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.
A vector which expresses the herpes simplex virus type 1 (HSV-1) (strain 17) DNA polymerase gene was constructed by ligating two separately cloned HSV DNA restriction fragments into an intermediate plasmid and then mobilizing the intact polymerase gene-encoding sequence into a pSV2 derivative. The expression vector (pD7) contains a functional simian virus 40 replication origin and early enhancer-promoter upstream from the HSV DNA polymerase-encoding sequence. COS-1 cells transfected with pD7 contained an RNA species, shown by Northern blot analysis to hybridize specifically with an HSV DNA pol probe and to be the same size (4.3 kilobases) as the pol mRNA found in HSV-1-infected COS-1 cells. A genetic complementation test was used to establish that pD7 expresses a functional pol gene product. COS-1 cells transfected with pD7 were able to partially complement the growth defect of an HSV-1 (KOS) temperature-sensitive mutant, tsC7, in the DNA polymerase gene at the nonpermissive temperature.  相似文献   

3.
The analysis of the deduced amino acid sequence of the herpes simplex virus type 1 (HSV-1) DNA polymerase reported here suggests that the polymerase structure consists of domains carrying separate biological functions. The HSV-1 enzyme is known to possess 5'-3'-exonuclease (RNase H), 3'-5'-exonuclease, and DNA polymerase catalytic activities. Sequence analysis suggests an arrangement of these activities into distinct domains resembling the organization of Escherichia coli polymerase I. In order to more precisely define the structure and C-terminal limits of a putative catalytic domain responsible for the DNA polymerization activity of the HSV-1 enzyme, we have undertaken in vitro mutagenesis and computer modeling studies of the HSV-1 DNA polymerase gene. Sequence analysis predicts that the major DNA polymerization domain of the HSV-1 enzyme will be contained between residues 690 and 1100, and we present a three-dimensional model of this region, on the basis of the X-ray crystallographic structure of the E. coli polymerase I. Consistent with these structural and modeling studies, deletion analysis by in vitro mutagenesis of the HSV-1 DNA polymerase gene expressed in Saccharomyces cerevisiae has confirmed that certain amino acids from the C terminus (residues 1073 to 1144 and 1177 to 1235) can be deleted without destroying HSV-1 DNA polymerase catalytic activity and that the extreme N-terminal 227 residues are also not required for this activity.  相似文献   

4.
Herpes simplex virus type 1 (HSV-1) DNA polymerase contains several conserved regions within the polymerase domain. The conserved regions I, II, III, V, and VII have been shown to have functional roles in the interaction with deoxynucleoside triphosphates (dNTPs) and DNA. However, the role of conserved region VI in DNA replication has remained unclear due, in part, to the lack of a well-characterized region VI mutant. In this report, recombinant viruses containing a point mutation (L774F) within the conserved region VI were constructed. These recombinant viruses were more susceptible to aphidicolin and resistant to both foscarnet and acyclovir, compared to the wild-type KOS strain. Marker transfer experiments demonstrated that the L774F mutation conferred the altered drug sensitivities. Furthermore, mutagenesis assays demonstrated that L774F recombinant viruses containing the supF marker gene, which was integrated within the thymidine kinase locus (tk), exhibited increased fidelity of DNA replication. These data indicate that conserved region VI, together with other conserved regions, forms the polymerase active site, has a role in the interaction with deoxyribonucleotides, and regulates DNA replication fidelity. The possible effect of the L774F mutation in altering the polymerase structure and activity is discussed.  相似文献   

5.
6.
Properties of herpes simplex virus type 1 and type 2 DNA polymerase   总被引:25,自引:0,他引:25  
Herpes simplex virus type 1 (HSV-1) and type 2 (HSV-2) DNA polymerases were highly purified from infected HeLa BU cells by DEAE cellulose, phosphocellulose and DNA cellulose column chromatography. DNA exonuclease activity but not endonuclease activity was found associated with both types of DNA polymerase. Both DNA polymerase activities could be activated by salt in a similar fashion with the optimal activity in the range of ionic strength between 0.22 and 0.29 alpha. At an ionic strength of 0.14, spermidine and putrescine in the concentration range (0--5 mM) studied could mimic the action of KCI in stimulating DNA polymerase activity. Spermine, in the same concentration range, had a biphasic effect. At an ionic strength of 0.29 all three polyamines were inhibitory. HSV-1 and HSV-2 DNA polymerase are similar in their column chromatographic behavior, sedimentation rate in sucrose gradient centrifugation, and activation energy, but they differ in their heat stability at 45 degrees C with the HSV-2 enzyme more stable than the HSV-1 enzyme. Kinetic behavior of both enzymes is similar, with Km values for deoxyribonucleoside triphosphates in the range of 5 . 10(-7) to 1.8 . 10(-8) M. IdUTP and dUTP served as apparent competitive inhibitors with respect to dTTP, and AraATP acted as an apparent competitive inhibitor with respect to dATP. AraATP could not replace dATP in the DNA polymerization reaction; in contrast, IdUTP could replace TTP. Phosphonoformic acid behaved as an uncompetitive inhibitor with respect to DNA. The ID(50) value estimated was foind to be dependent on the purity of the DNA polymerase used and the ionic strength of the assay condition. Each DNA-polymerase associated DNA exonuclease had the same stability at 45 degrees C as its DNA polymerase. The associated DNAase activity was inhibited by phosphonoformic acid and high ionic strength of the assay condition.  相似文献   

7.
The biochemical properties of four acyclovir-resistant mutants are described. Two of these mutants, PAAr5 and BWr, specified nucleotidyl transferase (DNA polymerase) activities which were less sensitive to inhibition by acyclovir triphosphate than their wild-type counterparts. Another mutant, IUdRr, exhibited reduced ability to phosphorylate acyclovir. The fourth mutant, ACGr4, both induced an altered DNA polymerase and failed to phosphorylate appreciable amounts of acyclovir. BWr, a new acyclovir-resistant mutant derived from the Patton strain of herpes simplex virus type 1, induced a DNA polymerase resistant to inhibition by acyclovir triphosphate, but, unlike the polymerases induced by PAAr5 and ACGr4, still sensitive to phosphonoacetic acid. Resistance of BWr to acyclovir mapped close to the PAAr locus and was separable from mutations in the herpes simplex virus thymidine kinase gene by recombination analysis.  相似文献   

8.
9.
10.
Fifteen temperature-sensitive mutants of herpes simplex virus type 1 were studied with regard to the relationship between their ability to synthesize viral DNA and to induce viral DNA polymerase (DP) activity at permissive (34 C) and nonpermissive (39 C) temperatures. At 34 C, all mutants synthesized viral DNA, while at 39 C four mutants demonstrated a DNA+ phenotype, three were DNA+/-, and eight were DNA-. DNA+ mutants induced levels of DP activity similar to thhose of the wild-type virus at both temperatures, and DNA+/- mutants induced reduced levels of DP activity at 39 C but not at 34 C. Among the DNA- mutants three were DP+, two were DP+/-, and three showed reduced DP activity at 34 C with no DP activity at 39 C. DNA-, DP- mutants induced the synthesis of a temperature-sensitive DP as determined by in vivo studies.  相似文献   

11.
The effect of exonuclease activity of the herpes simplex virus DNA polymerase (Pol) on DNA replication fidelity was examined by using the supF mutagenesis assay. The recombinants with exonuclease-deficient Pol, containing an integrated supF gene in the thymidine kinase locus (tk), exhibited supF mutation frequencies ranging from 0.14 to 5.6%, consistent with the tk mutation frequencies reported previously (Y. T. Hwang, B.-Y. Liu, D. M. Coen, and C. B. C. Hwang, J. Virol. 71:7791-7798, 1997). The increased mutation frequencies were 10- to 500-fold higher than those observed for wild-type Pol recombinants. The increased mutation frequencies also were significantly higher than those of supF mutant replicated by exonuclease-deficient Pols in the plasmid-borne assay. Furthermore, characterization of supF mutants demonstrated that recombinants with a defective exonuclease induced types and distributions of supF mutations different from those induced by wild-type Pol recombinants. The types of supF mutations induced by exonuclease-deficient recombinants differed between the plasmid- and genome-based assays. The spectra of supF mutations also differed between the two assays. In addition, exonuclease-defective viruses also induced different spectra of supF and tk mutations. Therefore, both the assay methods and the target genes used for mutagenesis studies can affect the repication fidelity of herpes simplex virus type 1 Pol with defective exonuclease activity.  相似文献   

12.
Three herpes simplex virus mutants that contain drug resistance mutations in the DNA polymerase gene exhibited no significant reduction in replication in the ears of mice compared with the wild type after inoculation at that site but were attenuated for pathogenicity after intracerebral inoculation. Cataracts were common sequelae in mice that survived mutant infections.  相似文献   

13.
During their life cycles, viruses typically undergo many transport events throughout the cell. These events depend on a variety of both viral and host proteins and are often not fully understood. Such studies are often complicated by asynchronous infections and the concurrent presence of various viral intermediates in the cells, making it difficult to molecularly define each step. In the case of the herpes simplex virus type 1, the etiological agent of cold sores and many other illnesses, the viral particles undergo an intricate series of transport steps during its life cycle. Upon entry by fusion with a cellular membrane, they travel to the host cell nucleus where the virus replicates and assembles new viral particles. These particles then travel across the two nuclear envelopes and transit through the trans-Golgi network before finally being transported to and released at the cell surface. Though viral components and some host proteins modulating these numerous transport events have been identified, the details of these processes remain to be elucidated. To specifically address how the virus escapes the nucleus, we set up an in vitro model that reproduces the unconventional route used by herpes simplex type 1 virus to leave nuclei. This has not only allowed us to clarify the route of capsid egress of the virus but is now useful to define it at the molecular level.  相似文献   

14.
Herpesviruses are the second leading cause of human viral diseases. Herpes Simplex Virus types 1 and 2 and Varicella-zoster virus produce neurotropic infections such as cutaneous and genital herpes, chickenpox, and shingles. Infections of a lymphotropic nature are caused by cytomegalovirus, HSV-6, HSV-7, and Epstein-Barr virus producing lymphoma, carcinoma, and congenital abnormalities. Yet another series of serious health problems are posed by infections in immunocompromised individuals. Common therapies for herpes viral infections employ nucleoside analogs, such as Acyclovir, and target the viral DNA polymerase, essential for viral DNA replication. Although clinically useful, this class of drugs exhibits a narrow antiviral spectrum, and resistance to these agents is an emerging problem for disease management. A better understanding of herpes virus replication will help the development of new safe and effective broad spectrum anti-herpetic drugs that fill an unmet need. Here, we present the first crystal structure of a herpesvirus polymerase, the Herpes Simplex Virus type 1 DNA polymerase, at 2.7 A resolution. The structural similarity of this polymerase to other alpha polymerases has allowed us to construct high confidence models of a replication complex of the polymerase and of Acyclovir as a DNA chain terminator. We propose a novel inhibition mechanism in which a representative of a series of non-nucleosidic viral polymerase inhibitors, the 4-oxo-dihydroquinolines, binds at the polymerase active site interacting non-covalently with both the polymerase and the DNA duplex.  相似文献   

15.
During herpes simplex virus infection, expression of the viral DNA polymerase (pol) gene is regulated temporally as an early (beta) gene and is additionally down-regulated at late times at the level of translation (D. R. Yager, A. I. Marcy, and D. M. Coen, J. Virol. 64:2217-2225, 1990). To examine the role of viral DNA synthesis in pol regulation, we studied pol expression during infections in which viral DNA synthesis was blocked, either by using drugs that inhibit Pol or ribonucleotide reductase or by using viral mutants with lesions in either the pol or a primase-helicase subunit gene. Under any of these conditions, the level of cytoplasmic pol mRNA was reduced. This reduction was first seen at approximately the time DNA synthesis begins and, when normalized to levels of other early mRNAs, became as great as 20-fold late in infection. The reduction was also observed in the absence of the adjacent origin of replication, oriL. Thus, although pol mRNA accumulated as expected for an early gene in terms of temporal regulation, it behaved more like that of a late (gamma) gene in its response to DNA synthesis inhibition. Surprisingly, despite the marked decrease in pol mRNA in the absence of DNA synthesis, the accumulation of Pol polypeptide was unaffected. This was accompanied by loss of the normal down-regulation of translation of pol mRNA at late times. We suggest a model to explain these findings.  相似文献   

16.
17.
The genome structures of herpes simplex virus type 1 (HSV-1)/HSV-2 intertypic recombinants have been previously determined by restriction endonuclease analysis, and these recombinants and their parental strains have been employed to demonstrate that mutations within the HSV DNA polymerase locus induce an altered HSV DNA polymerase activity, exhibiting resistance to three inhibitors of DNA polymerase. The viral DNA polymerases induced by two recombinants and their parental strains were purified and shown to possess similar molecular weights (142,000 to 144,000) and similar sensitivity to compounds which distinguish viral and cellular DNA polymerases. The HSV DNA polymerases induced by the resistant recombinant and the resistant parental strain were resistant to inhibition by phosphonoacetic acid, acycloguanosine triphosphate, and the 2',3'-dideoxynucleoside triphosphates. The resistant recombinant (R6-34) induced as much acycloguanosine triphosphate as did the sensitive recombinant (R6-26), but viral DNA synthesis in infected cells and the viral DNA polymerase activity were not inhibited. The 2',3'-dideoxynucleoside-triphosphates were effective competitive inhibitors for the HSV DNA polymerase, and the Ki values for the four 2',3'-dideoxynucleoside triphosphates were determined for the four viral DNA polymerases. The polymerases of the resistant recombinant and the resistant parent possessed a much higher Ki for the 2',3'-dideoxynucleoside triphosphates and for phosphonoacetic acid than did the sensitive strains. A 1.3-kilobase-pair region of HSV-1 DNA within the HSV DNA polymerase locus contained mutations which conferred resistance to three DNA polymerase inhibitors. This region of DNA sequences encoded for an amino acid sequence of 42,000 molecular weight and defined an active center of the HSV DNA polymerase enzyme.  相似文献   

18.
Translational regulation of herpes simplex virus DNA polymerase.   总被引:2,自引:0,他引:2       下载免费PDF全文
D R Yager  A I Marcy    D M Coen 《Journal of virology》1990,64(5):2217-2225
  相似文献   

19.
Fourteen mutants known or likely to contain mutations in the herpes simplex virus DNA polymerase gene were examined for their sensitivity to aphidicolin in plaque reduction assays. Eleven of these exhibited some degree of hypersensitivity to the drug; altered aphidicolin-sensitivity correlated with altered sensitivity to the pyrophosphate analog, phosphonoacetic acid. The DNA polymerase specified by one of these mutants, PAAr5, required roughly seven-fold less aphidicolin to inhibit its activity by 50% than did polymerase specified by its parental strain. Mutations responsible for the aphidicolin-hypersensitivity phenotype of PAAr5 were mapped to an 0.8 kbp region in the herpes simplex virus DNA polymerase locus. These data taken together indicate that 1) mutations in the herpes simplex virus DNA polymerase gene can confer altered sensitivity to aphidicolin, 2) that the HSV polymerase is sensitive to aphidicolin in vivo, and 3) that amino acid alterations which affect aphidicolin binding may affect the pyrophosphate exchange-release site as well, suggesting that aphidicolin binds in close proximity to this site.  相似文献   

20.
The herpes simplex virus type 1 (HSV-1) (strain 17) DNA polymerase gene has been cloned into an Escherichia coli-yeast shuttle vector fused to the galactokinase gene (GAL-1) promoter. Genes controlled by the GAL-1 promoter are induced by galactose, uninduced by raffinose, and repressed by glucose. Cell extracts from a strain of Saccharomyces cerevisiae harboring this vector (Y-MH202, expresser cells) grown in the presence of galactose and assayed in high salt (100 mM ammonium sulfate) contained a novel DNA polymerase activity. No significant high-salt DNA polymerase activity was detected in extracts from expresser cells grown in the presence of raffinose or in extracts from control cells containing the E. coli-yeast shuttle vector without the HSV-1 DNA polymerase gene grown in the presence of raffinose of galactose. Immunoblot analysis of the cell extracts by using a polyclonal rabbit antiserum prepared against a highly purified HSV-1 DNA polymerase preparation revealed the specific induction of the HSV-1 approximately 140-kilodalton DNA polymerase polypeptide in expresser cells grown in galactose. Extracts from the same cells grown in raffinose or control cells grown in either raffinose or galactose did not contain this immunoreactive polypeptide. The high-salt DNA polymerase activity in the extracts from expresser cells grown in galactose was inhibited greater than 90% by either acyclovir triphosphate or aphidicolin, as expected for HSV-1 DNA polymerase. In addition, the high-salt polymerase enzyme activity could be depleted from extracts by immunoprecipitation by using purified immunoglobulin G from this same polyclonal rabbit antiserum. These results demonstrate the successful expression of functional HSV-1 DNA polymerase enzyme in S. cerevisiae.  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