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.
2.
3.
4.
Short fragments of DNA (5 S) isolated by denaturation from polyoma replicative intermediates pulse-labeled in vitro were shown to have RNA covalently attached by three criteria: (1) such fragments were slightly denser than bulk viral DNA. (2) They could be labeled directly with α-32P-labeled ribotriphosphates. (3) Alkaline hydrolysis of fragments labeled with α-32P-labeled deoxynucleoside triphosphates showed 32P transfer to 3′ ribonucleoside monophosphates. Except for a preference of transfer from dC, the link showed little sequence specificity. The data are compatible with the notion that all short fragments in replicating viral DNA are initiated by an RNA primer. This RNA is maximally 30 bases long and is rather short-lived.  相似文献   

5.
The binding sites of calf thymus RNA polymerase II on polyoma DNA were monitored by electron microscopy. Six discrete binding sites were located at positions 0.06, 0.25, 0.57, 0.66, 0.85 and 0.98 on the physical map of polyoma DNA. Although most of these sites are located in easily denaturable regions of the DNA, the strongest binding sites do not overlap with the major A + T-rich regions. In addition, the same binding sites were observed on superhelical or linear polyoma DNA. These results suggest that the eucaryotic RNA polymerase II can recognize specific sequences on double-stranded DNA and not only easily denaturable regions. At least five of these sites correspond to the binding and initiation sites mapped previously for the Escherichia coli RNA polymerase (Lescure et al., 1976).Stable initiation complexes can be formed with both E. coli and calf thymus RNA polymerases in the presence of a single dinucleotide (GpU) and a specific ribotriphosphate (CTP). Under these conditions, the binding of both enzymes to the sites in positions 0.06 and 0.57 is stimulated whereas the binding in positions 0.65 and 0.84 is partially suppressed. Both eucaryotic and procaryotic RNA polymerases may recognize similar sequences of the viral DNA in vitro.  相似文献   

6.
The structure of the polyoma virus defective species D74 (74% the size of full-length polyoma virus DNA) has been determined and compared with that of polyoma virus A2 DNA. D74 appears to be composed entirely of viral DNA sequences. (No host DNA sequences have been detected.) It is made up of three DNA segments, each about 24, 24 and 27% in size. The two 24% segments appear to be identical and the 27% segment contains one copy of all the sequences found in the 24% fragments as well as a duplication of some of the sequences. When related to the physical map of A2 DNA, each segment is found to be composed of viral sequences from 1 to about 19 map units, 67 to 69 map units and 70 to 72 map units.Three features found in other polyoma virus defective species (Lund et al., 1977) are also present in D74. (1) Sequences from the region around 67 map units are linked to other (non-contiguous) viral sequences. (2) Sequences at about 72 map units are linked to sequences at 1 map unit. (3) Multiple copies of sequences from around the origin of viral DNA replication are present. From studies on other polyoma defective molecules (Griffin &; Fried, 1975; Lund et al., 1977), the origin of DNA replication for polyoma virus has been defined to lie within the sequences from 67 to 72 map units. Since D74 replicates efficiently but lacks the sequences between 69 to 70 map units, the origin of DNA replication appears to be further defined as lying within 67 and 69 map units and/or 70 to 72 map units.  相似文献   

7.
The phosphoesterase (PE) domain of the bacterial DNA repair enzyme LigD possesses distinctive manganese-dependent 3′-phosphomonoesterase and 3′-phosphodiesterase activities. PE exemplifies a new family of DNA end-healing enzymes found in all phylogenetic domains. Here, we determined the structure of the PE domain of Pseudomonas aeruginosa LigD (PaePE) using solution NMR methodology. PaePE has a disordered N-terminus and a well-folded core that differs in instructive ways from the crystal structure of a PaePE•Mn2+• sulfate complex, especially at the active site that is found to be conformationally dynamic. Chemical shift perturbations in the presence of primer-template duplexes with 3′-deoxynucleotide, 3′-deoxynucleotide 3′-phosphate, or 3′ ribonucleotide termini reveal the surface used by PaePE to bind substrate DNA and suggest a more efficient engagement in the presence of a 3′-ribonucleotide. Spectral perturbations measured in the presence of weakly catalytic (Cd2+) and inhibitory (Zn2+) metals provide evidence for significant conformational changes at and near the active site, compared to the relatively modest changes elicited by Mn2+.  相似文献   

8.
Polyoma virus transcription in vitro.   总被引:5,自引:0,他引:5  
  相似文献   

9.
The rate of synthesis of cellular DNA is stimulated in stationary phase mouse embryo cells infected with polyoma virus. Nascent cellular DNA strands pulselabeled with [3H]thymidine in the presence of replicating viral DNA are smaller, by an average of 2·1 × 107 daltons, than DNA made under similar conditions in uninfected cells. Previous work (Cheevers et al., 1972) has indicated that this observation is the consequence of activation in infected cells of cellular DNA initiation sites not in operation during a similar pulse-labeling interval in uninfected cells. Similar results were obtained using cells infected with the temperature-sensitive Ts-a mutant of polyoma at 32 °C, which permits both the induction of cellular DNA synthesis and replication of viral DNA. However, at a temperature of 39 °C, which permits only the induction of cellular DNA replication in Ts-a-infected cells, the size of newly synthesized DNA is not different from that of uninfected cells. Similarly, in rat embryo cells abortively infected with polyoma (wild-type), stimulation of cellular DNA synthesis occurs but viral DNA replication is restricted, and no difference is apparent in the size of newly formed DNA as compared to uninfected cells. These results are interpreted to mean that in productively infected cells, polyoma DNA and some regions of the host genome may be co-ordinately replicated.  相似文献   

10.
11.
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   

12.
Intermediate in SV40 DNA Chain Growth   总被引:19,自引:0,他引:19  
PREVIOUS studies of the DNA replication of simian virus 40 (SV40), an oncogenic member of the papoyavirus group, have been concerned with separation and characterization of replicative intermediates1–4. Circular replicating intermediates have been identified for SV401–3, as well as for the similar replication system of polyoma viral DNA5,6. The replicative intermediates of SV40 DNA have been observed by electron microscopy to contain two forks, three branches and no free ends1–3 as is the case for the circular replicating molecules of polyoma, bacteriophage λ7, Escherichia coli8 and colicin E1 in mini-cells9,10. An important property of replicative intermediates of SV40 DNA that has also been observed in replicating molecules of colicin E110 is that most molecules contain a superhelical region in the unreplicated portion of the molecule1.  相似文献   

13.
Chronic psychological stress is associated with accelerated aging, but the underlying biological mechanisms are not known. Prolonged elevations of the stress hormone cortisol is suspected to play a critical role. Through its actions, cortisol may potentially induce oxidatively generated damage to cellular constituents such as DNA and RNA, a phenomenon which has been implicated in aging processes. We investigated the relationship between 24 h excretion of urinary cortisol and markers of oxidatively generated DNA and RNA damage, 8-oxo-7,8-dihydro-2′-deoxyguanosine and 8-oxo-7,8-dihydroguanosine, in a sample of 220 elderly men and women (age 65 – 83 years). We found a robust association between the excretion of cortisol and the oxidation markers (R2 = 0.15, P<0.001 for both markers). Individuals in the highest quartile of cortisol excretion had a 57% and 61% higher median excretion of the DNA and RNA oxidation marker, respectively, than individuals in the lowest quartile. The finding adds support to the hypothesis that cortisol-induced damage to DNA/RNA is an explanatory factor in the complex relation between stress, aging and disease.  相似文献   

14.
Three related polyoma virus species, designated D92 (92% the size of full-length polyoma virus DNA), D91 (91%) and D76 (76%) have been analysed and their structures compared with that of polyoma virus A2 DNA. Three independent methods (restriction endonuclease cleavage, depurination fingerprinting and DNA-DNA hybridization) were used in the analysis.The defective DNAs appear to be: (1) entirely composed of viral sequences (no host DNA sequences were detected): (2) made up in part of long continuous sequences of DNA which appear identical to sequences of A2 DNA (D92 contains continuous sequences from 1 to 72 map units on the physical map of A2 DNA; that is, it contains the entire late region and part of the early region of the viral DNA. D91 and D76 contain those same sequences except for a 1% deletion around 18 map units): (3) made up in part of rearranged viral sequences.Several interesting features were noted about the rearranged sequences present in the defective DNAs. Sequences from the region around 67 map units were found linked to other (non-contiguous) regions of the DNA. Sequences from about 72 map units were linked to sequences from about 1 map unit. Multiple copies of sequences from 67 to 72 map units (from around the origin of DNA replication) were found (4 copies in D91 and D92, and 2 copies in D76).  相似文献   

15.
Structure of polyoma virus late nuclear RNA   总被引:18,自引:0,他引:18  
  相似文献   

16.
17.
18.
In addition to their capacity for template-directed 5′ → 3′ DNA synthesis at the polymerase (pol) site, DNA polymerases have a separate 3′ → 5′ exonuclease (exo) editing activity that is involved in assuring the fidelity of DNA replication. Upon misincorporation of an incorrect nucleotide residue, the 3′ terminus of the primer strand at the primer-template (P/T) junction is preferentially transferred to the exo site, where the faulty residue is excised, allowing the shortened primer to rebind to the template strand at the pol site and incorporate the correct dNTP. Here we describe the conformational changes that occur in the primer strand as it shuttles between the pol and exo sites of replication-competent Klenow and Klentaq DNA polymerase complexes in solution and use these conformational changes to measure the equilibrium distribution of the primer between these sites for P/T DNA constructs carrying both matched and mismatched primer termini. To this end, we have measured the fluorescence and circular dichroism spectra at wavelengths of >300 nm for conformational probes comprising pairs of 2-aminopurine bases site-specifically replacing adenine bases at various positions in the primer strand of P/T DNA constructs bound to DNA polymerases. Control experiments that compare primer conformations with available x-ray structures confirm the validity of this approach. These distributions and the conformational changes in the P/T DNA that occur during template-directed DNA synthesis in solution illuminate some of the mechanisms used by DNA polymerases to assure the fidelity of DNA synthesis.Escherichia coli DNA polymerase (DNAP)2 I is a single subunit polymerase that is organized into three functional domains: an N-terminal domain that is associated with 5′ → 3′ exonuclease activity, an intermediate domain that carries the 3′ → 5′ proofreading activity, and a C-terminal domain that is associated with the 5′ → 3′ template-directed polymerization activity. An important role of DNAP I is to remove the RNA primers of the Okazaki fragments formed during lagging strand DNA synthesis in E. coli replication and to fill in the resulting gaps by template-directed DNA synthesis (1). An N-terminal deletion mutant of DNAP I, known as the “large fragment” or Klenow form of the enzyme, contains only the polymerase (pol) and the 3′ → 5′ exonuclease (exo) domains. The Klenow polymerase has served and continues to serve as an excellent model system for isolating and defining general structure-function relationships in polymerases and in the supporting machinery of DNA replication.The main function of the 3′ → 5′ exonuclease activity of DNAP I is to remove misincorporated nucleotide residues from the 3′-end of the primer (2), thus contributing significantly to the overall fidelity of DNA replication (3). Contrary to initial expectations, crystallographic studies showed that the pol and exo active sites are quite far apart in replication polymerases, about 30 Å in Klenow (4). As a consequence, the ability of polymerases to “shuttle” the 3′-end of the primer strand efficiently between the pol and the exo sites in order to rectify misincorporation events during polymerization is critical to maintaining the overall accuracy of template-directed replication. Elucidation of the mechanisms of this shuttling and determination of the factors that control the rates (and equilibria) of the active site switching reaction will certainly increase our understanding of fidelity control by DNA polymerases.An early crystallographic study of the Klenow polymerase complexed with fully paired primer-template (P/T) DNA revealed that 3–4 nt of the 3′-primer terminus had been unwound from the template stand and partitioned into the exo site and that an extended single-stranded DNA (ssDNA) binding pocket of the exo site appeared to make position-specific hydrophobic contacts with the unstacked bases at the 3′-end of the primer (4). A separate crystallographic study of an editing complex confirmed that an ssDNA fragment 4 nt in length was bound at the exo site in the same conformation as seen for the single-stranded 3′-primer sequence unwound from P/T DNA (5). A structure of Klenow polymerase with the DNA bound at the pol site has not yet been reported, although such structures have been obtained for other homologous polymerases, including Klentaq (the “large fragment” of Thermus aquaticus (Taq) DNAP), Bacillus stearothermophilus (Bst) “large fragment” polymerase, and the T7 DNAP (68), all of which are members of the polymerase family that includes Klenow.The amino acid residues involved in the binding of DNA at the pol site in these polymerases (determined from co-crystal structures) and those of Klenow (determined by site-directed mutagenesis studies (9, 10)) are highly conserved, suggesting that a similar DNA binding mode at the pol site may apply to all of the DNAP I polymerases. The crystal structure of Klenow revealed that the polymerization domain has a shape reminiscent of a right hand in which the palm, fingers, and thumb domains form the DNA-binding crevice. Structural studies with various DNAP I polymerases in the presence of P/T DNA constructs yielded an “open” binary complex, whereas the addition of the next correct dNTP (as a chain-terminating dideoxy-NTP) resulted in the formation of a catalytically competent “closed” ternary complex (68). In the latter complex, the 3′-primer terminus was base-paired with the template DNA, and the templating base was poised for incorporation of the next correct nucleotide. These structures showed that the conformation of the DNA primer terminus bound at the pol site is markedly different from that of the “frayed open” primer observed at the exo site in Klenow (4, 5).Although crystallographic studies have provided a wealth of information about the conformations of the DNA substrates bound at the active sites of DNAP, replication itself is a dynamic process (reviewed in Ref. 11), and it is critical to be able to distinguish between various forms of DNA-polymerase complexes in solution in order to fully understand the mechanistic details of the replication process. A solution approach used by Millar and co-workers (reviewed in Ref. 12) for studying the conformation of DNA in these complexes involved measuring the time-resolved fluorescence anisotropy properties of a dansyl fluorophore attached to a DNA base located 8 bp upstream of the P/T DNA junction. The changes in the lifetime of the fluorophore, which appeared to depend mostly on the local environment occupied by the probe within the protein (i.e. buried versus partially exposed), were correlated with specific binding conformations of the primer to provide an estimate of the fractional occupancy of the pol and the exo sites. Reha-Krantz and co-workers (13) more recently used a related approach, here involving the monitoring of changes in the fluorescent lifetimes of a single 2-aminopurine (2-AP) base (a fluorescent analogue of adenine) site-specifically substituted in the template strand at the P/T junction, to make similar fractional occupancy measurements. However, we note that structural interpretations of these fluorescence experiments relied heavily on the available crystal structures, and it remained to be shown directly that the 3′-end of the primer in P/T DNA constructs assumes the same distribution of conformations when bound to the protein in solution.To get around this problem, as well as to directly investigate the conformations of the primer DNA in both active sites of the Klenow and Klentaq polymerases, we have used a novel CD spectroscopic approach to characterize the solution conformations of primer DNA bound to Klenow and Klentaq DNAPs. Previously, we had shown that CD spectroscopy, in conjunction with fluorescence measurements, can be used to examine changes in local DNA and RNA conformations at 2-AP dimer probes inserted at specified positions within the nucleic acid frameworks of a variety of macromolecular machines functioning in solution (1416). 2-AP is a structural isomer of adenine that forms base pairs with thymine in DNA (and uridine in RNA), and the substitution of 2-AP for adenine in such bp does not significantly perturb the structure or stability of the resultant double helix. Furthermore, when these probes are used as dimer pairs, the CD spectrum primarily reflects the interaction of the transition dipoles of the two probes themselves and thus the local conformation of the DNA at those positions within the P/T DNA. The characteristic CD and fluorescence signals for 2-AP probes in nucleic acids occur at wavelengths of >300 nm, a spectral region in which the protein and the canonical nucleic acid components of the “macromolecular machines of gene expression” are otherwise transparent. In this study, we have examined the binding of Klenow and Klentaq polymerases to P/T DNA constructs that were designed to be comparable with the nucleic acid components of functioning replication complexes. By examining the low energy CD spectra of site-specifically placed 2-AP probes, we have been able to characterize base conformations at defined positions within the DNA to reveal conformational features of specific DNA bases bound at and near both the pol and the exo active sites of these polymerases. These measurements, in that they directly reflect the actual conformations of the DNA chains bound within the active sites of the functioning polymerase, have also provided a direct means to estimate the equilibrium distributions of primer ends between the two active sites for various P/T DNA constructs.  相似文献   

19.
Fine structure of polyoma virus DNA.   总被引:7,自引:0,他引:7  
A fine structure map of polyoma DNA has been made based on cleavage with a number of restriction endonucleases (including HaeII and III, BamI, HindII and III, BumI, HpaII, and in part, HphI) and depurination of wild-type DNA, the eight HpaII restriction fragments and some HaeIII fragments. This analysis has made possible some correlation with simian virus 40 DNA, and has facilitated detailed examination of various polyoma strains and variants. Sequences from the region of the origin of DNA replication have been examined.  相似文献   

20.
Prior studies have shown that 2′-O methyltransferase activity of flaviviruses, coronaviruses, and poxviruses promotes viral evasion of Ifit1, an interferon-stimulated innate immune effector protein. Viruses lacking 2′-O methyltransferase activity exhibited attenuation in primary macrophages that was rescued in cells lacking Ifit1 gene expression. Here, we examined the role of Ifit1 in restricting pathogenesis in vivo of wild type WNV (WNV-WT) and a mutant in the NS5 gene (WNV-E218A) lacking 2′-O methylation of the 5′ viral RNA cap. While deletion of Ifit1 had marginal effects on WNV-WT pathogenesis, WNV-E218A showed increased replication in peripheral tissues of Ifit1 −/− mice after subcutaneous infection, yet this failed to correlate with enhanced infection in the brain or lethality. In comparison, WNV-E218A was virulent after intracranial infection as judged by increased infection in different regions of the central nervous system (CNS) and a greater than 16,000-fold decrease in LD50 values in Ifit1 −/− compared to wild type mice. Ex vivo infection experiments revealed cell-type specific differences in the ability of an Ifit1 deficiency to complement the replication defect of WNV-E218A. In particular, WNV-E218A infection was impaired in both wild type and Ifit1 −/− brain microvascular endothelial cells, which are believed to participate in blood-brain barrier (BBB) regulation of virus entry into the CNS. A deficiency of Ifit1 also was associated with increased neuronal death in vivo, which was both cell-intrinsic and mediated by immunopathogenic CD8+ T cells. Our results suggest that virulent strains of WNV have largely evaded the antiviral effects of Ifit1, and viral mutants lacking 2′-O methylation are controlled in vivo by Ifit1-dependent and -independent mechanisms in different cell types.  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