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.
Ribosome-targeting antibiotics block protein synthesis by binding at functionally important regions of the bacterial rRNA. Resistance is often conferred by addition of a methyl group at the antibiotic binding site within an rRNA region that is already highly modified with several nucleotide methylations. In bacterial rRNA, each methylation requires its own specific methyltransferase enzyme, and this raises the question as to how an extra methyltransferase conferring antibiotic resistance can be accommodated and how it can gain access to its nucleotide target within a short and functionally crowded stretch of the rRNA sequence. Here, we show that the Sgm methyltransferase confers resistance to 4,6-disubstituted deoxystreptamine aminoglycosides by introducing the 16S rRNA modification m7G1405 within the ribosomal A site. This region of Escherichia coli 16S rRNA already contains several methylated nucleotides including m4Cm1402 and m5C1407. Modification at m5C1407 by the methyltransferase RsmF is impeded as Sgm gains access to its adjacent G1405 target on the 30S ribosomal subunit. An Sgm mutant (G135A), which is impaired in S-adenosylmethionine binding and confers lower resistance, is less able to interfere with RsmF methylation on the 30S subunit. The two methylations at 16S rRNA nucleotide m4Cm1402 are unaffected by both the wild-type and the mutant versions of Sgm. The data indicate that interplay between resistance methyltransferases and the cell''s own indigenous methyltransferases can play an important role in determining resistance levels.  相似文献   

2.
3.
4.
Methylation at the 5-position of cytosine [m5C (5-methylcytidine)] occurs at three RNA nucleotides in Escherichia coli. All these modifications are at highly conserved nucleotides in the rRNAs, and each is catalyzed by its own m5C methyltransferase enzyme. Two of the enzymes, RsmB and RsmF, are already known and methylate 16S rRNA at nucleotides C967 and C1407, respectively. Here, we report the identity of the third E. coli m5C methyltransferase. Analysis of rRNAs by matrix-assisted laser desorption/ionization mass spectrometry showed that inactivation of the yccW gene leads to loss of m5C methylation at nucleotide 1962 in E. coli 23S rRNA. This methylation is restored by complementing the knockout strain with a plasmid-encoded copy of the yccW gene. Purified recombinant YccW protein retains its specificity for C1962 in vitro and methylates naked 23S rRNA isolated from the yccW knockout strain. However, YccW does not methylate assembled 50S subunits, and this is somewhat surprising as the published crystal structures show nucleotide C1962 to be fully accessible at the subunit interface. YccW-directed methylation at nucleotide C1962 is conserved in bacteria, and loss of this methylation in E. coli marginally reduces its growth rate. YccW had previously eluded identification because it displays only limited sequence similarity to the m5C methyltransferases RsmB and RsmF and is in fact more similar to known m5U (5-methyluridine) RNA methyltransferases. In keeping with the previously proposed nomenclature system for bacterial rRNA methyltransferases, yccW is now designated as the rRNA large subunit methyltransferase gene rlmI.  相似文献   

5.
Precursor and mature ribosomal RNA molecules from Xenopus laevis were examined by electron microscopy. A reproducible arrangement of hairpin loops was observed in these molecules. Maps based on this secondary structure were used to determine the arrangement of sequences in precursor RNA molecules and to identify the position of mature rRNAs within the precursors. A processing scheme was derived in which the 40 S rRNA is cleaved to 38 S RNA, which then yields 34 S plus 18 S RNA. The 34 S RNA is processed to 30 S, and finally to 28 S rRNA. The pathway is analogous to that of L-cell rRNA but differs from HeLa rRNA in that no 20 S rRNA intermediate was found. X. laevis 40 S rRNA (Mr = 2.7 × 106) is much smaller than HeLa or L-cell 45 8 rRNA (Mr = 4.7 × 106), but the arrangement of mature rRNA sequences in all precursors is very similar. Experiments with ascites cell 3′-exonuclease show that the 28 S region is located at or close to the 5′-end of the 40 S rRNA.Secondary structure maps were obtained also for single-stranded molecules of ribosomal DNA. The region in the DNA coding for the 40 S rRNA could be identified by its regular structure, which closely resembles that of the RNA. Regions corresponding to the 40 S RNA gene alternate with non-transcribed spacer regions along strands of rDNA. The latter have a large amount of irregular secondary structure and vary in length between different repeating units. A detailed map of the rDNA repeating unit was derived from these experiments.Optical melting studies are presented, showing that rRNAs with a high (G + C) content exhibit significant hypochromicity in the formamide/urea-containing solution that was used for spreading.  相似文献   

6.
Ribosomal RNA undergoes various modifications to optimize ribosomal structure and expand the topological potential of RNA. The most common nucleotide modifications in ribosomal RNA (rRNA) are pseudouridylations and 2′-O methylations (Nm), performed by H/ACA box snoRNAs and C/D box snoRNAs, respectively. Furthermore, rRNAs of both ribosomal subunits also contain various base modifications, which are catalysed by specific enzymes. These modifications cluster in highly conserved areas of the ribosome. Although most enzymes catalysing 18S rRNA base modifications have been identified, little is known about the 25S rRNA base modifications. The m1A modification at position 645 in Helix 25.1 is highly conserved in eukaryotes. Helix formation in this region of the 25S rRNA might be a prerequisite for a correct topological framework for 5.8S rRNA to interact with 25S rRNA. Surprisingly, we have identified ribosomal RNA processing protein 8 (Rrp8), a nucleolar Rossman-fold like methyltransferase, to carry out the m1A base modification at position 645, although Rrp8 was previously shown to be involved in A2 cleavage and 40S biogenesis. In addition, we were able to identify specific point mutations in Rrp8, which show that a reduced S-adenosyl-methionine binding influences the quality of the 60S subunit. This highlights the dual functionality of Rrp8 in the biogenesis of both subunits.  相似文献   

7.
Ribosomal RNA guanine-(N2)-methyltransferases and their targets   总被引:3,自引:0,他引:3  
Five nearly universal methylated guanine-(N2) residues are present in bacterial rRNA in the ribosome. To date four out of five ribosomal RNA guanine-(N2)-methyltransferases are described. RsmC(YjjT) methylates G1207 of the 16S rRNA. RlmG(YgjO) and RlmL(YcbY) are responsible for the 23S rRNA m2G1835 and m2G2445 formation, correspondingly. RsmD(YhhF) is necessary for methylation of G966 residue of 16S rRNA. Structure of Escherichia coli RsmD(YhhF) methyltransferase and the structure of the Methanococcus jannaschii RsmC ortholog were determined. All ribosomal guanine-(N2)-methyltransferases have similar AdoMet-binding sites. In relation to the ribosomal substrate recognition, two enzymes that recognize assembled subunits are relatively small single domain proteins and two enzymes that recognize naked rRNA are larger proteins containing separate methyltransferase- and RNA-binding domains. The model for recognition of specific target nucleotide is proposed. The hypothetical role of the m2G residues in rRNA is discussed.  相似文献   

8.
The chloroplast ribosomal unit of Chlamydomonas reinhardii displays two features which are not shared by other chloroplast ribosomal units. These include the presence of an intron in the 23 S ribosomal RNA gene and of two small genes coding for 3 S and 7 S rRNA in the spacer between the 16 S and 23 S rRNA genes (Rochaix & Malnoë, 1978). Sequencing of the 7 S and 3 S rRNAs as well as their genes and neighbouring regions has shown that: (1) the 7 S and 3 S rRNA genes are 282 and 47 base-pairs long, respectively, and are separated by a 23 base-pair A + T-rich spacer. (2) A sequence microheterogeneity exists within the 3 S RNA genes. (3) The sequences of the 7 S and 3 S rRNAs are homologous to the 5′ termini of prokaryotic and other chloroplast 23 S rRNAs, indicating that the C. reinhardii counterparts of 23 S rRNA have a composite structure. (4) The sequences of the 7 S and 3 S rRNAs are related to that of cytoplasmic 5.8 S rRNA, suggesting that these RNAs may perform similar functions in the ribosome. (5) Partial nucleotide sequence complementarity is observed between the 5′ ends of the 7 S and 3 S RNAs on one hand and the 23 S rRNA sequences which flank the ribosomal intron on the other. These data are compatible with the idea that these small rRNAs may play a role in the processing of the 23 S rRNA precursor.  相似文献   

9.
Ribosomal RNAs (rRNAs), assisted by ribosomal proteins, form the basic structure of the ribosome, and play critical roles in protein synthesis. Compared to prokaryotic ribosomes, eukaryotic ribosomes contain elongated rRNAs with several expansion segments and larger numbers of ribosomal proteins. To investigate architectural evolution and functional capability of rRNAs, we employed a Tn5 transposon system to develop a systematic genetic insertion of an RNA segment 31 nt in length into Escherichia coli rRNAs. From the plasmid library harboring a single rRNA operon containing random insertions, we isolated surviving clones bearing rRNAs with functional insertions that enabled rescue of the E. coli strain (Δ7rrn) in which all chromosomal rRNA operons were depleted. We identified 51 sites with functional insertions, 16 sites in 16S rRNA and 35 sites in 23S rRNA, revealing the architecture of E. coli rRNAs to be substantially flexible. Most of the insertion sites show clear tendency to coincide with the regions of the expansion segments found in eukaryotic rRNAs, implying that eukaryotic rRNAs evolved from prokaryotic rRNAs suffering genetic insertions and selections.  相似文献   

10.
Circular Vicia faba (broad bean) chloroplast DNA was hybridized to the restriction fragment BamHI B from the DNA of the transducing phage lambda rifd18, which carries the Escherichia coli ribosomal RNA operon rrnB. Cytochrome spreadings of the heteroduplexes show homologies in the 16 S and 23 S rRNA regions, but none in the spacer. The same lambda rifd18 fragment was hybridized to the Vicia cpDNA 2SalI fragment 3, which contains the Vicia rBNA operon, resulting in an analogous heteroduplex configuration. Cytochrome spreadings of this heteroduplex in increasing concentrations of formamide reveal regions of incomplete homologies. Heteroduplexes between the E. coli rrnD operon, obtained from the recombinant plasmid pBK8, and circular Vicia cpDNA revealed homologies in the spacer region as well as in the 16 S and 23 S rRNA region. Hybrids between all three types of rDNA and their homologous rRNAs were prepared using the mica adsorption technique. They show that the 23 S, 16 S, and 5 S rRNAs are transcribed from the same strand of Vicia cpDNA. The positions of the rRNAs were measured and compared to the heteroduplex structure. It was observed that the E. coli rrnD operon in the plasmid pBK8 contains two 5 S rRNA sequences near the distal end.  相似文献   

11.
12.
In bacterial 16S rRNAs, methylated nucleosides are clustered within the decoding center, and these nucleoside modifications are thought to modulate translational fidelity. The N4, 2′-O-dimethylcytidine (m4Cm) at position 1402 of the Escherichia coli 16S rRNA directly interacts with the P-site codon of the mRNA. The biogenesis and function of this modification remain unclear. We have identified two previously uncharacterized genes in E. coli that are required for m4Cm formation. mraW (renamed rsmH) and yraL (renamed rsmI) encode methyltransferases responsible for the N4 and 2′-O-methylations of C1402, respectively. Recombinant RsmH and RsmI proteins employed the 30S subunit (not the 16S rRNA) as a substrate to reconstitute m4Cm1402 in the presence of S-adenosylmethionine (Ado-Met) as the methyl donor, suggesting that m4Cm1402 is formed at a late step during 30S assembly in the cell. A luciferase reporter assay indicated that the lack of N4 methylation of C1402 increased the efficiency of non-AUG initiation and decreased the rate of UGA read-through. These results suggest that m4Cm1402 plays a role in fine-tuning the shape and function of the P-site, thus increasing decoding fidelity.  相似文献   

13.
An aerobic, Gram-staining negative, non-motile, and rod-shaped bacterial strain, SS011A0-7#2-2T, was isolated from the sediment of South China Sea with the depth of 1,500 m. Optimum growth occurred at pH 8.0, 30 °C, and 6 % (w/v) NaCl. Strain SS011A0-7#2-2T did not synthesize bacteriochlorophyll a or carotenoid, neither possess photosynthesis genes. Its genome DNA G+C content was 67.9 mol%. It contained Q-10 as the predominant ubiquinone and C18:1 ω7c (52.3 %) as the major fatty acid. The major polar lipids were phosphatidylcholine, phosphatidylglycerol, diphosphatidylglycerol, phosphatidylethanolamine, unidentified phospholipid, and unidentified aminolipid. The 16S rRNA gene sequence analysis revealed that it was closely related to Seohaeicola saemankumensis SD-15T, Phaeobacter gallaeciensis BS 107T and Roseovarius pacificus 81-2T in Rhodobacteraceae, with the 16S rRNA gene sequence similarities being 96.5, 95.7, and 95.6 %, respectively. However, the phylogeny of the 16S rRNA gene sequences revealed that strain SS011A0-7#2-2T was a member of the genus Seohaeicola. Strain SS011A0-7#2-2T was moderately halophilic which was different from Seohaeicola saemankumensis SD-15T, and it showed the enzyme activities and carbon source spectrum significantly different from Seohaeicola saemankumensis SD-15T. As its physiological and chemotaxinomic properties were different from those of Seohaeicola saemankumensis SD-15T, strain SS011A0-7#2-2T represents a novel species of the genus Seohaecola. The name Seohaeicola nanhaiensis sp. nov. is proposed, with strain SS011A0-7#2-2T (=LMG 27733T = CGMCC 1.12759T) as the type strain.  相似文献   

14.
Eight strains of Saccharomycopsis species were examined for their partial base sequences of 18S and 26S rRNAs. The base sequences on the fingerprint segment showed three kinds of ACUAU, UCUAU, and AUAU. In the partial base sequences (positions 1451–1618, 168 bases) of 18S rRNA, all the strains of Saccharomycopsis species examined had 1–0 base difference except for the strains of Saccharomycopsis fibuligera (base differences, 5–4). In the partial base sequences on the positions 1611–1835 region (225 bases) of 26S rRNA, however, S’copsis vini had 15–11 base differences. The other species had 7–2 base differences. In the partial base sequences on the positions 493–622 region (130 bases) of 26S rRNA, the calculated percent similarities were 62–77. Discussion was made phylogenetically and taxonomically, especially on the separation of S’copsis fibuligera (≡ Endomyces fibuliger) from the genus Saccharomycopsis.  相似文献   

15.
The potential of a solution-based hybridization assay using peptide nucleic acid (PNA) molecular beacon (MB) probes to quantify 16S rRNA of specific populations in RNA extracts of environmental samples was evaluated by designing PNA MB probes for the genera Dechloromonas and Dechlorosoma. In a kinetic study with 16S rRNA from pure cultures, the hybridization of PNA MB to target 16S rRNA exhibited a higher final hybridization signal and a lower apparent rate constant than the hybridizations to nontarget 16S rRNAs. A concentration of 10 mM NaCl in the hybridization buffer was found to be optimal for maximizing the difference between final hybridization signals from target and nontarget 16S rRNAs. Hybridization temperatures and formamide concentrations in hybridization buffers were optimized to minimize signals from hybridizations of PNA MB to nontarget 16S rRNAs. The detection limit of the PNA MB hybridization assay was determined to be 1.6 nM of 16S rRNA. To establish proof for the application of PNA MB hybridization assays in complex systems, target 16S rRNA from Dechlorosoma suillum was spiked at different levels to RNA isolated from an environmental (bioreactor) sample, and the PNA MB assay enabled effective quantification of the D. suillum RNA in this complex mixture. For another environmental sample, the quantitative results from the PNA MB hybridization assay were compared with those from clone libraries.  相似文献   

16.
Methylation of the bacterial small ribosomal subunit (16S) rRNA on the N1 position of A1408 confers exceptionally high-level resistance to a broad spectrum of aminoglycoside antibiotics. Here, we present a detailed structural and functional analysis of the Catenulisporales acidiphilia 16S rRNA (m1A1408) methyltransferase (‘CacKam’). The apo CacKam structure closely resembles other m1A1408 methyltransferases within its conserved SAM-binding fold but the region linking core β strands 6 and 7 (the ‘β6/7 linker’) has a unique, extended structure that partially occludes the putative 16S rRNA binding surface, and sequesters the conserved and functionally critical W203 outside of the CacKam active site. Substitution of conserved residues in the SAM binding pocket reveals a functional dichotomy in the 16S rRNA (m1A1408) methyltransferase family, with two apparently distinct molecular mechanisms coupling cosubstrate/ substrate binding to catalytic activity. Our results additionally suggest that CacKam exploits the W203-mediated remodeling of the β6/7 linker as a novel mechanism to control 30S substrate recognition and enzymatic turnover.  相似文献   

17.
Ribosomal RNAs (rRNAs) have long been known to carry chemical modifications, including 2′O-methylation, pseudouridylation, N6-methyladenosine (m6A), and N6,6-dimethyladenosine. While the functions of many of these modifications are unclear, some are highly conserved and occur in regions of the ribosome critical for mRNA decoding. Both 28S rRNA and 18S rRNA carry single m6A sites, and while the methyltransferase ZCCHC4 has been identified as the enzyme responsible for the 28S rRNA m6A modification, the methyltransferase responsible for the 18S rRNA m6A modification has remained unclear. Here, we show that the METTL5-TRMT112 methyltransferase complex installs the m6A modification at position 1832 of human 18S rRNA. Our work supports findings that TRMT112 is required for METTL5 stability and reveals that human METTL5 mutations associated with microcephaly and intellectual disability disrupt this interaction. We show that loss of METTL5 in human cancer cell lines and in mice regulates gene expression at the translational level; additionally, Mettl5 knockout mice display reduced body size and evidence of metabolic defects. While recent work has focused heavily on m6A modifications in mRNA and their roles in mRNA processing and translation, we demonstrate here that deorphanizing putative methyltransferase enzymes can reveal previously unappreciated regulatory roles for m6A in noncoding RNAs.  相似文献   

18.
Discovery and characterization of functional RNA structures remains challenging due to deficiencies in de novo secondary structure modeling. Here we describe a dynamic programming approach for model-free sequence comparison that incorporates high-throughput chemical probing data. Based on SHAPE probing data alone, ribosomal RNAs (rRNAs) from three diverse organisms – the eubacteria E. coli and C. difficile and the archeon H. volcanii – could be aligned with accuracies comparable to alignments based on actual sequence identity. When both base sequence identity and chemical probing reactivities were considered together, accuracies improved further. Derived sequence alignments and chemical probing data from protein-free RNAs were then used as pseudo-free energy constraints to model consensus secondary structures for the 16S and 23S rRNAs. There are critical differences between these experimentally-informed models and currently accepted models, including in the functionally important neck and decoding regions of the 16S rRNA. We infer that the 16S rRNA has evolved to undergo large-scale changes in base pairing as part of ribosome function. As high-quality RNA probing data become widely available, structurally-informed sequence alignment will become broadly useful for de novo motif and function discovery.  相似文献   

19.
A Gram-stain negative, short rod-shaped aerobic bacterium with flagella, designated strain Y32T, was isolated from coastal seawater in Xiamen, Fujian Province of China. 16S rRNA gene sequence comparisons showed that strain Y32T is a member of the family Oceanospirillaceae, forming a distinct lineage with species of the genus Litoribacillus. The 16S rRNA gene sequence similarities between strain Y32T and other strains were all less than 94.0 %. Strain Y32T was found to grow optimally at 28 °C, at pH 7.0–8.0 and in the presence of 4–5 % (w/v) NaCl. The major fatty acids were identified as Summed Feature 3 (comprising C16:1 ω7c and/or C16:1 ω6c, 49.4 %), C16:0 (17.7 %), C14:0 (6.9 %) and C18:1 ω9c (5.4 %). The major respiratory quinone was identified as ubiquinone-8 (Q-8). The major polar lipids were identified as diphosphatidylglycerol, phosphatidylethanolamine and phosphatidylglycerol. The DNA G+C content of strain Y32T was determined to be 55.6 mol%. According to its morphology, physiology, fatty acid composition, polar lipids composition and 16S rRNA gene sequence data, strain Y32T represents a novel species of a new genus in the family Oceanospirillaceae, for which the name Litoribrevibacter albus gen. nov. sp. nov. is proposed. The type strain of Litoribrevibacter albus is Y32T (=MCCC 1F01211T=NBRC 110071T).  相似文献   

20.
THE 16S and 23S rRNAs of E. coli are known to be synthesized through intermediates of marginally. higher S values (about 17S and 24S). Although various authors1–3 have reported the presence of these precursors, no sequence studies have been presented. Nevertheless, there is strong but indirect evidence3 that the precursor 16S rRNA (p16) is about 200 residues longer than 16S rRNA (ml6). In contrast, the evidence suggests that the precursor 23S rRNA has few or no further residues.  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