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.
Translation of in vitro-synthesized herpes simplex virus type 2 (HSV-2) gG-2 mRNA in a reticulocyte lysate system was used to study the processing of HSV-2 gG-2. In the presence of canine pancreatic microsomal membranes, a single species that is protected from trypsin digestion was detected. This product comigrates with the 104,000-Mr (104K) high mannose intermediate seen in HSV-2-infected-cell lysates. Endo-beta-N-acetylglucosaminidase H treatment of the in vitro-synthesized 104K protein yielded a single product migrating at 100 K. The 72K and 31K cleavage products of gG-2 were not observed in the in vitro system. These data show that the molecular weight of the nonglycosylated form of the gG-2 protein is 100,000 and that the cotranslational processing of this protein in the endoplasmic reticulum yields the 104K high-mannose intermediate.  相似文献   

2.
During the synthesis of glycoprotein G-2 (gG-2) of herpes simplex virus type 2, the 104,000-Da gG-2 precursor (104K precursor) is cleaved to generate the 72K and the 31K intermediates. The 72K product is processed to generate the mature gG-2 (molecular mass, 108,000 Da), while the 31K product is additionally processed and secreted into the extracellular medium as the 34K component (H. K. Su, R. Eberle, and R. J. Courtney, J. Virol. 61:1735-1737, 1987). In this study, the orientations of the 31K and 72K products on the 104K precursor were determined by using two antipeptide sera produced in rabbits and a monoclonal antibody, 13 alpha C6, directed against gG-2. The sera prepared against synthetic peptides corresponding to the terminal amino acid residues 67 to 78 and an internal peptide at amino acids 247 to 260 of gG-2 recognized the 104K precursor and the 31K cleavage product but not the 72K intermediate. In contrast, 13 alpha C6 detected the 72K cleavage product and the uncleaved precursor but not the 31K cleavage component. The epitope recognized by 13 alpha C6 was mapped within amino acids 486 to 566. These results suggest that the 31K cleavage product is derived from the amino-terminal portion of the 104K precursor molecule and that the 72K intermediate is derived from the carboxyl terminus. In support of our model described above for the synthesis of gG-2, antibodies recognizing either of the cleavage products reacted with the uncleaved precursor but not with the other cleavage product. By using partial endo-beta-N-acetylglucosaminidase H analysis, two N-linked glycosylation sites were found on each of the cleavage products. The distribution of the N-linked glycosylation sites and the reactivities of the antipeptide sera allowed the cleavage region on the precursor to be mapped to within amino acids 260 to 437.  相似文献   

3.
The gG-2 glycoprotein gene of herpes simplex virus type 2 (HSV-2) was cloned into the mammalian expression vector pMSG under the control of the inducible mouse mammary tumor virus promoter. Transfection of this cloned gG-2 construct into NIH 3T3 cells resulted in the stable expression of gG-2 upon induction with dexamethasone. In addition, the 104,000-molecular-weight (104K) and 72K gG-2 precursors as well as the 34K secreted component were generated in the transformed cells. The synthesis of gG-2 in these transformed cells appeared to follow the same cleavage-processing pathway as gG-2 synthesis during an HSV-2 infection. These results indicate that the processing of gG-2 can occur in the absence of an HSV-2 infection.  相似文献   

4.
The gene (US4) coding for herpes simplex virus type 2 (HSV-2) glycoprotein G (gG-2) was cloned and constitutively expressed in Chinese hamster ovary (CHO) cells. The expression vector containing the dihydrofolate reductase (dhfr) gene, and the HSV-2 US4 gene under the control of the Simian virus 40 early promoter (SV40 EP), was transfected into dhfr-deficient CHO cells. The transfected cells were selected and amplified using methotrexate (MTX). To demonstrate that the gG-2 produced in these transformed cells had antigenic determinants in common with the native glycoprotein, CHO cells expressing gG-2 were used in an immunofluorescent assay (IFA) for the detection of HSV-2 type-specific antibodies in human serum samples. Seven of eight serum samples from adults with prior episodes of culture proven HSV-2 infections were found to be positive by the IFA method whereas none of seven serum samples from young children with culture documented HSV-1 infections were positive by IFA. Thus the recombinant CHO : gG-2 cells have diagnostic utility in an HSV-2 specific serologic assay.  相似文献   

5.
M Suh 《Journal of virology》1982,41(3):1095-1098
Transformation of hamster embryo cells by herpes simplex virus stimulated the production of a 35-kilodalton (35K) protein that was specifically immunoprecipitated, along with other polypeptides, by rabbit hyperimmune serum. This 35K polypeptide was further analyzed by partial digestion with Staphylococcus aureus V8 protease in parallel with a 35K polypeptide from herpes simplex virus type 2-infected cells. These polypeptides had almost identical partial-proteolytic cleavage maps, indicating that they are probably the same or that they are very similar polypeptides.  相似文献   

6.
Evidence is presented that the herpes simplex virus type 2 glycoprotein previously designated gF is antigenically related to herpes simplex virus type 1 gC (gC-1). An antiserum prepared against type 1 virion envelope proteins immunoprecipitated gF of type 2 (gF-2), and competition experiments revealed that the anti-gC-1 component of the antiserum was responsible for the anti-gF-2 cross-reactivity. An antiserum prepared against fully denatured purified gF-2, however, and three anti-gF-2 monoclonal antibodies failed to precipitate any type 1 antigen, indicating that the extent of cross-reactivity between gC-1 and gF-2 may be limited. Several aspects of gF-2 synthesis and processing were investigated. Use of the enzymes endo-beta-N-acetylglucosaminidase H and alpha-D-N-acetylgalactosaminyl oligosaccharidase revealed that the fully processed form of gF-2 (about 75,000 [75K] apparent molecular weight) had both complex-type N-linked and O-linked oligosaccharides, whereas newly synthesized forms (67K and 69K) had only high-mannose N-linked oligosaccharides. These last two forms were both reduced in size to 54K by treatment with endo-beta-N-acetylglucosaminidase H and therefore appear to differ only in the number of N-linked chains. Neutralization tests and radioiodination experiments revealed that gF-2 is exposed on the surfaces of virions and that the 75K form of gF-2 is exposed on cell surfaces. The similarities and differences of gF-2 and gC-1 are discussed in light of recent mapping results which suggest collinearity of their respective genes.  相似文献   

7.
Herpes simplex virus (HSV) codes for several envelope glycoproteins, including glycoprotein G-2 (gG-2) of HSV type 2 (HSV-2), which are dispensable for replication in cell culture. However, clinical isolates which are deficient in such proteins occur rarely. We describe here five clinical HSV-2 isolates which were found to be unreactive to a panel of anti-gG-2 monoclonal antibodies and therefore considered phenotypically gG-2 negative. These isolates were further examined for expression of the secreted amino-terminal and cell-associated carboxy-terminal portions of gG-2 by immunoblotting and radioimmunoprecipitation. The gG-2 gene was completely inactivated in four isolates, with no expression of the two protein products. For one isolate a normally produced secreted portion and a truncated carboxy-terminal portion of gG-2 were detected in virus-infected cell medium. Sequencing of the complete gG-2 gene identified a single insertion or deletion of guanine or cytosine nucleotides in all five strains, resulting in a premature termination codon. The frameshift mutations were localized within runs of five or more guanine or cytosine nucleotides and were dispersed throughout the gene. For the isolate for which a partially inactivated gG-2 gene was detected, the frameshift mutation was localized upstream of but adjacent to the nucleotides coding for the transmembranous region. Thus, this study demonstrates the existence of clinical HSV-2 isolates which do not express an envelope glycoprotein and identifies the underlying molecular mechanism to be a single frameshift mutation.  相似文献   

8.
Incubation of herpes simplex virus type 1-infected Vero and HEp-2 cells at a reduced temperature (34 degrees C) enhanced the detection of the nonglycosylated precursors (pgB97 and pgC75) to the gB and gC glycoproteins in the cytoplasmic and nuclear fractions. Relative to the fully glycosylated and high-mannose forms detected, the nonglycosylated precursors were the predominant components associated with the nuclear fraction of infected cells. Furthermore, addition of protease inhibitors to the fractionation buffers did not affect the distribution or abundance of the nonglycosylated precursors, suggesting that the presence of pgB97 and pgC75 was not the result of proteolysis. When infected Vero or HEp-2 cells were harvested at various times postinfection, the nonglycosylated precursors were detected after the initial appearance of the high mannose components (pgB110 and pgC105). In Vero cells, pgB97 and pgC75 were detected simultaneously at 8 h postinfection, whereas detection was not apparent in HEp-2 cells until 20 h postinfection. Conditions which favored detection of appreciable amounts of nonglycosylated precursors provided an unique approach to probe possible post-translational modifications in the absence of inhibitors of glycosylation. In nuclear fractions isolated from cycloheximide-treated HEp-2 or Vero cells, numerous discrete gC-immunoreactive bands migrating with decreased electrophoretic mobility relative to the nonglycosylated precursor pgC75 were observed. This series of one to four additional bands was eliminated by digestion with endoglycosidase H, and the appearance of these bands was blocked by the addition of tunicamycin. Collectively, the data suggest that high-mannose core oligosaccharides may be added to the nonglycosylated precursor of the gC glycoprotein of herpes simplex virus type 1 in a post-translational fashion.  相似文献   

9.
Routine serodiagnosis of herpes simplex virus (HSV) infections is currently performed using recombinant glycoprotein G (gG) antigens from herpes simplex virus 1 (HSV-1) and HSV-2. This is a single-antigen test and has only one diagnostic application. Relatively little is known about HSV antigenicity at the proteome-wide level, and the full potential of mining the antibody repertoire to identify antigens with other useful diagnostic properties and candidate vaccine antigens is yet to be realized. To this end we produced HSV-1 and -2 proteome microarrays in Escherichia coli and probed them against a panel of sera from patients serotyped using commercial gG-1 and gG-2 (gGs for HSV-1 and -2, respectively) enzyme-linked immunosorbent assays. We identified many reactive antigens in both HSV-1 and -2, some of which were type specific (i.e., recognized by HSV-1- or HSV-2-positive donors only) and others of which were nonspecific or cross-reactive (i.e., recognized by both HSV-1- and HSV-2-positive donors). Both membrane and nonmembrane virion proteins were antigenic, although type-specific antigens were enriched for membrane proteins, despite being expressed in E. coli.  相似文献   

10.
The ribonucleotide reductase (ribonucleoside-diphosphate reductase; EC 1.17.4.1) induced by herpes simplex virus type 2 infection of serum-starved BHK-21 cells was purified to provide a preparation practically free of both eucaryotic ribonucleotide reductase and contaminating enzymes that could significantly deplete the substrates. Certain key properties of the herpes simplex virus type 2 ribonucleotide reductase were examined to define the extent to which it resembled the herpes simplex virus type 1 ribonucleotide reductase. The herpes simplex virus type 2 ribonucleotide reductase was inhibited by ATP and MgCl2 but only weakly inhibited by the ATP X Mg complex. Deoxynucleoside triphosphates were at best only weak inhibitors of this enzyme. ADP was a competitive inhibitor (K'i, 11 microM) of CDP reduction (K'm, 0.5 microM), and CDP was a competitive inhibitor (K'i, 0.4 microM) of ADP reduction (K'm, 8 microM). These key properties closely resemble those observed for similarly purified herpes simplex virus type 1 ribonucleotide reductase and serve to distinguish these virally induced enzymes from other ribonucleotide reductases.  相似文献   

11.
The virion host shutoff (vhs) gene of herpes simplex virus encodes a virion polypeptide that induces degradation of host mRNAs at early times and rapid turnover of viral mRNAs throughout infection. To better investigate the vhs function, an in vitro mRNA degradation system was developed, consisting of cytoplasmic extracts from HeLa cells infected with wild-type herpes simplex virus type 1 or a mutant encoding a defective vhs polypeptide. Host and viral mRNAs were degraded rapidly in extracts from cells productively infected with wild-type herpes simplex virus type 1 but not in extracts from mock-infected cells or cells infected with the mutant vhs1. In contrast, 28S rRNA was stable in all three kinds of extract. Accelerated turnover of host mRNAs was also observed in extracts from cells infected with wild-type virus in the presence of dactinomycin, indicating that the activity was induced by a structural component of the infecting virions. The in vitro vhs activity was inactivated by heat or proteinase K digestion but was insensitive to brief treatment of the extracts with micrococcal nuclease. It was not inhibited by placental RNase inhibitor, it exhibited a strong dependence upon added Mg2+, it was active at concentrations of K+ up to 200 mM, and it did not require the components of an energy-generating system. In summary, the in vitro mRNA degradation system appears to accurately reproduce the vhs-mediated decay of host and viral mRNAs and should be useful for studies of the mechanism of vhs action.  相似文献   

12.
Bcl-2 oncogene expression plays a role in the establishment of persistent viral infection by blocking virus-induced apoptosis. This might be achieved by preventing virus-induced activation of caspase-3, an IL-1beta-converting enzyme (ICE)-like cysteine protease that has been implicated in the death effector phase of apoptosis. Contrary to this model, we show that three cell types highly overexpressing functional Bcl-2 displayed caspase-3 activation and underwent apoptosis in response to infection with alphaviruses Semliki Forest and Sindbis as efficiently as vector control counterparts. In all three cell types, overexpressed 26 kDa Bcl-2 was cleaved into a 23 kDa protein. Antibody epitope mapping revealed that cleavage occurred at one or two target sites for caspases within the amino acid region YEWD31 (downward arrow) AGD34 (downward arrow) A, removing the N-terminal BH4 region known to be essential for the death-protective activity of Bcl-2. Preincubation of cells with the caspase inhibitor Z-VAD prevented Bcl-2 cleavage and partially restored the protective activity of Bcl-2 against virus-induced apoptosis. Moreover, a murine Bcl-2 mutant having Asp31, Asp34 and Asp36 substituted by Glu was resistant to proteolytic cleavage and abrogated apoptosis following virus infection. These findings indicate that alphaviruses can trigger a caspase-mediated inactivation of Bcl-2 in order to evade the death protection imposed by this survival factor.  相似文献   

13.
The metabolism and mode of action of the anti-herpes compound buciclovir [R)-9-(3,4-dihydroxybutyl)-guanine, BCV) has been studied in herpes simplex virus-infected and uninfected Vero cells. In uninfected cells, a low and constant concentration of intracellular BCV was found, while in herpes simplex virus-infected cells, an increasing concentration of BCV phosphates was found due to metabolic trapping. The major phosphorylation product was BCV triphosphate (BCVTP) which was 92% of the total amount of BCV phosphates. BCV phosphates were accumulated to the same extent in cells infected with either a herpes simplex virus type 1 or a herpes simplex virus type 2 strain while thymidine kinase-deficient mutants of herpes simplex virus type 1 were 10 times less efficient in accumulating BCV phosphates. In uninfected Vero cells, the concentration of the phosphorylated forms of BCV was less than 1% of that found in herpes simplex virus-infected cells. The BCVTP formed in herpes simplex virus-infected cells was highly stable, as 80% of the amount of BCVTP was still present even 17 h after removal of extracellular BCV. BCV was a good substrate for herpes simplex virus type 1- and type 2-induced thymidine kinases but not for the cellular cytosol or mitochondrial thymidine kinases. BCV monophosphate could be phosphorylated by cellular guanylate kinase to BCV diphosphate. BCVTP was a selective and competitive inhibitor to deoxyguanosine triphosphate of the purified herpes simplex virus type 1- and type 2-induced DNA polymerases. BCVTP could neither act as an alternative substrate in the herpes simplex virus type 2 or cellular DNA polymerase reactions, nor could [3H]BCV monophosphate be detected in DNA formed by herpes simplex virus type 2 DNA polymerase, or be detected in nucleic acids extracted from herpes simplex virus type 1-infected cells. These data indicate that BCVTP may inhibit the herpes simplex virus-induced DNA polymerase without being incorporated into DNA.  相似文献   

14.
Monoclonal antibody 13 alpha C5-1-A11 immunoprecipitated two major polypeptides of molecular weights 108,000 and 120,000 from extracts of herpes simplex virus type 2-infected BHK-21 cells labeled with [35S]methionine or [3H]glucosamine. In pulse-chase experiments, both labels were chased from the 120,000-molecular-weight peptide (120K peptide) into the 108K molecule. Endoglycosidase H (endo H) reduced the 120K peptide to a 112K peptide but did not affect the 108K peptide. Similar profiles were obtained with monoclonal antibody AP-1 which reacts with a 92K glycoprotein, gG, which maps to the short unique region of the genome. Cross-absorption experiments indicated that both antibodies reacted with the same peptides, suggesting that the 120K peptide is a partially glycosylated high-mannose-type precursor of gG (pgG1). Immunoprecipitation from monensin-treated cells indicated that pgG1(120K) may undergo peptide cleavage to form a 74K high-mannose-type peptide (pgG2) and that this 74K peptide may be further processed into an endo H-resistant 110K to 116K peptide. In the presence of tunicamycin, gG(108K) was replaced by 110K and 105K peptides which were resistant to both endo H and endoglycosidase F. The 105K peptide was the only molecule labeled by [3H]galactose or [3H]glucosamine in the presence of tunicamycin, and none of the peptides were labeled with [3H]mannose, indicating the probable presence of O-linked sugars in the 105K peptide. Our results imply that cotranslational glycosylation of the unglycosylated precursor 110K peptide results in the high-mannose-type pgG1(120K), which probably undergoes peptide cleavage. This putative cleavage product may then mature into gG (108K) by the trimming of sugars and the addition of complex and probably O-linked sugars; the high-mannose-type pgG2(74K) is probably an intermediate peptide formed in this process.  相似文献   

15.
Although cell surface chondroitin sulfate (CS) is regarded as an auxiliary receptor for binding of herpes simplex virus to cells, and purified CS chain types A, B, and C are known to interfere poorly or not at all with the virus infection of cells, we have found that CS type E (CS-E), derived from squid cartilage, exhibited potent antiviral activity. The IC(50) values ranged from 0.06 to 0.2 mug/ml and substantially exceeded the antiviral potency of heparin, the known inhibitor of virus binding to cells. Furthermore, in mutant gro2C cells that express CS but not heparan sulfate, CS-E showed unusually high anti-herpes virus activity with IC(50) values of <1 ng/ml. Enzymatic degradation of CS-E with chondroitinase ABC abolished its antiviral activity. CS-E inhibited the binding to cells of the purified virus attachment protein gC. A direct interaction of gC with immobilized CS-E and inhibition of this binding by CS-E oligosaccharide fragments greater than octasaccharide were demonstrated. Likewise, the gro2C-specific CS chains interfered with the binding of viral gC to these cells and were found to contain a considerable proportion (13%) of the E-disaccharide unit, suggesting that this unit is an essential component of the CS receptor for herpes simplex virus on gro2C cells and that the antiviral activity of CS-E was due to interference with the binding of viral gC to a CS-E-like receptor on the cell surface. Knowledge of the determinants of antiviral properties of CS-E will help in the development of inhibitors of herpes simplex virus infections in humans.  相似文献   

16.
To facilitate detection of glycoprotein K (gK) specified by herpes simplex virus, a 12-amino-acid epitope tag was inserted within gK domain III. Recombinant virus gKprotC-DIII, expressing the tagged gK, was isolated. This virus formed wild-type plaques and replicated as efficiently as the wild-type KOS virus in Vero cells. Anti-protein C MAb detected high-mannose and Golgi complex-dependent glycosylated gK within cells as well as on purified virions. The gK-null virus DeltagK (gK(-/-)) entered Vero cells substantially more slowly than the wild-type KOS (gK(+/+)), while DeltagK virus grown in complementing VK302 cells (gK(-/+)) entered with entry kinetics similar to those of the KOS virus.  相似文献   

17.
Chemical and enzymatic methods were employed for the synthesis of the title compound, 2'F-Guo 7. High antiviral activity of 2'F-Guo was established in chick embryo cells infected with influenza virus FPV/Rostock/34 (H7N1) and herpes simplex virus (HSV) type I (1C strain).  相似文献   

18.
Previous results have indicated that the herpes simplex virus 1 UL31 and UL34 proteins interact and form a complex at the inner nuclear membranes of infected cells, where both play important roles in the envelopment of nucleocapsids at the inner nuclear membrane. In the work described here, mapping studies using glutathione S-transferase pull-down assays indicated that amino acids 137 to 181 of the UL34 protein are sufficient to mediate an interaction with the UL31 protein. A recombinant virus (v3480) lacking UL34 codons 138 to 181 was constructed. Similar to a UL34 null virus, v3480 failed to replicate on Vero cells and grew to a limited extent on rabbit skin cells. A UL34-expressing cell line restored v3480 growth and plaque formation. Similar to the localization of UL31 protein in cells infected with a UL34 null virus, the UL31 protein was present in the nuclei of Hep2 cells infected with v3480. Hep2 cells infected with v3480 contained the UL34 protein in the cytoplasm, the nucleus, and the nuclear membrane, and this was noted to be similar to the appearance of cells infected with a UL31 null virus. In transient expression assays, the interaction between UL34 amino acids 137 to 181 and the UL31 protein was sufficiently robust to target green fluorescent protein and emerin to intranuclear sites that contained the UL31 protein. These data indicate that amino acids 137 to 181 of the UL34 protein are (i) sufficient to mediate interactions with the UL31 protein in vitro and in vivo, (ii) necessary for the colocalization of UL31 and UL34 in infected cells, and (iii) essential for normal viral replication.  相似文献   

19.
High-mannose type N-linked glycan with 6 mannosyl residues, termed "M6Gn2", displayed clear binding to the same M6Gn2, conjugated with ceramide mimetic (cer-m) and incorporated in liposome, or coated on polystyrene plates. However, the conjugate of M6Gn2-cer-m did not interact with complex-type N-linked glycan with various structures having multiple GlcNAc termini, conjugated with cer-m. The following observations indicate that hamster embryonic fibroblast NIL-2 K cells display homotypic autoadhesion, mediated through the self-recognition capability of high-mannose type glycans expressed on these cells: (i) NIL-2 K cells display clear binding to lectins capable of binding to high-mannose type glycans (e.g., ConA), but not to other lectins capable of binding to other carbohydrates (e.g. GS-II). (ii) NIL-2 K cells adhere strongly to plates coated with M6Gn2-cer-m, but not to plates coated with complex-type N-linked glycans having multiple GlcNAc termini, conjugated with cer-m; (iii) degree of NIL-2 K cell adhesion to plates coated with M6Gn2-cer-m showed a clear dose-dependence on the amount of M6Gn2-cer-m; and (iv) the degree of NIL-2 K adhesion to plates coated with M6Gn2-cer-m was inhibited in a dose-dependent manner by α1,4-L-mannonolactone, the specific inhibitor in high-mannose type glycans addition. These data indicate that adhesion of NIL-2 K is mediated by self-aggregation of high mannose type glycan. Further studies are to be addressed on auto-adhesion of other types of cells based on self interaction of high mannose type glycans.  相似文献   

20.
将伪狂犬病病毒TK^-/gG^-/LacZ^ 突变株的基因组DNA与含有缺失的gG基因的转移质粒pUSKBB共转染猪肾传代细胞PK-15,待完全病变后收获病毒进行空斑试验,用PCR筛选gG缺失的重组病毒。空斑纯化3次后,随机挑取空斑进行PCR扩增,证实所获得的病毒为均一的TK^-/gG^-缺失株。遗传稳定性试验表明该重组病毒能在PK-15细胞上稳定遗传,动物试验表明该缺失株对Balb/c小鼠极为安全且能保护Balb/c小鼠抵抗致死量PRV强毒的攻击。该突变株的获得为我国伪狂犬病的控制和根除奠定了基础。  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