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.
We report that photo-excitation of one-electron-oxidized adenine [A(-H)•] in dAdo and its 2′-deoxyribonucleotides leads to formation of deoxyribose sugar radicals in remarkably high yields. Illumination of A(-H)• in dAdo, 3′-dAMP and 5′-dAMP in aqueous glasses at 143 K leads to 80-100% conversion to sugar radicals at C5′ and C3′. The position of the phosphate in 5′- and 3′-dAMP is observed to deactivate radical formation at the site of substitution. In addition, the pH has a crucial influence on the site of sugar radical formation; e.g. at pH ~5, photo-excitation of A(-H)• in dAdo at 143 K produces mainly C5′• whereas only C3′• is observed at high pH ~12. 13C substitution at C5′ in dAdo yields 13C anisotropic couplings of (28, 28, 84) G whose isotropic component 46.7 G identifies formation of the near planar C5′•. A β-13C 16 G isotropic coupling from C3′• is also found. These results are found to be in accord with theoretically calculated 13C couplings at C5′ [DFT, B3LYP, 6-31(G) level] for C5′• and C3′•. Calculations using time-dependent density functional theory [TD-DFT B3LYP, 6-31G(d)] confirm that transitions in the near UV and visible induce hole transfer from the base radical to the sugar group leading to sugar radical formation.  相似文献   

2.
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   

3.
Dengue virus multifunctional proteins NS3 protease/helicase and NS5 methyltransferase/RNA-dependent RNA polymerase form part of the viral replication complex and are involved in viral RNA genome synthesis, methylation of the 5′-cap of viral genome, and polyprotein processing among other activities. Previous studies have shown that NS5 residue Lys-330 is required for interaction between NS3 and NS5. Here, we show by competitive NS3-NS5 interaction ELISA that the NS3 peptide spanning residues 566–585 disrupts NS3-NS5 interaction but not the null-peptide bearing the N570A mutation. Small angle x-ray scattering study on NS3(172–618) helicase and covalently linked NS3(172–618)-NS5(320–341) reveals a rigid and compact formation of the latter, indicating that peptide NS5(320–341) engages in specific and discrete interaction with NS3. Significantly, NS3:Asn-570 to alanine mutation introduced into an infectious DENV2 cDNA clone did not yield detectable virus by plaque assay even though intracellular double-stranded RNA was detected by immunofluorescence. Detection of increased negative-strand RNA synthesis by real time RT-PCR for the NS3:N570A mutant suggests that NS3-NS5 interaction plays an important role in the balanced synthesis of positive- and negative-strand RNA for robust viral replication. Dengue virus infection has become a global concern, and the lack of safe vaccines or antiviral treatments urgently needs to be addressed. NS3 and NS5 are highly conserved among the four serotypes, and the protein sequence around the pinpointed amino acids from the NS3 and NS5 regions are also conserved. The identification of the functionally essential interaction between the two proteins by biochemical and reverse genetics methods paves the way for rational drug design efforts to inhibit viral RNA synthesis.  相似文献   

4.
We have previously reported that the NS3 helicase (N3H) and NS5B-to-3′X (N5BX) regions are important for the efficient replication of hepatitis C virus (HCV) strain JFH-1 and viral production in HuH-7 cells. In the current study, we investigated the relationships between HCV genome replication, virus production, and the structure of N5BX. We found that the Q377R, A450S, S455N, R517K, and Y561F mutations in the NS5B region resulted in up-regulation of J6CF NS5B polymerase activity in vitro. However, the activation effects of these mutations on viral RNA replication and virus production with JFH-1 N3H appeared to differ. In the presence of the N3H region and 3′ untranslated region (UTR) of JFH-1, A450S, R517K, and Y561F together were sufficient to confer HCV genome replication activity and virus production ability to J6CF in cultured cells. Y561F was also involved in the kissing-loop interaction between SL3.2 in the NS5B region and SL2 in the 3′X region. We next analyzed the 3′ structure of HCV genome RNA. The shorter polyU/UC tracts of JFH-1 resulted in more efficient RNA replication than J6CF. Furthermore, 9458G in the JFH-1 variable region (VR) was responsible for RNA replication activity because of its RNA structures. In conclusion, N3H, high polymerase activity, enhanced kissing-loop interactions, and optimal viral RNA structure in the 3′UTR were required for J6CF replication in cultured cells.  相似文献   

5.
New natural intergenotypic (2/5) recombinant of hepatitis C virus   总被引:3,自引:1,他引:3       下载免费PDF全文
A 9.2-kb sequence from a hepatitis C virus (HCV) strain found in southwest France was compared to sequences from reference strains in HCV sequence databases. We found a recombinant virus with genotype 2 at the 5′ end and genotype 5 at the 3′ end. The crossover point was located between genes NS2 and NS3. Recombination between HCV genotypes must now be considered in studies on HCV epidemiology and evolution and in predictions of the virus response to antiviral therapy. Knowing the location of the recombination point may also be useful for constructing infectious chimeric viruses.  相似文献   

6.
Synthetic oligodeoxyribonucleotides containing CpG-dinucleotides (CpG DNA) in specific sequence contexts activate the vertebrate immune system. We have examined the effect of 3′-deoxy-2′–5′-ribonucleoside (3′-deoxynucleoside) incorporation into CpG DNA on the immunostimulatory activity. Incorporation of 3′-deoxynucleosides results in the formation of 2′5′-internucleotide linkages in an otherwise 3′–5′-linked CpG DNA. In studies, both in vitro and in vivo, CpG DNA containing unnatural 3′-deoxynucleoside either within the CpG-dinucleotide or adjacent to the CpG-dinucleotide failed to induce immunostimulatory activity, suggesting that the modification was not recognized by the receptors. Incorporation of the same modification distal to the CpG-dinucleotide in the 5′-flanking sequence potentiated the immunostimulatory activity of the CpG DNA. The same modification when incorporated in the 3′-flanking sequence had an insignificant effect on immunostimulatory activity of CpG DNA. Interestingly, substitution of a 3′-deoxynucleoside in the 5′-flanking sequence distal to the CpG-dinucleotide resulted in increased IL-6 and IL-10 secretion with similar levels of IL-12 compared with parent CpG DNA. The incorporation of the same modification in the 3′-flanking sequence resulted in lower IL-6 and IL-10 secretion with similar levels of IL-12 compared with parent CpG DNA. These results suggest that site-specific incorporation of 3′-deoxynucleotides in CpG DNA modulates immunostimulatory properties.  相似文献   

7.
Activation of pre-messenger RNA (pre-mRNA) splicing requires 5′ splice site recognition by U1 small nuclear RNA (snRNA), which is replaced by U5 and U6 snRNA. Here we use crosslinking to investigate snRNA interactions with the 5′ exon adjacent to the 5′ splice site, prior to the first step of splicing. U1 snRNA was found to interact with four different 5′ exon positions using one specific sequence adjacent to U1 snRNA helix 1. This novel interaction of U1 we propose occurs before U1-5′ splice site base pairing. In contrast, U5 snRNA interactions with the 5′ exon of the pre-mRNA progressively shift towards the 5′ end of U5 loop 1 as the crosslinking group is placed further from the 5′ splice site, with only interactions closest to the 5′ splice site persisting to the 5′ exon intermediate and the second step of splicing. A novel yeast U2 snRNA interaction with the 5′ exon was also identified, which is ATP dependent and requires U2-branchpoint interaction. This study provides insight into the nature and timing of snRNA interactions required for 5′ splice site recognition prior to the first step of pre-mRNA splicing.  相似文献   

8.
The isomerization of phosphodiester functionality of nucleic acids from 3′,5′- to a less common 2′,5′-linkage influences the complex interplay of stereoelectronic effects that drive pseudorotational equilibrium of sugar rings and thus affect the conformational propensities for compact or more extended structures. The present study highlights the subtle balance of non-covalent forces at play in structural equilibrium of 2′,5′-linked RNA analogue, 3′-O-(2-methoxyethyl) substituted dodecamer *CG*CGAA*U*U*CG*CG, 3′-MOE-2′,5′-RNA, where all cytosines and uracils are methylated at C5. The NMR and UV spectroscopic studies have shown that 3′-MOE-2′,5′-RNA adopts both hairpin and duplex secondary structures, which are involved in a dynamic exchange that is slow on the NMR timescale and exhibits strand and salt concentration as well as pH dependence. Unusual effect of pH over a narrow physiological range is observed for imino proton resonances with exchange broadening observed at lower pH and relatively sharp lines observed at higher pH. The solution structure of 3′-MOE-2′,5′-RNA hairpin displays a unique and well-defined loop, which is stabilized by Watson–Crick A5·*U8 base pair and by n → π* stacking interactions of O4′ lone-pair electrons of A6 and *U8 with aromatic rings of A5 and *U7, respectively. In contrast, the stem region of 3′-MOE-2′,5′-RNA hairpin is more flexible. Our data highlight the important feature of backbone modifications that can have pronounced effects on interstrand association of nucleic acids.  相似文献   

9.
Infection with Japanese encephalitis virus (JEV) can induce the expression of pro-inflammatory cytokines and cause acute encephalitis in humans. β-oxidation breaks down fatty acids for ATP production in mitochondria, and impaired β-oxidation can induce pro-inflammatory cytokine expression. To address the role of fatty-acid β-oxidation in JEV infection, we measured the oxygen consumption rate of mock- and JEV-infected cells cultured with or without long chain fatty acid (LCFA) palmitate. Cells with JEV infection showed impaired LCFA β-oxidation and increased interleukin 6 (IL-6) and tumor necrosis factor α (TNF-α) expression. JEV nonstructural protein 5 (NS5) interacted with hydroxyacyl-CoA dehydrogenase α and β subunits, two components of the mitochondrial trifunctional protein (MTP) involved in LCFA β-oxidation, and NS5 proteins were detected in mitochondria and co-localized with MTP. LCFA β-oxidation was impaired and higher cytokines were induced in cells overexpressing NS5 protein as compared with control cells. Deletion and mutation studies showed that the N-terminus of NS5 was involved in the MTP association, and a single point mutation of NS5 residue 19 from methionine to alanine (NS5-M19A) reduced its binding ability with MTP. The recombinant JEV with NS5-M19A mutation (JEV-NS5-M19A) was less able to block LCFA β-oxidation and induced lower levels of IL-6 and TNF-α than wild-type JEV. Moreover, mice challenged with JEV-NS5-M19A showed less neurovirulence and neuroinvasiveness. We identified a novel function of JEV NS5 in viral pathogenesis by impairing LCFA β-oxidation and inducing cytokine expression by association with MTP.  相似文献   

10.
The alternatively spliced 290-nucleotide NS2-specific exon of the parvovirus minute virus of mice (MVM), which is flanked by a large intron upstream and a small intron downstream, constitutively appears both in the R1 mRNA as part of a large 5′-terminal exon (where it is translated in open reading frame 3 [ORF3]), and in the R2 mRNA as an internal exon (where it is translated in ORF2). We have identified a novel bipartite exon enhancer element, composed of CA-rich and purine-rich elements within the 5′ and 3′ regions of the exon, respectively, that is required to include NS2-specific exon sequences in mature spliced mRNA in vivo. These two compositionally different enhancer elements are somewhat redundant in function: either element alone can at least partially support exon inclusion. They are also interchangeable: either element can function at either position. Either a strong 3′ splice site upstream (i.e., the exon 5′ terminus) or a strong 5′ splice site downstream (i.e., the exon 3′ terminus) is sufficient to prevent skipping of the NS2-specific exon, and a functional upstream 3′ splice site is required for inclusion of the NS2-specific exon as an internal exon into the mature, doubly spliced R2 mRNA. The bipartite enhancer functionally strengthens these termini: the requirement for both the CA-rich and purine-rich elements can be overcome by improvements to the polypyrimidine tract of the upstream intron 3′ splice site, and the purine-rich element also supports exon inclusion mediated through the downstream 5′ splice sites. In summary, a suboptimal large-intron polypyrimidine tract, sequences within the downstream small intron, and a novel bipartite exonic enhancer operate together to yield the balanced levels of R1 and R2 observed in vivo. We suggest that the unusual bipartite exonic enhancer functions to mediate proper levels of inclusion of the NS2-specific exon in both singly spliced R1 and doubly spliced R2.  相似文献   

11.
12.
This study identified specific and avid RNA aptamers consisting of 2′-hydroxyl- or 2′-fluoropyrimidines against hepatitis C virus (HCV) NS5B replicase, an enzyme that is essential for HCV replication. These aptamers acted as potent decoys to competitively impede replicase-catalyzed RNA synthesis activity. Cytoplasmic expression of the 2′-hydroxyl aptamer efficiently inhibited HCV replicon replication in human liver cells through specific interaction with, and sequestration of, the target protein without either off-target effects or escape mutant generation. A selected 2′-fluoro aptamer could be truncated to a chemically manufacturable length of 29 nucleotides (nt), with increase in the affinity to HCV NS5B. Noticeably, transfection of the truncated aptamer efficiently suppressed HCV replication in cells without escape mutant appearance. The aptamer was further modified through conjugation of a cholesterol or galactose-polyethylene glycol ligand for in vivo availability and liver-specific delivery. The conjugated aptamer efficiently entered cells and inhibited genotype 1b subgenomic and genotype 2a full-length HCV JFH-1 RNA replication without toxicity and innate immunity induction. Importantly, a therapeutically feasible amount of the conjugated aptamer was delivered in vivo to liver tissue in mice. Therefore, cytoplasmic expression of 2′-hydroxyl aptamer or direct administration of chemically synthesized and ligand-conjugated 2′-fluoro aptamer against HCV NS5B could be a potent anti-HCV approach.  相似文献   

13.
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   

14.
15.
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   

16.
Conformational properties of trimeric and tetrameric 2′,5′-linked oligonucleotides, 3′-MOE-A32′,5′ (1) and 3′-MOE-A42′,5′ (2), and their 3′,5′-linked analogs, 2′-MOE-A33′,5′ (3) and 2′-MOE-A43′,5′ (4), were examined with the use of heteronuclear NMR spectroscopy. The temperature-dependent 3JHH, 3JHP and 3JCP coupling constants, acquired in the range of 273–343 K, gave insight into the conformation of sugar rings in terms of a two-state North ↔ South (N ↔ S) pseudorotational equilibrium and into the conformation of the sugar–phosphate backbone in the model antisense oligonucleotides 1–4. 2′,5′-linked oligomers 3′-MOE-A32′,5′ (1) and 3′-MOE-A42′,5′ (2) show preference for N-type conformers and indication of A-type conformational features, which is prerequisite for antisense hybridization. The drive of N ↔ S equilibrium in 1–4 has been rationalized with the competing gauche effects of 2′/3′-phosphodiester and 3′/2′-MOE groups, anomeric and steric effects. Furthermore, the pairwise comparisons of 3′-MOE with 3′-OH and 3′-deoxy 2′,5′-linked adenine trimers emphasized the fine tuning of N ↔ S equilibrium in 3′-MOE-A32′,5′ (1) and 3′-MOE-A42′,5′ (2) by the steric effects of 3′-MOE group and the possibility of water-mediated H-bonds with vicinal phosphodiester functionality. In full correspondence, the drive of N ↔ S equilibrium towards N by 2′-MOE in 3′,5′-linked analogs 2′-MOE-A33′,5′ (3) and 2′-MOE-A43′,5′ (4) is weaker in comparison with 3′-OH group in the corresponding ribo analogs. βt, γ+ and ε rotamers are preferred in both 2′,5′- and in 3′,5′-linked oligonucleotides 1–4.  相似文献   

17.
Barley (Hordeum vulgare L.) leaves were used to isolate and characterize the chloroplast NAD(P)H dehydrogenase complex. The stroma fraction and the thylakoid fraction solubilized with sodium deoxycholate were analyzed by native polyacrylamide gel electrophoresis, and the enzymes detected with NADH and nitroblue tetrazolium were electroeluted. The enzymes electroeluted from band S from the stroma fraction and from bands T1 (ET1) and T2 from the thylakoid fraction solubilized with sodium deoxycholate had ferredoxin-NADP oxidoreductase (FNR; EC 1.18.1.2) and NAD(P)H-FeCN oxidoreductase (NAD[P]H-FeCNR) activities. Their NADPH-FeCNR activities were inhibited by 2′-monophosphoadenosine-5′-diphosphoribose and by enzyme incubation with p-chloromercuriphenylsulfonic acid (p-CMPS), NADPH, and p-CMPS plus NADPH. They presented Michaelis constant NADPH values that were similar to those of FNRs from several sources. Their NADH-FeCNR activities, however, were not inhibited by 2′-monophosphoadenosine-5′-diphosphoribose but were weakly inhibited by enzyme incubation with NADH, p-CMPS, and p-CMPS plus NADH. We found that only ET1 contained two polypeptides of 29 and 35 kD, which reacted with the antibodies raised against the mitochondrial complex I TYKY subunit and the chloroplast ndhA gene product, respectively. However, all three enzymes contained two polypeptides of 35 and 53 kD, which reacted with the antibodies raised against barley FNR and the NADH-binding 51-kD polypeptide of the mitochondrial complex I, respectively. The results suggest that ET1 is the FNR-containing thylakoidal NAD(P)H dehydrogenase complex.  相似文献   

18.
A procedure was devised to detect and assay uridine 5′-pyrophosphate (UDP)-glucuronic acid pyrophosphorylase in plant extracts. Substrates are UDP-glucuronic acid and 32P-pyrophosphate, and the 32P-uridine 5′-triphosphate produced is selectively adsorbed to charcoal. The charcoal adsorption procedure is a modification of that used to determine 32P-adenosine 5′-triphosphate produced by adenosine 5′-pyrophosphate glucose pyrophosphorylase, and the modification greatly improves the retention of uridine 5′-triphosphate.  相似文献   

19.
Evidence is presented for the occurrence of 5′-terminal capping structures in the polyadenylated RNA of oat (Avena sativa) coleoptiles. These structures are composed of an inverted terminal nucleoside containing the modified base 7-methylguanine which is joined 5′ to 5′ with a second (penultimate) nucleoside by means of three phosphate groups in two pyrophosphate linkages. The penultimate nucleoside is joined to the remainder of the RNA molecule by a conventional 3′,5′ phosphodiester bond. A significant difference between the cap structures of oat coleoptile RNA and those of previously described higher eucaryotic cellular mRNAs is the lack of ribose methylations in the penultimate nucleosides of the plant RNA.  相似文献   

20.
Owing to the markedly increased reactivity of amino functional groups versus hydroxyls, the 5′-amino-5′-deoxy nucleoside and nucleotide analogs have proven widely useful in biological, pharmaceutical and genomic applications. However, synthetic procedures leading to these analogs have not been fully explored, which may possibly have limited the scope of their utility. Here we describe the synthesis of the 5′-amino-2′,5′-dideoxy analogs of adenosine, cytidine, guanosine, inosine and uridine from their respective naturally occurring nucleosides via the reduction of 5′-azido-2′,5′-dideoxy intermediates using the Staudinger reaction, and the high yield conversion of these modified nucleosides and 5′-amino-5′-deoxythymidine to the corresponding 5′-N-triphosphates through reaction with trisodium trimetaphosphate in the presence of tris(hydroxymethyl)aminomethane (Tris). We also show that each of these nucleotide analogs can be efficiently incorporated into DNA by the Klenow fragment of Escherichia coli DNA polymerase I when individually substituted for its naturally occurring counterpart. Mild acid treatment of the resulting DNA generates polynucleotide fragments that arise from specific cleavage at each modified nucleotide, providing a sequence ladder for each base. Because the ladders are generated after the extension, the corresponding products may be manipulated by enzymatic and/or purification processes. The potential utility of this extension–cleavage procedure in genomic sequence analysis is discussed.  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