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.
The RegA proteins from the bacteriophage T4 and RB69 are translational repressors that control the expression of multiple phage mRNAs. RegA proteins from the two phages share 78% sequence identity; however, in vivo expression studies have suggested that the RB69 RegA protein binds target RNAs with a higher affinity than T4 RegA protein. To study the RNA binding properties of T4 and RB69 RegA proteins more directly, the binding sites of RB69 RegA protein on synthetic RNAs corresponding to the translation initiation region of two RB69 target genes were mapped by RNase protection assays. These assays revealed that RB69 RegA protein protects nucleotides –9 to –3 (relative to the start codon) on RB69 gene 44, which contains the sequence GAAAAUU. On RB69 gene 45, the protected site (nucleotides –8 to –3) contains a similar purine-rich sequence: GAAAUA. Interestingly, T4 RegA protein protected the same nucleotides on these RNAs. To examine the specificity of RNA binding, quantitative RNA gel shift assays were performed with synthetic RNAs corresponding to recognition elements (REs) in three T4 and three RB69 mRNAs. Comparative gel shift assays demonstrated that RB69 RegA protein has an ~7-fold higher affinity for T4 gene 44 RE RNA than T4 RegA protein. RB69 RegA protein also binds RB69 gene 44 RE RNA with a 4-fold higher affinity than T4 RegA protein. On the other hand, T4 RegA exhibited a higher affinity than RB69 RegA protein for RB69 gene 45 RE RNA. With respect to their affinities for cognate RNAs, both RegA proteins exhibited the following hierarchy of affinities: gene 44 > gene 45 > regA. Interestingly, T4 RegA exhibited the highest affinity towards RB69 gene 45 RE RNA, whereas RB69 RegA protein had the highest affinity for T4 gene 44 RE RNA. The helix–loop groove RNA binding motif of T4 RegA protein is fully conserved in RB69 RegA protein. However, homology modeling of the structure of RB69 RegA protein reveals that the divergent residues are clustered in two areas of the surface, and that there are two large areas of high conservation near the helix–loop groove, which may also play a role in RNA binding.  相似文献   

2.
Hypoxanthine (H), the deamination product of adenine, has been implicated in the high frequency of A to G transitions observed in retroviral and other RNA genomes. Although H·C base pairs are thermodynamically more stable than other H·N pairs, polymerase selection may be determined in part by kinetic factors. Therefore, the hypoxanthine induced substitution pattern resulting from replication by viral polymerases may be more complex than that predicted from thermodynamics. We have examined the steady-state kinetics of formation of base pairs opposite template H in RNA by HIV-RT, and for the incorporation of dITP during first- and second-strand synthesis. Hypoxanthine in an RNA template enhances the k2app for pairing with standard dNTPs by factors of 10–1000 relative to adenine at the same sequence position. The order of base pairing preferences for H in RNA was observed to be H·C >> H·T > H·A > H·G. Steady-state kinetics of insertion for all possible mispairs formed with dITP were examined on RNA and DNA templates of identical sequence. Insertion of dITP opposite all bases occurs 2–20 times more frequently on RNA templates. This bias for higher insertion frequencies on RNA relative to DNA templates is also observed for formation of mispairs at template A. This kinetic advantage afforded by RNA templates for mismatches and pairing involving H suggests a higher induction of mutations at adenines during first-strand synthesis by HIV-RT.  相似文献   

3.
Drosophila melanogaster is a valuable invertebrate model for viral infection and antiviral immunity, and is a focus for studies of insect-virus coevolution. Here we use a metagenomic approach to identify more than 20 previously undetected RNA viruses and a DNA virus associated with wild D. melanogaster. These viruses not only include distant relatives of known insect pathogens but also novel groups of insect-infecting viruses. By sequencing virus-derived small RNAs, we show that the viruses represent active infections of Drosophila. We find that the RNA viruses differ in the number and properties of their small RNAs, and we detect both siRNAs and a novel miRNA from the DNA virus. Analysis of small RNAs also allows us to identify putative viral sequences that lack detectable sequence similarity to known viruses. By surveying >2,000 individually collected wild adult Drosophila we show that more than 30% of D. melanogaster carry a detectable virus, and more than 6% carry multiple viruses. However, despite a high prevalence of the Wolbachia endosymbiont—which is known to be protective against virus infections in Drosophila—we were unable to detect any relationship between the presence of Wolbachia and the presence of any virus. Using publicly available RNA-seq datasets, we show that the community of viruses in Drosophila laboratories is very different from that seen in the wild, but that some of the newly discovered viruses are nevertheless widespread in laboratory lines and are ubiquitous in cell culture. By sequencing viruses from individual wild-collected flies we show that some viruses are shared between D. melanogaster and D. simulans. Our results provide an essential evolutionary and ecological context for host–virus interaction in Drosophila, and the newly reported viral sequences will help develop D. melanogaster further as a model for molecular and evolutionary virus research.  相似文献   

4.
Mutations of the pore-region residue T442 in Shaker channels result in large effects on channel kinetics. We studied mutations at this position in the backgrounds of NH2-terminal–truncated Shaker H4 and a Shaker -NGK2 chimeric channel having high conductance (Lopez, G.A., Y.N. Jan, and L.Y. Jan. 1994. Nature (Lond.). 367: 179–182). While mutations of T442 to C, D, H, V, or Y resulted in undetectable expression in Xenopus oocytes, S and G mutants yielded functional channels having deactivation time constants and channel open times two to three orders of magnitude longer than those of the parental channel. Activation time courses at depolarized potentials were unaffected by the mutations, as were first-latency distributions in the T442S chimeric channel. The mutant channels show two subconductance levels, 37 and 70% of full conductance. From single-channel analysis, we concluded that channels always pass through the larger subconductance state on the way to and from the open state. The smaller subconductance state is traversed in ∼40% of activation time courses. These states apparently represent kinetic intermediates in channel gating having voltage-dependent transitions with apparent charge movements of ∼1.6 e0. The fully open T442S chimeric channel has the conductance sequence Rb+ > NH4 + > K+. The opposite conductance sequence, K+ > NH4 + > Rb+, is observed in each of the subconductance states, with the smaller subconductance state discriminating most strongly against Rb+.  相似文献   

5.
Kim YG  Maas S  Rich A 《Nucleic acids research》2001,29(5):1125-1131
Human immunodeficiency virus type 1 (HIV-1) and human T cell leukemia virus type II (HTLV-2) use a similar mechanism for –1 translational frameshifting to overcome the termination codon in viral RNA at the end of the gag gene. Previous studies have identified two important RNA signals for frameshifting, the slippery sequence and a downstream stem–loop structure. However, there have been somewhat conflicting reports concerning the individual contributions of these sequences. In this study we have performed a comprehensive mutational analysis of the cis-acting RNA sequences involved in HIV-1 gagpol and HTLV-2 gagpro frameshifting. Using an in vitro translation system we determined frameshifting efficiencies for shuffled HIV-1/HTLV-2 RNA elements in a background of HIV-1 or HTLV-2 sequences. We show that the ability of the slippery sequence and stem–loop to promote ribosomal frameshifting is influenced by the flanking upstream sequence and the nucleotides in the spacer element. A wide range of frameshift efficiency rates was observed for both viruses when shuffling single sequence elements. The results for HIV-1/HTLV-2 chimeric constructs represent strong evidence supporting the notion that the viral wild-type sequences are not designed for maximal frameshifting activity but are optimized to a level suited to efficient viral replication.  相似文献   

6.
RNA molecules with internal 2′,5′-branches are intermediates in RNA splicing, and branched RNAs have recently been proposed as retrotransposition intermediates. A broadly applicable in vitro synthetic route to branched RNA that does not require self-splicing introns or spliceosomes would substantially improve our ability to study biochemical processes that involve branched RNA. We recently described 7S11, a deoxyribozyme that was identified by in vitro selection and has general RNA branch-forming ability. However, an important restriction for 7S11 is that the branch-site RNA nucleotide must be a purine (A or G), because a pyrimidine (U or C) is not tolerated. Here, we describe the compact 6CE8 deoxyribozyme (selected using a 20 nt random region) that synthesizes 2′,5′-branched RNA with any nucleotide at the branch site. The Mn2+-dependent branch-forming ligation reaction is between an internal branch-site 2′-hydroxyl nucleophile on one RNA substrate with a 5′-triphosphate on another RNA substrate. The preference for the branch-site nucleotide is U > C A > G, although all four nucleotides are tolerated with useful ligation rates. Nearly all other nucleotides elsewhere in both RNA substrates allow ligation activity, except that the sequence requirement for the RNA strand with the 5′-triphosphate is 5′-pppGA, with 5′-pppGAR (R = purine) preferred. These characteristics permit 6CE8 to prepare branched RNAs of immediate practical interest, such as the proposed branched intermediate of Ty1 retrotransposition. Because this branched RNA has two strands with identical sequence that emerge from the branch site, we developed strategies to control which of the two strands bind with the deoxyribozyme during the branch-forming reaction. The ability to synthesize the proposed branched RNA of Ty1 retrotransposition will allow us to explore this important biochemical pathway in greater detail.  相似文献   

7.
Genomes and antigenomes of many positive-strand RNA viruses contain 3′-poly(A) and 5′-poly(U) tracts, respectively, serving as mutual templates. Mechanism(s) controlling the length of these homopolymeric stretches are not well understood. Here, we show that in coxsackievirus B3 (CVB3) and three other enteroviruses the poly(A) tract is ~80–90 and the poly(U) tract is ~20 nt-long. Mutagenesis analysis indicate that the length of the CVB3 3′-poly(A) is determined by the oriR, a cis-element in the 3′-noncoding region of viral RNA. In contrast, while mutations of the oriR inhibit initiation of (−) RNA synthesis, they do not affect the 5′-poly(U) length. Poly(A)-lacking genomes are able to acquire genetically unstable AU-rich poly(A)-terminated 3′-tails, which may be generated by a mechanism distinct from the cognate viral RNA polyadenylation. The aberrant tails ensure only inefficient replication. The possibility of RNA replication independent of oriR and poly(A) demonstrate that highly debilitated viruses are able to survive by utilizing ‘emergence’, perhaps atavistic, mechanisms.  相似文献   

8.
Based on our recent studies of RNA cleavage by oligonucleotide–terpyridine·Cu(II) complex 5′- and/or 3′-conjugates, we designed 2′-O-methyloligonucleotides with two terpyridine-attached nucleosides at contiguous internal sites. To connect the 2′-terpyridine-modified uridine residue at the 5′-side to the 5′-O-terpyridyl nucleoside residue at the 3′-side, a dimethoxytrityl derivative of 5-hydroxypropyl-5′-O-terpyridyl-2′-deoxyuridine-3′-phosphoramidite was newly synthesized. Using this unit, we constructed two terpyridine conjugates, with either an unusual phophodiester bond or the bond extended by a propanediol(s)-containing linker. Cleavage reactions of the target RNA oligomer, under the conditions of conjugate excess in the presence of Cu(II), indicated that the conjugates precisely cleaved the RNA at the predetermined site and that one propanediol-containing linker was the most appropriate for inducing high cleavage activity. Furthermore, a comparison of the activity of the propanediol agent with those of the control conjugates with one complex confirmed that the two complexes are required for efficient RNA cleavage. The reaction of the novel cleaver revealed a bell-shaped pH–rate profile with a maximum at pH ~7.5, which is a result of the cooperative action of the complexes. In addition, we demonstrated that the agent catalytically cleaves an excess of the RNA, with the kinetic parameter kcat/Km = 0.118 nM–1 h–1.  相似文献   

9.
Binding of monovalent and divalent cations to two adenine–adenine platform structures from the Tetrahymena group I intron ribozyme has been studied using continuum solvent models based on the generalised Born and the finite-difference Poisson–Boltzmann approaches. The adenine–adenine platform RNA motif forms an experimentally characterised monovalent ion binding site important for ribozyme folding and function. Qualitative agreement between calculated and experimental ion placements and binding selectivity was obtained. The inclusion of solvation effects turned out to be important to obtain low energy structures and ion binding placements in agreement with the experiment. The calculations indicate that differences in solvation of the isolated ions contribute to the calculated ion binding preference. However, Coulomb attraction and van der Waals interactions due to ion size differences and RNA conformational adaptation also influence the calculated ion binding affinity. The calculated alkali ion binding selectivity for both platforms followed the order K+ > Na+ > Rb+ > Cs+ > Li+ (Eisenman series VI) in the case of allowing RNA conformational relaxation during docking. With rigid RNA an Eisenman series V was obtained (K+ > Rb+ > Na+ > Cs+ > Li+). Systematic energy minimisation docking simulations starting from several hundred initial placements of potassium ions on the surface of platform containing RNA fragments identified a coordination geometry in agreement with the experiment as the lowest energy binding site. The approach could be helpful to identify putative ion binding sites in nucleic acid structures determined at low resolution or with experimental methods that do not allow identification of ion binding sites.  相似文献   

10.
Viruses are replication competent genomes which are relatively gene-poor. Even the largest viruses (i.e. Herpesviruses) encode only slightly >200 open reading frames (ORFs). However, because viruses replicate obligatorily inside cells, and considering that evolution may be driven by a principle of economy of scale, it is reasonable to surmise that many viruses have evolved the ability to co-opt cell-encoded proteins to provide needed surrogate functions. An in silico survey of viral sequence databases reveals that most positive-strand and double-stranded RNA viruses have ORFs for RNA helicases. On the other hand, the genomes of retroviruses are devoid of virally-encoded helicase. Here, we review in brief the notion that the human immunodeficiency virus (HIV-1) has adopted the ability to use one or more cellular RNA helicases for its replicative life cycle.  相似文献   

11.
SARS-CoV-2 is a betacoronavirus with a linear single-stranded, positive-sense RNA genome, whose outbreak caused the ongoing COVID-19 pandemic. The ability of coronaviruses to rapidly evolve, adapt, and cross species barriers makes the development of effective and durable therapeutic strategies a challenging and urgent need. As for other RNA viruses, genomic RNA structures are expected to play crucial roles in several steps of the coronavirus replication cycle. Despite this, only a handful of functionally-conserved coronavirus structural RNA elements have been identified to date. Here, we performed RNA structure probing to obtain single-base resolution secondary structure maps of the full SARS-CoV-2 coronavirus genome both in vitro and in living infected cells. Probing data recapitulate the previously described coronavirus RNA elements (5′ UTR and s2m), and reveal new structures. Of these, ∼10.2% show significant covariation among SARS-CoV-2 and other coronaviruses, hinting at their functionally-conserved role. Secondary structure-restrained 3D modeling of these segments further allowed for the identification of putative druggable pockets. In addition, we identify a set of single-stranded segments in vivo, showing high sequence conservation, suitable for the development of antisense oligonucleotide therapeutics. Collectively, our work lays the foundation for the development of innovative RNA-targeted therapeutic strategies to fight SARS-related infections.  相似文献   

12.
The RNA-dependent RNA polymerase of the severe acute respiratory syndrome coronavirus 2 is an important target in current drug development efforts for the treatment of coronavirus disease 2019. Molnupiravir is a broad-spectrum antiviral that is an orally bioavailable prodrug of the nucleoside analogue β-D-N4-hydroxycytidine (NHC). Molnupiravir or NHC can increase G to A and C to U transition mutations in replicating coronaviruses. These increases in mutation frequencies can be linked to increases in antiviral effects; however, biochemical data of molnupiravir-induced mutagenesis have not been reported. Here we studied the effects of the active compound NHC 5’-triphosphate (NHC-TP) against the purified severe acute respiratory syndrome coronavirus 2 RNA-dependent RNA polymerase complex. The efficiency of incorporation of natural nucleotides over the efficiency of incorporation of NHC-TP into model RNA substrates followed the order GTP (12,841) > ATP (424) > UTP (171) > CTP (30), indicating that NHC-TP competes predominantly with CTP for incorporation. No significant inhibition of RNA synthesis was noted as a result of the incorporated monophosphate in the RNA primer strand. When embedded in the template strand, NHC-monophosphate supported the formation of both NHC:G and NHC:A base pairs with similar efficiencies. The extension of the NHC:G product was modestly inhibited, but higher nucleotide concentrations could overcome this blockage. In contrast, the NHC:A base pair led to the observed G to A (G:NHC:A) or C to U (C:G:NHC:A:U) mutations. Together, these biochemical data support a mechanism of action of molnupiravir that is primarily based on RNA mutagenesis mediated via the template strand.  相似文献   

13.
We have determined the crystal structure of the RNA octamer duplex r(guguuuac)/r(guaggcac) with a tandem wobble pair, G·G/U·U (motif III), to compare it with U·G/G·U (motif I) and G·U/U·G (motif II) and to better understand their relative stabilities. The crystal belongs to the rhombohedral space group R3. The hexagonal unit cell dimensions are a = b = 41.92 Å, c = 56.41 Å, and γ = 120°, with one duplex in the asymmetric unit. The structure was solved by the molecular replacement method at 1.9 Å resolution and refined to a final R factor of 19.9% and Rfree of 23.3% for 2862 reflections in the resolution range 10.0–1.9 Å with F ≥ 2σ(F). The final model contains 335 atoms for the RNA duplex and 30 water molecules. The A-RNA stacks in the familiar head-to-tail fashion forming a pseudo-continuous helix. The uridine bases of the tandem U·G pairs have slipped towards the minor groove relative to the guanine bases and the uridine O2 atoms form bifurcated hydrogen bonds with the N1 and N2 of guanines. The N2 of guanine and O2 of uridine do not bridge the ‘locked’ water molecule in the minor groove, as in motifs I and II, but are bridged by water molecules in the major groove. A comparison of base stacking stabilities of motif III with motifs I and II confirms the result of thermodynamic studies, motif I > motif III > motif II.  相似文献   

14.
The key step in bacterial translation is formation of the pre-initiation complex. This requires initial contacts between mRNA, fMet-tRNA and the 30S subunit of the ribosome, steps that limit the initiation of translation. Here we report a method for improving translational initiation, which allows expression of several previously non-expressible genes. This method has potential applications in heterologous protein synthesis and high-throughput expression systems. We introduced a synthetic RNA stem–loop (stem length, 7 bp; ΔG0 = –9.9 kcal/mol) in front of various gene sequences. In each case, the stem–loop was inserted 15 nt downstream from the start codon. Insertion of the stem–loop allowed in vitro expression of five previously non-expressible genes and enhanced the expression of all other genes investigated. Analysis of the RNA structure proved that the stem–loop was formed in vitro, and demonstrated that stabilization of the ribosome binding site is due to stem–loop introduction. By theoretical RNA structure analysis we showed that the inserted RNA stem–loop suppresses long-range interactions between the translation initiation domain and gene-specific mRNA sequences. Thus the inserted RNA stem–loop supports the formation of a separate translational initiation domain, which is more accessible to ribosome binding.  相似文献   

15.
To reveal the clonal architecture of melanoma and associated driver mutations, whole genome sequencing (WGS) and targeted extension sequencing were used to characterize 124 melanoma cases. Significantly mutated gene analysis using 13 WGS cases and 15 additional paired extension cases identified known melanoma genes such as BRAF, NRAS, and CDKN2A, as well as a novel gene EPHA3, previously implicated in other cancer types. Extension studies using tumors from another 96 patients discovered a large number of truncation mutations in tumor suppressors (TP53 and RB1), protein phosphatases (e.g., PTEN, PTPRB, PTPRD, and PTPRT), as well as chromatin remodeling genes (e.g., ASXL3, MLL2, and ARID2). Deep sequencing of mutations revealed subclones in the majority of metastatic tumors from 13 WGS cases. Validated mutations from 12 out of 13 WGS patients exhibited a predominant UV signature characterized by a high frequency of C->T transitions occurring at the 3′ base of dipyrimidine sequences while one patient (MEL9) with a hypermutator phenotype lacked this signature. Strikingly, a subclonal mutation signature analysis revealed that the founding clone in MEL9 exhibited UV signature but the secondary clone did not, suggesting different mutational mechanisms for two clonal populations from the same tumor. Further analysis of four metastases from different geographic locations in 2 melanoma cases revealed phylogenetic relationships and highlighted the genetic alterations responsible for differential drug resistance among metastatic tumors. Our study suggests that clonal evaluation is crucial for understanding tumor etiology and drug resistance in melanoma.  相似文献   

16.
17.
The active site of a protein folding reaction is in domain V of the 23S rRNA in the bacterial ribosome and its homologs in other organisms. This domain has long been known as the peptidyl transferase center. Domain V of Bacillus subtilis is split into two segments, the more conserved large peptidyl transferase loop (RNA1) and the rest (RNA2). These two segments together act as a protein folding modulator as well as the complete domain V RNA. A number of site-directed mutations were introduced in RNA1 and RNA2 of B.subtilis, taking clues from reports of these sites being involved in various steps of protein synthesis. For example, sites like G2505, U2506, U2584 and U2585 in Escherichia coli RNA1 region are protected by deacylated tRNA at high Mg2+ concentration and A2602 is protected by amino acyl tRNA when the P site remains occupied already. Mutations A2058G and A2059G in the RNA1 region render the ribosome Eryr in E.coli and Lncr in tobacco chloroplast. Sites in P loop G2252 and G2253 in E.coli are protected against modification by the CCA end of the P site bound tRNA. Mutations were introduced in corresponding nucleotides in B.subtilis RNA1 and RNA2 of domain V. The mutants were tested for refolding using unfolded protein binding assays with unfolded carbonic anhydrase. In the protein folding assay, the mutants showed partial to complete loss of this activity. In the filter binding assay for the RNA–refolding protein complex, the mutants showed an extent of protein binding that agreed well with their protein folding activity.  相似文献   

18.
A G Polson  B L Bass 《The EMBO journal》1994,13(23):5701-5711
Double-stranded RNA adenosine deaminase (dsRAD), previously called the double-stranded RNA (dsRNA) unwinding/modifying activity, modifies adenosines to inosines within dsRNA. We used ribonuclease U2 and a mutant of ribonuclease T1 to map the sites of modification in several RNA duplexes. We found that dsRAD had a 5' neighbor preference (A = U > C > G) but no apparent 3' neighbor preference. Further, the proximity of the strand termini affected whether an adenosine was modified. Most importantly, dsRAD exhibited selectivity, modifying a minimal number of adenosines in short dsRNAs. Our results suggest that the specific editing of glutamate receptor subunit B mRNA could be performed in vivo by dsRAD without the aid of specificity factors, and support the hypothesis that dsRAD is responsible for hypermutations in certain RNA viruses.  相似文献   

19.
Remdesivir (RDV), a broadly acting nucleoside analogue, is the only FDA approved small molecule antiviral for the treatment of COVID-19 patients. To date, there are no reports identifying SARS-CoV-2 RDV resistance in patients, animal models or in vitro. Here, we selected drug-resistant viral populations by serially passaging SARS-CoV-2 in vitro in the presence of RDV. Using high throughput sequencing, we identified a single mutation in RNA-dependent RNA polymerase (NSP12) at a residue conserved among all coronaviruses in two independently evolved populations displaying decreased RDV sensitivity. Introduction of the NSP12 E802D mutation into our SARS-CoV-2 reverse genetics backbone confirmed its role in decreasing RDV sensitivity in vitro. Substitution of E802 did not affect viral replication or activity of an alternate nucleoside analogue (EIDD2801) but did affect virus fitness in a competition assay. Analysis of the globally circulating SARS-CoV-2 variants (>800,000 sequences) showed no evidence of widespread transmission of RDV-resistant mutants. Surprisingly, we observed an excess of substitutions in spike at corresponding sites identified in the emerging SARS-CoV-2 variants of concern (i.e., H69, E484, N501, H655) indicating that they can arise in vitro in the absence of immune selection. The identification and characterisation of a drug resistant signature within the SARS-CoV-2 genome has implications for clinical management and virus surveillance.  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