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 10 毫秒
1.
2.
We compared the levels of gene expression obtained after herpes simplex virus (HSV) superinfection of cell lines containing integrated human beta-interferon (IFN) or chloramphenicol acetyltransferase (CAT) genes under the control of HSV immediate-early (IE) or delayed-early class promoters. DNA-transfected mouse Ltk+ cell lines harboring coselected IE175-IFN or thymidine kinase (TK)-IFN hybrid genes gave only low basal expression of human IFN. However, infection of both cell types with HSV type 1 or HSV type 2 produced abundant synthesis of IFN-specific RNA and biologically active IFN protein product. The IE175-IFN cell lines consistently gave 20- to 150-fold increases in IFN titers, and several TK-IFN cell lines yielded 100- to 500-fold induction. In the IE175-IFN cells, expression of IFN RNA also increased up to 200-fold and was detectable within 30 to 60 min after virus infection. Qualitatively similar results were obtained with hybrid G418-resistant Ltk- or Vero cell lines containing coselected IE175-CAT and TK-CAT constructs, except that there was relatively high basal expression of IE175-CAT. All three sets of IE cell lines (but not the delayed-early cell lines) responded to virus infection both in the presence of cycloheximide and with mutants defective in IE gene expression, demonstrating specific trans-activation by the pre-IE virion factor. In contrast, activation in the TK hybrid cell types required viral gene expression and the presence of a functional IE175 gene product. Up to 30-fold amplification in the copy number of the resident IFN or CAT DNA sequences also occurred within 20 h after HSV infection in IE175 hybrid cells but not in TK hybrid cells. Amplification was abolished either by treatment with phosphonacetate or by superinfection with a ts mutant unable to synthesize viral DNA, demonstrating specific HSV activation of the viral DNA replication origin (oriS) present in the IE hybrid constructs.  相似文献   

3.
Null mutants and attenuated mutants of herpes simplex virus (HSV) have been shown to induce immunity against challenge from wild-type virus. Null viruses with a defect in late gene products would be expected to express more viral genes than viruses with defects in essential early gene products and thus induce a better immune response. Herpesviruses encode a late gene product (serine protease) that is autocatalytic and cleaves the capsid assembly protein during viral replication. To determine whether a virus with a mutation in this gene could induce immunity, we constructed a recombinant virus containing the gusA reporter gene in the protease domain of the HSV type 1 UL26 open reading frame (ORF). Consistent with previous results (M. Gao, L. Matusick-Kumar, W. Hurlburt, S. F. DiTusa, W. W. Newcomb, J. C. Brown, P. J. McCann, I. Deckman, and R. J. Colonno, J. Virol. 68:3702-3712, 1994), recombinant virus could be isolated only from helper cell lines expressing the product of the UL26 ORF. Mice inoculated with the recombinant virus were unaffected by doses of virus that were lethal to mice infected with wild-type virus. Mice which were previously inoculated with the recombinant virus were also protected by a subsequent challenge with wild-type virus in a dose-dependent manner. These results indicate that recombinant viruses lacking the protease gene are avirulent but render protection from subsequent challenge.  相似文献   

4.
The herpes simplex virus type 1 (HSV-1) immediate-early protein ICP27 is an essential regulatory protein which is required for virus replication. Transfection experiments have demonstrated that ICP27 along with the HSV-1 transactivators ICP4 and ICP0 can positively regulate the expression of some late HSV-1 target plasmids and can negatively regulate the expression of some immediate-early and early target plasmids. We previously showed that mutants defective in the activation of a late target plasmid mapped to the carboxy-terminal half of the protein, whereas mutants defective in the repression of an early target plasmid mapped within the C-terminal 78 amino acids of ICP27 (M. A. Hardwicke, P. J. Vaughan, R. E. Sekulovich, R. O'Conner, and R. M. Sandri-Goldin, J. Virol. 63:4590-4602, 1989). In this study, we cotransfected ICP27 activator and repressor mutants along with wild-type ICP27 plasmid to determine whether these mutants could interfere with the wild-type activities. Mutants which were defective only in the activation function were dominant to the wild-type protein and inhibited the activation of the late target plasmid pVP5-CAT, whereas mutants defective in the repressor function did not inhibit either the activation of pVP5-CAT or the repression of the early target plasmid pTK-CAT. Furthermore, cell lines which stably carried three different activator mutants were impaired in their ability to support the growth of wild-type HSV-1 strain KOS, resulting in virus yields 5- to 40-fold lower than in control cells. The defect in virus replication appeared to stem from a decrease in the expression of HSV-1 late gene products during infection as measured by steady-state mRNA levels and by immunoprecipitation analysis of specific polypeptides. These results indicate that ICP27 activator mutations specifically interfere with the activation function of the protein both in transfection and during infection. Moreover, these results suggest that the repressor region may be important for binding of the polypeptide, since mutations in this region did not interfere with the activities of wild-type ICP27 and therefore presumably could not compete for binding.  相似文献   

5.
Three different recombinant plasmids containing the entire 15-kilobase L and S inverted repeat sequence of herpes simplex virus type 2 DNA have been introduced into cultured Ltk- or BSC cells by both the calcium and DEAE-dextran transfection procedures. In each case, after 24 h approximately 1% of the cells gave strongly positive nuclear staining when assayed by immunofluorescence with hyperimmune antisera made against early and immediate-early infected-cell polypeptides. The nuclear fluorescence pattern and intensity mimicked that observed within 2 to 3 h after infection of Ltk- cells with either herpes simplex virus type 1 or type 2 wild-type virus. Herpes simplex virus type 1 (KOStsB2)-infected Ltk- cells under nonpermissive conditions did not express these antigens in the nucleus. Therefore, we conclude that either one or both of the 185,000- and 110,000-molecular-weight immediate early proteins, or some other as yet unknown gene product encoded entirely within the inverted repeats, can be transiently expressed in large amounts in transfected cells in the absence of other viral genes or accompanying virion components. Permanent mouse cell lines derived from transfection with these plasmids by using the thymidine kinase coselection procedure did not express sufficient nuclear antigen to be detectable by immunofluorescence.  相似文献   

6.
Gene 12 of equine herpesvirus 1 (EHV-1), the homolog of herpes simplex virus (HSV) VP16 (alpha TIF, Vmw65), was cloned into a eukaryotic expression vector by PCR and used in transactivation studies of both the EHV-1 and HSV-1 IE1 promoters. Results demonstrated that the product of gene 12 is a potent transactivator of immediate-early gene expression of both viruses, which requires sequences in the upstream HSV-1 promoter for activity. Mutational analysis of the gene 12 open reading frame indicated that removal of the C-terminal 7 amino acids, which contain a short region of homology with the extreme C terminus of VP16, inactivated the protein. Within this region, only a single methionine residue appeared to be essential for activity, implying that gene 12 may have a modular array of organization similar to that of VP16. However, fusion of the gene 12 C terminus to a truncated form of VP16, which contained the complex formation domain, did not restore activity to the HSV-1 protein. These data demonstrate that the EHV-1 immediate-early transactivator may not be functionally colinear with VP16, with transactivation requiring both the C terminus and another region(s) present within the N-terminal portion.  相似文献   

7.
8.
The herpes simplex virus type 1 (HSV-1) mutant d109 does not express any of the immediate-early (IE) proteins and persists in cells for a prolonged length of time. As has been shown by Nicholl et al. (J. Gen. Virol. 81:2215-2218, 2000) and Mossman et al. (J. Virol. 75:750-758, 2001) using other mutants defective for IE gene expression, infection with d109 induced the expression of a number of interferon-stimulated genes. Induction of these genes was significantly greater at multiplicities of infection (MOI) of 10 PFU/cell or greater, and the resulting antiviral effect was only seen at MOIs greater than 10 PFU/cell. Using mutants defective for sets of IE genes established that the lack of ICP0 expression was necessary for high levels of interferon-stimulated gene expression in HEL cells. The induction of interferon-stimulated genes by d109 could also be inhibited by infection with an E1-:E3-:E4- adenovirus expressing levels of ICP0 that are comparable to those expressed within the first hour of wild-type virus infection. Lastly, the addition of the proteasome inhibitor MG132 to cells infected with a mutant that expresses ICP0, d106, also resulted in the induction of interferon-stimulated genes. Thus, ICP0 may function through the proteasome very early in HSV infection to inhibit a cellular antiviral response induced by the virion.  相似文献   

9.
10.
11.
F Yao  P A Schaffer 《Journal of virology》1994,68(12):8158-8168
The herpes simplex virus type 1 immediate-early protein ICP0 enhances expression of a spectrum of viral genes alone and synergistically with ICP4. To test whether ICP0 and ICP4 interact physically, we performed far-Western blotting analysis of proteins from mock-, wild-type-, and ICP4 mutant virus-infected cells with in vitro-synthesized [35S]Met-labeled ICP0 and ICP4 as probes. The ICP4 and ICP0 polypeptides synthesized in vitro exhibited molecular weights similar to those of their counterparts in herpes simplex virus type 1-infected cells, and the in vitro-synthesized ICP4 was able to bind to a probe containing the ICP4 consensus binding site. Far-Western blotting experiments demonstrated that ICP0 interacts directly and specifically with ICP4 and with itself. To further define the interaction between ICP0 and ICP4, we generated a set of glutathione S-transferase (GST)-ICP0 fusion proteins that contain GST and either ICP0 N-terminal amino acids 1 to 244 or 1 to 394 or C-terminal amino acids 395 to 616 or 395 to 775. Using GST-ICP0 fusion protein affinity chromatography and in vitro-synthesized [35S]Met-labeled ICP0 and ICP4, ICP4 was shown to interact preferentially with the fusion protein containing ICP0 C-terminal amino acids 395 to 775, whereas ICP0 interacted efficiently with both the N-terminal GST-ICP0 fusion proteins and the C-terminal GST-ICP0 fusion proteins containing amino acids 395 to 775. Fusion protein affinity chromatography also demonstrated that the C-terminal 235 amino acid residues of ICP4 are important for efficient interaction with ICP0. Collectively, these results reveal a direct and specific physical interaction between ICP0 and ICP4.  相似文献   

12.
A detailed analysis of the expression of the bacterial chloramphenicol acetyltransferase gene controlled by the herpes simplex virus major capsid protein (VP5) promoter showed that this promoter can be functionally separated into an 80-base core region, which has the minimal information required to serve as a pol II promoter but which is not fully activated by viral superinfection or by cotransfections with plasmids bearing functional alpha (immediate-early) genes, and an approximately 100-base regulatory region upstream of the core, which allowed full induction of VP5 promoter-driven chloramphenicol acetyltransferase activity but which repressed the ability of the VP5 core promoter to be cis activated by the simian virus 40 enhancer. This was in distinct contrast to the situation with the alkaline exonuclease promoter (a model early promoter) and defined the regions of this promoter which can be used to study the interaction between viral promoters and putative regulatory proteins induced by viral infection.  相似文献   

13.
14.
Regulation of cellular genes transduced by herpes simplex virus.   总被引:12,自引:10,他引:2       下载免费PDF全文
Previous studies demonstrated that the rabbit beta-globin gene is transcribed from its own promoter and regulated as a herpes simplex virus (HSV) early gene following insertion into the early HSV thymidine kinase gene in the intact viral genome (J. R. Smiley, C. Smibert, and R. D. Everrett, J. Virol. 61:2368-2377, 1987). We report here that the beta-globin promoter remained under early control after insertion into the late HSV gene encoding glycoprotein C. On the basis of these findings, we concluded that the beta-globin promoter is functionally equivalent to an HSV early-control region. We found that a transduced human alpha-globin gene was also regulated as an early HSV gene, while two linked Alu elements mimicked the behavior of HSV late genes. These results demonstrate that certain aspects of HSV temporal regulation can be duplicated by cellular elements and provide strong support for the hypothesis that the regulation of HSV gene expression can occur through mechanisms that do not rely on recognition of virus-specific temporal control signals.  相似文献   

15.
16.
ICP0, a herpes simplex virus immediate-early gene product, is a highly phosphorylated nuclear protein that is a potent activator of virus and host genes. Using biochemical and genetic assays employing plasmids encoding mutant forms of ICP0 and a recombinant adenovirus that expresses ICP0, we mutant forms of ICP0 and a recombinant adenovirus that expresses ICP0, we provide evidence that the protein multimerizes. Some mutant forms of ICP0 were transdominant and interfered with activation of a target reporter gene or with complementation of an ICP0-minus virus.  相似文献   

17.
The varicella-zoster virus (VZV) open reading frame 10 (ORF10) protein is the homolog of the herpes simplex virus type 1 (HSV-1) protein VP16. These are two virion tegument proteins that have extensive amino acid sequence identity in their amino-terminal and middle domains. ORF10, however, lacks the acidic carboxy terminus which is critical for transactivation by VP16. Earlier studies showed that VZV ORF10 does not form a tertiary complex with the TAATGARAT regulatory element (where R is a purine) with which HSV-1 VP16 interacts, suggesting that ORF10 may not have transactivating ability. Using transient-expression assays, we show that VZV ORF10 is able to transactivate VZV immediate-early (IE) gene (ORF62) and HSV-1 IE gene (ICP4 and ICP0) promoters. Furthermore, cell lines stably expressing ORF10 complement the HSV-1 mutant in1814, which lacks the transactivating function of VP16, and enhance the de novo synthesis of infectious virus following transfection of HSV-1 virion DNA. These results indicate that ORF10, like its HSV-1 homolog VP16, is a transactivating protein despite the absence of sequences similar to the VP16 carboxy-terminal domain. The transactivating function of the VZV ORF10 tegument protein may be critical for efficient initiation of viral infection.  相似文献   

18.
We showed that the expression of a single protein, glycoprotein D (gD-1), specified by herpes simplex virus type 1 (HSV-1) renders cells resistant to infection by HSV but not to infection by other viruses. Mouse (LMtk-) and human (HEp-2) cell lines containing the gene for gD-1 under control of the human metallothionein promoter II expressed various levels of gD-1 constitutively and could be induced to express higher levels with heavy metal ions. Radiolabeled viruses bound equally well to gD-1-expressing and control cell lines. Adsorbed viruses were unable to penetrate cells expressing sufficient levels of gD-1, based on lack of any cytopathic effects of the challenge virus and on failure to detect either the induction of viral protein synthesis or the shutoff of host protein synthesis normally mediated by a virion-associated factor. The resistance to HSV infection conferred by gD-1 expression was not absolute and depended on several variables, including the amount of gD-1 expressed, the dosage of the challenge virus, the serotype of the challenge virus, and the properties of the cells themselves. The interference activity of gD-1 is discussed in relation to the role of gD-1 in virion infectivity and its possible role in permitting escape of progeny HSV from infected cells.  相似文献   

19.
20.
Although NK cells can kill both malignant cells and virus-infected cells without prior sensitization, it has remained unclear whether the mechanism by which an NK cell is activated in the presence of a tumor cell is similar to that induced by the presence of a virus-infected cell. In our experimental system using homogeneous populations of cloned human CD16+ NK cells, we found that HSV-infected target cells do not induce in the NK cells the same pharmacologically-active second messengers elicited by NK-sensitive tumor cells. Although phosphoinositide turnover and calcium signaling were generated in NK cells exposed to NK-sensitive tumor cells, the recognition of HSV-infected cells by NK cells did not result in similar transmembrane signaling. Furthermore, depending on the cell type infected by HSV, alternative mechanisms of cytotoxicity were employed. HSV-infected foreskin fibroblasts were rapidly and selectively killed by cloned NK cells without a requirement for IFN or accessory cells. In contrast to this direct cytotoxicity against HSV-infected foreskin fibroblasts, NK cell-mediated cytotoxicity against an HSV-infected fibrosarcoma cell line (1591) was dependent on IFN-alpha production by accessory cells. Importantly, in both systems of cytotoxicity, IFN-alpha activation of NK cells resulted in augmented killing against both infected and uninfected targets. These results suggest that NK cell activation induced during antiviral immunity is distinct from activation elicited during an antitumor response. These differences include the utilization of alternative forms of signal transduction and alternative mechanisms of cytotoxicity.  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