.
The inactivation of bacterial endospores by hydrostatic pressure requires the combined application of heat and pressure. We have determined the resistance of spores of 14 food isolates and 5 laboratory strains of Bacillus subtilis, B. amyloliquefaciens, and B. licheniformis to treatments with pressure and temperature (200 to 800 MPa and 60 to 80°C) in mashed carrots. A large variation in the pressure resistance of spores was observed, and their reduction by treatments with 800 MPa and 70°C for 4 min ranged from more than 6 log units to no reduction. The sporulation conditions further influenced their pressure resistance. The loss of dipicolinic acid (DPA) from spores that varied in their pressure resistance was determined, and spore sublethal injury was assessed by determination of the detection times for individual spores. Treatment of spores with pressure and temperature resulted in DPA-free, phase-bright spores. These spores were sensitive to moderate heat and exhibited strongly increased detection times as judged by the time required for single spores to grow to visible turbidity of the growth medium. The role of DPA in heat and pressure resistance was further substantiated by the use of the DPA-deficient mutant strain B. subtilis CIP 76.26. Taken together, these results indicate that inactivation of spores by combined pressure and temperature processing is achieved by a two-stage mechanism that does not involve germination. At a pressure between 600 and 800 MPa and a temperature greater than 60°C, DPA is released predominantly by a physicochemical rather than a physiological process, and the DPA-free spores are inactivated by moderate heat independent of the pressure level. Relevant target organisms for pressure and temperature treatment of foods are proposed, namely, strains of B. amyloliquefaciens, which form highly pressure-resistant spores.  相似文献   

19.
Berberine, an alkaloid originally extracted from the plant Coptis chinensis and other herb plants, has been used as a pharmacological substance for many years. The therapeutic effect of berberine has been attributed to its interaction with nucleic acids and blocking cell division. However, levels of berberine entering individual microbial cells minimal for growth inhibition and its effects on bacterial spores have not been determined. In this work the kinetics and levels of berberine accumulation by individual dormant and germinated spores were measured by laser tweezers Raman spectroscopy and differential interference and fluorescence microscopy, and effects of berberine on spore germination and outgrowth and spore and growing cell viability were determined. The major conclusions from this work are that: (1) colony formation from B. subtilis spores was blocked ~ 99% by 25 μg/mL berberine plus 20 μg/mL INF55 (a multidrug resistance pump inhibitor); (2) 200 μg/mL berberine had no effect on B. subtilis spore germination with L-valine, but spore outgrowth was completely blocked; (3) berberine levels accumulated in single spores germinating with ≥ 25 μg/mL berberine were > 10 mg/mL; (4) fluorescence microscopy showed that germinated spores accumulated high-levels of berberine primarily in the spore core, while dormant spores accumulated very low berberine levels primarily in spore coats; and (5) during germination, uptake of berberine began at the time of commitment (T1) and reached a maximum after the completion of CaDPA release (Trelease) and spore cortex lysis (Tlysis).  相似文献   

20.
The infectious agent of the disease anthrax is the spore of Bacillus anthracis. Bacterial spores are extremely resistant to environmental stresses, which greatly hinders spore decontamination efforts. The spore cortex, a thick layer of modified peptidoglycan, contributes to spore dormancy and resistance by maintaining the low water content of the spore core. The cortex is degraded by germination-specific lytic enzymes (GSLEs) during spore germination, rendering the cells vulnerable to common disinfection techniques. This study investigates the relationship between SleB, a GSLE in B. anthracis, and YpeB, a protein necessary for SleB stability and function. The results indicate that ΔsleB and ΔypeB spores exhibit similar germination phenotypes and that the two proteins have a strict codependency for their incorporation into the dormant spore. In the absence of its partner protein, SleB or YpeB is proteolytically degraded soon after expression during sporulation, rather than escaping the developing spore. The three PepSY domains of YpeB were examined for their roles in the interaction with SleB. YpeB truncation mutants illustrate the necessity of a region beyond the first PepSY domain for SleB stability. Furthermore, site-directed mutagenesis of highly conserved residues within the PepSY domains resulted in germination defects corresponding to reduced levels of both SleB and YpeB in the mutant spores. These results identify residues involved in the stability of both proteins and reiterate their codependent relationship. It is hoped that the study of GSLEs and interacting proteins will lead to the use of GSLEs as targets for efficient activation of spore germination and facilitation of spore cleanup.  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