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.
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   

2.
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   

3.
The sequence-specific protease (termed GPR) that degrades small, acid-soluble proteins (SASP) during germination of spores of Bacillus species is synthesized during sporulation as an inactive precursor termed P46. Approximately 2 h later in sporulation, P46 is converted proteolytically to a smaller form, termed P41, which is active in vitro, but which does not act significantly on SASP in vivo until spore germination is initiated. While it appears likely that P46-->P41 conversion is an autoprocessing event, the mechanisms regulating P46-->P41 conversion in vivo are not clear. In this work we found that P46-->P41 conversion in vitro was stimulated tremendously in an allosteric manner by pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) plus Ca2+ but not by Ca2+ in combination with a variety of DPA analogs. The processing reaction stimulated by Ca(2+)-DPA was seen at pH 5.1 but not at pH 6.2 or 7, and under these conditions P46-->P41 conversion exhibited a linear time course and was a first-order reaction, indicative of an intramolecular autoprocessing reaction. At pH 5.1, P46-->P41 conversion was stimulated markedly by very high ionic strength. At pHs from 5.1 to 6.6, P46-->P41 conversion also occurred when P46 was dehydrated to approximately 54% relative humidity. This processing was stimulated markedly when dehydration was carried out in the presence of DPA and NaCl but was greatly decreased when dehydration was to 10, 33, or 75% relative humidity. Since previous work has shown that P(46)-->P(41) processing in vivo takes place (i) after a fall in forespore pH to 6.3 to 6.9 and approximately in parallel with (ii) DPA accumulation by the forespore and (iii) dehydration of the forespore, out new finings in vitro suggest that these three changes may synergistically trigger P(46)-->P(41) autoprocessing in the developing forespore. Presumably the conditions in vivo during this authoprocessing preclude significant attack of the P(41) generated on its SASP substrates.  相似文献   

4.
The protease which initiates the massive protein degradation early in bacterial spore germination has been purified from Bacillus megaterium spores. The enzyme has a molecular weight of 160,000 and contains four apparently identical subunits, but only the tetramer is enzymatically active. A radioimmunoassay has been developed for this enzyme and has been used to show that the protease is absent from growing cells, but appears early in sporulation within the developing forespore. In contrast, the protease antigen disappears rapidly during spore germination, in parallel with the loss in enzyme activity.  相似文献   

5.
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   

6.
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   

7.
8.
Analysis of protein degradation during the life cycle of Blastocladiella emersonii showed that (i) protein degradation is especially high during two phases of differentiation (sporulation, 12%/h and germination, 5%/h) in contrast with a much smaller degradation rate in the other phases (growth and zoospores, less than 1%/hr); (ii) protein degradation during germination in growth medium, as well as most of the germination process, is quantitatively unaffected by cycloheximide; (iii) a caseinolytic protease (pH optimum 5.5, apparent molecular weight 55,000 to 60,000) is present in extracts of zoospores and germinating cells; (iv) this protease activity is very low (perhaps absent) in extracts of late growth phase cells, but reappears during induced sporulation; (v) a different class of caseinolytic protease activity (pH optima 7 and 10; apparent molecular weight 25,000 to 30,000) is found in cellular extracts of late growth phase and early phases of sporulation; (vi) the latter class of enzyme activity is released into the medium during later phases of sporulation and is replaced in the cells by the former class. Speculations as to the roles of protein degradation in cell differentiation are discussed.  相似文献   

9.
The pepsin-like aspartyl proteases consist of a single polypeptide chain with topologically similar amino- and carboxyl-terminal domains, each of which contributes 1 aspartic acid residue to the active site. This structure has been proposed to have evolved by gene duplication and fusion from a dimeric enzyme composed of two identical polypeptide chains, such as the aspartyl protease (PRT) of human immunodeficiency virus type 1 (HIV-1). To determine if a single polypeptide form of the HIV-1 protease would be enzymatically active, two protease coding regions were linked to form a dimeric gene (pFGGP). Expression of this gene in Escherichia coli yielded a protein with the expected molecular mass of 22 kDa. The in vitro kinetic parameters of PRT and FGGP (where FGGP is the single polypeptide form of the HIV-1 protease with 2 glycine residues connecting the two subunits) for three peptide substrates are similar. Construction and analysis of a CheY-GAG-FGGP fusion protein demonstrated that FGGP is capable of precursor processing in vivo. Mutation of one or both of the active site aspartates to either asparagine or glutamate rendered the enzyme inactive, demonstrating that both active site aspartate residues are required for enzymatic activity.  相似文献   

10.
Degradation of small, acid-soluble spore proteins during germination of Bacillus subtilis spores is initiated by a sequence-specific protease called GPR. Western blot (immunoblot) analysis of either Bacillus megaterium or B. subtilis GPR expressed in B. subtilis showed that GPR is synthesized at about the third hour of sporulation in a precursor form and is processed to an approximately 2- to 5-kDa-smaller species 2 to 3 h later, at or slightly before the time of accumulation of dipicolinic acid by the forespore. This was found with both normal levels of expression of B. subtilis and B. megaterium GPR in B. subtilis, as well as when either protein was overexpressed up to 100-fold. The sporulation-specific processing of GPR was blocked in all spoIII, -IV, and -V mutants tested (none of which accumulated dipicolinic acid), but not in a spoVI mutant which accumulated dipicolinic acid. The amino-terminal sequences of the B. megaterium and B. subtilis GPR initially synthesized in sporulation were identical to those predicted from the coding genes' sequences. However, the processed form generated in sporulation lacked 15 (B. megaterium) or 16 (B. subtilis) amino-terminal residues. The amino acid sequence surrounding this proteolytic cleavage site was very homologous to the consensus sequence recognized and cleaved by GPR in its small, acid-soluble spore protein substrates. This observation, plus the efficient processing of overproduced GPR during sporulation, suggests that the GPR precursor may autoproteolyze itself during sporulation. During spore germination, the GPR from either species expressed in B. subtilis was further processed by removal of one additional amino-terminal amino acid (leucine), generating the mature protease which acts during spore germination.  相似文献   

11.
Acetohydroxyacid synthase I from Escherichia coli K-12 has been purified to near homogeneity. Analysis of the purified enzyme by sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed the presence of two polypeptides, one with a molecular weight of 60,000 and one with a molecular weight of 9,500. These two polypeptides were present in constant proportion to each other and to enzyme activity. The molar ratio of the two polypeptides (Mr 9,500:60,000), estimated from stained polyacrylamide gels, was 1. Antisera raised against the 60,000 Mr polypeptide precipitated both the 60,000 and the 9,500 Mr polypeptides from extracts of cells labeled with [35S]methionine. The addition of sodium dodecyl sulfate before immunoprecipitation eliminated the smaller polypeptide, and only the larger one was recovered. The hydrodynamic properties of the native enzyme confirmed a previous report that the largest enzymatically active species has a molecular weight of about 200,000; this species contains both the 60,000- and 9,500-molecular-weight polypeptides.  相似文献   

12.
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   

13.
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   

14.
The DNA sequences encoding the large subunit of the mRNA-capping enzyme of vaccinia virus were located on the viral genome. The formation of an enzyme-guanylate covalent intermediate labeled with [alpha-32P]GTP allowed the identification of the large subunit of the capping enzyme and was used to monitor the appearance of the enzyme during the infectious cycle. This assay confirmed that after vaccinia infection, a novel 84,000-molecular-weight polypeptide corresponding to the large subunit was rapidly synthesized before viral DNA replication. Hybrid-selected cell-free translation of early viral mRNA established that vaccinia virus encoded a polypeptide identical in molecular weight with the 32P-labeled 84,000-molecular-weight polypeptide found in vaccinia virions. Like the authentic capping enzyme, this virus-encoded cell-free translation product bound specifically to DNA-cellulose. A comparison of the partial proteolytic digestion fragments generated by V8 protease, chymotrypsin, and trypsin demonstrated that the 32P-labeled large subunit and the [35S]methionine-labeled cell-free translation product were identical. The mRNA encoding the large subunit of the capping enzyme was located 3.1 kilobase pairs to the left of the HindIII D restriction fragment of the vaccinia genome. Furthermore, the mRNA was determined to be 3.0 kilobases in size, and its 5' and 3' termini were precisely located by S1 nuclease analysis.  相似文献   

15.
Ethylenediamine-soluble glycoproteins were extracted from isolated Microsporum gypseum hyphal walls during sporulation and from spore coats before and after germination. This study was carried out to identify a sporulation-specific cell wall protein that possibly served as a substrate for the alkaline protease which initiated the macroconidial germination of this fungus. Analyses revealed that water-insoluble glycoprotein accounted for 10% of the ungerminated spore coat but only for 4 to 5% of the mycelial wall dry weight. This fraction was modified in its amino acid composition during sporulation, and it decreased in protein content during spore germination. Water-soluble glycoprotein, which accounted for approximately 3 to 3.5% of either the spore coat or mycelial wall dry weight, was of similar amino acid composition from both sources and did not decrease in protein content upon spore germination. The water-insoluble glycoprotein was found to be rich in leucine, aspartic acid, glycine, glutamic acid, and phenylalanine residues. The water-soluble glycoprotein was rich in proline, threonine, glycine, serine, glutamic acid, and alanine.  相似文献   

16.
A spore cortex-lytic enzyme of Clostridium perfringens S40 which is encoded by sleC is synthesized at an early stage of sporulation as a precursor consisting of four domains. After cleavage of an N-terminal presequence and a C-terminal prosequence during spore maturation, inactive proenzyme is converted to active enzyme by processing of an N-terminal prosequence with germination-specific protease (GSP) during germination. The present study was undertaken to characterize GSP. In the presence of 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonic acid (CHAPS), a nondenaturing detergent which was needed for the stabilization of GSP, GSP activity was extracted from germinated spores. The enzyme fraction, which was purified to 668-fold by column chromatography, contained three protein components with molecular masses of 60, 57, and 52 kDa. The protease showed optimum activity at pH 5.8 to 8.5 in the presence of 0.1% CHAPS and retained activity after heat treatment at 55 degrees C for 40 min. GSP specifically cleaved the peptide bond between Val-149 and Val-150 of SleC to generate mature enzyme. Inactivation of GSP by phenylmethylsulfonyl fluoride and HgCl(2) indicated that the protease is a cysteine-dependent serine protease. Several pieces of evidence demonstrated that three protein components of the enzyme fraction are processed forms of products of cspA, cspB, and cspC, which are positioned in a tandem array just upstream of the 5' end of sleC. The amino acid sequences deduced from the nucleotide sequences of the csp genes showed significant similarity and showed a high degree of homology with those of the catalytic domain and the oxyanion binding region of subtilisin-like serine proteases. Immunochemical studies suggested that active GSP likely is localized with major cortex-lytic enzymes on the exterior of the cortex layer in the dormant spore, a location relevant to the pursuit of a cascade of cortex hydrolytic reactions.  相似文献   

17.
Purification of Bacteroides amylophilus protease   总被引:2,自引:0,他引:2       下载免费PDF全文
A protease was released by Bacteroides amylophilus cells in late stationary phase, approximately 12 hr after maximum cell density was reached. The protease was concentrated by adsorption on diethylaminoethyl (DEAE)-Sephadex and was purified 532-fold by DEAE-Sephadex chromatography, by G-200-Sephadex gel filtration, and by isoelectric focusing. The purified protease was active between pH 4.5 and 12.0 with optima at pH 6.0 and 11.5. Evidence against there being a single protease was given by the differential inhibition of esterase and protease activities by some inhibitors. There was some evidence that only a single protease was present as the ratio of protease activity at various pH values did not alter significantly during purification or when the purified protease was partially heat-inactivated or treated with two specific trypsin-type protease inhibitors: N-alpha-tosyl-l-lysylchloromethyl ketone or phenylmethane-sulfonyl fluoride. Two forms of the same protease were found by acrylamide gel electrophoresis. Gel filtration confirmed the presence of protease in 30,000 and 60,000 molecular-weight forms. Treatment with 1 mm ethylenediaminetetraacetic acid or with 4 m urea failed to convert the 60,000-molecular-weight to the 30,000-molecular-weight species.  相似文献   

18.
Pyruvate oxidase, a tetrameric enzyme consisting of 4 identical subunits, dissociates into apoenzyme monomers and free FAD when treated with acid ammonium sulfate in the presence of high concentrations of potassium bromide. Reconstitution of the native enzymatically active protein can be accomplished by incubating equimolar concentrations of apomonomers and FAD at pH 6.5. The kinetics of the reconstitution reaction have been measured by 1) enzyme activity assays, 2) spectrophotometric assays to measure FAD binding, and 3) high performance liquid chromatography analysis measuring the distribution of monomeric, dimeric, and tetrameric species during reconstitution. The kinetic analysis indicates that the second order reaction of apomonomers with FAD to form an initial monomer-FAD complex is fast. The rate-limiting step for enzymatic reactivation appears to be the folding of the polypeptide chain in the monomer-FAD complex to reconstitute the three-dimensional FAD binding site prior to subunit reassociation. The subsequent formation of native tetramers appears to proceed via an essentially irreversible dimer assembly pathway.  相似文献   

19.
The zymogen of the protease (GPR) that initiates protein degradation during spore germination in Bacillus species is not activated in vitro under normal physiological conditions. However, there is rapid, acid-pH-dependent, zero-order, proteolytic activation of the purified zymogen in high concentrations of dimethyl sulfoxide. These findings provide further evidence that GPR activates itself during sporulation.  相似文献   

20.
The spoVM gene encodes a 26-amino-acid polypeptide that is essential for spore formation in Bacillus subtilis. A transposon insertion within the spoVM open reading frame has been shown to encode a chimeric protein which is biologically inactive and produces a phenotype identical to that of a deletion and insertion mutation. A genetic approach was used to identify possible interacting proteins, and the membrane-bound FtsH protease was identified. Mutations in ftsH suppressed the sporulation defect of certain spoVM mutants but not others. However, production of the mother cell sigma factors, sigmaE and sigmaK, was abnormal in the suppressed strains, and mutations in either spoVM or ftsH alone impaired sigma factor production and sporulation gene expression. Using FtsH purified from Escherichia coli, we demonstrated that in vitro (i) SpoVM inhibits FtsH protease activity and (ii) SpoVM is a substrate for the FtsH protease. We propose that during sporulation, SpoVM serves as a competitive inhibitor of FtsH activity. This interaction appears to be important for completion of the prespore engulfment step of sporulation, based on the phenotype of certain spoVM ftsH double mutants.  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