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.
RNA modification plays an important role in modulating host-pathogen interaction. Flavivirus NS5 protein encodes N-7 and 2′-O methyltransferase activities that are required for the formation of 5′ type I cap (m7GpppAm) of viral RNA genome. Here we reported, for the first time, that flavivirus NS5 has a novel internal RNA methylation activity. Recombinant NS5 proteins of West Nile virus and Dengue virus (serotype 4; DENV-4) specifically methylates polyA, but not polyG, polyC, or polyU, indicating that the methylation occurs at adenosine residue. RNAs with internal adenosines substituted with 2′-O-methyladenosines are not active substrates for internal methylation, whereas RNAs with adenosines substituted with N6-methyladenosines can be efficiently methylated, suggesting that the internal methylation occurs at the 2′-OH position of adenosine. Mass spectroscopic analysis further demonstrated that the internal methylation product is 2′-O-methyladenosine. Importantly, genomic RNA purified from DENV virion contains 2′-O-methyladenosine. The 2′-O methylation of internal adenosine does not require specific RNA sequence since recombinant methyltransferase of DENV-4 can efficiently methylate RNAs spanning different regions of viral genome, host ribosomal RNAs, and polyA. Structure-based mutagenesis results indicate that K61-D146-K181-E217 tetrad of DENV-4 methyltransferase forms the active site of internal methylation activity; in addition, distinct residues within the methyl donor (S-adenosyl-L-methionine) pocket, GTP pocket, and RNA-binding site are critical for the internal methylation activity. Functional analysis using flavivirus replicon and genome-length RNAs showed that internal methylation attenuated viral RNA translation and replication. Polymerase assay revealed that internal 2′-O-methyladenosine reduces the efficiency of RNA elongation. Collectively, our results demonstrate that flavivirus NS5 performs 2′-O methylation of internal adenosine of viral RNA in vivo and host ribosomal RNAs in vitro.  相似文献   

2.
RNA molecules that bind tightly and specifically to a Rex fusion protein have been isolated from a conformationally constrained pool of random sequence RNAs. The anti-Rex aptamers effectively mimic several features of the wild-type Rex-binding element (XBE). The highest-affinity aptamers effectively compete with the wild-type XBE for binding to the RNA-binding domain of Rex, an arginine-rich motif (ARM), but do not bind to the functionally analogous Rev protein or its ARM. However, characteristic sequence and structural motifs found in some of the anti-Rex aptamers may provide insights into how the Rex protein can interact with other viral RNAs, such as the Rev-responsive element. The anti-Rex aptamers can functionally substitute for the XBE in vivo, a result which supports a previously proposed model for mRNA transport in which the viral genome serves as a platform for assembling a nucleoprotein complex that can co-opt the cellular transport apparatus. Overall, these studies suggest that anti-Rex aptamers may serve as RNA decoys of the Rex protein.  相似文献   

3.
4.
Hepatitis C virus (HCV)-encoded nonstructural protein 3 (NS3) possesses protease, NTPase, and helicase activities, which are considered essential for viral proliferation. Thus, HCV NS3 is a good putative therapeutic target protein for the development of anti-HCV agents. In this study, we isolated specific RNA aptamers to the helicase domain of HCV NS3 from a combinatorial RNA library with 40-nucleotide random sequences using in vitro selection techniques. The isolated RNAs were observed to very avidly bind the HCV helicase with an apparent Kd of 990 pM in contrast to original pool RNAs with a Kd of >1 microM. These RNA ligands appear to impede binding of substrate RNA to the HCV helicase and can act as potent decoys to competitively inhibit helicase activity with high efficiency compared with poly(U) or tRNA. The minimal binding domain of the ligands was determined to evaluate the structural features of the isolated RNA molecules. Interestingly, part of binding motif of the RNA aptamers consists of similar secondary structure to the 3'-end of HCV negative-strand RNA. Moreover, intracellular NS3 protein can be specifically detected in situ with the RNA aptamers, indicating that the selected RNAs are very specific to the HCV NS3 helicase. Furthermore, the RNA aptamers partially inhibited RNA synthesis of HCV subgenomic replicon in Huh-7 hepatoma cell lines. These results suggest that the RNA aptamers selected in vitro could be useful not only as therapeutic and diagnostic agents of HCV infection but also as a powerful tool for the study of HCV helicase mechanism.  相似文献   

5.
TRBP is a human cellular protein that binds the human immunodeficiency virus type 1 TAR RNA. Here, we show that the intact presence of amino acids 247 to 267 in TRBP correlates with its ability to bind RNA. This region contains a lysine- and arginine-rich motif, KKLAKRNAAAKMLLRVHTVPLDAR. A 24-amino-acid synthetic peptide (TR1) of this sequence bound TAR RNA with affinities similar to that of the entire TRBP, thus suggesting that this short motif contains a sufficient RNA-binding activity. Using RNA probe-shift analysis, we determined that TR1 does not bind all double-stranded RNAs but prefers TAR and other double-stranded RNAs with G+C-rich characteristics. Immunoprecipitation of TRBP from human immunodeficiency virus type 1-infected T lymphocytes recovered TAR RNA. This is consistent with a TRBP-TAR ribonucleoprotein during viral infection. Computer alignment revealed that TR1 is highly homologous to the RNA-binding domain of human P1/dsI protein kinase and two regions within Drosophila Staufen. We suggest that these proteins are related by virtue of sharing a common RNA-binding moiety.  相似文献   

6.
Genomes of RNA viruses encounter a continual threat from host cellular ribonucleases. Therefore, viruses have evolved mechanisms to protect the integrity of their genomes. To study the mechanism of 3′-end repair in dengue virus-2 in mammalian cells, a series of 3′-end deletions in the genome were evaluated for virus replication by detection of viral antigen NS1 and by sequence analysis. Limited deletions did not cause any delay in the detection of NS1 within 5 d. However, deletions of 7–10 nucleotides caused a delay of 9 d in the detection of NS1. Sequence analysis of RNAs from recovered viruses showed that at early times, virus progenies evolved through RNA molecules of heterogeneous lengths and nucleotide sequences at the 3′ end, suggesting a possible role for terminal nucleotidyl transferase activity of the viral polymerase (NS5). However, this diversity gradually diminished and consensus sequences emerged. Template activities of 3′-end mutants in the synthesis of negative-strand RNA in vitro by purified NS5 correlate well with the abilities of mutant RNAs to repair and produce virus progenies. Using the Mfold program for RNA structure prediction, we show that if the 3′ stem–loop (3′ SL) structure was abrogated by mutations, viruses eventually restored the 3′ SL structure. Taken together, these results favor a two-step repair process: non-template-based nucleotide addition followed by evolutionary selection of 3′-end sequences based on the best-fit RNA structure that can support viral replication.  相似文献   

7.
Jeong YY  Kim SH  Jang SI  You JC 《BMB reports》2008,41(7):511-515
The nucleocapsid (NC) protein of the Human Immunodeficiency Virus-1 plays a key role in viral genomic packaging by specifically recognizing the Psi(Psi) RNA sequence within the HIV-1 genome RNA. Recently, a novel cell-based assay was developed to probe the specific interactions in vivo between the NC and Psi-RNA using E.coli cells (J. Virol. 81: 6151-55, 2007). In order to examine the extendibility of this cell-based assay to RNAs other than Psi-RNA, this study tested the RNA aptamers isolated in vitro using the SELEX method, but whose specific binding ability to NC in a living cellular environment has not been established. The results demonstrate for the first time that each of those aptamer RNAs can bind specifically to NC in a NC zinc finger motif dependent manner within the cell. This confirms that the cell-based assay developed for NC-Psi interaction can be further extended and applied to NC-binding RNAs other than Psi-RNA.  相似文献   

8.
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   

9.
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   

10.
In order to find small RNA molecules that are specific and high-affinity ligands of nonstructural 5B (NS5B) polymerase, we screened by SELEX (systematic evolution of ligands by exponential amplification) a structurally constrained RNA library with an NS5BDeltaC55 enzyme carrying a C-terminal biotinylation sequence. Among the selected clones, two aptamers appeared to be high-affinity ligands of NS5B, with apparent dissociation constants in the low nanomolar range. They share a sequence that can assume a stem-loop structure. By mutation analysis, this structure has been shown to correspond to the RNA motif responsible for the tight interaction with NS5B. The aptamers appeared to be highly specific for the hepatitis C virus (HCV) polymerase since interaction with the GB virus B (GBV-B) NS5B protein cannot be observed. This is consistent with the observation that the activity of the HCV NS5B polymerase is efficiently inhibited by the selected aptamers, while neither GBV-B nor poliovirus 3D polymerases are affected. The mechanism of inhibition of the NS5B activity turned out to be noncompetitive with respect to template RNA, suggesting that aptamers and template RNA do not bind to the same site. As a matter of fact, mutations introduced in a basic exposed surface of the thumb domain severely impaired both the binding of and activity inhibition by the RNA aptamers.  相似文献   

11.
The RIG-I like receptor (RLR) comprises three homologues: RIG-I (retinoic acid-inducible gene I), MDA5 (melanoma differentiation-associated gene 5), and LGP2 (laboratory of genetics and physiology 2). Each RLR senses different viral infections by recognizing replicating viral RNA in the cytoplasm. The RLR contains a conserved C-terminal domain (CTD), which is responsible for the binding specificity to the viral RNAs, including double-stranded RNA (dsRNA) and 5′-triphosphated single-stranded RNA (5′ppp-ssRNA). Here, the solution structures of the MDA5 and LGP2 CTD domains were solved by NMR and compared with those of RIG-I CTD. The CTD domains each have a similar fold and a similar basic surface but there is the distinct structural feature of a RNA binding loop; The LGP2 and RIG-I CTD domains have a large basic surface, one bank of which is formed by the RNA binding loop. MDA5 also has a large basic surface that is extensively flat due to open conformation of the RNA binding loop. The NMR chemical shift perturbation study showed that dsRNA and 5′ppp-ssRNA are bound to the basic surface of LGP2 CTD, whereas dsRNA is bound to the basic surface of MDA5 CTD but much more weakly, indicating that the conformation of the RNA binding loop is responsible for the sensitivity to dsRNA and 5′ppp-ssRNA. Mutation study of the basic surface and the RNA binding loop supports the conclusion from the structure studies. Thus, the CTD is responsible for the binding affinity to the viral RNAs.  相似文献   

12.
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   

13.
The antiviral and antitumor functions of RNase L are enabled by binding to the allosteric effectors 5′-phosphorylated, 2′,5′-linked oligoadenylates (2-5A). 2-5A is produced by interferon-inducible 2′,5′-oligoadenylate synthetases (OAS) upon activation by viral double-stranded RNA (dsRNA). Because mutations in RNase L have been implicated as risk factors for prostate cancer, we sought to determine if OAS activators are present in prostate cancer cells. We show that prostate cancer cell lines (PC3, LNCaP and DU145), but not normal prostate epithelial cells (PrEC), contain RNA fractions capable of binding to and activating OAS. To identify the RNA activators, we developed a cDNA cloning strategy based on stringent affinity of RNAs for OAS. We thus identified mRNAs for Raf kinase inhibitor protein (RKIP) and poly(rC)-binding protein 2 (PCBP2) that bind and potently activate OAS. In addition, human endogenous retrovirus (hERV) envelope RNAs were present in PC3 cells that bind and activate OAS. Analysis of several gene expression profiling studies indicated that PCBP2 RNA was consistently elevated in metastatic prostate cancer. Results suggest that OAS activation may occur in prostate cancer cells in vivo stimulated by cellular mRNAs for RKIP and PCBP2.  相似文献   

14.
Flavivirus NS3 and NS5 are required in viral replication and 5′-capping. NS3 has NS2B-dependent protease, RNA helicase, and 5′-RNA triphosphatase activities. NS5 has 5′-RNA methyltransferase (MT)/guanylyltransferase (GT) activities within the N-terminal 270 amino acids and the RNA-dependent RNA polymerase (POL) activity within amino acids 271–900. A chimeric NS5 containing the D4MT/D4GT and the D2POL domains in the context of wild-type (WT) D2 RNA was constructed. RNAs synthesized in vitro were transfected into baby hamster kidney cells. The viral replication was analyzed by an indirect immunofluorescence assay to monitor NS1 expression and by quantitative real-time PCR. WT D2 RNA-transfected cells were NS1- positive by day 5, whereas the chimeric RNA-transfected cells became NS1-positive ∼30 days post-transfection in three independent experiments. Sequence analysis covering the entire genome revealed the appearance of a single K74I mutation within the D4MT domain ∼16 days post-transfection in two experiments. In the third, D290N mutation in the conserved NS3 Walker B motif appeared ≥16 days post-transfection. A time course study of serial passages revealed that the 30-day supernatant had gradually evolved to gain replication fitness. Trans-complementation by co-expression of WT D2 NS5 accelerated viral replication of chimeric RNA without changing the K74I mutation. However, the MT and POL activities of NS5 WT D2 and the chimeric NS5 proteins with or without the K74I mutation are similar. Taken together, our results suggest that evolution of the functional interactions involving the chimeric NS5 protein encoded by the viral genome species is essential for gain of viral replication fitness.  相似文献   

15.
16.
17.
Telomeres protect chromosome ends from being recognized as double-stranded breaks. Telomeric function is ensured by the shelterin complex in which TRF2 protein is an essential player. The G-rich strand of telomere DNA can fold into G-quadruplex (G4) structure. Small molecules stabilizing G4 structures, named G4 ligands, have been shown to alter telomeric functions in human cells. In this study, we show that a guanine-rich RNA sequence located in the 5′-UTR region of the TRF2 mRNA (hereafter 91TRF2G) is capable of forming a stable quadruplex that causes a 2.8-fold decrease in the translation of a reporter gene in human cells, as compared to a mutant 5′-UTR unable to fold into G4. We also demonstrate that several highly selective G4 ligands, the pyridine dicarboxamide derivative 360A and bisquinolinium compounds Phen-DC(3) and Phen-DC(6), are able to bind the 91TRF2G:RNA sequence and to modulate TRF2 protein translation in vitro. Since the naturally occurring 5′-UTR TRF2:RNA G4 element was used here, which is conserved in several vertebrate orthologs, the present data substantiate a potential translational mechanism mediated by a G4 RNA motif for the downregulation of TRF2 expression.  相似文献   

18.
19.
The envelope glycoprotein of human immunodeficiency virus (HIV) consists of an exterior glycoprotein (gp120) and a trans-membrane domain (gp41) and has an important role in viral entry into cells. HIV-1 entry has been validated as a clinically relevant anti-viral strategy for drug discovery. In the present work, several 2′-F substituted RNA aptamers that bind to the HIV-1BaL gp120 protein with nanomole affinity were isolated from a RNA library by the SELEX (Systematic Evolution of Ligands by EXponential enrichment) procedure. From two of these aptamers we created a series of new dual inhibitory function anti-gp120 aptamer–siRNA chimeras. The aptamers and aptamer–siRNA chimeras specifically bind to and are internalized into cells expressing HIV gp160. The Dicer-substrate siRNA delivered by the aptamers is functionally processed by Dicer, resulting in specific inhibition of HIV-1 replication and infectivity in cultured CEM T-cells and primary blood mononuclear cells (PBMCs). Moreover, we have introduced a ‘sticky’ sequence onto a chemically synthesized aptamer which facilitates attachment of the Dicer substrate siRNAs for potential multiplexing. Our results provide a set of novel inhibitory agents for blocking HIV replication and further validate the use of aptamers for delivery of Dicer substrate siRNAs.  相似文献   

20.
Hu antigen C (HuC) has three RNA-binding domains (RBDs). The N-terminal two, RBD1 and RBD2, are linked in tandem and bind to the AU-rich elements (AREs) in the 3′-untranslated region of particular mRNAs. The solution structures of HuC RBD1 and RBD2 were determined by NMR methods. The HuC RBD1 and RBD2 structures are quite similar to those of Sxl RBD1 and RBD2, respectively. The individual RBDs of HuC, RBD1 and RBD2 in isolation can interact rather weakly with the minimal ARE motif, AUUUA, while the didomain fragment, RBD1–RBD2, of HuC binds more tightly to a longer ARE RNA, UAUUUAUUUU. Chemical shift perturbations by the longer RNA on HuC RBD1–RBD2 were mapped on and around the two β-sheets and on the C-terminal region of RBD1. The HuC RBD1–RBD2 residues that exhibited significant chemical shift perturbations coincide with those conserved in Sxl RBD1–RBD2. These data indicate that the RNA-binding characteristics of the HuC and Sxl didomain fragments are similar, even though the target RNAs and the biological functions of the proteins are different.  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