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.
In bacteria, the free amino group of the methionylated initiator tRNA is specifically modified by the addition of a formyl group. The functional relevance of such a formylation for the initiation of translation is not yet precisely understood. Advantage was taken here of the availability of the fmt gene, encoding the Escherichia coli Met-tRNA(fMet) formyltransferase, to measure the influence of variations in the level of formyltransferase activity on the involvement of various mutant tRNA(fMet) and tRNA(mMet) species in either initiation or elongation in vivo. The data obtained established that formylation plays a dual role, firstly, by dictating tRNA(fMet) to engage in the initiation of translation, and secondly, by preventing the misappropriation of this tRNA by the elongation apparatus. The importance of formylation in the initiator identity of tRNA(fMet) was further shown by the demonstration that elongator tRNA(fMet) may be used in initiation and no longer in elongation, provided that it is mutated into a formylatable species and is given the three G.C base pairs characteristic of the anticodon stem of initiator tRNAs.  相似文献   

2.
Protein synthesis in eukaryotic organelles such as mitochondria and chloroplasts is widely believed to require a formylated initiator methionyl tRNA (fMet-tRNA(fMet)) for initiation. Here we show that initiation of protein synthesis in yeast mitochondria can occur without formylation of the initiator methionyl-tRNA (Met-tRNA(fMet)). The formylation reaction is catalyzed by methionyl-tRNA formyltransferase (MTF) located in mitochondria and uses N(10)-formyltetrahydrofolate (10-formyl-THF) as the formyl donor. We have studied yeast mutants carrying chromosomal disruptions of the genes encoding the mitochondrial C(1)-tetrahydrofolate (C(1)-THF) synthase (MIS1), necessary for synthesis of 10-formyl-THF, and the methionyl-tRNA formyltransferase (open reading frame YBL013W; designated FMT1). A direct analysis of mitochondrial tRNAs using gel electrophoresis systems that can separate fMet-tRNA(fMet), Met-tRNA(fMet), and tRNA(fMet) shows that there is no formylation in vivo of the mitochondrial initiator Met-tRNA in these strains. In contrast, the initiator Met-tRNA is formylated in the respective "wild-type" parental strains. In spite of the absence of fMet-tRNA(fMet), the mutant strains exhibited normal mitochondrial protein synthesis and function, as evidenced by normal growth on nonfermentable carbon sources in rich media and normal frequencies of generation of petite colonies. The only growth phenotype observed was a longer lag time during growth on nonfermentable carbon sources in minimal media for the mis1 deletion strain but not for the fmt1 deletion strain.  相似文献   

3.
An Escherichia coli strain with thermosensitive expression of the gene encoding peptide deformylase (fms) has been constructed. At nonpermissive temperatures, this strain fails to grow. The essential character of the fms gene was further used to clone by heterologous complementation the locus corresponding to Thermus thermophilus peptide deformylase. The cloned fragment also carries the methionyl-tRNA(fMet) formyltransferase gene (fmt). It is located immediately downstream from the fms gene, as in E. coli. Further sequence analysis of the region surrounding the E. coli fms-fmt locus indicates that the genes bordering the fms-fmt region are not conserved in T. thermophilus.  相似文献   

4.
In Escherichia coli, the free amino group of the aminoacyl moiety of methionyl-tRNA(fMet) is specifically modified by a transformylation reaction. To identify the nucleotides governing the recognition of the tRNA substrate by the formylase, initiator tRNA(fMet) was changed into an elongator tRNA with the help of an in vivo selection method. All the mutations isolated were in the tRNA acceptor arm, at positions 72 and 73. The major role of the acceptor arm was further established by the demonstration of the full formylability of a chimaeric tRNA(Met) containing the acceptor stem of tRNA(fMet) and the remaining of the structure of tRNA(mMet). In addition, more than 30 variants of the genes encoding tRNA(mMet) or tRNA(fMet) have been constructed, the corresponding mutant tRNA products purified and the parameters of the formylation reaction measured. tRNA(mMet) became formylatable by the only change of the G1.C72 base-pair into C1-A72. It was possible to render tRNA(mMet) as good a substrate as tRNA(fMet) for the formylase by the introduction of a limited number of additional changes in the acceptor stem. In conclusion, A73, G2.C71, C3.G70 and G4.C69 are positive determinants for the specific processing of methionyl-tRNA(fMet) by the formylase while the occurrence of a G.C or C.G base-pair between positions 1 and 72 acts as a major negative determinant. This pattern appears to account fully for the specificity of the formylase and the lack of formylation of any aminoacylated tRNA, excepting the methionyl-tRNA(fMet).  相似文献   

5.
Mutations in the fmt gene (encoding formyl methionine transferase) that eliminate formylation of initiator tRNA (Met‐tRNAi) confer resistance to the novel antibiotic class of peptide deformylase inhibitors (PDFIs) while concomitantly reducing bacterial fitness. Here we show in Salmonella typhimurium that novel mutations in initiation factor 2 (IF2) located outside the initiator tRNA binding domain can partly restore fitness of fmt mutants without loss of antibiotic resistance. Analysis of initiation of protein synthesis in vitro showed that with non‐formylated Met‐tRNAi IF2 mutants initiated much faster than wild‐type IF2, whereas with formylated fMet‐tRNAi the initiation rates were similar. Moreover, the increase in initiation rates with Met‐tRNAi conferred by IF2 mutations in vitro correlated well with the increase in growth rate conferred by the same mutations in vivo, suggesting that the mutations in IF2 compensate formylation deficiency by increasing the rate of in vivo initiation with Met‐tRNAi. IF2 mutants had also a high propensity for erroneous initiation with elongator tRNAs in vitro, which could account for their reduced fitness in vivo in a formylation‐proficient strain. More generally, our results suggest that bacterial protein synthesis is mRNA‐limited and that compensatory mutations in IF2 could increase the persistence of PDFI‐resistant bacteria in clinical settings.  相似文献   

6.
The composition and structure of lipopolysaccharides (LPS) of three isogenic strains of Yersinia pseudotuberculosis serovar O:1b (without plasmids (82-) and with plasmids pVM82 (82+) or p57 (57+)) grown at 8 or 37 degrees C were studied by chemical and immunochemical methods, SDS-polyacrylamide gel electrophoresis, and 13C-NMR spectroscopy. At the lower temperature, the (82-) and (82+) strains synthesized S-form of LPS with similar structure characterized by high acylation and immunochemical activity. On the other hand, LPS of the (82+) strain had shorter carbohydrate chains than LPS of the (82-) strain. The contents of LPS were decreased in cells of the plasmid-free strain grown at the higher temperature. LPS isolated from these cells were of the R-form and had low acylation and immunochemical activity. Total LPS content in cells of the (82+) strain did not significantly depend on the growth temperature. LPS of the warm variant of these bacteria contained a polysaccharide fragment and had moderate immunochemical activity. The cells of the (57+) strain at both growth temperatures had low LPS contents and produced LPS of low acylation without O-specific chains (cold variant) or containing O-polysaccharide with low polymerization degree (bacteria grown at 37 degrees C). The data indicate that in the absence of the plasmids, LPS synthesis is encoded by the chromosomal genes in pseudotuberculosis bacteria. Expression of the genes involved in LPS synthesis is regulated by the temperature of bacterial growth. Genes responsible for temperature-dependent regulation of LPS biosynthesis are located on chromosomal DNA. The pVM82 plasmid includes two gene groups; one group is localized in a 57-mD fragment of DNA and inhibits LPS synthesis, suppressing temperature-dependent regulation of the synthesis. The genes located in a 25-mD fragment of the pVM82 plasmid are de-repressors of the 57-mD fragment, and they restore the ability of pseudotuberculosis bacteria to synthesize relatively long LPS at both growth temperatures.  相似文献   

7.
8.
Host functions required for replication of progeny double-stranded DNA of bacteriophage G4 were examined by using metabolic inhibitors and Escherichia coli dna mutants. In dna+ bacteria, synthesis of the progeny replicative form (RF) was relatively resistant to 30 microgram/ml of chloramphenicol, but considerably sensitive to 200 microgram/ml of rifampicin. The RF replication was severely inhibited by 50 microgram/ml of mitomycin C, 50 microgram/ml of nalidixic acid, or 200 microgram/ml of novobiocin. At 41 degrees C, synthesis of G4 progeny RF was distinctly affected in a dnaC(D) mutant and in a dnaG host. The progeny RF replication was prevented at 42 degrees C in a dnaE strain as well as in a dnaB mutant. In a dnaZ strain, the synthetic rate of the progeny RF was markedly reduced at 42 degrees C. At 43 degrees C, the rate of G4 progeny RF synthesis was reduced even in dna+ or dnaA bacteria, but significant amounts of the progeny RF were still synthesized in these hosts at the high temperature. In addition to five dna gene products, host rep function was essential for the RF replication.  相似文献   

9.
Temperature is an important factor regulating microbial activity and shaping the soil microbial community. Little is known, however, on how temperature affects the most important groups of the soil microorganisms, the bacteria and the fungi, in situ. We have therefore measured the instantaneous total activity (respiration rate), bacterial activity (growth rate as thymidine incorporation rate) and fungal activity (growth rate as acetate-in-ergosterol incorporation rate) in soil at different temperatures (0-45 degrees C). Two soils were compared: one was an agricultural soil low in organic matter and with high pH, and the other was a forest humus soil with high organic matter content and low pH. Fungal and bacterial growth rates had optimum temperatures around 25-30 degrees C, while at higher temperatures lower values were found. This decrease was more drastic for fungi than for bacteria, resulting in an increase in the ratio of bacterial to fungal growth rate at higher temperatures. A tendency towards the opposite effect was observed at low temperatures, indicating that fungi were more adapted to low-temperature conditions than bacteria. The temperature dependence of all three activities was well modelled by the square root (Ratkowsky) model below the optimum temperature for fungal and bacterial growth. The respiration rate increased over almost the whole temperature range, showing the highest value at around 45 degrees C. Thus, at temperatures above 30 degrees C there was an uncoupling between the instantaneous respiration rate and bacterial and fungal activity. At these high temperatures, the respiration rate closely followed the Arrhenius temperature relationship.  相似文献   

10.
Here we report experimental evidence that the translation initiation apparatus accepts D-amino acids ((D)aa), as opposed to only L-methionine, as initiators. Nineteen (D)aa, as the stereoisomers to their natural L-amino acids, were charged onto initiator tRNA(fMet)(CAU) using flexizyme technology and tested for initiation in a reconstituted Escherichia coli translation system lacking methionine, i.e., the initiator was reprogrammed from methionine to (D)aa. Remarkably, all (D)aa could initiate translation while the efficiency of initiation depends upon the type of side chain. The peptide product initiated with (D)aa was generally in a nonformylated form, indicating that methionyl-tRNA formyltransferase poorly formylated the corresponding (D)aa-tRNA(fMet)(CAU). Although the inefficient formylation of (D)aa-tRNA(fMet)(CAU) resulted in modest expression of the corresponding peptide, preacetylation of (D)aa-tRNA(fMet)(CAU) dramatically increased expression level, implying that the formylation efficiency is one of the critical determinants of initiation efficiency with (D)aa. Our findings provide not only the experimental evidence that translation initiation tolerates (D)aa, but also a new means for the mRNA-directed synthesis of peptides capped with (D)aa or acyl-(D)aa at the N terminus.  相似文献   

11.
Deep-sea bacteria were isolated from the digestive tract of animals inhabiting depths of 5900 m in the Puerto Rico Trench and 4300 m near the Walvis Ridge. Growth of two bacterial strains was measured in marine broth and in solid media under a range of pressures and temperatures. Both strains were barophilic at 2 degrees C (+/- 1 degrees C) with an optimal growth rate of 0.22 h-1 at a pressure 30% lower than that encountered in situ. At 1 atm they grew at temperatures ranging from 1.2 to 18.2 degrees C (+/- 0.3 degrees C), while in situ pressures increased the upper temperature limit to 23.3 degrees C. Both strains were identified as members of the genus Vibrio, based on standard taxonomic tests and mol% G + C values (47.0 and 47.1). Ribonucleotide sequences determined for 5S ribosomal RNA from each strain confirmed relationship to the Vibrio-Photobacterium group, as represented by V. harveyi and P. phosphoreum, but the barophiles were clearly distinct from these species. Secondary structure conformed to the established model for eubacterial 5S rRNA.  相似文献   

12.
Starting from a p-aminobenzoate-requiring strain of Escherichia coli (E. coli K-12 AB3292), we have isolated mutants that can grow in the absence of p-aminobenzoate (and thus tetrahydrofolate). The following lines of evidence suggest that at least one of these mutants is capable of initiating protein synthesis without formylation of methionyl-transfer ribonucleic acid (methionyl-tRNA(fMet)). (i) tRNA isolated (and charged in vivo with [(35)S]methionine) from this mutant grown in a p-aminobenzoate-free medium contained less than 0.4% of the total methionine charged to the tRNA as formylmethionine. However, when the mutant was grown in the presence of p-aminobenzoate, 40 to 50% of the total [(35)S]methionine was detected as formylmethionine. (ii) Extracts of the mutant grown in the absence of p-aminobenzoate contained no formyl-tetrahydrofolate, but such extracts did contain formylatable methionyl-tRNA and a functional transformylase. (iii) Tetrahydrofolate-free extracts of the mutant were capable of supporting protein synthesis with viral RNA (from f2) as messenger, but the resulting synthesized proteins contained no formylmethionine, and methionine residues were detected where formylmethionine residues are normally found. In the presence of formyl-tetrahydrofolate, use of a similar extract resulted in the detection of 30 to 40% of the total polypeptide methionine as formylmethionine. (iv) Initiation of protein synthesis in vitro occurred more readily with formyl-tetrahydrofolate-free extracts of the mutant than with similar extracts prepared from the parent strain. However, in the presence of formyl-tetrahydrofolate, initiation of protein synthesis proceeded equally well with both kinds of extracts. tRNA from this mutant and another spontaneously derived mutant was found to be partially deficient in the modified nucleoside ribothymidine (rT). Analysis of extracts showed that the mutants contained decreased levels of the methylase that results in the formation of ribothymidine. In vivo studies with an independently isolated rT(-) strain suggest that the lack of rT in tRNA facilitates the growth of E. coli under conditions where protein synthesis is forced to take place without formylation.  相似文献   

13.
GroEL/S is a molecular chaperone system in Escherichia coli which not only assists the folding of intracellular proteins but also affects the cellular activity against the change of environmental condition. Here we show that the growth rate of E. coli DH5alpha can be improved at low temperature by expressing a GroEL/S variant achieved through irrational protein engineering approach. The GroELS variant (GroELS(var)) accelerating the growth of E. coli DH5alpha was screened through enrichment culture of the mutant libraries obtained by random mutagenesis. E. coli DH5alpha harboring the groELS(var) gene exhibited approximately 1.5-2 times higher growth rate compared to the strain with wild-type GroELS at 15-30 degrees C. At 10 degrees C, a temperature that the growth of E. coli DH5alpha almost stops, the GroELS(var) triggered the growth of E. coli DH5alpha. We identified that seven nucleotides of groELS gene and six amino acids of the GroELS were changed through the mutagenesis and screening. Site directed mutagenic analysis revealed that H360 in GroEL(var) is the most crucial residue determining the activity of GroELS(var) and more than one of the other residues in GroEL(var) may be additionally involved in the activity of GroELS(var). The improvement of growth rate induced by the GroELS(var) was observed only in the strain DH5alpha and not detected in other E. coli strains, such as BL21, BW25113, codon+, JM110, Top10, and XL1-blue.  相似文献   

14.
15.
To clarify the characteristics of thermophilic bacteria in cattle manure compost, enzymatic activity and species diversity of cultivated bacteria were investigated at 54, 60, 63, 66 and 70 degrees C, which were dependent on composting temperature. The highest level of thermophilic bacterial activity was observed at 54 degrees C. Following an increase in temperature to 63 degrees C, a reduction in bacterial diversity was observed. At 66 degrees C, bacterial diversity increased again, and diverse bacteria including Thermus spp. and thermophilic Bacillus spp. appeared to adapt to the higher temperature. At 70 degrees C, bacterial activity measured as superoxide dismutase and catalase activity was significantly higher than at 66 degrees C. However, the decomposition rate of protein in the compost was lower than the rate at 66 degrees C due to the higher compost temperature.  相似文献   

16.
This report describes a mutant of Listeria monocytogenes strain 10403S (serotype 1/2a) with a defective response to conditions of high osmolarity, an environment that L. monocytogenes encounters in some ready-to-eat foods. A library of L. monocytogenes clones mutagenized with Tn917 was generated and scored for sensitivity to 4% NaCl in order to identify genes responsible for growth or survival in elevated-NaCl environments. One of the L. monocytogenes Tn917 mutants, designated strain OSM1, was selected, and the gene interrupted by the transposon was sequenced. A BLAST search with the putative translated amino acid sequence indicated that the interrupted gene product was a homolog of htrA (degP), a gene coding for a serine protease identified as a stress response protein in several gram-positive and gram-negative bacteria. An htrA deletion strain, strain LDW1, was constructed, and the salt-sensitive phenotype of this strain was complemented by introduction of a plasmid carrying the wild-type htrA gene, demonstrating that htrA is necessary for optimal growth under conditions of osmotic stress. Additionally, strain LDW1 was tested for its response to temperature and H(2)O(2) stresses. The results of these growth assays indicated that strain LDW1 grew at a lower rate than the wild-type strain at 44 degrees C but at a rate similar to that of the wild-type strain when incubated at 4 degrees C. In addition, strain LDW1 was significantly more sensitive to a 52 degrees C heat shock than the wild-type strain. Strain LDW1 was also defective in its response to H(2)O(2) challenge at 37 degrees C, since 100 or 150 micro g of H(2)O(2) was more inhibitory for the growth of strain LDW1 than for that of the parent strain. The stress response phenotype observed for strain LDW1 is similar to that observed for other HtrA(-) organisms, which suggests that L. monocytogenes HtrA may play a role in degrading misfolded proteins that accumulate under stress conditions.  相似文献   

17.
Previous work by Schulman and Pelka (1975) J. Biol. Chem. 250, 542-547, indicated that the absence of a pairing between the bases 1 and 72 in initiator tRNA(fMet) explained the relatively small activity of peptidyl-tRNA hydrolase towards N-acetyl-methionyl-tRNA(fMet). In the present study, the structural requirements for the sensitivity of an N-acetyl-aminoacyl-tRNA to Escherichia coli peptidyl-tRNA hydrolase activity have been further investigated. Ten derivatives of tRNA(fMet) with various combinations of bases at positions 1 and 72 in the acceptor stem have been produced, aminoacylated and chemically acetylated. The release of the aminoacyl moiety from these tRNA derivatives was assayed in the presence of peptidyl-tRNA hydrolase purified from an overproducing strain. tRNA(fMet) derivatives with either C1A72, C1C72, U1G72, U1C72 or A1C72 behaved as poor substrates of the enzyme, as compared to those with C1G72, U1A72, G1C72, A1U72 or G1U72. With the exception of U1G72, it could be therefore concluded that the relative resistance of tRNA(fMet) to peptidyl-tRNA hydrolase did not depend on a particular combination of nucleotides at positions 1 and 72, but rather reflected the absence of a base pairing at these positions. In a second series of experiments, the unpairing of the 1 and 72 bases, created with C-A or A-C bases, instead of G-C in methionyl-tRNA(mMet) or in valyl-tRNA(Val1), was shown to markedly decrease the rate of hydrolysis catalysed by peptidyl-tRNA hydrolase. Altogether, the data indicate that the stability of the 1-72 pair governs the degree of sensitivity of a peptidyl-tRNA to peptidyl-tRNA hydrolase.  相似文献   

18.
The infB gene encodes translation initiation factor IF2. We have determined the entire sequence of infB from two cold-sensitive Escherichia coli strains IQ489 and IQ490. These two strains have been isolated as suppressor strains for the temperature-sensitive secretion mutation secY24. The mutations causing the suppression phenotype are located within infB. The only variations from the wild-type (wt) infB found in the two mutant strains are a replacement of Asp409 with Glu in strain IQ489 and an insertion of Gly between Ala421 and Gly422 in strain IQ490. Both positions are located in the GTP-binding G-domain of IF2. A model of the G-domain of E.coli IF2 is presented in. Physiological quantities of the recombinant mutant proteins were expressed in vivo in E.coli strains from which the chromosomal infB gene has been inactivated. At 42 degrees C, the mutants sustained normal cell growth, whereas a significant decrease in growth rate was found at 25 degrees C for both mutants as compared to wt IF2 expressed in the control strain. Circular dichroism spectra were recorded of the wt and the two mutant proteins to investigate the structural properties of the proteins. The spectra are characteristic of alpha-helix dominated structure, and reveal a significant different behavior between the wt and mutant IF2s with respect to temperature-induced conformational changes. The temperature-induced conformational change of the wt IF2 is a two-state process. In a ribosome-dependent GTPase assay in vitro the two mutants showed practically no activity at temperatures below 10 degrees C and a reduced activity at all temperatures up to 45 degrees C, as compared to wt IF2. The results indicate that the amino acid residues, Asp409 and Gly422, are located in important regions of the IF2 G-domain and demonstrate the importance of GTP hydrolysis in translation initiation for optimal cell growth.  相似文献   

19.
Initiation of protein synthesis in mitochondria and chloroplasts is widely believed to require a formylated initiator methionyl-tRNA (fMet-tRNAfMet) in a process involving initiation factor 2 (IF2). However, yeast strains disrupted at the FMT1 locus, encoding mitochondrial methionyl-tRNA formyltransferase, lack detectable fMet-tRNAfMet but exhibit normal mitochondrial function as evidenced by normal growth on non-fermentable carbon sources. Here we show that mitochondrial translation products in Saccharomyces cerevisiae were synthesized in the absence of formylated initiator tRNA. ifm1 mutants, lacking the mitochondrial initiation factor 2 (mIF2), are unable to respire, indicative of defective mitochondrial protein synthesis, but their respiratory defect could be complemented by plasmid-borne copies of either the yeast IFM1 gene or a cDNA encoding bovine mIF2. Moreover, the bovine mIF2 sustained normal respiration in ifm1 fmt1 double mutants. Bovine mIF2 supported the same pattern of mitochondrial translation products as yeast mIF2, and the pattern did not change in cells lacking formylated Met-tRNAfMet. Mutant yeast lacking any mIF2 retained the ability to synthesize low levels of a subset of mitochondrially encoded proteins. The ifm1 null mutant was used to analyze the domain structure of yeast mIF2. Contrary to a previous report, the C terminus of yeast mIF2 is required for its function in vivo, whereas the N-terminal domain could be deleted. Our results indicate that formylation of initiator methionyl-tRNA is not required for mitochondrial protein synthesis. The ability of bovine mIF2 to support mitochondrial translation in the yeast fmt1 mutant suggests that this phenomenon may extend to mammalian mitochondria as well.  相似文献   

20.
The low ethanol tolerance of thermophilic anaerobic bacteria (<2%, v/v) is a major obstacle for their industrial exploitation for ethanol production. The ethanol tolerance of the thermophilic anaerobic ethanol-producing strain Thermoanaerobacter A10 was studied during batch tests of xylose fermentation at a temperature range of 50-70 degrees C with exogenously added ethanol up to approximately 6.4% (v/v). At the optimum growth temperature of 70 degrees C, the strain was able to tolerate 4.7% (v/v) ethanol, and growth was completely inhibited at 5.6% (v/v). A higher ethanol tolerance was found at lower temperatures. At 60 degrees C, the strain was able to tolerate at least 5.1% (v/v) ethanol. A generalized form of Monod kinetic equation proposed by Levenspiel was used to describe the ethanol (product) inhibition. The model predicted quite well the experimental data for the temperature interval 50-70 degrees C, and the maximum specific growth rate and the toxic power (n), which describes the order of ethanol inhibition at each temperature, were estimated. The toxic power (n) was 1.33 at 70 degrees C, and corresponding critical inhibitory product concentration (P(crit)) above which no microbial growth occurs was determined to be 5.4% (v/v). An analysis of toxic power (n) and P(crit) showed that the optimum temperature for combined microbial growth and ethanol tolerance was 60 degrees C. At this temperature, the toxic power (n), and P(crit) were 0.50, and 6.5% (v/v) ethanol, respectively. From a practical point of view, the model may be applied to compare the ethanol inhibition (ethanol tolerance) on microbial growth of different thermophilic anaerobic bacterial strains.  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