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 234 毫秒
1.
Evidence is accumulating that small, noncoding RNAs are important regulatory molecules. Computational and experimental searches have led to the identification of ~60 small RNA genes in Escherichia coli. However, most of these studies focused on the intergenic regions and assumed that small RNAs were >50 nt. Thus, the previous screens missed small RNAs encoded on the antisense strand of protein-coding genes and small RNAs of <50 nt. To identify additional small RNAs, we carried out a cloning-based screen focused on RNAs of 30–65 nt. In this screen, we identified RNA species corresponding to fragments of rRNAs, tRNAs and known small RNAs. Several of the small RNAs also corresponded to 5′- and 3′-untranslated regions (UTRs) and internal fragments of mRNAs. Four of the 3′-UTR-derived RNAs were highly abundant and two showed expression patterns that differed from the corresponding mRNAs, suggesting independent functions for the 3′-UTR-derived small RNAs. We also detected three previously unidentified RNAs encoded in intergenic regions and RNAs from the long direct repeat and hok/sok elements. In addition, we identified a few small RNAs that are expressed opposite protein-coding genes and could base pair with 5′ or 3′ ends of the mRNAs with perfect complementarity.  相似文献   

2.
3.
Uniformly 32P-labeled, double-stranded genome RNA isolated from purified reovirus contains two types of 5′-terminal sequences. One strand contains a phosphatase-resistant 5′-terminal structure, XpppG*pCpU, which is also present in the viral mRNA. The 5′ blocking group, X, is removed by β-elimination indicating that it is a nucleoside containing free 2′,3′-hydroxyls. G*pC is an alkaline-resistant, 2′-O-methylated sequence. The other strand contains a phosphatase-sensitive 5′ sequence, ppGpPupPyp. The results are discussed in relation to blocked 5′-terminal structures in other viral and cellular RNAs.  相似文献   

4.
5.
Song BH  Yun SI  Choi YJ  Kim JM  Lee CH  Lee YM 《RNA (New York, N.Y.)》2008,14(9):1791-1813
Tertiary or higher-order RNA motifs that regulate replication of positive-strand RNA viruses are as yet poorly understood. Using Japanese encephalitis virus (JEV), we now show that a key element in JEV RNA replication is a complex RNA motif that includes a string of three discontinuous complementary sequences (TDCS). The TDCS consists of three 5-nt-long strands, the left (L) strand upstream of the translation initiator AUG adjacent to the 5′-end of the genome, and the middle (M) and right (R) strands corresponding to the base of the Flavivirus-conserved 3′ stem–loop structure near the 3′-end of the RNA. The three strands are arranged in an antiparallel configuration, with two sets of base-pairing interactions creating L-M and M-R duplexes. Disrupting either or both of these duplex regions of TDCS completely abolished RNA replication, whereas reconstructing both duplex regions, albeit with mutated sequences, fully restored RNA replication. Modeling of replication-competent genomes recovered from a large pool of pseudorevertants originating from six replication-incompetent TDCS mutants suggests that both duplex base-pairing potentials of TDCS are required for RNA replication. In all cases, acquisition of novel sequences within the 3′M-R duplex facilitated a long-range RNA–RNA interaction of its 3′M strand with either the authentic 5′L strand or its alternative (invariably located upstream of the 5′ initiator), thereby restoring replicability. We also found that a TDCS homolog is conserved in other flaviviruses. These data suggest that two duplex base-pairings defined by the TDCS play an essential regulatory role in a key step(s) of Flavivirus RNA replication.  相似文献   

6.

Background

SELEX is a well established in vitro selection tool to analyze the structure of ligand-binding nucleic acid sequences called aptamers. Genomic SELEX transforms SELEX into a tool to discover novel, genomically encoded RNA or DNA sequences binding a ligand of interest, called genomic aptamers. Concerns have been raised regarding requirements imposed on RNA sequences undergoing SELEX selection.

Methodology/Principal Findings

To evaluate SELEX and assess the extent of these effects, we designed and performed a Neutral SELEX experiment omitting the selection step, such that the sequences are under the sole selective pressure of SELEX''s amplification steps. Using high-throughput sequencing, we obtained thousands of full-length sequences from the initial genomic library and the pools after each of the 10 rounds of Neutral SELEX. We compared these to sequences obtained from a Genomic SELEX experiment deriving from the same initial library, but screening for RNAs binding with high affinity to the E. coli regulator protein Hfq. With each round of Neutral SELEX, sequences became less stable and changed in nucleotide content, but no sequences were enriched. In contrast, we detected substantial enrichment in the Hfq-selected set with enriched sequences having structural stability similar to the neutral sequences but with significantly different nucleotide selection.

Conclusions/Significance

Our data indicate that positive selection in SELEX acts independently of the neutral selective requirements imposed on the sequences. We conclude that Genomic SELEX, when combined with high-throughput sequencing of positively and neutrally selected pools, as well as the gnomic library, is a powerful method to identify genomic aptamers.  相似文献   

7.
Small noncoding RNAs (sRNAs) regulate the response of bacteria to environmental stress in conjunction with the Sm-like RNA binding protein Hfq. DsrA sRNA stimulates translation of the RpoS stress response factor in Escherichia coli by base-pairing with the 5′ leader of the rpoS mRNA and opening a stem–loop that represses translation initiation. We report that rpoS leader sequences upstream of this stem–loop greatly increase the sensitivity of rpoS mRNA to Hfq and DsrA. Native gel mobility shift assays show that Hfq increases the rate of DsrA binding to the full 576 nt rpoS leader as much as 50-fold. By contrast, base-pairing with a 138-nt RNA containing just the repressor stem–loop is accelerated only twofold. Deletion and mutagenesis experiments showed that sensitivity to Hfq requires an upstream AAYAA sequence. Leaders long enough to contain this sequence bind Hfq tightly and form stable ternary complexes with Hfq and DsrA. A model is proposed in which Hfq recruits DsrA to the rpoS mRNA by binding both RNAs, releasing the self-repressing structure in the mRNA. Once base-pairing between DsrA and rpoS mRNA is established, interactions between Hfq and the mRNA may stabilize the RNA complex by removing Hfq from the sRNA.  相似文献   

8.
9.
10.
Small interfering RNAs regulate gene expression in diverse biological processes, including heterochromatin formation and DNA elimination, developmental regulation, and cell differentiation. In the single-celled eukaryote Entamoeba histolytica, we have identified a population of small RNAs of 27 nt size that (i) have 5′-polyphosphate termini, (ii) map antisense to genes, and (iii) associate with an E. histolytica Piwi-related protein. Whole genome microarray expression analysis revealed that essentially all genes to which antisense small RNAs map were not expressed under trophozoite conditions, the parasite stage from which the small RNAs were cloned. However, a number of these genes were expressed in other E. histolytica strains with an inverse correlation between small RNA and gene expression level, suggesting that these small RNAs mediate silencing of the cognate gene. Overall, our results demonstrate that E. histolytica has an abundant 27 nt small RNA population, with features similar to secondary siRNAs from C. elegans, and which appear to regulate gene expression. These data indicate that a silencing pathway mediated by 5′-polyphosphate siRNAs extends to single-celled eukaryotic organisms.  相似文献   

11.
The slime mold Physarum polycephalum is a morphologically simple organism with a large and complex genome. The exon–intron organization of its genes exhibits features typical for protists and fungi as well as those characteristic for the evolutionarily more advanced species. This indicates that both the taxonomic position as well as the size of the genome shape the exon–intron organization of an organism. The average gene has 3.7 introns which are on average 138 bp, with a rather narrow size distribution. Introns are enriched in AT base pairs by 13% relative to exons. The consensus sequences at exon–intron boundaries resemble those found for other species, with minor differences between short and long introns. A unique feature of P.polycephalum introns is the strong preference for pyrimidines in the coding strand throughout their length, without a particular enrichment at the 3′-ends.  相似文献   

12.
13.
Previous work has demonstrated that iron-dependent variations in the steady-state concentration and translatability of sodB mRNA are modulated by the small regulatory RNA RyhB, the RNA chaperone Hfq and RNase E. In agreement with the proposed role of RNase E, we found that the decay of sodB mRNA is retarded upon inactivation of RNase E in vivo, and that the enzyme cleaves within the sodB 5′-untranslated region (5′-UTR) in vitro, thereby removing the 5′ stem–loop structure that facilitates Hfq and ribosome binding. Moreover, RNase E cleavage can also occur at a cryptic site that becomes available upon sodB 5′-UTR/RyhB base pairing. We show that while playing an important role in facilitating the interaction of RyhB with sodB mRNA, Hfq is not tightly retained by the RyhB–sodB mRNA complex and can be released from it through interaction with other RNAs added in trans. Unlike turnover of sodB mRNA, RyhB decay in vivo is mainly dependent on RNase III, and its cleavage by RNase III in vitro is facilitated upon base pairing with the sodB 5′-UTR. These data are discussed in terms of a model, which accounts for the observed roles of RNase E and RNase III in sodB mRNA turnover.  相似文献   

14.
The bacterial protein Hfq participates in the regulation of translation by small noncoding RNAs (sRNAs). Several mechanisms have been proposed to explain the role of Hfq in the regulation by sRNAs binding to the 5′-untranslated mRNA regions. However, it remains unknown how Hfq affects those sRNAs that target the coding sequence. Here, the contribution of Hfq to the annealing of three sRNAs, RybB, SdsR, and MicC, to the coding sequence of Salmonella ompD mRNA was investigated. Hfq bound to ompD mRNA with tight, subnanomolar affinity. Moreover, Hfq strongly accelerated the rates of annealing of RybB and MicC sRNAs to this mRNA, and it also had a small effect on the annealing of SdsR. The experiments using truncated RNAs revealed that the contributions of Hfq to the annealing of each sRNA were individually adjusted depending on the structures of interacting RNAs. In agreement with that, the mRNA structure probing revealed different structural contexts of each sRNA binding site. Additionally, the annealing of RybB and MicC sRNAs induced specific conformational changes in ompD mRNA consistent with local unfolding of mRNA secondary structure. Finally, the mutation analysis showed that the long AU-rich sequence in the 5′-untranslated mRNA region served as an Hfq binding site essential for the annealing of sRNAs to the coding sequence. Overall, the data showed that the functional specificity of Hfq in the annealing of each sRNA to the ompD mRNA coding sequence was determined by the sequence and structure of the interacting RNAs.  相似文献   

15.
Success with genome editing by the RNA-programmed nuclease Cas9 has been limited by the inability to predict effective guide RNAs and DNA target sites. Not all guide RNAs have been successful, and even those that were, varied widely in their efficacy. Here we describe and validate a strategy for Caenorhabditis elegans that reliably achieved a high frequency of genome editing for all targets tested in vivo. The key innovation was to design guide RNAs with a GG motif at the 3′ end of their target-specific sequences. All guides designed using this simple principle induced a high frequency of targeted mutagenesis via nonhomologous end joining (NHEJ) and a high frequency of precise DNA integration from exogenous DNA templates via homology-directed repair (HDR). Related guide RNAs having the GG motif shifted by only three nucleotides showed severely reduced or no genome editing. We also combined the 3′ GG guide improvement with a co-CRISPR/co-conversion approach. For this co-conversion scheme, animals were only screened for genome editing at designated targets if they exhibited a dominant phenotype caused by Cas9-dependent editing of an unrelated target. Combining the two strategies further enhanced the ease of mutant recovery, thereby providing a powerful means to obtain desired genetic changes in an otherwise unaltered genome.  相似文献   

16.
The Hfq protein, which shares sequence and structural homology with the Sm and Lsm proteins, binds to various RNAs, primarily recognizing AU-rich single-stranded regions. In this paper, we study the ability of the Escherichia coli Hfq protein to bind to a polyadenylated fragment of rpsO mRNA. Hfq exhibits a high specificity for a 100-nucleotide RNA harboring 18 3′-terminal A-residues. Structural analysis of the adenylated RNA–Hfq complex and gel shift assays revealed the presence of two Hfq binding sites. Hfq binds primarily to the poly(A) tail, and to a lesser extent a U-rich sequence in a single-stranded region located between two hairpin structures. The oligo(A) tail and the interhelical region are sensitive to 3′–5′ exoribonucleases and RNase E hydrolysis, respectively, in vivo. In vitro assays demonstrate that Hfq protects poly(A) tails from exonucleolytic degradation by both PNPase and RNase II. In addition, RNase E processing, which occurred close to the U-rich sequence, is impaired by the presence of Hfq. These data suggest that Hfq modulates the sensitivity of RNA to ribonucleases in the cell.  相似文献   

17.
18.
19.
Tomato Genomic Resources Database (TGRD) allows interactive browsing of tomato genes, micro RNAs, simple sequence repeats (SSRs), important quantitative trait loci and Tomato-EXPEN 2000 genetic map altogether or separately along twelve chromosomes of tomato in a single window. The database is created using sequence of the cultivar Heinz 1706. High quality single nucleotide polymorphic (SNP) sites between the genes of Heinz 1706 and the wild tomato S. pimpinellifolium LA1589 are also included. Genes are classified into different families. 5′-upstream sequences (5′-US) of all the genes and their tissue-specific expression profiles are provided. Sequences of the microRNA loci and their putative target genes are catalogued. Genes and 5′-US show presence of SSRs and SNPs. SSRs located in the genomic, genic and 5′-US can be analysed separately for the presence of any particular motif. Primer sequences for all the SSRs and flanking sequences for all the genic SNPs have been provided. TGRD is a user-friendly web-accessible relational database and uses CMAP viewer for graphical scanning of all the features. Integration and graphical presentation of important genomic information will facilitate better and easier use of tomato genome. TGRD can be accessed as an open source repository at http://59.163.192.91/tomato2/.  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