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 8 毫秒
1.
2.
Salmonella typhimurium LT2 contains intervening sequences (IVSs) of 90–110 nt within all its 23S rRNA that are cleaved out by RNase III, resulting in rRNA fragmentation. In order to determine the functionality of 23S rRNA that contains unexcised IVSs, we constructed an S. typhimurium RNase III (rnc) deficient strain by transducing a mini-Tn10 (rnc-14::Tn10) from Escherichia coli K-12. The resulting strain of S. typhimurium was viable, contained IVSs within all of its 23S rRNA, and showed a growth reduction similar to that observed for the RNase III deficient strain of E. coli. These results indicate that ribosomes containing 23S rRNA in which IVSs are not excised are functional in translation, and make it unlikely that RNase III excision of IVSs from strain LT2 23S rRNA is dictated by a selective pressure to uphold the functional integrity of ribosomes.  相似文献   

3.
The rrl genes for 23S rRNA of Salmonella typhimurium LT2 are known to carry intervening sequences (IVSs) at two sites, helix-25 and helix-45, which are excised by RNase III during rRNA maturation, resulting in rRNA which is fragmented but nevertheless functional. We isolated DNA fragments containing the seven rrl genes from BlnI, I-CeuI, and SpeI genomic digests following pulsed-field gel electrophoresis and used these DNA fragments as templates for PCRs utilizing primers upstream and downstream of helix-25 and helix-45. Variance in amplicon length and cycle sequencing indicated that rrlG and rrlH have IVSs in helix-25 of approximately 110 bp which are only 56% identical. rrnA, rrnB, rrnC, rrnD, rrnE, and rrnH have IVSs of approximately 90 bp in helix-45, and all have the same nucleotide sequence. Twenty-one independent wild-type strains of S. typhimurium from Salmonella Reference Collection A were analyzed for IVSs by using PCRs with genomic DNAs and by denaturing agarose electrophoresis of RNAs. Many strains resemble LT2, but some have no IVSs in helix-25 and others have IVSs in helix-45 in all seven rrl genes. However, the IVSs in individual wild-type lines are relatively stable, for several LT2 isolates separated over many years by many single-colony isolations are indistinguishable from one another, with the exception of line LB5010, which differs by one helix-25 IVS. We postulate that IVSs have entered strain LT2 by three independent lateral-transfer events and that the IVS in helix-45 was dispersed to and maintained in the same sequence in six of the seven rrl genes by the mechanism of gene conversion.  相似文献   

4.
In Rhodobacter capsulatus and Rhizobium leguminosarum, an internal transcribed spacer consisting of helices 9 and 10 is removed during 23S rRNA processing, which leads to the occurrence of a 5.8S-like rRNA. The particular rRNA maturation steps are not known, with exception of the initial RNase III cleavage in helix 9. We found that GC-rich stem-loop structures of helix 9, which are released by RNase III, are immediately degraded. The degradation of helix 10 is slower and its kinetics differs in both species. Nevertheless, the helix 10 processing mechanism is conserved and includes cleavages by RNase E.  相似文献   

5.
E Kordes  S Jock  J Fritsch  F Bosch    G Klug 《Journal of bacteriology》1994,176(4):1121-1127
In Rhodobacter capsulatus wild-type strains, the 23S rRNA is cleaved into [16S] and [14S] rRNA molecules. Our data show that a region predicted to form a hairpin-loop structure is removed from the 23S rRNA during this processing step. We have analyzed the processing of rRNA in the wild type and in the mutant strain Fm65, which does not cleave the 23S rRNA. In addition to the lack of 23S rRNA processing, strain Fm65 shows impeded processing of a larger 5.6-kb rRNA precursor and slow maturation of 23S and 16S rRNAs from pre-23S and pre-16S rRNA species. Similar effects have also been described previously for Escherichia coli RNase III mutants. Processing of the 5.6-kb precursor was independent of protein synthesis, while the cleavage of 23S rRNA to generate 16S and 14S rRNA required protein synthesis. We identified a DNA fragment of the wild-type R. capsulatus chromosome that conferred normal processing of 5.6-kb rRNA and 23S rRNA when it was expressed in strain Fm65.  相似文献   

6.
The excision of intervening sequences from Salmonella 23S ribosomal RNA   总被引:31,自引:0,他引:31  
A B Burgin  K Parodos  D J Lane  N R Pace 《Cell》1990,60(3):405-414
Novel, approximately 90 bp intervening sequences (IVs) were discovered within the 23S rRNA genes of S. typhimurium and S. arizonae. These non-rRNA sequences are transcribed and then excised during rRNA maturation. The rRNA fragments that result from the excision of the extra sequences are not religated. This results in fragmented 23S rRNAs. The excision of one IVS was shown to be catalyzed in vivo and in vitro by ribonuclease III. These IVSs are highly volatile evolutionarily, sometimes occurring in only some of the multiple rRNA operons of a particular cell. The sporadic nature of the occurrence of fragmented rRNAs among closely related organisms argues that such fragmentation is a derived state, not a primitive one. Possible sources of these IVSs, their parallels with internal transcribed spacers and introns in eukaryotes, and their possible roles in the evolutionary process are discussed.  相似文献   

7.
Intervening sequences (IVSs) were originally identified in the rrl genes for 23S rRNA (rrl genes, for large ribosomal subunit, part of rrn operon encoding rRNA) of Salmonella enterica serovars Typhimurium LT2 and Arizonae. These sequences are transcribed but later removed during RNase III processing of the rRNA, resulting in fragmentation of the 23S species; IVSs are uncommon, but have been reported in at least 10 bacterial genera. Through PCR amplification of IVS-containing regions of the rrl genes we showed that most Proteus and Providencia strains contain IVSs similar to those of serovar Typhimurium in distribution and location in rrl genes. By extraction and Northern blotting of rRNA, we also found that these IVSs result in rRNA fragmentation. We report the first finding of two very different sizes of IVS (113 bp and 183 to 187 bp) in different rrl genes in the same strain, in helix 25 of Proteus and Providencia spp.; IVSs from helix 45 are 113 to 123 bp in size. Analysis of IVS sequence and postulated secondary structure reveals striking similarities of Proteus and Providencia IVSs to those of serovar Typhimurium, with the stems of the smaller IVSs from helix 25 being similar to those of Salmonella helix 25 IVSs and with both the stem and the central loop domain of helix 45 IVSs being similar. Thus, IVSs of related sequences are widely distributed throughout the Enterobacteriaceae, in Salmonella, Yersinia, Proteus, and Providencia spp., but we did not find them in Escherichia coli, Citrobacter, Enterobacter, Klebsiella, or Morganella spp.; the sporadic distribution of IVSs of related sequence indicates that lateral genetic transfer has occurred.  相似文献   

8.
Although the absence of intervening sequences (IVSs) within the 23S rRNA genes in Campylobacter lari isolates has been described, there are apparently no reports regarding correlations between the nucleotide sequences of 23S rRNA genes and erythromycin (Ery) susceptibility in C. lari isolates. Here, we determined the minimum inhibitory concentrations of 35 C. lari isolates [n?=?19 for urease-positive thermophilic Campylobacter (UPTC); n?=?16 urease-negative (UN) C. lari] obtained from Asia, Europe, and North America. We found that the 18 isolates were resistant to the Ery (defined as ≧8 μg/mL), and three isolates, UPTC A1, UPTC 92251, and UPTC 504, showed increased resistance (16 μg/mL). No correlations between the IVSs in the helix 45 region within the 23S rRNA gene sequences and Ery resistance were identified in the C. lari isolates examined. In addition, no point mutations occurred at any expected or putative position within the V domain in the isolates. In conclusion, antibiotic resistance against the macrolide erythromycin is mediated through an alternative pathway to that described above.  相似文献   

9.
10.
Ramakanth Madhugiri 《FEBS letters》2009,583(14):2339-2342
Sinorhizobium meliloti harbours genes encoding orthologs of ribonuclease (RNase) E and RNase J, the principle endoribonucleases in Escherichia coli and Bacillus subtilis, respectively. To analyse the role of RNase J in S. meliloti, RNA from a mutant with miniTn5-insertion in the RNase J-encoding gene was compared to the wild-type and a difference in the length of the 5.8S-like ribosomal RNA (rRNA) was observed. Complementation of the mutant, Northern blotting and primer extension revealed that RNase J is necessary for the 5′-end maturation of 16S rRNA and of the two 23S rRNA fragments, but not of 5S rRNA.  相似文献   

11.
Rhodobacter sphaeroides showed chemotaxis towards L-alanine but not towards the analog 2-aminoisobutyrate. 2-Aminoisobutyrate and alanine were shown to share a common transport system, but 2-aminoisobutyrate was not metabolized. Chemotaxis towards alanine was inhibited by structurally unrelated metabolites, suggesting cross-inhibition by common metabolic intermediates.  相似文献   

12.
13.
Rhodobacter sphaeroides has a complex chemosensory system comprising two classic CheAs, two atypical CheAs, and eight response regulators (six CheYs and two CheBs). The classic CheAs, CheA(1) and CheA(2), have similar domain structures to Escherichia coli CheA, whereas the atypical CheAs, CheA(3) and CheA(4), lack some of the domains found in E. coli CheA. CheA(2), CheA(3), and CheA(4) are all essential for chemotaxis. Here we demonstrate that CheA(3) and CheA(4) are both unable to undergo ATP-dependent autophosphorylation, however, CheA(4) is able to phosphorylate CheA(3). The in vitro kinetics of this phosphorylation reaction were consistent with a reaction mechanism in which CheA(3) associates with a CheA(4) dimer forming a complex, CheA(3)A(4). To the best of our knowledge, CheA(3)A(4) is the first characterized histidine protein kinase where the subunits are encoded by distinct genes. Selective phosphotransfer was observed from CheA(3)-P to the response regulators CheY(1), CheY(6), and CheB(2). Using phosphorylation site and kinase domain mutants of CheA we show that phosphosignaling involving CheA(2), CheA(3), and CheA(4) is essential for chemotaxis in R. sphaeroides. Interestingly, CheA(3) was not phosphorylated in vitro by CheA(1) or CheA(2), although CheA(1) and CheA(2) mutants with defective kinase domains were phosphorylated by CheA(4). Because in vivo CheA(3) and CheA(4) localize to the cytoplasmic chemotaxis cluster, while CheA(2) localizes to the polar chemotaxis cluster, it is likely that the physical separation of CheA(2) and CheA(4) prevents unwanted cross-talk between these CheAs.  相似文献   

14.
15.
G Afseth  Y Y Mo    L P Mallavia 《Journal of bacteriology》1995,177(10):2946-2949
Characterization of the rRNA operon from the obligate intracellular bacterium Coxiella burnetii has determined the order of the rRNA genes to be 16S-23S-5S. A 444-bp intervening sequence (IVS) was identified to interrupt the 23S rRNA gene beginning at position 1176. The IVS is predicted to form a stem-loop structure formed by flanking inverted repeats, and the absence of intact 23S rRNA molecules suggests that the loop is removed. An open reading frame in the IVS has been identified that shows 70% similarity at the amino acid level to IVS open reading frames characterized from four species of Leptospira.  相似文献   

16.
Ordered processing of Escherichia coli 23S rRNA in vitro.   总被引:6,自引:2,他引:4       下载免费PDF全文
In an RNase III-deficient strain of E. coli 23S pre-rRNA accumulates unprocessed in 50S ribosomes and in polysomes. These ribosomes provide a substrate for the analysis of rRNA maturation in vitro. S1 nuclease protection analysis of the products obtained in in vitro processing reactions demonstrates that 23S rRNA processing is ordered. The double stranded stem of 23S rRNA is cleaved by RNase III in vitro to two intermediate RNAs at the 5' end and one at the 3' end. Mature termini are then produced by other enzyme(s) in a soluble protein fraction from wild-type cells. The nature of the reaction at the 5' end is not clear, but the reaction at the 3' end is exonucleolytic, producing three heterogeneous mature termini. The two reactions are coordinated; 3' end maturation progresses concurrently with cleavages at the 5' end. Two results suggest a possible link between final maturation and translation: in vitro, mature termini are formed efficiently in the presence of additives required for protein synthesis; and all the processing intermediates detected from in vitro reactions are also found in polysomes from wild-type cells.  相似文献   

17.
Rhodobacter sphaeroides is a motile bacterium that has multiple chemotaxis genes organized predominantly in three major operons (cheOp(1), cheOp(2), and cheOp(3)). The chemoreceptor proteins are clustered at two distinct locations, the cell poles and in one or more cytoplasmic clusters. One intriguing possibility is that the physically distinct chemoreceptor clusters are each composed of a defined subset of specific chemotaxis proteins, including the chemoreceptors themselves plus specific CheW and CheA proteins. Here we report the subcellular localization of one such protein, CheA(2), under aerobic and photoheterotrophic growth conditions. CheA(2) is predominantly clustered and localized at the cell poles under both growth conditions. Furthermore, its localization is dependent upon one or more genes in cheOp(2) but not those of cheOp(1) or cheOp(3). In E. coli, the polar localization of CheA depends upon CheW. The R. sphaeroides cheOp(2) contains two cheW genes. Interestingly, CheW(2) is required under both aerobic and photoheterotrophic conditions, whereas CheW(3) is not required under aerobic conditions but appears to play a modest role under photoheterotrophic conditions. This suggests that R. sphaeroides contains at least two distinct chemotaxis complexes, possibly composed of proteins dedicated for each subcellular location. Furthermore, the composition of these spatially distinct complexes may change under different growth conditions.  相似文献   

18.
19.
The Escherichia coli two-component chemosensory pathway has been extensively studied, and its response regulator, CheY, has become a paradigm for response regulators. However, unlike E. coli, most chemotactic nonenteric bacteria have multiple CheY homologues. The roles and cellular localization of the CheYs in Rhodobacter sphaeroides were determined. Only two CheYs were required for chemotaxis, CheY(6) and either CheY(3) or CheY(4). These CheYs were partially localized to either of the two chemotaxis signaling clusters, with the remaining protein delocalized. Interestingly, mutation of the CheY(6) phosphorylatable aspartate to asparagine produced a stopped motor, caused by phosphorylation on alternative site Ser-83 by CheA. Extensive mutagenesis of E. coli CheY has identified a number of activating mutations, which have been extrapolated to other response regulators (D13K, Y106W, and I95V). Analogous mutations in R. sphaeroides CheYs did not cause activation. These results suggest that although the R. sphaeroides and E. coli CheYs are similar in that they require phosphorylation for activation, they may differ in both the nature of the phosphorylation-induced conformational change and their subsequent interactions with the flagellar motor. Caution should therefore be used when projecting from E. coli CheY onto novel response regulators.  相似文献   

20.
The H/ACA RNAs represent an abundant, evolutionarily conserved and functionally diverse class of non‐coding RNAs. Many H/ACA RNAs direct pseudouridylation of rRNAs and snRNAs, while members of the rapidly growing group of ‘orphan’ H/ACA RNAs participate in pre‐rRNA processing, telomere synthesis and probably, in other nuclear processes. The yeast snR30 ‘orphan’ H/ACA snoRNA has long been known to function in the nucleolytic processing of 18S rRNA, but its molecular role remained unknown. Here, we provide biochemical and genetic evidence demonstrating that during pre‐rRNA processing, two evolutionarily conserved sequence elements in the 3′‐hairpin of snR30 base‐pair with short pre‐rRNA sequences located in the eukaryote‐specific internal region of 18S rRNA. The newly discovered snR30‐18S base‐pairing interactions are essential for 18S rRNA production and they constitute a complex snoRNA target RNA transient structure that is novel to H/ACA RNAs. We also demonstrate that besides the 18S recognition motifs, the distal part of the 3′‐hairpin of snR30 contains an additional snoRNA element that is essential for 18S rRNA processing and that functions most likely as a snoRNP protein‐binding site.  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