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 953 毫秒
1.
Retinoic acid-inducible gene I (RIG-I) recognizes specific molecular patterns of viral RNAs for inducing type I interferon. The C-terminal domain (CTD) of RIG-I binds to double-stranded RNA (dsRNA) with the 5′-triphosphate (5′-PPP), which induces a conformational change in RIG-I to an active form. It has been suggested that RIG-I detects infection of influenza A virus by recognizing the 5′-triphosphorylated panhandle structure of the viral RNA genome. Influenza panhandle RNA has a unique structure with a sharp helical bending. In spite of extensive studies of how viral RNAs activate RIG-I, whether the structural elements of the influenza panhandle RNA confer the ability to activate RIG-I signaling has been poorly explored. Here, we investigated the dynamics of the influenza panhandle RNA in complex with RIG-I CTD using NMR spectroscopy and showed that the bending structure of the panhandle RNA negates the requirement of a 5′-PPP moiety for RIG-I activation.  相似文献   

2.
The 5′ untranslated region (5′UTR) of the dengue virus (DENV) genome contains two defined elements essential for viral replication. At the 5′ end, a large stem-loop (SLA) structure functions as the promoter for viral polymerase activity. Next to the SLA, there is a short stem-loop that contains a cyclization sequence known as the 5′ upstream AUG region (5′UAR). Here, we analyzed the secondary structure of the SLA in solution and the structural requirements of this element for viral replication. Using infectious DENV clones, viral replicons, and in vitro polymerase assays, we defined two helical regions, a side stem-loop, a top loop, and a U bulge within SLA as crucial elements for viral replication. The determinants for SLA-polymerase recognition were found to be common in different DENV serotypes. In addition, structural elements within the SLA required for DENV RNA replication were also conserved among different mosquito- and tick-borne flavivirus genomes, suggesting possible common strategies for polymerase-promoter recognition in flaviviruses. Furthermore, a conserved oligo(U) track present downstream of the SLA was found to modulate RNA synthesis in transfected cells. In vitro polymerase assays indicated that a sequence of at least 10 residues following the SLA, upstream of the 5′UAR, was necessary for efficient RNA synthesis using the viral 3′UTR as template.  相似文献   

3.
Sequences and structures within the terminal genomic regions of plus-strand RNA viruses are targets for the binding of host proteins that modulate functions such as translation, RNA replication, and encapsidation. Using murine norovirus 1 (MNV-1), we describe the presence of long-range RNA-RNA interactions that were stabilized by cellular proteins. The proteins potentially responsible for the stabilization were selected based on their ability to bind the MNV-1 genome and/or having been reported to be involved in the stabilization of RNA-RNA interactions. Cell extracts were preincubated with antibodies against the selected proteins and used for coprecipitation reactions. Extracts treated with antibodies to poly(C) binding protein 2 (PCBP2) and heterogeneous nuclear ribonucleoprotein (hnRNP) A1 significantly reduced the 5′-3′ interaction. Both PCBP2 and hnRNP A1 recombinant proteins stabilized the 5′-3′ interactions and formed ribonucleoprotein complexes with the 5′ and 3′ ends of the MNV-1 genomic RNA. Mutations within the 3′ complementary sequences (CS) that disrupt the 5′-3′-end interactions resulted in a significant reduction of the viral titer, suggesting that the integrity of the 3′-end sequence and/or the lack of complementarity with the 5′ end is important for efficient virus replication. Small interfering RNA-mediated knockdown of PCBP2 or hnRNP A1 resulted in a reduction in virus yield, confirming a role for the observed interactions in efficient viral replication. PCBP2 and hnRNP A1 induced the circularization of MNV-1 RNA, as revealed by electron microscopy. This study provides evidence that PCBP2 and hnRNP A1 bind to the 5′ and 3′ ends of the MNV-1 viral RNA and contribute to RNA circularization, playing a role in the virus life cycle.  相似文献   

4.
5.
The subgenomic RNA 2 of tobacco necrosis virus A (TNV sgRNA2) encodes the viral coat protein, is unpolyadenylated and presumably uncapped. Here, we show that TNV sgRNA2 is translated cap independently. This cap-independent translation requires the leader and a 140 nt element of the trailer both in wheat germ extract and in tobacco protoplasts. Similar to barley yellow dwarf virus (BYDV), the TNV 5′ and 3′ elements stimulate translation synergistically. Computer-aided phylogenetic analysis of the secondary structure of the TNV trailer revealed that the 3′ translation element is part of a major conserved stem–loop that contains similarities to structures in the BYDV 3′ translation element. These data suggest that the translation mechanisms of TNV sgRNA2 and BYDV RNA are related. To further characterize this relationship, we tested whether cooperativity exists between TNV sgRNA2 and BYDV 5′ and 3′ elements. We found that the TNV sgRNA2 5′ element stimulates translation synergistically with the BYDV 3′ element in vitro. This finding is the first evidence for conservation of structures that enable a 5′–3′ interaction stimulating cap-independent translation.  相似文献   

6.
RNA viruses recruit the host translational machinery by different mechanisms that depend partly on the structure of their genomes. In this regard, the plus-strand RNA genomes of several different pathogenic plant viruses do not contain traditional translation-stimulating elements, i.e., a 5′-cap structure and a 3′-poly(A) tail, and instead rely on a 3′-cap-independent translational enhancer (3′CITE) located in their 3′ untranslated regions (UTRs) for efficient synthesis of viral proteins. We investigated the structure and function of the I-shaped class of 3′CITE in tombusviruses—also present in aureusviruses and carmoviruses—using biochemical and molecular approaches and we determined that it adopts a complex higher-order RNA structure that facilitates translation by binding simultaneously to both eukaryotic initiation factor (eIF) 4F and the 5′ UTR of the viral genome. The specificity of 3′CITE binding to eIF4F is mediated, at least in part, through a direct interaction with its eIF4E subunit, whereas its association with the viral 5′ UTR relies on complementary RNA–RNA base-pairing. We show for the first time that this tripartite 5′ UTR/3′CITE/eIF4F complex forms in vitro in a translationally relevant environment and is required for recruitment of ribosomes to the 5′ end of the viral RNA genome by a mechanism that shares some fundamental features with cap-dependent translation. Notably, our results demonstrate that the 3′CITE facilitates the initiation step of translation and validate a molecular model that has been proposed to explain how several different classes of 3′CITE function. Moreover, the virus–host interplay defined in this study provides insights into natural host resistance mechanisms that have been linked to 3′CITE activity.  相似文献   

7.
Hepatitis C is a major public health concern, with an estimated 170 million people infected worldwide and an urgent need for new drug development. An attractive therapeutic approach is to prevent the ‘cap-independent’ translation initiation of the viral proteins by interfering with both the structure and function of the hepatitis C viral internal ribosomal entry site (HCV IRES). Towards this goal, we report the design, synthesis and purification of novel bi-functional molecules containing DNA or RNA antisenses attached to functional groups performing RNA hydrolysis. These 5′ or 3′-coupled conjugates bind the HCV IRES with affinity and specificity and elicit targeted hydrolysis of the viral genomic RNA after short (1 h) incubation at low (500 nM) concentration at 37°C in vitro. Additional secondary cleavage sites are induced and their mapping within the RNA structure indicates that functional domains IIIb-e are excised from the IRES that, based on cryo-EM studies, becomes incapable of binding the small ribosomal subunit and initiation factor 3 (eIF3). All these molecules inhibit, in a dose-dependent manner, the ‘IRES-dependent’ translation in vitro. The 5′-coupled imidazole conjugate reduces viral protein synthesis by half at a 300 nM concentration (IC50), corresponding to a 4-fold increase of activity when compared to the naked oligonucleotide. These new conjugates are now being tested for activity on infected hepatic cell lines.  相似文献   

8.
SARS-CoV-2 is a positive-sense RNA virus responsible for the Coronavirus Disease 2019 (COVID-19) pandemic, which continues to cause significant morbidity, mortality and economic strain. SARS-CoV-2 can cause severe respiratory disease and death in humans, highlighting the need for effective antiviral therapies. The RNA synthesis machinery of SARS-CoV-2 is an ideal drug target and consists of non-structural protein 12 (nsp12), which is directly responsible for RNA synthesis, and numerous co-factors involved in RNA proofreading and 5′ capping of viral RNAs. The formation of the 5′ 7-methylguanosine (m7G) cap structure is known to require a guanylyltransferase (GTase) as well as a 5′ triphosphatase and methyltransferases; however, the mechanism of SARS-CoV-2 RNA capping remains poorly understood. Here we find that SARS-CoV-2 nsp12 is involved in viral RNA capping as a GTase, carrying out the addition of a GTP nucleotide to the 5′ end of viral RNA via a 5′ to 5′ triphosphate linkage. We further show that the nsp12 NiRAN (nidovirus RdRp-associated nucleotidyltransferase) domain performs this reaction, and can be inhibited by remdesivir triphosphate, the active form of the antiviral drug remdesivir. These findings improve understanding of coronavirus RNA synthesis and highlight a new target for novel or repurposed antiviral drugs against SARS-CoV-2.  相似文献   

9.
Subdomain 5BSL3.2 of hepatitis C virus RNA lies at the core of a network of distal RNA–RNA contacts that connect the 5′ and 3′ regions of the viral genome and regulate the translation and replication stages of the viral cycle. Using small-angle X-ray scattering and NMR spectroscopy experiments, we have determined at low resolution the structural models of this subdomain and its distal complex with domain 3′X, located at the 3′-terminus of the viral RNA chain. 5BSL3.2 adopts a characteristic ‘L’ shape in solution, whereas the 5BSL3.2–3′X distal complex forms a highly unusual ‘Y’-shaped kissing junction that blocks the dimer linkage sequence of domain 3′X and promotes translation. The structure of this complex may impede an effective association of the viral polymerase with 5BSL3.2 and 3′X to start negative-strand RNA synthesis, contributing to explain the likely mechanism used by these sequences to regulate viral replication and translation. In addition, sequence and shape features of 5BSL3.2 are present in functional RNA motifs of flaviviruses, suggesting conserved regulatory processes within the Flaviviridae family.  相似文献   

10.
11.
The rapid spread of COVID-19 is motivating development of antivirals targeting conserved SARS-CoV-2 molecular machinery. The SARS-CoV-2 genome includes conserved RNA elements that offer potential small-molecule drug targets, but most of their 3D structures have not been experimentally characterized. Here, we provide a compilation of chemical mapping data from our and other labs, secondary structure models, and 3D model ensembles based on Rosetta''s FARFAR2 algorithm for SARS-CoV-2 RNA regions including the individual stems SL1-8 in the extended 5′ UTR; the reverse complement of the 5′ UTR SL1-4; the frameshift stimulating element (FSE); and the extended pseudoknot, hypervariable region, and s2m of the 3′ UTR. For eleven of these elements (the stems in SL1–8, reverse complement of SL1–4, FSE, s2m and 3′ UTR pseudoknot), modeling convergence supports the accuracy of predicted low energy states; subsequent cryo-EM characterization of the FSE confirms modeling accuracy. To aid efforts to discover small molecule RNA binders guided by computational models, we provide a second set of similarly prepared models for RNA riboswitches that bind small molecules. Both datasets (‘FARFAR2-SARS-CoV-2’, https://github.com/DasLab/FARFAR2-SARS-CoV-2; and ‘FARFAR2-Apo-Riboswitch’, at https://github.com/DasLab/FARFAR2-Apo-Riboswitch’) include up to 400 models for each RNA element, which may facilitate drug discovery approaches targeting dynamic ensembles of RNA molecules.  相似文献   

12.
Positive-strand RNA viruses use diverse mechanisms to regulate viral and host gene expression for ensuring their efficient proliferation or persistence in the host. We found that a small viral noncoding RNA (0.4 kb), named SR1f, accumulated in Red clover necrotic mosaic virus (RCNMV)-infected plants and protoplasts and was packaged into virions. The genome of RCNMV consists of two positive-strand RNAs, RNA1 and RNA2. SR1f was generated from the 3′ untranslated region (UTR) of RNA1, which contains RNA elements essential for both cap-independent translation and negative-strand RNA synthesis. A 58-nucleotide sequence in the 3′ UTR of RNA1 (Seq1f58) was necessary and sufficient for the generation of SR1f. SR1f was neither a subgenomic RNA nor a defective RNA replicon but a stable degradation product generated by Seq1f58-mediated protection against 5′→3′ decay. SR1f efficiently suppressed both cap-independent and cap-dependent translation both in vitro and in vivo. SR1f trans inhibited negative-strand RNA synthesis of RCNMV genomic RNAs via repression of replicase protein production but not via competition of replicase proteins in vitro. RCNMV seems to use cellular enzymes to generate SR1f that might play a regulatory role in RCNMV infection. Our results also suggest that Seq1f58 is an RNA element that protects the 3′-side RNA sequences against 5′→3′ decay in plant cells as reported for the poly(G) tract and stable stem-loop structure in Saccharomyces cerevisiae.  相似文献   

13.
The segmented double-stranded (ds) RNA genome of the rotaviruses is replicated asymmetrically, with viral mRNA serving as the template for the synthesis of minus-strand RNA. Previous studies with cell-free replication systems have shown that the highly conserved termini of rotavirus gene 8 and 9 mRNAs contain cis-acting signals that promote the synthesis of dsRNA. Based on the location of the cis-acting signals and computer modeling of their secondary structure, the ends of the gene 8 or 9 mRNAs are proposed to interact in cis to form a modified panhandle structure that promotes the synthesis of dsRNA. In this structure, the last 11 to 12 nucleotides of the RNA, including the cis-acting signal that is essential for RNA replication, extend as a single-stranded tail from the panhandled region, and the 5′ untranslated region folds to form a stem-loop motif. To understand the importance of the predicted secondary structure in minus-strand synthesis, mutations were introduced into viral RNAs which affected the 3′ tail and the 5′ stem-loop. Analysis of the RNAs with a cell-free replication system showed that, in contrast to mutations which altered the structure of the 5′ stem-loop, mutations which caused complete or near-complete complementarity between the 5′ end and the 3′ tail significantly inhibited (≥10-fold) minus-strand synthesis. Likewise, incubation of wild-type RNAs with oligonucleotides which were complementary to the 3′ tail inhibited replication. Despite their replication-defective phenotype, mutant RNAs with complementary 5′ and 3′ termini were shown to competitively interfere with the replication of wild-type mRNA and to bind the viral RNA polymerase VP1 as efficiently as wild-type RNA. These results indicate that the single-strand nature of the 3′ end of rotavirus mRNA is essential for efficient dsRNA synthesis and that the specific binding of the RNA polymerase to the mRNA template is required but not sufficient for the synthesis of minus-strand RNA.  相似文献   

14.
To identify cis-acting elements in the foamy virus (FV) RNA pregenome, we developed a transient-vector-production system based on cotransfection of indicator gene-bearing vector and gag-pol and env expression plasmids. Two elements which were critical for vector transfer were found and mapped approximately. The first element was located in the RU5 leader and the 5′ gag region (approximately up to position 650 of the viral RNA). The second element was located in an approximately 2-kb sequence in the 3′ pol region. Although small 5′ and 3′ deletions, as well as internal deletions of the latter element, were tolerated, both elements were found to be absolutely required for vector transfer. The functional characterization of the pol region-located cis-acting element revealed that it is essential for efficient incorporation or the stability of particle-associated virion RNA. Furthermore, virions derived from a vector lacking this sequence were found to be deficient in the cleavage of the Gag protein by the Pol precursor protease. Our results suggest that during the formation of infectious virions, complex interactions between FV Gag and Pol and the viral RNA take place.  相似文献   

15.
An unexpectedly high number of regulatory RNAs have been recently discovered that fine-tune the function of genes at all levels of expression. We employed Genomic SELEX, a method to identify protein-binding RNAs encoded in the genome, to search for further regulatory RNAs in Escherichia coli. We used the global regulator protein Hfq as bait, because it can interact with a large number of RNAs, promoting their interaction. The enriched SELEX pool was subjected to deep sequencing, and 8865 sequences were mapped to the E. coli genome. These short sequences represent genomic Hfq-aptamers and are part of potential regulatory elements within RNA molecules. The motif 5′-AAYAAYAA-3′ was enriched in the selected RNAs and confers low-nanomolar affinity to Hfq. The motif was confirmed to bind Hfq by DMS footprinting. The Hfq aptamers are 4-fold more frequent on the antisense strand of protein coding genes than on the sense strand. They were enriched opposite to translation start sites or opposite to intervening sequences between ORFs in operons. These results expand the repertoire of Hfq targets and also suggest that Hfq might regulate the expression of a large number of genes via interaction with cis-antisense RNAs.  相似文献   

16.
The 3′ noncoding region (NCR) of the negative-strand RNA [3′(−)NCR RNA] of the arterivirus simian hemorrhagic fever virus (SHFV) is 209 nucleotides (nt) in length. Since this 3′ region, designated 3′(−)209, is the site of initiation of full-length positive-strand RNA and is the template for the synthesis of the 5′ leader sequence, which is found on both full-length and subgenomic mRNAs, it is likely to contain cis-acting signals for RNA synthesis and to interact with cellular and viral proteins to form replication complexes. Gel mobility shift assays showed that cellular proteins in MA104 S100 cytoplasmic extracts formed two complexes with the SHFV 3′(−)209 RNA, and results from competition gel mobility shift assays demonstrated that these interactions were specific. Four proteins with molecular masses of 103, 86, 55, and 36 kDa were detected in UV-induced cross-linking assays, and three of these proteins (103, 55, and 36 kDa) were also detected by Northwestern blotting assays. Identical gel mobility shift and UV-induced cross-linking patterns were obtained with uninfected and SHFV-infected extracts, indicating that the four proteins detected are cellular, not viral, proteins. The binding sites for the four cellular proteins were mapped to the region between nt 117 and 184 (68-nt sequence) from the 3′ end of the SHFV negative-strand RNA. This 68-nt sequence was predicted to form two stem-loops, SL4 and SL5. The 3′(−)NCR RNA of another arterivirus, lactate dehydrogenase-elevating virus C (LDV-C), competed with the SHFV 3′(−)209 RNA in competition gel mobility shift assays. UV-induced cross-linking assays showed that four MA104 cellular proteins with the same molecular masses as those that bind to the SHFV 3′(−)209 RNA also bind to the LDV-C 3′(−)NCR RNA and equine arteritis virus 3′(−)NCR RNA. However, each of these viral RNAs also bound to an additional MA104 protein. The binding sites for the MA104 cellular proteins were shown to be located in similar positions in the LDV-C 3′(−)NCR and SHFV 3′(−)209 RNAs. These data suggest that the binding sites for a set of the cellular proteins are conserved in all arterivirus RNAs and that these cell proteins may be utilized as components of viral replication complexes.  相似文献   

17.
RNA structures present throughout RNA virus genomes serve as scaffolds to organize multiple factors involved in the initiation of RNA synthesis. Several of these RNA elements play multiple roles in the RNA replication pathway. An RNA structure formed around the 5′- end of the poliovirus genomic RNA has been implicated in the initiation of both negative- and positive-strand RNA synthesis. Dissecting the roles of these multifunctional elements is usually hindered by the interdependent nature of the viral replication processes and often pleiotropic effects of mutations. Here, we describe a novel approach to examine RNA elements with multiple roles. Our approach relies on the duplication of the RNA structure so that one copy is dedicated to the initiation of negative-strand RNA synthesis, while the other mediates positive-strand synthesis. This allows us to study the function of the element in promoting positive-strand RNA synthesis, independently of its function in negative-strand initiation. Using this approach, we demonstrate that the entire 5′-end RNA structure that forms on the positive-strand is required for initiation of new positive-strand RNAs. Also required to initiate positive-strand RNA synthesis are the binding sites for the viral polymerase precursor, 3CD, and the host factor, PCBP. Furthermore, we identify specific nucleotide sequences within “stem a” that are essential for the initiation of positive-strand RNA synthesis. These findings provide direct evidence for a trans-initiation model, in which binding of proteins to internal sequences of a pre-existing positive-strand RNA affects the synthesis of subsequent copies of that RNA, most likely by organizing replication factors around the initiation site.  相似文献   

18.
DExD/H proteins catalyze structural rearrangements in RNA by coupling ATP hydrolysis to the destabilization of RNA helices or RNP complexes. The Escherichia coli DExD/H protein DbpA specifically recognizes a region within the catalytic core of 23S rRNA. To better characterize the interaction of DbpA with this region and to identify changes in the complex between different nucleotide-bound states of the enzyme, RNase T1, RNase T2, kethoxal and DMS footprinting of DbpA on a 172 nt fragment of 23S rRNA were performed. A number of protections identified in helices 90 and 92 were consistent with biochemical experiments measuring the RNA binding and ATPase activity of DbpA with truncated RNAs. When DbpA was bound with AMPPNP, but not ADP, several additional footprints were detected in helix 93 and the single-stranded region 5′ of helix 90, suggesting binding of the helicase domains of DbpA at these sites. These results propose that DbpA can act at multiple sites and hint at the targets of its biological activity on rRNA.  相似文献   

19.
RNA function is determined by its structural organization. The RNA structure consists of the combination of distinct secondary structure motifs connected by junctions that play an essential role in RNA folding. Selective 2′-hydroxyl acylation analyzed by primer extension (SHAPE) probing is an established methodology to analyze the secondary structure of long RNA molecules in solution, which provides accurate data about unpaired nucleotides. However, the residues located at the junctions of RNA structures usually remain undetected. Here we report an RNA probing method based on the use of a novel open-paddlewheel diruthenium (OPW-Ru) compound [Ru2Cl2(µ-DPhF)3(DMSO)] (DPhF = N,N′-diphenylformamidinate). This compound has four potential coordination sites in a singular disposition to establish covalent bonds with substrates. As a proof of concept, we have analyzed the reactivity of OPW-Ru toward RNA using two viral internal ribosome entry site (IRES) elements whose function depends on the structural organization of the molecule. Our study suggests that the compound OPW-Ru preferentially attacks at positions located one or two nucleotides away from junctions or bulges of the RNA structure. The OPW-Ru fingerprinting data differ from that obtained by other chemical reagents and provides new information about RNA structure features.  相似文献   

20.
We have isolated 2′-Fluoro-substituted RNA aptamers that bind to streptavidin (SA) with an affinity around 7 ± 1.8 nM, comparable with that of recently described peptide aptamers. Binding to SA was not prevented by prior saturation with biotin, enabling nucleic acid aptamers to form useful ternary complexes. Mutagenesis, secondary structure analysis, ribonuclease footprinting and deletion analysis provided evidence for the essential structural features of SA-binding aptamers. In order to provide a general method for the exploitation of these aptamers, we produced derivatives in which they were fused to the naturally structured RNA elements, CopT or CopA. In parallel, we produced derivatives of CD4-binding aptamers fused to the complementary CopA or CopT elements. When mixed, these two chimeric aptamers rapidly hybridized, by virtue of CopA–CopT complementarity, to form stable, bi-functional aptamers that we called ‘adaptamers’. We show that a CD4–SA-binding adaptamer can be used to capture CD4 onto a SA-derivatized surface, illustrating their general utility as indirect affinity ligands.  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