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.
The content of dipicolinic acid (DPA) was assayed in the spores of Bacillus anthracoides 96 during various stages of its growth. The content of DPA was ca. 10.7 per cent of the dry biomass weight in the seven-day-old culture containing 96 to 99 per cent of the spores in a "starvation" medium. The morphology of the culture was modified, and the content of DPA in the spores fell to 3.6 per cent half an hour after the inoculation into the medium favourable for the growth (MPA). During the following one to four hours of the germination, the refraction index of the spores and the content of DPA in them decreased (the content of DPA to 2 per cent).  相似文献   

2.
Non-enzymatic formation of dipicolinic acid (DPA) from diketopimelic acid and ammonia was clearly demonstrated using a new method for DPA analysis. The reaction rates of DPA formation were almost the same under aerobic and anaerobic conditions. Nearly equimolecular quantities of DPA and tetrahydrodipicolinic acid were detected in spontaneous reaction mixture. The spontaneous reaction seemed to be due to dismutation of dihydrodipicolinic acid, resulting in DPA and tetrahydrodipicolinic acid. The apparent optimum pH of the spontaneous reaction was 8.2 and the maximal rate of DPA formation was observed with a 1 : 4 molar ratio of diketopimelic acid to ammonia. The rate of the spontaneous reaction was stimulated by ferrous sulfate, FMN, and riboflavin. Dihydrodipicolinate reductase catalyzes the reduction of dihydrodipicolinate, prepared from pyruvate and aspartic beta-semialdehyde, with NADPH as reductant. The reductase was isolated from Bacillus subtilis, and found to stimulate DPA formation from diketopimelic acid and ammonia. The enzymatic DPA formation was absolutely dependent on oxygen, and optimum pH was 6.4. The catalytic action of the enzyme was similar to that of the oxidase. Possible mechanisms of DPA formation from diketopimelic acid and ammonia are proposed.  相似文献   

3.
AIMS: To determine the mechanism of killing of Bacillus subtilis spores by hydrogen peroxide. METHODS AND RESULTS: Killing of spores of B. subtilis with hydrogen peroxide caused no release of dipicolinic acid (DPA) and hydrogen peroxide-killed spores were not appreciably sensitized for DPA release upon a subsequent heat treatment. Hydrogen peroxide-killed spores appeared to initiate germination normally, released DPA and hydrolysed significant amounts of their cortex. However, the germinated killed spores did not swell, did not accumulate ATP or reduced flavin mononucleotide and the cores of these germinated spores were not accessible to nucleic acid stains. CONCLUSIONS: These data indicate that treatment with hydrogen peroxide results in spores in which the core cannot swell properly during spore germination. SIGNIFICANCE AND IMPACT OF THE STUDY: The results provide further information on the mechanism of killing of spores of Bacillus species by hydrogen peroxide.  相似文献   

4.
AIMS: To determine the mechanisms of killing of Bacillus subtilis spores by ethanol or strong acid or alkali. METHODS AND RESULTS: Killing of B. subtilis spores by ethanol or strong acid or alkali was not through DNA damage and the spore coats did not protect spores against these agents. Spores treated with ethanol or acid released their dipicolinic acid (DPA) in parallel with spore killing and the core wet density of ethanol- or acid-killed spores fell to a value close to that for untreated spores lacking DPA. The core regions of spores killed by these two agents were stained by nucleic acid stains that do not penetrate into the core of untreated spores and acid-killed spores appeared to have ruptured. Spores killed by these two agents also did not germinate in nutrient and non-nutrient germinants and were not recovered by lysozyme treatment. Spores killed by alkali did not lose their DPA, did not exhibit a decrease in their core wet density and their cores were not stained by nucleic acid stains. Alkali-killed spores released their DPA upon initiation of spore germination, but did not initiate metabolism and degraded their cortex very poorly. However, spores apparently killed by alkali were recovered by lysozyme treatment. CONCLUSIONS: The data suggest that spore killing by ethanol and strong acid involves the disruption of a spore permeability barrier, while spore killing by strong alkali is due to the inactivation of spore cortex lytic enzymes.SIGNIFICANCE AND IMPACT OF THE STUDY: The results provide further information on the mechanisms of spore killing by various chemicals.  相似文献   

5.
The ability of ethylenediaminetetraacetic acid (EDTA) and dipicolinic acid (DPA) to rupture Azotobacter cysts was directly related to the molarity of the reaction mixture. Cysts were not ruptured at concentrations above 0.15 m (NaCl or phosphate). These higher salt concentrations prevented the formation of metallo:EDTA (or DPA) chelate complexes. Sudden shifts in the tonicity of NaCl or phosphate mixtures induced cyst rupture in the absence of EDTA. The effect of these mechanisms on the survival and germination of Azotobacter cysts in the soil is discussed.  相似文献   

6.
The role of dipicolinic acid (DPA) in determining the resistance of Bacillus cereus spores to u.v. and gamma radiation was investigated. B. cereus BIS-59 spores containing varying amounts of DPA were prepared by appropriate compositional adjustments in the secondary media. Compared with spores containing 6 per cent DPA (dry weight) those containing 0.8 per cent DPA were far more sensitive to u.v. radiation. Similar u.v. radiation sensitivity was also found in respect of a DPA-less mutant of B. cereus T 6A 1. Pre-treatment of DPA deficient spores (of wild type or mutant B. cereus) with DPA or the presence of DPA during irradiation resulted in increased resistance of these spores to u.v. radiation. In the range 0.2 to 1 per cent DPA content of spores of B. cereus BIS-59, a striking inverse relationship could be discerned between the DPA content and the number of spore photo-products (5-thymidyl, 5,6-dihydrothymine) formed in DNA and spore viability. The resistance of B. cereus spores to gamma radiation did not seem to be influenced by their DPA content.  相似文献   

7.
A potentiometric chemosensor for selective determination of dipicolinic acid (2,6-pyridinedicarboxylic acid, DPA) was developed based on the surface imprinting technique coupled with a nanoscale transducer: an indium tin oxide (ITO)-coated glass plate. The sensor fabrication conditions, optimal recognition condition, as well as selectivity, sensitivity, and stability of the DPA sensor have been investigated. The DPA sensor could recognize DPA from 3,5-pyridinedicarboxylic acid. Potentiometric measurements demonstrated selective detection of DPA in a concentration range of 1.5 x 10(-6) to 0.0194 M. The response time of DPA sensor for 4 x 10(-4) M DPA was 25 s. The potentiometric response of the DPA sensor to DPA is at 90% of its initial magnitude after 550 times measurement. The viability of such a modified ITO electrode in the presence of other inorganic, organic, and biological materials was probed.  相似文献   

8.
A Bacillus subtilis mutant is described which forms heat-resistant spores only in the presence of external dipicolinic acid (DPA). The mutation, dpa-1, is localized in a new sporulation locus, linked to pyrA. The dpa-1 strain is unable to synthesize DPA but can incorporate external DPA. The amount of DPA incorporated, the frequency of heat-resistant spores and their degree of resistance are all dependent on the concentration of external DPA. Spores of dpa- 1 strains exhibit normal resistance to most chemicals, including octanol and chloroform, but not to ethanol, pyridine, phenol and trichloroacetic acid. Complete resistance to the latter group depends on DPA. DPA incorporation is slow and apparently requires an energy supply but not protein synthesis. Direct involvement of DPA in the heat-resistance of the spores is suggested. Thin sections of DPA-less spores exhibit clearly visible cytoplasmic membranes and ribosomes. These structures are absent or less visible in the core of spores obtained with added DPA.  相似文献   

9.
Lysine biosynthetic pathway enzymes of Bacillus brevis ATCC 1068 were studied as a function of stage of development (growth and sporulation). The synthesis of aspartic-2-eemialdehyde dehydrogenase (ASA-dehydrogenase), dihydrodipicolinate synthase (DHDPA-synthase), DHPA-reductase and diaminopimelate decarboxylase (DAP-decarboxylase) was found not to be co-regulated, since lysine was not a co-repressor for these enzymes. Unlike the aspartokinase isoenzymes, the other enzymes of the lysine pathway were not derepressed in thiosine-resistant, lysine-excreting mutants. Thus, the aspartokinase isoenzymes were the key enzymes during growth and regulation of lysine biosynthesis through restriction of l-ASA synthesis via feedback control by lysine on the aspartokinases was therefore suggested.In contrast to other Bacillus species, the levels of the lysine biosynthetic pathway enzymes of strain ATCC 10068 were not derepressed during the change from vegetative growth to sporulation. Two control mechanisms, enabling the observed preferential channelling of carbon for the synthesis of spore-specific diaminopimelic acid (DAP) and dipicolinic acid (DPA) were a) loss of DAP-decarboxylase, b) inhibition of DHDPA-reductase by DPA. Increase in the level of the DAP pool during sporulation, as a consequence of the loss of DAP-decarboxylase, and its relevance to the non-enzymatic formation of DPA has been discussed.Abbreviations l-ASA l-aspartic-2-semialdehyde - DAP diaminopimelic acid - DPA dipicolinic acid - DHDPA dihydrodipicolinate - AGM aspargine-glycerol medium - PY peptone-yeast extract - NB+NSM nutrient broth plus nutrient sporulation medium  相似文献   

10.
When they were subjected to a range of physical and chemical treatments, spores of Pasteuria penetrans showed properties similar to those of other endospore-forming bacteria. The spores did not take up some stains, were resistant to desiccation and sonication and showed extrusion of spore contents ('spore popping') on prolonged exposure to 0.1% KMnO4 in 0.3 n HNO3. Calcium and dipicolinic acid (DPA) were present at concentrations of 0.28% and 0.96% of the spore dry weight respectively, giving a Ca: DPA molar ratio of 1.2. The infectivity of P. penetrans spores was reduced to a low level after heating at 100°C for 5 min, but spore attachment was not markedly affected by heating at 100°C for 15 min. Evidence for the presence of catalase in P. penetrans spores was equivocal because the low levels of catalase activity observed in spore suspensions may have been due to contamination from catalase-positive nematode tissue. When P. penetrans spores were exposed to a range of substances known to act as germinants for spores of Bacillus spp., germination or loss of refractility was not observed by phase microscopy. In vitro culture of P. penetrans was attempted by inoculating either spores or vegetative mycelial bodies onto a diverse range of simple and complex media and incubating them in aerobic, reduced oxygen, anaerobic and increased CO2 environments. Signs of spore germination or growth of vegetative stages were never observed.  相似文献   

11.
Bacillus subtilis cells with mutations in the spoVA operon do not complete sporulation. However, a spoVA strain with mutations that remove all three of the spore's functional nutrient germinant receptors (termed the ger3 mutations) or the cortex lytic enzyme SleB (but not CwlJ) did complete sporulation. ger3 spoVA and sleB spoVA spores lack dipicolinic acid (DPA) and have lower core wet densities and levels of wet heat resistance than wild-type or ger3 spores. These properties of ger3 spoVA and sleB spoVA spores are identical to those of ger3 spoVF and sleB spoVF spores that lack DPA due to deletion of the spoVF operon coding for DPA synthetase. Sporulation in the presence of exogenous DPA restored DPA levels in ger3 spoVF spores to 53% of the wild-type spore levels, but there was no incorporation of exogenous DPA into ger3 spoVA spores. These data indicate that one or more products of the spoVA operon are involved in DPA transport into the developing forespore during sporulation.  相似文献   

12.
AIMS: To measure rates of release of small molecules during pressure germination of Bacillus subtilis spores, and the role of SpoVA proteins in dipicolinic acid (DPA) release. METHODS AND RESULTS: Rates of DPA release during B. subtilis spore germination with pressures of 150 or 500 megaPascals were much higher in spores with elevated levels of SpoVA proteins, and spores with a temperature-sensitive mutation in the spoVA operon were temperature-sensitive in DPA release during pressure germination. Spores also released arginine and glutamic acid, but not AMP, during pressure germination. CONCLUSIONS: Pressure germination of B. subtilis spores causes release of many small molecules including DPA. SpoVA proteins are involved in the release of DPA, perhaps because SpoVA proteins are a component of a DPA channel in the spore's inner membrane. SIGNIFICANCE AND IMPACT OF THE STUDY: This work provides new insight into the mechanism of pressure germination of spores of Bacillus species, a process that has significant potential for usage in the food industry.  相似文献   

13.
Rough and smooth variants of Bacillus stearothermophilus strains ATCC 12976 and 12980 were isolated. These variants showed morphologically different colonies. In both strains the rough variants were less heat-resistant than the smooth ones, and their activation occurred in a shorter time; on the other hand, their dipicolinic acid (DPA) content was higher. These results indicate that a relationship between higher DPA content and higher thermal resistance does not exist. Therefore, this evidence supports the hypothesis that DPA is not the determining factor in the degree of spore heat resistance but that it could, instead, have a role in maintaining thermal resistance obtained by other means.  相似文献   

14.
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   

15.
A novel method is described that allows a direct determination of the location of a portion of the dipicolinic acid (DPA) in spores. Ultraviolet (UV) irradiation is shown to cause cross-linkage of DPA to spore proteins which have the characteristics of membrane proteins. We suggest that DPA resides in the inner forespore membrane (IFSM) and spore cytoplasm (i.e., the spore protoplast). Only that portion of the DPA in the vicinity of the IFSM appeared to form UV-induced DPA-protein adducts.  相似文献   

16.
A high-performance liquid chromatographic method with indirect fluorescence detection has been developed for quantification of dipicolinic acid, a major constituent of bacterial endospores. After separation on a reversed-phase column, a post-column reagent of sodium acetate at 1 mol l(-1) with 50 micromol l(-1) terbium chloride was added for complexation of dipicolinic acid. Terbium monodipicolinate complexes formed were quantified by measuring the fluorescence emission maximum at 548 nm after excitation with UV light at 270 nm wavelength. Parameters of post-column complexation were optimized to achieve a detection limit of 0.5 nmol DPA l(-1), corresponding to about 10(3) Desulfosporosinus orientis endospores per ml. The method was applied to the analysis of spore contamination in tuna and for estimating the endospore numbers in marine sediments.  相似文献   

17.
The spectroscopic and electrochemiluminescence (ECL) properties of dipicolinic acid (DPA), (bpy)(2)Ru(2+) (bpy = 2,2'-bipyridine) and the species formed when DPA and (bpy)(2)Ru(2+) [abbreviated to (bpy)(2)Ru(DPA)(+)] are allowed to react are reported. The UV-Vis absorption maxima for (bpy)(2)Ru(2+) and (bpy)(2)Ru(DPA)(+) are 493 and 475 nm, respectively, indicating the in situ formation of a complex between DPA and (bpy)(2)Ru(2+). DPA, (bpy)(2)Ru(2+) and (bpy)(2)Ru(DPA)(+) display ECL upon oxidation in the presence of the oxidative-reductive co-reactant tri-n-propylamine (TPrA). The ECL of (bpy)(2)Ru(DPA)(+) is at least two-fold higher than either of the parent species. An ECL spectrum of (bpy)(2)Ru(DPA)(+) displays a peak maximum 40 nm red-shifted from the photoluminescence peak maximum, suggesting that the excited state formed electrochemically is different from that formed spectroscopically.  相似文献   

18.
19.
从人血中提取红细胞膜,用注射器加压推打的方法首次获得了包含80mmol/L吡啶二羧酸(DPA)的封闭完好的内翻外囊泡(IOVs).离心除去囊泡外DPA,即可按Newton法测其阴离子转运活性.此法在红细胞膜内翻外囊泡体系上成功地建立了带3蛋白(Band 3)测活方法,具有简便迅速,重复性好等优点.  相似文献   

20.
Spores prepared from different sporulating media containing varying amounts of Ca and dipicolinic acid (DPA), exhibited differential responses to germination in L-alanine (0.25 M). Ca-spores with moderately high Ca and DPA contents could be triggered to germination by L-alanine, whereas P-spores with low contents of Ca and DPA could not be germinated by L-alanine unless Ca2+ or DPA was exogenously added. The initiation of L-alanine induced germination by P-spores in the presence of 45CaCl2 was associated with a marked uptake of 45Ca2+. Experiments involving stepwise extraction of 45Ca from prelabelled spores indicated that a part of the spore calcium may be involved in L-alanine induced germination. Both Ca2+ and DPA seemed to have a stimulatory effect on the incorporation of 14C-L-alanine.  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