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.
Inactivation of the Bacillus subtilis sspF gene had no effect on sporulation, spore resistance, or germination in a wild-type strain or one lacking DNA protective alpha/beta-type small, acid-soluble proteins (SASP). Overexpression of SspF in wild-type spores or in spores lacking major alpha/beta-type SASP (alpha- beta- spores) had no effect on sporulation but slowed spore outgrowth and restored a small amount of UV and heat resistance to alpha- beta- spores. In vitro analyses showed that SspF is a DNA binding protein and is cleaved by the SASP-specific protease (GPR) at a site similar to that cleaved in alpha/beta-type SASP. SspF was also degraded during spore germination and outgrowth, and this degradation was initiated by GPR.  相似文献   

2.
Three mutant forms of the protease (GPR) that initiates degradation of small, acid-soluble spore proteins (SASP) during germination of spores of Bacillus species have been generated. In one variant (GPR delta), the putative pro sequence removed in conversion of the GPR zymogen (termed P46) to the active enzyme (termed P41) was deleted. GPR delta was expressed in both Escherichia coli and Bacillus subtilis as a polypeptide of 41 kDa (P41) which was active both in vivo and in vitro. The other two variants had changes in the sequence around the site where the pro sequence is removed, making this sequence even more like that recognized and cleaved by GPR in its SASP substrates. One of these variants (GPRS) was synthesized as P46S in both B. subtilis and E. coli, but P46S was processed to P41S earlier in B. subtilis sporulation than was wild-type P46. The second variant (GPREI) was made as P46EI but underwent extremely rapid processing to P41EI in both E. coli and B. subtilis. Expression of elevated (> 100-fold) levels of GPR delta or GPREI blocked sporulation at the time of synthesis of glucose dehydrogenase. Expression of elevated levels of GPRS or low levels (< 20% of the wild-type level) of GPR delta or GPREI did not retard sporulation, but the SASP level in the resultant spores was greatly reduced. Prolonged incubation of P41 delta, P41EI, or wild-type P41, either in vivo or with purified proteins in vitro, resulted in a second self-cleavage event generating a 39-kDa polypeptide termed P39. The sequence in the P(41)-->P(39) cleavage site was also quite similar to that recognized and cleaved by GPR in SASP. Together, these results strongly support a model in which activation of GPR during sporulation by conversion of P(46) to P(41) is a self-processing event triggered by a change in the spore core environment (i.e., dehydration) which precludes attack of the active P(41) on its SASP substrates. However, in the first minutes of spore germination, rapid spore core hydration allows rapid attack of active GPR on SASP.  相似文献   

3.
We have determined the amounts of a number of small molecules and enzymes in the mother cell compartment and the developing forespore during sporulation of Bacillus megaterium. Significant amounts of adenosine 5'-triphosphate and reduced nicotinamide adenine dinucleotide were present in the forespore compartment before accumulation of dipicolinic acid (DPA), but these compounds disappeared as DPA was accumulated. 3-Phosphoglyceric acid (3-PGA) accumulated only within the developing forespore, beginning 1 to 2 h before DPA accumulation. Throughout its development the forespore contained constant levels of enzymes of both 3-PGA synthesis (phosphoglycerate kinase and glyceraldehyde-3-phosphate dehydrogenase) and 3-PGA utilization (phosphoglycerate mutase, enolase, and pyruvate kinase) at levels similar to those in the mother cell and the dormant spore. Despite the presence of enzymes for 3-PGA utilization, this compound was stable within isolated forespores. Two acid-soluble proteins (A and B proteins) also accumulated only in the forespore, beginning 1 to 2 h before DPA accumulation. At this time the specific protease involved in degradation of the A and B proteins during germination also appeared, but only in the forespore compartment. Nevertheless, the A and B proteins were stable within isolated forespores. Arginine and glutamic acid accumulated within the forespore in parallel with DPA accumulation. The forespore also contained the enzyme arginase at a level similar to that in the mother cell and a level of glutamic acid decarboxylase 2- to 25-fold higher than that in the mother cell, depending on when in sporulation the forespores were isolated. The specific activities of several other enzymes (protease active on hemoglobin, ornithine transcarbamylase, malate dehydrogenase, aconitase, and isocitrate dehydrogenase) in forespores were about 10% or less of the values in the mother cell. Aminopeptidase was present at similar levels in both compartments; threonine deaminase was not found in either compartment.  相似文献   

4.
The DNA in the core of spores of Bacillus species is saturated with a group of small, acid-soluble proteins (SASP) that protect DNA from a variety of harsh treatments and play a major role in spore resistance and long-term spore survival. During spore germination, SASPs are rapidly degraded to amino acids and this degradation is initiated by a sequence-specific protease called germination protease (GPR), which exhibits no obvious mechanistic or amino acid sequence similarity to any known class of proteases. GPR is synthesized during sporulation as an inactive tetrameric zymogen termed P(46), which later autoprocesses to a smaller form termed P(41), which is active only during spore germination. Here, we report the crystal structure of P(46) from Bacillus megaterium at 3.0 A resolution and the fact that P(46) monomer adopts a novel fold. The asymmetric unit contains two P(46) monomers and the functional tetramer is a dimer of dimers, with an approximately 9 A channel in the center of the tetramer. Analysis of the P(46) structure and site-directed mutagenesis studies have provided some insight into the mechanism of zymogen activation as well as the zymogen's lack of activity and the inactivity of P(41) in the mature spore.  相似文献   

5.
Four genes (ssp genes) coding for small, acid-soluble spore proteins of Bacillus megaterium and the gene for the protease that cleaves them during germination were cloned in the integratable plasmid pJH101. Each plasmid was integrated into the B. megaterium chromosome by a Campbell-type mechanism, allowing mapping of all five genes. The gene for the small, acid-soluble spore protein-specific protease (gpr) mapped near rib, and the sspA gene mapped between argA and hisA. The three other genes of the spp gene family (sspB, -D, and -F) all mapped near metC/D, with the order: sspF-sspD-metC/D-hemA-argO-sspB. While neither gpr nor sspF has been mapped in B. subtilis, the positions of the sspA, -B, and -D loci are similar in B. megaterium and B. subtilis, suggesting that the members of this multigene family have not recently undergone significant movement on the chromosome. It appears that more gene rearrangement has occurred in the flanking genes than has occurred in the ssp family of genes producing the small, acid-soluble spore proteins.  相似文献   

6.
7.
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   

8.
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   

9.
The small acid-soluble spore proteins alpha and beta were not detected during stationary-phase growth of asporogenous Bacillus subtilis mutants blocked in stages 0, II, or III, but mutants blocked in stages IV or V accumulated nearly wild-type levels of these small acid-soluble spore proteins. Similar results were obtained when production of Bacillus megaterium C protein (also a small acid-soluble spore protein), as well as alpha and beta, were monitored in these mutants containing a recombinant plasmid carrying the B. megaterium C protein gene. The only exception was a spo0H mutant which synthesized a small amount of C protein, but no alpha or beta.  相似文献   

10.
Dormant spores of Bacillus megaterium were found to contain approximately 850 pmol of coenzyme A (CoA) per milligram of dry weight. Of this total, less than 1.5% was acetyl-CoA, 25% was CoA-disulfide, 43% was in disulfide linkage to protein, and the remainder was the free thiol. Dormand spores of Bacillus cereus and Clostridium bifermentans contained 700 and 600 pmol of CoA per milligram of dry weight, respectively; in both species approximately 45% of the CoA 45% of the CoA was in disulfide linkage to protein. During germination of spores of all three species, greater than 75% of the CoA-protein disulfides were cleaved. In B. megaterium, cleavage of these disulfides during spore germination did not require exogenous metabolites and occurred at about the same time as the initiation of germination. Much of the CoA was converted to acetyl-CoA at this time. Dormant spores also contained reduced nicotinamide adenine dinucleotide-dependent CoA-disulfide reductase at levels higher than those in other stages of growth. The level of total CoA in the growing cells was two- to three-fold higher than in spores. This level remained constant throughout growth and sporulation, but less than 2% of the total cellular CoA was in disulfide linkage to protein until late in sporulation. The CoA-protein disulfides accumulated exclusively within the developing spore at about the time when dipicolinic acid was accumulated.  相似文献   

11.
12.
Spores of Bacillus subtilis with a mutation in spoVF cannot synthesize dipicolinic acid (DPA) and are too unstable to be purified and studied in detail. However, the spores of a strain lacking the three major germinant receptors (termed Deltager3), as well as spoVF, can be isolated, although they spontaneously germinate much more readily than Deltager3 spores. The Deltager3 spoVF spores lack DPA and have higher levels of core water than Deltager3 spores, although sporulation with DPA restores close to normal levels of DPA and core water to Deltager3 spoVF spores. The DPA-less spores have normal cortical and coat layers, as observed with an electron microscope, but their core region appears to be more hydrated than that of spores with DPA. The Deltager3 spoVF spores also contain minimal levels of the processed active form (termed P(41)) of the germination protease, GPR, a finding consistent with the known requirement for DPA and dehydration for GPR autoprocessing. However, any P(41) formed in Deltager3 spoVF spores may be at least transiently active on one of this protease's small acid-soluble spore protein (SASP) substrates, SASP-gamma. Analysis of the resistance of wild-type, Deltager3, and Deltager3 spoVF spores to various agents led to the following conclusions: (i) DPA and core water content play no role in spore resistance to dry heat, dessication, or glutaraldehyde; (ii) an elevated core water content is associated with decreased spore resistance to wet heat, hydrogen peroxide, formaldehyde, and the iodine-based disinfectant Betadine; (iii) the absence of DPA increases spore resistance to UV radiation; and (iv) wild-type spores are more resistant than Deltager3 spores to Betadine and glutaraldehyde. These results are discussed in view of current models of spore resistance and spore germination.  相似文献   

13.
During germination of spores of Bacillus species the degradation of the spore's pool of small, acid-soluble proteins (SASP) is initiated by a protease termed GPR, the product of the gpr gene. Bacillus megaterium and B. subtilis mutants with an inactivated gpr gene grew, sporulated, and triggered spore germination as did gpr+ strains. However, SASP degradation was very slow during germination of gpr mutant spores, and in rich media the time taken for spores to return to vegetative growth (defined as outgrowth) was much longer in gpr than in gpr+ spores. Not surprisingly, gpr spores had much lower rates of RNA and protein synthesis during outgrowth than did gpr+ spores, although both types of spores had similar levels of ATP. The rapid decrease in the number of negative supertwists in plasmid DNA seen during germination of gpr+ spores was also much slower in gpr spores. Additionally, UV irradiation of gpr B. subtilis spores early in germination generated significant amounts of spore photoproduct and only small amounts of thymine dimers (TT); in contrast UV irradiation of germinated gpr+ spores generated almost no spore photoproduct and three to four times more TT. Consequently, germinated gpr spores were more UV resistant than germinated gpr+ spores. Strikingly, the slow outgrowth phenotype of B. subtilis gpr spores was suppressed by the absence of major alpha/beta-type SASP. These data suggest that (i) alpha/beta-type SASP remain bound to much, although not all, of the chromosome in germinated gpr spores; (ii) the alpha/beta-type SASP bound to the chromosome in gpr spores alter this DNA's topology and UV photochemistry; and (iii) the presence of alpha/beta-type SASP on the chromosome is detrimental to normal spore outgrowth.  相似文献   

14.
The amino acid sequence-specific protease (termed GPR) in the bacterium Bacillus megaterium initiates the rapid degradation of small, acid-soluble spore proteins during the germination of spores of this organism. GPR is synthesized during spore formation as an inactive zymogen termed P46, which later autoprocesses to a smaller active form termed P41, which acts during spore germination. However, GPR exhibits no obvious mechanistic or amino acid sequence similarity to any of the known classes of proteases. To initiate the determination of the mechanisms of P46 to P41 conversion, P46 inactivity, and P41 catalysis, B. megaterium GPR has been overexpressed in Escherichia coli and purified to homogeneity by anion-exchange and size exclusion chromatography, and crystals of both P46 and P41 have been obtained by the vapor diffusion method. P46 crystals diffracted x rays to 3.5 A but the crystals of P41 diffracted x rays to only 6.5 A. A native x-ray diffraction data set of P46 has been collected; the unit cell parameters are a = b = 76.8, c = 313.1 A, alpha = beta = gamma = 90 degrees; the space group is tetragonal P41212 or P43212. The asymmetric unit contains two monomeric molecules with a crystal volume per unit protein mass of 2. 85 A3/Da and a solvent content of about 57%. An isomorphous heavy atom derivative data set has also been obtained for P46 crystals with potassium dicyanoaurate (I).  相似文献   

15.
After a few minutes of germination, nucleoids in the great majority of spores of Bacillus subtilis and Bacillus megaterium were ring shaped. The major spore DNA binding proteins, the alpha/beta-type small, acid-soluble proteins (SASP), colocalized to these nucleoid rings early in spore germination, as did the B. megaterium homolog of the major B. subtilis chromosomal protein HBsu. The percentage of ring-shaped nucleoids was decreased in germinated spores with lower levels of alpha/beta-type SASP. As spore outgrowth proceeded, the ring-shaped nucleoids disappeared and the nucleoid became more compact. This change took place after degradation of most of the spores' pool of major alpha/beta-type SASP and was delayed when alpha/beta-type SASP degradation was delayed. Later in spore outgrowth, the shape of the nucleoid reverted to the diffuse lobular shape seen in growing cells.  相似文献   

16.
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   

17.
Germination protease (GPR) initiates the degradation of small, acid-soluble spore proteins (SASP) during germination of spores of Bacillus and Clostridium species. The GPR amino acid sequence is not homologous to members of the major protease families, and previous work has not identified residues involved in GPR catalysis. The current work has focused on identifying catalytically essential amino acids by mutagenesis of Bacillus megaterium gpr. A residue was selected for alteration if it (i) was conserved among spore-forming bacteria, (ii) was a potential nucleophile, and (iii) had not been ruled out as inessential for catalysis. GPR variants were overexpressed in Escherichia coli, and the active form (P41) was assayed for activity against SASP and the zymogen form (P46) was assayed for the ability to autoprocess to P41. Variants inactive against SASP and unable to autoprocess were analyzed by circular dichroism spectroscopy and multi-angle laser light scattering to determine whether the variant's inactivity was due to loss of secondary or quaternary structure, respectively. Variation of D127 and D193, but no other residues, resulted in inactive P46 and P41, while variants of each form were well structured and tetrameric, suggesting that D127 and D193 are essential for activity and autoprocessing. Mapping these two aspartate residues and a highly conserved lysine onto the B. megaterium P46 crystal structure revealed a striking similarity to the catalytic residues and propeptide lysine of aspartic acid proteases. These data indicate that GPR is an atypical aspartic acid protease.  相似文献   

18.
The products of the hexacistronic spoVA operon of Bacillus subtilis may be involved in the transport of dipicolinic acid into the forespore during sporulation and its release during spore germination. The major hydrophilic coding region of B. subtilis spoVAD was cloned, the protein was expressed in Escherichia coli as a His tag fusion protein, and a rabbit antiserum was raised against the purified protein. Western blot analyses of fractions from B. subtilis spores showed that SpoVAD is an integral inner membrane protein present at levels >50-fold higher than those of the spore's nutrient germinant receptors that are also present in the inner membrane. SpoVAD also persisted in outgrowing spores.  相似文献   

19.
The protease that initiates rapid proteolysis during germination of Bacillus megaterium spores is synthesized during sporulation as a 46,000-molecular-weight polypeptide (P46) and is processed later in sporulation to a 41,000-molecular-weight polypeptide (P41), which is converted to a 40,000-molecular-weight polypeptide (P40) early in spore germination. P40 is known to be both tetrameric and enzymatically active. In this work, we show that P46 and P41 are both tetrameric, but that only P41 is enzymatically active. The identification of a zymogen form (P46) of this protease explains in part the regulation of the activity of this enzyme.  相似文献   

20.
P M Hauser  D Karamata 《Biochimie》1992,74(7-8):723-733
A reliable method for measuring the spore DNA content, based on radioactive DNA labelling, spore germination in absence of DNA replication and diphenylamine assay, was developed. The accuracy of the method, within 10-15%, is adequate for determining the number of chromosomes per spore, provided that the genome size is known. B subtilis spores were shown to be invariably monogenomic, while those of larger bacilli Bacillus megaterium, Bacillus cereus and Bacillus thuringiensis, often, if not invariably, contain two genomes. Attempts to modify the spore DNA content of B subtilis by altering the richness of the sporulation medium, the sporulation conditions (liquid or solid medium), or by mutation, were apparently unsuccessful. An increase of spore size with medium richness, not accompanied by an increase in DNA content, was observed. The implication of the apparently species-specific spore ploidy and the influence of the sporulation conditions on spore size and shape are discussed.  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