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 15 毫秒
1.
Lytic Replication of Coliphage Lambda in Salmonella typhosa Hybrids   总被引:2,自引:0,他引:2       下载免费PDF全文
Hybrids between Escherichia coli K-12 and Salmonella typhosa which conserved a continuous K-12 chromosomal diploid segment extending from pro through ara to the strA locus were sensitive to plaque formation by wild-type λ. These partially diploid S. typhosa hybrids could be lysogenized with λ and subsequently induced to produce infectious phage particles. When the K-12 genes were segregated from a lysogenic S. typhosa hybrid, phage-productive ability was no longer detectable due to loss of a genetic region necessary for vegetative replication of λ. However, λ prophage was shown to persist in a quiescent state in the S. typhosa hybrid segregant with phage-productive ability being reactivated after replacement of the essential K-12 λ replication region. Low-frequency transduction and high-frequency transduction lysates containing the gal+ genes of S. typhosa were prepared by induction of λ-lysogenic S. typhosa hybrids indicating that the attλ site is chromosomally located in S. typhosa in close proximity to the gal locus as in E. coli K-12. After propagation in S. typhosa hybrids, λ was subject to restriction by E. coli K-12 recipients, thus establishing that S. typhosa does not perform the K-12 modification of λ deoxyribonucleic acid. Hybrids of S. typhosa, however, did not restrict λ grown previously on E. coli K-12. The K-12 genetic region required for λ phage production in S. typhosa was located within min 66 to min 72 on the genetic map of the E. coli chromosome. Transfer of an F-merogenote encompassing the 66 to 72 min E. coli chromosomal region to λ-insensitive S. typhosa hybrids enabled them to replicate wild-type λ. The λ-insensitive S. typhosa hybrid, WR4255, which blocks λ replication, can be mutagenized to yield mutant strains sensitive to λvir and λimm434. These WR4255 mutants remained insensitive to plaque formation by wild-type λ.  相似文献   

2.
A mutant of Escherichia coli temperature-sensitive for deoxyribonucleic acid synthesis, dnaD, was found to have temperature-sensitive modification and restriction phenotypes. In contrast to the original observation by Carl (1970), the mutant could support the growth of λ phage at 41 C. However, the λ phages thus produced were able to form plaques with normal plating efficiency only on E. coli C, a restriction-less strain, but not on E. coli K. Since the λ phages produced in the mutant at 30 C could form plaques equally well on both E. coli strains, it was concluded that the dnaD mutant has a temperature-sensitive modification phenotype. Furthermore, since the dnaD mutant allowed some growth of unmodified λ·C phages at 41 C but less at 30 C, the mutant is also temperature sensitive in restriction. The relationship, if any, between temperature-sensitive deoxyribonucleic acid synthesis and temperature-sensitive modification-restriction in the dnaD mutant is not known. Similar experiments were done with three dnaC mutants and one dnaA mutant. Two dnaC mutants were found to have altered restriction phenotypes at 41 C, but none of the mutants were defective in modification.  相似文献   

3.
The lon mutants of Escherichia coli grow apparently normally except that, after temporary periods of inhibition of deoxyribonucleic acid synthesis, septum formation is specifically inhibited. Under these conditions, long, multinucleate, nonseptate filaments result. The lon mutation also creates a defect such that wild-type bacteriophage λ fails to lysogenize lon mutants efficiently and consequently forms clear plaques on a lon host. Two lines of evidence suggest that this failure probably results from interference with expression of the λcI gene, which codes for repressor, or with repressor action:-(i) when a lon mutant was infected with a λcII, cIII, or c Y mutant, there was an additive effect between the lon mutation and the λc mutations upon reduction of lysogenization frequency; and (ii) lon mutants permitted the growth of the λcro mutant under conditions in which the repressor was active. The isolation of λ mutants (λtp) which gained the ability to form turbid plaques on lon cells is also reported.  相似文献   

4.
1953 was a historical year for biology, as it marked the birth of the DNA helix, but also a report by Bertani and Weigle on ‘a barrier to infection’ of bacteriophage λ in its natural host, Escherichia coli K-12, that could be lifted by ‘host-controlled variation’ of the virus. This paper lay dormant till Nobel laureate Arber and PhD student Dussoix showed that the λ DNA was rejected and degraded upon infection of different bacterial hosts, unless it carried host-specific modification of that DNA, thus laying the foundations for the phenomenon of restriction and modification (R-M). The restriction enzyme of E.coli K-12, EcoKI, was purified in 1968 and required S-adenosylmethionine (AdoMet) and ATP as cofactors. By the end of the decade there was substantial evidence for a chromosomal locus hsdK with three genes encoding restriction (R), modification (M) and specificity (S) subunits that assembled into a large complex of >400 kDa. The 1970s brought the message that EcoKI cut away from its DNA recognition target, to which site the enzyme remained bound while translocating the DNA past itself, with concomitant ATP hydrolysis and subsequent double-strand nicks. This translocation event created clearly visible DNA loops in the electron microscope. EcoKI became the archetypal Type I R-M enzyme with curious DNA translocating properties reminiscent of helicases, recognizing the bipartite asymmetric site AAC(N6)GTGC. Cloning of the hsdK locus in 1976 facilitated molecular understanding of this sophisticated R-M complex and in an elegant ‘pas de deux’ Murray and Dryden constructed the present model based on a large body of experimental data plus bioinformatics. This review celebrates the golden anniversary of EcoKI and ends with the exciting progress on the vital issue of restriction alleviation after DNA damage, also first reported in 1953, which involves intricate control of R subunit activity by the bacterial proteasome ClpXP, important results that will keep scientists on the EcoKI track for another 50 years to come.  相似文献   

5.
Many enteropathogenic bacteria target the mammalian gut. The mechanisms protecting the host from infection are poorly understood. We have studied the protective functions of secretory antibodies (sIgA) and the microbiota, using a mouse model for S. typhimurium diarrhea. This pathogen is a common cause of diarrhea in humans world-wide. S. typhimurium (S. tm att, sseD) causes a self-limiting gut infection in streptomycin-treated mice. After 40 days, all animals had overcome the disease, developed a sIgA response, and most had cleared the pathogen from the gut lumen. sIgA limited pathogen access to the mucosal surface and protected from gut inflammation in challenge infections. This protection was O-antigen specific, as demonstrated with pathogens lacking the S. typhimurium O-antigen (wbaP, S. enteritidis) and sIgA-deficient mice (TCRβ−/−δ−/−, JH −/−, IgA−/−, pIgR−/−). Surprisingly, sIgA-deficiency did not affect the kinetics of pathogen clearance from the gut lumen. Instead, this was mediated by the microbiota. This was confirmed using ‘L-mice’ which harbor a low complexity gut flora, lack colonization resistance and develop a normal sIgA response, but fail to clear S. tm att from the gut lumen. In these mice, pathogen clearance was achieved by transferring a normal complex microbiota. Thus, besides colonization resistance ( = pathogen blockage by an intact microbiota), the microbiota mediates a second, novel protective function, i.e. pathogen clearance. Here, the normal microbiota re-grows from a state of depletion and disturbed composition and gradually clears even very high pathogen loads from the gut lumen, a site inaccessible to most “classical” immune effector mechanisms. In conclusion, sIgA and microbiota serve complementary protective functions. The microbiota confers colonization resistance and mediates pathogen clearance in primary infections, while sIgA protects from disease if the host re-encounters the same pathogen. This has implications for curing S. typhimurium diarrhea and for preventing transmission.  相似文献   

6.
We have found previously that, in contrast to the free O initiator protein of λ phage or plasmid rapidly degraded by the Escherichia coli ClpP/ClpX protease, the λO present in the replication complex (RC) is protected from proteolysis. However, in cells growing in a complete medium, a temperature shift from 30 to 43°C resulted in the decay of the λO fraction, which indicated disassembly of RC. This process occurred due to heat shock induction of the groE operon, coding for molecular chaperones of the Hsp60 system. Here we demonstrate that an increase in the cellular concentration of GroEL and GroES proteins is not in itself sufficient to cause RC disassembly. Another requirement is a DNA gyrase-mediated negative resupercoiling of λ plasmid DNA, which counteracts DNA relaxation and starts to dominate 10 min after the temperature upshift. We presume that RC dissociates from λ DNA during the negative resupercoiling, becoming susceptible to the subsequent action of GroEL/S and ClpP/ClpX proteins. In contrast to λcro+, in λcro plasmid-harboring cells, the RC reveals heat shock resistance. After temperature upshift of the λcrots plasmid-harboring cells, a Cro repressor-independent control of λ DNA replication and heat shock resistance of RC are established before the period of DNA gyrase-mediated negative supercoiling. We suggest that the tight binding of RC to λ DNA is due to interaction of RC with other DNA-bound proteins, and is related to the molecular basis of the λcro plasmid replication control.  相似文献   

7.
8.
Phenethyl β-d-galactopyranoside (PEG) was hydrolyzed by the β-galactosidase of Escherichia coli to form the toxic product phenethyl alcohol. Salmonella typhimurium did not hydrolyze PEG. In mixed culture, the ratio of S. typhimurium to E. coli was increased by growing the organisms in lactose broth containing 2.5% PEG. The high concentration of PEG required for inhibition of E. coli can be attributed to inadequate cell permeability rather than to prevention of β-galactosidase induction.  相似文献   

9.
The appearance of lytic bacteriophage against newly introduced starter strains used during commercial cheese manufacture occurs rapidly, and their origin is not well understood. In this study, members of the group N streptococci were examined for the presence of bacteriophage restriction and modification systems. Two streptococcal phages from Streptococcus cremoris TR and Streptococcus lactis C2 (phage designations tr and c2) showed restricted lytic development on S. cremoris 799 and KH, respectively. Efficiency of plaquing was 1.9 × 10−7 for tr plaqued on 799 and 2.1 × 10−7 for c2 plaqued on KH. After passage through the restrictive hosts, these phages demonstrated high lytic ability for formerly restrictive hosts. Stress of the restrictive host strains at temperatures of 40 to 50°C resulted in a significant increase in the efficiency of plaquing of restricted bacteriophages. Elevated temperatures are encountered during commercial cheese manufacture. The results suggested that the temporary loss of host restriction activity with the resulting modification of nonspecific bacteriophage may contribute directly to the appearance of lytic phage against new starter strains.  相似文献   

10.
Neonatal meningitis, caused by Escherichia coli K1, is a serious central nervous system disease. We have established that macrophages serve as permissive niches for E. coli K1 to multiply in the host and for attaining a threshold level of bacterial load, which is a prerequisite for the onset of the disease. Here, we demonstrate experimentally that three N-glycans in FcγRIa interact with OmpA of E. coli K1 for binding to and entering the macrophages. Adoptive transfer of FcγRIa−/− bone marrow-derived macrophages transfected with FcγRIa into FcγRIa−/− newborn mice renders them susceptible to E. coli K1-induced meningitis. In contrast, mice that received bone marrow-derived macrophages transfected with FcγRIa in which N-glycosylation sites 1, 4, and 5 are mutated to alanines exhibit resistance to E. coli K1 infection. Our molecular dynamics and simulation studies predict that N-glycan 5 exhibits strong binding at the barrel site of OmpA formed by loops 3 and 4, whereas N-glycans 1 and 4 interact with loops 1, 3, and 4 of OmpA at tip regions. Molecular modeling data also suggest no role for the IgG binding site in the invasion process. In agreement, experimental mutations in IgG binding site had no effect on the E. coli K1 entry into macrophages in vitro or on the onset of meningitis in newborn mice. Together, this integration of experimental and computational studies reveals how the N-glycans in FcγRIa interact with the OmpA of E. coli K1 for inducing the disease pathogenesis.  相似文献   

11.
Along the ribosome assembly pathway, various ribosomal RNA processing and modification reactions take place. Stem–loop 69 in the large subunit of Escherichia coli ribosomes plays a substantial role in ribosome functioning. It contains three highly conserved pseudouridines synthesized by pseudouridine synthase RluD. One of the pseudouridines is further methylated by RlmH. In this paper we show that RlmH has unique substrate specificity among rRNA modification enzymes. It preferentially methylates pseudouridine and less efficiently uridine. Furthermore, RlmH is the only known modification enzyme that is specific to 70S ribosomes. Kinetic parameters determined for RlmH are the following: The apparent KM for substrate 70S ribosomes is 0.51 ± 0.06 μM, and for cofactor S-adenosyl-L-methionine 27 ± 3 μM; the kcat values are 4.95 ± 1.10 min−1 and 6.4 ± 1.3 min−1, respectively. Knowledge of the substrate specificity and the kinetic parameters of RlmH made it possible to determine the kinetic parameters for RluD as well. The KM value for substrate 50S subunits is 0.98 ± 0.18 μM and the kcat value is 1.97 ± 0.46 min−1. RluD is the first rRNA pseudouridine synthase to be kinetically characterized. The determined rates of RluD- and RlmH-directed modifications of 23S rRNA are compatible with the rate of 50S assembly in vivo. The fact that RlmH requires 30S subunits demonstrates the dependence of 50S subunit maturation on the simultaneous presence of 30S subunits.  相似文献   

12.
As hosts acquire resistance to viruses, viruses must overcome that resistance to re-establish infectivity, or go extinct. Despite the significant hurdles associated with adapting to a resistant host, viruses are evolutionarily successful and maintain stable coevolutionary relationships with their hosts. To investigate the factors underlying how pathogens adapt to their hosts, we performed a deep mutational scan of the region of the λ tail fiber tip protein that mediates contact with the receptor on λ’s host, Escherichia coli. Phages harboring amino acid substitutions were subjected to selection for infectivity on wild type E. coli, revealing a highly restrictive fitness landscape, in which most substitutions completely abrogate function. A subset of positions that are tolerant of mutation in this assay, but diverse over evolutionary time, are associated with host range expansion. Imposing selection for phage infectivity on three λ-resistant hosts, each harboring a different missense mutation in the λ receptor, reveals hundreds of adaptive variants in λ. We distinguish λ variants that confer promiscuity, a general ability to overcome host resistance, from those that drive host-specific infectivity. Both processes may be important in driving adaptation to a novel host.Subject terms: Bacteriophages, Molecular evolution, Viral genetics  相似文献   

13.
Inversion of the G segment in bacteriophage Mu DNA occurs by a site-specific recombination event and determines the host specificity of Mu phage particles produced. Inversion is mediated by a Mu function (Gin). The gin gene has been placed under control of the inducible λ pL promoter and a synthetic Shine-Dalgarno linker upstream of the initiation codon. The Gin protein content in induced cells is boosted to ˜10% of total protein. Partially purified extracts from overproducing strains promote efficient inversion of the G DNA segment in vitro which is visualized by agarose gel electrophoresis of the substrate DNA after cutting with appropriate restriction endonucleases. The in vitro reaction requires Mg2+, a super-coiled DNA substrate and occurs in the absence of exogenous ATP. Inversion from the G(+) to the G(−) orientation is as efficient as the switch from G(−) to G(+).  相似文献   

14.
Amyloids are highly aggregated proteinaceous fibers historically associated with neurodegenerative conditions including Alzheimers, Parkinsons, and prion-based encephalopathies. Polymerization of amyloidogenic proteins into ordered fibers can be accelerated by preformed amyloid aggregates derived from the same protein in a process called seeding. Seeding of disease-associated amyloids and prions is highly specific and cross-seeding is usually limited or prevented. Here we describe the first study on the cross-seeding potential of bacterial functional amyloids. Curli are produced on the surface of many Gram-negative bacteria where they facilitate surface attachment and biofilm development. Curli fibers are composed of the major subunit CsgA and the nucleator CsgB, which templates CsgA into fibers. Our results showed that curli subunit homologs from Escherichia coli, Salmonella typhimurium LT2, and Citrobacter koseri were able to cross-seed in vitro. The polymerization of Escherichia coli CsgA was also accelerated by fibers derived from a distant homolog in Shewanella oneidensis that shares less than 30% identity in primary sequence. Cross-seeding of curli proteins was also observed in mixed colony biofilms with E. coli and S. typhimurium. CsgA was secreted from E. coli csgB− mutants assembled into fibers on adjacent S. typhimurium that presented CsgB on its surfaces. Similarly, CsgA was secreted by S. typhimurium csgB− mutants formed curli on CsgB-presenting E. coli. This interspecies curli assembly enhanced bacterial attachment to agar surfaces and supported pellicle biofilm formation. Collectively, this work suggests that the seeding specificity among curli homologs is relaxed and that heterogeneous curli fibers can facilitate multispecies biofilm development.  相似文献   

15.
Pseudouridines in the stable RNAs of Bacteria are seldom subjected to further modification. There are 11 pseudouridine (Ψ) sites in Escherichia coli rRNA, and further modification is found only at Ψ1915 in 23S rRNA, where the N-3 position of the base becomes methylated. Here, we report the identity of the E. coli methyltransferase that specifically catalyzes methyl group addition to form m3Ψ1915. Analyses of E. coli rRNAs using MALDI mass spectrometry showed that inactivation of the ybeA gene leads to loss of methylation at nucleotide Ψ1915. Methylation is restored by complementing the knockout strain with a plasmid-encoded copy of ybeA. Homologs of the ybeA gene, and thus presumably the ensuing methylation at nucleotide m3Ψ1915, are present in most bacterial lineages but are essentially absent in the Archaea and Eukaryota. Loss of ybeA function in E. coli causes a slight slowing of the growth rate. Phylogenetically, ybeA and its homologs are grouped with other putative S-adenosylmethionine-dependent, SPOUT methyltransferase genes in the Cluster of Orthologous Genes COG1576; ybeA is the first member to be functionally characterized. The YbeA methyltransferase is active as a homodimer and docks comfortably into the ribosomal A site without encroaching into the P site. YbeA makes extensive interface contacts with both the 30S and 50S subunits to align its active site cofactor adjacent to nucleotide Ψ1915. Methylation by YbeA (redesignated RlmH for rRNA large subunit methyltransferase H) possibly functions as a stamp of approval signifying that the 50S subunit has engaged in translational initiation.  相似文献   

16.
The competition for glucose between Escherichia coli ML30, a typical copiotrophic enterobacterium and Chelatobacter heintzii ATCC29600, an environmentally successful strain, was studied in a carbon-limited culture at low dilution rates. First, as a base for modelling, the kinetic parameters μmax and Ks were determined for growth with glucose. For both strains, μmax was determined in batch culture after different precultivation conditions. In the case of C. heintzii, μmax was virtually independent of precultivation conditions. When inoculated into a glucose-excess batch culture medium from a glucose-limited chemostat run at a dilution rate of 0.075 h−1 C. heintzii grew immediately with a μmax of 0.17±0.03 h−1. After five transfers in batch culture, μmax had increased only slightly to 0.18±0.03 h−1. A different pattern was observed in the case of E. coli. Inoculated from a glucose-limited chemostat at D=0.075 h−1 into glucose-excess batch medium E. coli grew only after an acceleration phase of ∼3.5 h with a μmax of 0.52 h−1. After 120 generations and several transfers into fresh medium, μmax had increased to 0.80±0.03 h−1. For long-term adapted chemostat-cultivated cells, a Ks for glucose of 15 μg l−1 for C. heintzii, and of 35 μg l−1 for E. coli, respectively, was determined in 14C-labelled glucose uptake experiments. In competition experiments, the population dynamics of the mixed culture was determined using specific surface antibodies against C. heintzii and a specific 16S rRNA probe for E. coli. C. heintzii outcompeted E. coli in glucose-limited continuous culture at the low dilution rates of 0.05 and 0.075 h−1. Using the determined pure culture parameter values for Ks and μmax, it was only possible to simulate the population dynamics during competition with an extended form of the Monod model, which includes a finite substrate concentration at zero growth rate (smin). The values estimated for smin were dependent on growth rate; at D=0.05 h−1, it was 12.6 and 0 μg l−1 for E. coli and C. heintzii, respectively. To fit the data at D=0.075 h−1, smin for E. coli had to be raised to 34.9 μg l−1 whereas smin for C. heintzii remained zero. The results of the mathematical simulation suggest that it is not so much the higher Ks value, which is responsible for the unsuccessful competition of E. coli at low residual glucose concentration, but rather the existence of a significant smin.  相似文献   

17.
The ςS- and ς70-associated forms of RNA polymerase core enzyme (E) of Escherichia coli have very similar promoter recognition specificities in vitro. Nevertheless, the in vivo expression of many stress response genes is strongly dependent on ςS. Based on in vitro assays, it has recently been proposed that the disaccharide trehalose specifically stimulates the formation and activity of EςS and thereby contributes to promoter selectivity (S. Kusano and A. Ishihama, J. Bacteriol. 179:3649–3654, 1997). However, we demonstrate here that a trehalose-free otsA mutant exhibits growth phase-related and osmotic induction of various ςS-dependent genes which is indistinguishable from that of an otherwise isogenic wild-type strain and that stationary-phase cells do not accumulate trehalose (even though the trehalose-synthesizing enzymes are induced). We conclude that in vivo trehalose does not play a role in the expression of ςS-dependent genes and therefore also not in sigma factor selectivity at the promoters of these genes.  相似文献   

18.
Surface water and groundwater are continuously used as sources of drinking water in many metropolitan areas of the United States. The quality of water from these sources may be reduced due to increases in contaminants such as Escherichia coli from urban and agricultural runoffs. In this study, a multiplex fluorogenic PCR assay was used to quantify E. coli O157:H7 in soil, manure, cow and calf feces, and dairy wastewater in an artificial wetland. Primers and probes were designed to amplify and quantify the Shiga-like toxin 1 (stx1) and 2 (stx2) genes and the intimin (eae) gene of E. coli O157:H7 in a single reaction. Primer specificity was confirmed with DNA from 33 E. coli O157:H7 and related strains with and without the three genes. A direct correlation was determined between the fluorescence threshold cycle (CT) and the starting quantity of E. coli O157:H7 DNA. A similar correlation was observed between the CT and number of CFU per milliliter used in the PCR assay. A detection limit of 7.9 × 10−5 pg of E. coli O157:H7 DNA ml−1 equivalent to approximately 6.4 × 103 CFU of E. coli O157:H7 ml−1 based on plate counts was determined. Quantification of E. coli O157:H7 in soil, manure, feces, and wastewater was possible when cell numbers were ≥3.5 × 104 CFU g−1. E. coli O157:H7 levels detected in wetland samples decreased by about 2 logs between wetland influents and effluents. The detection limit of the assay in soil was improved to less than 10 CFU g−1 with a 16-h enrichment. These results indicate that the developed PCR assay is suitable for quantitative determination of E. coli O157:H7 in environmental samples and represents a considerable advancement in pathogen quantification in different ecosystems.  相似文献   

19.
The inhibitory activities of known microcins were evaluated against some diarrheagenic Escherichia coli strains. Some antibacterial properties of microcin J25, the most active one, were studied. A rapid two-step purification was performed. The MIC and the minimum bactericidal concentration of J25 against E. coli O157:H7 were 1 and 100 μg ml−1, respectively. A 104-CFU ml−1 contamination by this strain was destroyed in milk and meat extract by 6.25 μg of J25 ml−1 and in half-diluted egg yolk by 50 μg of J25 ml−1.  相似文献   

20.
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