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 62 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.
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   

3.
AIMS: To determine the mechanism of action of inhibitors of the germination of spores of Bacillus species, and where these inhibitors act in the germination process. METHODS AND RESULTS: Spores of various Bacillus species are significant agents of food spoilage and food-borne disease, and inhibition of spore germination is a potential means of reducing such problems. Germination of the following spores was studied: (i) wild-type B. subtilis spores; (ii) B. subtilis spores with a nutrient receptor variant allowing recognition of a novel germinant; (iii) B. subtilis spores with elevated levels of either the variant nutrient receptor or its wild-type allele; (iv) B. subtilis spores lacking all nutrient receptors and (v) wild-type B. megaterium spores. Spores were germinated with a variety of nutrient germinants, Ca2+-dipicolinic acid (DPA) and dodecylamine for B. subtilis spores, and KBr for B. megaterium spores. Compounds tested as inhibitors of germination included alkyl alcohols, a phenol derivative, a fatty acid, ion channel blockers, enzyme inhibitors and several other compounds. Assays used to assess rates of spore germination monitored: (i) the fall in optical density at 600 nm of spore suspensions; (ii) the release of the dormant spore's large depot of DPA; (iii) hydrolysis of the dormant spore's peptidoglycan cortex and (iv) generation of CFU from spores that lacked all nutrient receptors. The results with B. subtilis spores allowed the assignment of inhibitory compounds into two general groups: (i) those that inhibited the action of, or response to, one nutrient receptor and (ii) those that blocked the action of, or response to, several or all of the nutrient receptors. Some of the compounds in groups 1 and 2 also blocked action of at least one cortex lytic enzyme, however, this does not appear to be the primary site of their action in inhibiting spore germination. The inhibitors had rather different effects on germination of B. subtilis spores with nutrients or non-nutrients, consistent with previous work indicating that germination of B. subtilis spores by non-nutrients does not involve the spore's nutrient receptors. In particular, none of the compounds tested inhibited spore germination with dodecylamine, and only three compounds inhibited Ca2+-DPA germination. In contrast, all compounds had very similar effects on the germination of B. megaterium spores with either glucose or KBr. The effects of the inhibitors tested on spores of both Bacillus species were largely reversible. CONCLUSIONS: This work indicates that inhibitors of B. subtilis spore germination fall into two classes: (i) compounds (most alkyl alcohols, N-ethylmaleimide, nifedipine, phenols, potassium sorbate) that inhibit the action of, or response to, primarily one nutrient receptor and (ii) compounds [amiloride, HgCl2, octanoic acid, octanol, phenylmethylsulphonylfluoride (PMSF), quinine, tetracaine, tosyl-l-arginine methyl ester, trifluoperazine] that inhibit the action of, or response to, several nutrient receptors. Action of these inhibitors, is reversible. The similar effects of inhibitors on B. megaterium spore germination by glucose or KBr indicate that inorganic salts likely trigger germination by activating one or more nutrient receptors. The lack of effect of all inhibitors on dodecylamine germination suggests that this compound stimulates germination by creating channels in the spore's inner membrane allowing DPA release. SIGNIFICANCE AND IMPACT OF THE STUDY: This work provides new insight into the steps in spore germination that are inhibited by various chemicals, and the mechanism of action of these inhibitors. The work also provides new insights into the process of spore germination itself.  相似文献   

4.
Aims:  To determine conditions for generation and recovery of Bacillus subtilis spore populations heavily damaged by moist heat treatment.
Methods and Results:  Bacillus subtilis spores were treated with moist heat and spore viability was assessed on different media. A rich medium and several minimal media gave similar spore recoveries after moist heat treatment, but lack of glucose in minimal media greatly decreased spore recovery. High NaCl levels also greatly decreased the recovery of moist heat-treated spores on minimal media, and addition of good osmoprotectants reversed this effect. Moist heat treatment did not decrease spore recovery on minimal media with high salt through DNA damage or by eliminating spore germination, but by affecting spore outgrowth.
Conclusions:  Conditions for generating B. subtilis spore populations with high levels of conditional moist heat damage have been determined. The major conditional damage appears to be in spore outgrowth, perhaps because of damage to one or more important metabolic enzymes.
Significance and Impact of the Study:  This work has provided new insight into the mechanism of B. subtilis spore killing by moist heat.  相似文献   

5.
AIMS: To elucidate the factors that determine the rate of germination of Bacillus subtilis spores with very high pressure (VHP) and the mechanism of VHP germination. METHODS AND RESULTS: Spores of B. subtilis were germinated rapidly with a VHP of 500 MPa at 50 degrees C. This VHP germination did not require the spore's nutrient-germinant receptors, as found previously, and did not require diacylglycerylation of membrane proteins. However, the spore's pool of dipicolinic acid (DPA) was essential. Either of the two redundant enzymes that degrade the spore's peptidoglycan cortex, and thus allow completion of spore germination, was essential for completion of VHP germination. However, neither of these enzymes was needed for DPA release triggered by VHP treatment. Completion of spore germination as well as DPA release with VHP had an optimum temperature of approx. 60 degrees C, in contrast to an optimum temperature of 40 degrees C for germination with the moderately high pressure of 150 MPa. The rate of spore germination by VHP decreased approx. fourfold when the sporulation temperature increased from 23 degrees C to 44 degrees C, and decreased twofold when 1 mol l(-1) salt was present in sporulation. However, large variations in levels of unsaturated fatty acids in the spore's inner membranes did not affect rates of VHP germination. Complete germination of spores by VHP was not inhibited significantly by killing of spores with several oxidizing agents, and was not inhibited by ethanol, octanol or o-chlorophenol at concentrations that abolish nutrient germination. Completion of spore germination by VHP was also inhibited by Hg(2+), but this ion did not inhibit DPA release caused by VHP. In contrast, dodecylamine, a surfactant that can trigger spore germination, strongly inhibited DPA release caused by VHP treatment. CONCLUSIONS: VHP does not cause spore germination by acting upon the spore's nutrient-germinant receptors, but by directly causing DPA release. This DPA release then leads to subsequent completion of germination. VHP likely acts on the spore's inner membrane to cause DPA release, targeting either a membrane protein or the membrane itself. However, the precise identity of this target is not yet clear. SIGNIFICANCE AND IMPACT OF THE STUDY: There is significant interest in the use of VHP to eliminate or reduce levels of bacterial spores in foods. As at least partial spore germination by pressure is almost certainly essential for subsequent spore killing, knowledge of factors involved and the mechanism of VHP germination are crucial to the understanding of spore killing by VHP. This work provides new insight into factors that can affect the rate of B. subtilis spore germination by VHP, and into the mechanism of VHP germination itself.  相似文献   

6.
Aims: To determine effects of inner membrane lipid composition on Bacillus subtilis sporulation and spore properties. Methods and Results: The absence of genes encoding lipid biosynthetic enzymes had no effect on B. subtilis sporulation, although the expected lipids were absent from spores’ inner membrane. The rate of spore germination with nutrients was decreased c. 50% with mutants that lacked the major cardiolipin (CL) synthase and another enzyme for synthesis of a major phospholipid. Spores lacking the minor CL synthase or an enzyme essential for glycolipid synthesis exhibited 50–150% increases in rates of dodecylamine germination, while spores lacking enzymes for phosphatidylethanolamine (PE), phosphatidylserine (PS) and lysylphosphatidylglycerol (l‐PG) synthesis exhibited a 30–50% decrease. Spore sensitivity to H2O2 and tert‐butylhydroperoxide was increased 30–60% in the absence of the major CL synthase, but these spores’ sensitivity to NaOCl or Oxone? was unaffected. Spores of lipid synthesis mutants were less resistant to wet heat, with spores lacking enzymes for PE, PS or l‐PG synthesis exhibiting a two to threefold decrease and spores of other strains exhibiting a four to 10‐fold decrease. The decrease in spore wet heat resistance correlated with an increase in core water content. Conclusions: Changing the lipid composition of the B. subtilis inner membrane did not affect sporulation, although modest effects on spore germination and wet heat and oxidizing agent sensitivity were observed, especially when multiple lipids were absent. The increases in rates of dodecylamine germination were likely due to increased ability of this compound to interact with the spore’s inner membrane in the absence of some CL and glycolipids. The effects on spore wet heat sensitivity are likely indirect, because they were correlated with changes in core water content. Significance and Impact of the Study: The results of this study provide insight into roles of inner membrane lipids in spore properties.  相似文献   

7.
AIMS: To elucidate the factors influencing the sensitivity of Bacillus subtilis spores in killing and disrupting by mechanical abrasion, and the mechanism of stimulation of spore germination by abrasion. METHODS AND RESULTS: Spores of B. subtilis strains were abraded by shaking with glass beads in liquid or the dry state, and spore killing, disruption and germination were determined. Dormant spores were more resistant to killing and disruption by abrasion than were growing cells or germinated spores. However, dormant spores of the wild-type strain with or without most coat proteins removed, spores of strains with mutations causing spore coat defects, spores lacking their large depot of dipicolinic acid (DPA) and spores with defects in the germination process exhibited essentially identical rates of killing and disruption by abrasion. When spores lacking all nutrient germinant receptors were enumerated by plating directly on nutrient medium, abrasion increased the plating efficiency of these spores before killing them. Spores lacking all nutrient receptors and either of the two redundant cortex-lytic enzymes behaved similarly in this regard, but the plating efficiency of spores lacking both cortex-lytic enzymes was not stimulated by abrasion. CONCLUSIONS: Dormant spores are more resistant to killing and disruption by abrasion than are growing cells or germinated spores, and neither the complete coats nor DPA are important in spore resistance to such treatments. Germination is not essential for spore killing by abrasion, although abrasion can trigger spore germination by activation of either of the spore's cortex-lytic enzymes. SIGNIFICANCE AND IMPACT OF THE STUDY: This work provides new insight into the mechanisms of the killing, disruption and germination of spores by abrasion and makes the surprising finding that at least much of the spore coat is not important in spore resistance to abrasion.  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.
10.
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   

11.
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   

12.
Aims:  To determine the germination and inactivation of Bacillus cereus spores lacking various germination proteins using moderately high pressure (MHP) and heat.
Methods:  The inactivation and germination of wild-type B. cereus spores in buffer by MHP (150 MPa) at various temperatures, as well as the MHP inactivation and germination of B. cereus spores lacking individual germinant receptors and monovalent cation antiporters, was determined.
Results:  Loss of individual germinant receptors had no large effects on spore inactivation or germination, although germination of receptor-deficient spores was generally slightly decreased. Loss of the GerN in particular the GerN and GerT antiporters also decreased spore germination by MHP, especially at 40 and 50°C.
Conclusions:  Both inactivation and germination of B. cereus spores by MHP increased with rise of temperature; however, mutant strains lacking individual germinant receptor had similar levels of germination as compared to wild-type spores. To evaluate the role of germinant receptors in MHP, a strain lacking a large number of germinant receptors is needed.
Significance and Impact of the Study:  The results of this work may lead to a better understanding of how MHP causes germination of spores of B. cereus .  相似文献   

13.
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   

14.
Alpha/beta-type small, acid-soluble spore proteins (SASP) are essential for the resistance of DNA in spores of Bacillus species to damage. An alpha/beta-type SASP, Ssp2, from Clostridium perfringens was expressed at significant levels in B. subtilis spores lacking one or both major alpha/beta-type SASP (alpha- and alpha- beta- strains, respectively). Ssp2 restored some of the resistance of alpha- beta- spores to UV and nitrous acid and of alpha- spores to dry heat. Ssp2 also restored much of the resistance of alpha- spores to nitrous acid and restored full resistance of alpha- spores to UV and moist heat. These results further indicate the interchangeability of alpha/beta-type SASP in DNA protection in spores.  相似文献   

15.
Dipicolinic acid (DPA) comprises approximately 10% of the dry weight of spores of Bacillus species. Although DPA has long been implicated in spore resistance to wet heat and spore stability, definitive evidence on the role of this abundant molecule in spore properties has generally been lacking. Bacillus subtilis strain FB122 (sleB spoVF) produced very stable spores that lacked DPA, and sporulation of this strain with DPA yielded spores with nearly normal DPA levels. DPA-replete and DPA-less FB122 spores had similar levels of the DNA protective alpha/beta-type small acid-soluble spore proteins (SASP), but the DPA-less spores lacked SASP-gamma. The DPA-less FB122 spores exhibited similar UV resistance to the DPA-replete spores but had lower resistance to wet heat, dry heat, hydrogen peroxide, and desiccation. Neither wet heat nor hydrogen peroxide killed the DPA-less spores by DNA damage, but desiccation did. The inability to synthesize both DPA and most alpha/beta-type SASP in strain PS3664 (sspA sspB sleB spoVF) resulted in spores that lost viability during sporulation, at least in part due to DNA damage. DPA-less PS3664 spores were more sensitive to wet heat than either DPA-less FB122 spores or DPA-replete PS3664 spores, and the latter also retained viability during sporulation. These and previous results indicate that, in addition to alpha/beta-type SASP, DPA also is extremely important in spore resistance and stability and, further, that DPA has some specific role(s) in protecting spore DNA from damage. Specific roles for DPA in protecting spore DNA against damage may well have been a major driving force for the spore's accumulation of the high levels of this small molecule.  相似文献   

16.
17.
Aims: To compare physical properties of spores that were produced in broth sporulation media at greater than 108 spores ml−1. Methods and Results: Bacillus atrophaeus reproducibly sporulated in nutrient broth (NB) and sporulation salts. Microscopy measurements showed that the spores were 0·68 ± 0·11 μm wide and 1·21 ± 0·18 μm long. Coulter Multisizer (CM3) measurements revealed the spore volumes and volume-equivalent spherical diameters, which were 0·48 ± 0·38 μm3 and 0·97 ± 0·07 μm, respectively. Bacillus cereus reproducibly sporulated in NB, sporulation salts, 200 mmol l−1 glutamate and antifoam. Spores were 0·95 ± 0·11 μm wide and 1·31 ± 0·17 μm long. Spore volumes were 0·78 ± 0·61 μm3 and volume-equivalent spherical diameters were 1·14 ± 0·11 μm. Bacillus atrophaeus spores were hydrophilic and B. cereus spores were hydrophobic. However, spore hydrophobicity was significantly altered after treatment with pH-adjusted bleach. Conclusions: The utility of a CM3 for both quantifying Bacillus spores and measuring spore sizes was demonstrated, although the volume between spore exosporium and spore coat was not measured. This study showed fundamental differences between spores from a Bacillus subtilis- and B. cereus-group species. Significance and Impact of the Study: This is useful for developing standard methods for broth spore production and physical characterization of both living and decontaminated spores.  相似文献   

18.
AIMS: To determine the mechanisms of Bacillus subtilis spore resistance to and killing by a novel sporicide, dimethyldioxirane (DMDO) that was generated in situ from acetone and potassium peroxymonosulfate at neutral pH. METHODS AND RESULTS: Spores of B. subtilis were effectively killed by DMDO. Rates of killing by DMDO of spores lacking most DNA protective alpha/beta-type small, acid-soluble spore proteins (alpha- beta- spores) or the major DNA repair protein, RecA, were very similar to that of wild-type spore killing. Survivors of wild-type and alpha- beta- spores treated with DMDO also exhibited no increase in mutations. Spores lacking much coat protein due either to mutation or chemical decoating were much more sensitive to DMDO than were wild-type spores, but were more resistant than growing cells. Wild-type spores killed with this reagent retained their large pool of dipicolinic acid (DPA), and the survivors of spores treated with DMDO were sensitized to wet heat. The DMDO-killed spores germinated with nutrients, albeit more slowly than untreated spores, but germinated faster than untreated spores with dodecylamine. The killed spores were also germinated by very high pressures and by lysozyme treatment in hypertonic medium, but many of these spores lysed shortly after their germination, and none of these treatments were able to revive the DMDO-killed spores. CONCLUSIONS: DMDO is an effective reagent for killing B. subtilis spores. The spore coat is a major factor in spore resistance to DMDO, which does not kill spores by DNA damage or by inactivating some component needed for spore germination. Rather, this reagent appears to kill spores by damaging the spore's inner membrane in some fashion. SIGNIFICANCE AND IMPACT OF THE STUDY: This work demonstrates that DMDO is an effective decontaminant for spores of Bacillus species that can work under mild conditions, and the killed spores cannot be revived. Evidence has also been obtained on the mechanisms of spore resistance to and killing by this reagent.  相似文献   

19.
Sporulation of a Bacillus subtilis strain (termed alpha(-) beta(-)) lacking the majority of the alpha/beta-type small, acid-soluble spore proteins (SASP) that are synthesized in the developing forespore and saturate spore DNA exhibited a number of differences from that of the wild-type strain, including delayed forespore accumulation of dipicolinic acid, overexpression of forespore-specific genes, and delayed expression of at least one mother cell-specific gene turned on late in sporulation, although genes turned on earlier in the mother cell were expressed normally in alpha(-) beta(-) strains. The sporulation defects in alpha(-) beta(-) strains were corrected by synthesis of chromosome-saturating levels of either of two wild-type, alpha/beta-type SASP but not by a mutant SASP that binds DNA poorly. Spores from alpha(-) beta(-) strains also exhibited less glutaraldehyde resistance and slower outgrowth than did wild-type spores, but at least some of these defects in alpha(-) beta(-) spores were abolished by the synthesis of normal levels of alpha/beta-type SASP. These results indicate that alpha/beta-type SASP may well have global effects on gene expression during sporulation and spore outgrowth.  相似文献   

20.
DNA in dormant spores of Bacillus species is in an A-like conformation   总被引:7,自引:0,他引:7  
The DNA in dormant spores of Bacillus species is associated with alpha/beta-type small, acid-soluble proteins (SASP), which are double-stranded DNA-binding proteins whose amino acid sequence has been highly conserved in evolution. In vitro these proteins bind most strongly to DNA which readily adopts an A-like conformation, as binding of alpha/beta-type SASP causes DNA to assume an A-like conformation. As predicted by this conformational change in DNA, binding of alpha/beta-type SASP to relaxed but covalently closed plasmid DNA results in the introduction of a large number of negative supercoils. Associated with the conformational change in DNA brought about by alpha/beta-type SASP binding is a change in its photochemistry such that ultraviolet irradiation does not generate pyrimidine dimers, but rather a thyminyl-thymine adduct termed spore photoproduct (SP). The latter two properties of DNA complexed with alpha/beta-type SASP in vitro are similar to those of DNA in dormant spores of Bacillus species in which: (i) plasmid DNA has a much higher number of negative supercoils than plasmid in growing cells; and (ii) ultraviolet irradiation produces SP and no pyrimidine dimers, while only pyrimidine dimers are formed in growing cells. During sporulation these changes in the properties of spore DNA take place in parallel with synthesis of alpha/beta-type SASP, and the magnitude of the changes is greatly reduced in mutants that make low amounts of these proteins. A straightforward interpretation of these data is that DNA in dormant spores of Bacillus species is in an A-like conformation.(ABSTRACT TRUNCATED AT 250 WORDS)  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