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.
The ICP34.5 protein of herpes simplex virus (HSV) is involved in many aspects of viral pathogenesis; promoting neurovirulence, inhibiting interferon-induced shutoff of protein synthesis, interacting with PCNA and TBK1, inhibiting dendritic cell (DC) maturation, and binding to Beclin 1 to interfere with autophagy. Because of its key role in neuropathogenicity, the γ34.5 gene is deleted in all oncolytic HSVs (oHSVs) currently in clinical trial for treating malignant gliomas. Unfortunately, deletion of γ34.5 attenuates virus replication in cancer cells, especially human glioblastoma stem cells (GSCs). To develop new oHSVs for use in the brain and that replicate in GSCs, we explored the effect of deleting the γ34.5 Beclin 1 binding domain (BBD). To ensure cancer selectivity and safety, we inactivated the ICP6 gene (UL39, large subunit of ribonucleotide reductase), constructing ICP6 mutants with different γ34.5 genotypes: Δ68HR-6, intact γ34.5; Δ68H-6, γ34.5 BBD deleted; and 1716-6, γ34.5 deleted. Multimutated Δ68H-6 exhibited minimal neuropathogenicity in HSV-1-susceptible mice, as opposed to Δ68H and Δ68HR-6. It replicated well in human glioma cell lines and GSCs, effectively killing cells in vitro and prolonging survival of mice bearing orthotopic brain tumors. In contrast, 1716 and 1716-6 barely replicated in GSCs. Infection of glioma cells with Δ68H-6 and 1716-6 induced autophagy and increased phosphorylation of eIF2α, while inhibition of autophagy, by Beclin 1 short hairpin RNA (shRNA) knockdown or pharmacological inhibition, had no effect on virus replication or phosphorylated eIF2α (p-eIF2α) levels. Thus, Δ68H-6 represents a new oHSV vector that is safe and effective against a variety of brain tumor models.  相似文献   

2.
Herpes simplex virus (HSV) enters cells by fusion at plasma membranes or endosomes. Cellular factors route the virus to different pathways. αVβ3-integrin directs HSV to a lipid raft and acidic endosome pathway. We report that infection mediated by nectin1 plus αVβ3-integrin exhibits the same characteristics as entry mediated by raft-located forms of nectin. αVβ3-integrin relocalizes nectin1 to lipid rafts, independently of virus. Thus, HSV routing to the lipid raft-dependent pathway is consequent to the integrin-induced relocalization of nectin1. Inhibition by the Na+/H+ exchanger 5-(N-ethyl-N-isopropyl)amirolide suggests that αVβ3-integrin overexpression favors HSV macropinocytic uptake in some cells but not in others.  相似文献   

3.
The chimeric herpes simplex viruses (HSV) are Δγ134.5 vectors encoding the human cytomegalovirus (HCMV) IRS1 or TRS1 genes. They are capable of late viral protein synthesis and are superior to Δγ134.5 HSVs in oncolytic activity. The interferon (IFN) response limits efficient HSV gene expression and replication. HCMV TRS1 and IRS1 restore one γ134.5 gene function: evasion of IFN-inducible protein kinase R, allowing late viral protein synthesis. Here we show that, unlike wild-type HSV, the chimeric HSV do not restore another γ134.5 function, the suppression of early IFN signaling mediated by IFN regulatory factor 3 (IRF3).  相似文献   

4.
Jin H  Ma Y  Yan Z  Prabhakar BS  He B 《Journal of virology》2012,86(2):1059-1068
The γ134.5 protein of herpes simplex viruses (HSV) is essential for virulence. Accordingly, an HSV mutant lacking γ134.5 is attenuated in vivo. Despite its vaccine potential, the mechanism by which the γ134.5 null mutant triggers protective immunity is unknown. In this report we show that vaccination with the γ134.5 null mutant protects against lethal challenge from wild-type virus via IκB kinase in dendritic cells (DCs), which sense virus-associated molecular patterns. Unlike mock-treated DCs, DCs primed with the γ134.5 null mutant ex vivo mediate resistance to wild-type HSV after adoptive transfer into naïve mice. Furthermore, the γ134.5 null mutant activates IκB kinase, which facilitates p65/RelA phosphorylation and nuclear translocation, resulting in DC maturation. While unable to produce infectious virus in DCs, this mutant virus expresses early and late genes. In its abortive infection, the γ134.5 null mutant induces protective immunity more effectively in CD8+ DCs than in CD8 DCs. This is mirrored by a higher level of interleukin-6 (IL-6) and IL-12 secretion by CD8+ DCs than CD8 DCs. Remarkably, inhibition of p65/RelA phosphorylation or nuclear translocation in CD8+ DCs disrupts protective immunity. These results suggest that engagement of the γ134.5 null mutant with CD8+ DCs elicits innate immunity to activate NF-κB, which translates into protective immunity.  相似文献   

5.
We recently provided evidence that the ribonucleotide reductase R1 subunits of herpes simplex virus types 1 and 2 (HSV-1 and -2) protect cells against tumor necrosis factor alpha- and Fas ligand-induced apoptosis by interacting with caspase 8. Double-stranded RNA (dsRNA) is a viral intermediate known to initiate innate antiviral responses. Poly(I · C), a synthetic analogue of viral dsRNA, rapidly triggers caspase 8 activation and apoptosis in HeLa cells. Here, we report that HeLa cells after HSV-1 and HSV-2 infection were quickly protected from apoptosis caused by either extracellular poly(I · C) combined with cycloheximide or transfected poly(I · C). Cells infected with the HSV-1 R1 deletion mutant ICP6Δ were killed by poly(I · C), indicating that HSV-1 R1 plays a key role in antiapoptotic responses to poly(I · C). Individually expressed HSV R1s counteracted caspase 8 activation by poly(I · C). In addition to their binding to caspase 8, HSV R1s also interacted constitutively with receptor-interacting protein 1 (RIP1) when expressed either individually or with other viral proteins during HSV infection. R1(1-834)-green fluorescent protein (GFP), an HSV-2 R1 deletion mutant protein devoid of antiapoptotic activity, did not interact with caspase 8 and RIP1, suggesting that these interactions are required for protection against poly(I · C). HSV-2 R1 inhibited the interaction between the Toll/interleukin-1 receptor domain-containing adaptor-inducing beta interferon (IFN-β) (TRIF) and RIP1, an interaction that is essential for apoptosis triggered by extracellular poly(I · C) plus cycloheximide or TRIF overexpression. TRIF silencing reduced poly(I · C)-triggered caspase 8 activation in mock- and ICP6Δ-infected cells, confirming that TRIF is involved in poly(I · C)-induced apoptosis. Thus, by interacting with caspase 8 and RIP1, HSV R1s impair the apoptotic host defense mechanism prompted by dsRNA.  相似文献   

6.
We previously reported that HSV-2 R1, the R1 subunit (ICP10; UL39) of herpes simplex virus type-2 ribonucleotide reductase, protects cells against apoptosis induced by the death receptor (DR) ligands tumor necrosis factor-alpha- (TNFα) and Fas ligand (FasL) by interrupting DR-mediated signaling at, or upstream of, caspase-8 activation. Further investigation of the molecular mechanism underlying HSV-2 R1 protection showed that extracellular-regulated kinase 1/2 (ERK1/2), phosphatidylinositol 3-kinase (PI3-K)/Akt, NF-κB and JNK survival pathways do not play a major role in this antiapoptotic function. Interaction studies revealed that HSV-2 R1 interacted constitutively with caspase-8. The HSV-2 R1 deletion mutant R1(1-834)-GFP and Epstein–Barr virus (EBV) R1, which did not protect against apoptosis induced by DR ligands, did not interact with caspase-8, indicating that interaction is required for protection. HSV-2 R1 impaired caspase-8 activation induced by caspase-8 over-expression, suggesting that interaction between the two proteins prevents caspase-8 dimerization/activation. HSV-2 R1 bound to caspase-8 directly through its prodomain but did not interact with either its caspase domain or Fas-associated death domain protein (FADD). Interaction between HSV-2 R1 and caspase-8 disrupted FADD-caspase-8 binding. We further demonstrated that individually expressed HSV-1 R1 (ICP6) shares, with HSV-2 R1, the ability to bind caspase-8 and to protect cells against DR-induced apoptosis. Finally, as the long-lived Fas protein remained stable during the early period of infection, experiments with the HSV-1 UL39 deletion mutant ICP6∆ showed that HSV-1 R1 could be essential for the protection of HSV-1-infected cells against FasL.  相似文献   

7.
After infection with herpes simplex virus type 1 (HSV-1), latent infection persists for life in the trigeminal ganglion and reactivation results in an outbreak of cold sores around the mouth. Many previous studies have reported HSV-1 reactivation to be a risk factor for Alzheimer’s disease (AD). This study enrolled subjects with AD (n = 85), subjects with amnestic mild cognitive impairment (aMCI; a prodromal stage of AD) (n = 34), and healthy controls (n = 28). The avidity index of anti-HSV-1 IgG antibodies—a known indicator of HSV-1 reactivation—was measured in order to clarify the relationship between HSV-1 reactivation and symptoms of cognitive function in AD.Cognitive function in AD and aMCI were evaluated using scores from the mini-mental state examination (MMSE) and frontal assessment battery (FAB). The results showed that the subjects with aMCI, for which cerebral function is better preserved than subjects with AD, had a higher anti-HSV-1 IgG antibody avidity index than the AD subjects or healthy controls. Furthermore, the anti-HSV-1 IgG antibody avidity index was even higher in the subjects with high MMSE scores on orientation to time and three-step command subscores. We observed a negative correlation between the anti-HSV-1 IgG antibody avidity index and plasma BDNF concentration, which is an indicator of encephalitis. This suggests that HSV-1 reactivation, as observed through an increase in the anti-HSV-1 IgG avidity index, does not progress to encephalitis. These results suggest that HSV-1 reactivation occurs from the stage of aMCI, which is prodromal to AD, and can affect AD symptoms without an intermediary stage of severe encephalitis. The study demonstrates that the anti-HSV-1 IgG antibody avidity index could be a useful biomarker for the early diagnosis of aMCI as well as AD, and suggests that antiviral medication to treat HSV-1 could play a role in preventing the onset of AD.  相似文献   

8.
Synthesis and assembly of ribosomal proteins into mature ribosomes persist late after infection of cells with herpes simplex virus type 1, while synthesis of β-actin is drastically shut off. Since mRNAs encoding ribosomal proteins and β-actin undergo concomitant degradation in infected HeLa cells, we have advanced the hypothesis that translation of the remaining mRNAs is differentially controlled after infection. The behaviour of mRNAs for three ribosomal proteins and for β-actin was investigated during the course of infection. In uninfected cells, β-actin mRNAs are associated with large polyribosomes, while only a part of ribosomal protein mRNAs are present in polyribosomes. In the course of infection, β-actin mRNAs are released from the ribosomes and are sequestered with 40S ribosomal subunits. Simultaneously, ribosomal protein mRNAs become associated with an increased number of ribosomes, even late in infection. In addition, virally induced phosphorylation of ribosomal protein S6 is more efficient in pre-existing ribosomes than in newly assembled ribosomes. These results indicate that in infected cells (i) translation of β-actin mRNA is selectively inhibited at a step necessary for binding the 60S ribosomal subunits; (ii) the rate of initiation of translation of ribosomal protein mRNAs increases after infection; and (iii) it is likely that translation of ribosomal protein mRNAs takes place preferentially on pre-existing ribosomes. Received: 5 February 1997 / Accepted: 28 May 1997  相似文献   

9.
Earlier studies have described the alpha 4/c113 baby hamster kidney cell line which constitutively expresses the alpha 4 protein, the major regulatory protein of herpes simplex virus 1 (HSV-1). Introduction of the HSV-1 glycoprotein B (gB) gene, regulated as a gamma 1 gene, into these cells yielded a cell line which constitutively expressed both the alpha 4 and gamma 1 gB genes. The expression of the gB gene was dependent on the presence of functional alpha 4 protein. In this article we report that we introduced into the alpha 4/c113 and into the parental BHK cells, the HSV-1 BamHI J fragment, which encodes the domains of four genes, including those of glycoproteins D, G, and I (gD, gG, and gI), and most of the coding sequences of the glycoprotein E (gE) gene. In contrast to the earlier studies, we obtained significant constitutive expression of gD (also a gamma 1 gene) in a cell line (BJ) derived from parental BHK cells, but not in a cell line (alpha 4/BJ) which expresses functional alpha 4 protein. RNA homologous to the gD gene was present in significant amounts in the BJ cell line; smaller amounts of this RNA were detected in the alpha 4/BJ cell line. RNA homologous to gE, presumed to be polyadenylated from signals in the vector sequences, was present in the BJ cells but not in the alpha 4/BJ cells. The expression of the HSV-1 gD and gE genes was readily induced in the alpha 4/BJ cells by superinfection with HSV-2. The BJ cell line was, in contrast, resistant to expression of HSV-1 and HSV-2 genes. The BamHI J DNA fragment copy number was approximately 1 per BJ cell genome equivalent and 30 to 50 per alpha 4/BJ cell genome equivalent. We conclude that (i) the genes specifying gD and gB belong to different viral regulatory gene subsets, (ii) the gD gene is subject to both positive and negative regulation, (iii) both gD and gE mRNAs are subject to translational controls although they may be different, and (iv) the absence of expression of gD in the alpha 4/BJ cells reflects the expression of the alpha 4 protein in these cells.  相似文献   

10.
Fang J  Wang Y  Lv X  Shen X  Ni X  Ding K 《Glycoconjugate journal》2012,29(5-6):365-377
A soluble homogeneous β-glucan, GFPBW1, with a molecular mass of 300?kDa was purified from the fraction of the fruit bodies of Grifola frondosa extracted with 5?% NaOH. Using various methods, such as infrared spectroscopy, NMR, methylation and monosaccharide composition analysis, its structure was determined to be a β-D-(1-3)-linked glucan backbone with a single β-D-(1-6)-linked glucopyranosyl residue branched at C-6 on every third residue. It induced TNF-α and IL-6 production and the activation of Syk and NF-κB signaling in resident peritoneal macrophages from ICR mice, which could be significantly inhibited by the blocking reagent laminarin. A competitive phagocytosis assay with FITC-zymosan indicated that GFPBW1 could bind to DC-associated C-type lectin 1 (Dectin-1). The TNF-α secretion and activation of Syk/NF-κB signaling triggered by GFPBW1 were enhanced in RAW264.7 cells overexpressing wild but not mutant (Δ38 and Y15S) Dectin-1. Furthermore, GFPBW1 potentiated the Concanavalin A-induced proliferative response of splenocytes and inhibited Sarcoma-180 growth allografted in ICR mice but not in immunodeficient BALB/c nu/nu mice. These results suggested that the antitumor activity of GFPBW1 was partially associated with the activation of macrophages via the Dectin-1/Syk/NF-κB signaling pathway. This molecule could be a promising biological response modifier with clear application for antitumor therapies.  相似文献   

11.
12.
Toll-like receptors (TLRs) are a class of proteins (patterns recognition receptors-PRRs) capable of recognizing molecules frequently found in pathogens (that are so-called pathogen-associated molecular patterns-PAMPs), they play a key role in the initiation of innate immune response by detecting PAMPs. Our findings show that the functional effects of TLRs co-stimulation on megakaryocytopoiesis. A single cell may receive multiple signal inputs and we consider that multiple TLRs are likely triggered during infection by multiple PAMPs that, in turn, might be involved in infection driven megakaryocytopoiesis, and the present study provide the evidence for the megakaryocytic effects of TLRs co-stimulation.  相似文献   

13.
We have recently demonstrated that mammalian uracil-DNA glycosylase activity is undetectable in adult neurons. On the basis of this finding we hypothesized that uracil, derived either from oxidative deamination of cytosine or misincorporation of dUMP in place of dTMP during DNA repair by the unique nuclear DNA polymerase present in adult neurons, DNA polymerase β, might accumulate in neuronal DNA. Uracil residues could also arise in the herpes simplex 1 (HSV1) genome during latency in nerve cells. We therefore suggest a role for the virus encoded uracil-DNA glycosylase in HSV1 reactivation and in the first steps of DNA replication. We show here 1) that the viral DNA polymerase incorporates dUTP in place of dTTP with a comparable efficiencyin vitro; 2) that virus specific DNA/protein interactions between the virus encoded origin binding protein and its target DNA sequence is altered by the presence of uracil residues in its central region TCGCA. Thus uracil, present in viral OriS or other key sequences could hamper the process leading to viral reactivation. Hence, HSV1 uracil-DNA glycosylase, dispensable in viral proliferation in tissue culture, could be essential in neurons for the “cleansing” of the viral genome of uracil residues before the start of replication.  相似文献   

14.

Background

Despite effective radiotherapy for the initial stages of cancer, several studies have reported the recurrence of various cancers, including medulloblastoma. Here, we attempt to capitalize on the radiation-induced aggressive behavior of medulloblastoma cells by comparing the extracellular protease activity and the expression pattern of molecules, known to be involved in cell adhesion, migration and invasion, between non-irradiated and irradiated cells.

Methodology/Principal Findings

We identified an increase in invasion and migration of irradiated compared to non-irradiated medulloblastoma cells. RT-PCR analysis confirmed increased expression of uPA, uPAR, focal adhesion kinase (FAK), N-Cadherin and integrin subunits (e.g., α3, α5 and β1) in irradiated cells. Furthermore, we noticed a ∼2-fold increase in tyrosine phosphorylation of FAK in irradiated cells. Immunoprecipitation studies confirmed increased interaction of integrin β1 and FAK in irradiated cells. In addition, our results show that overexpression of uPAR in cancer cells can mimic radiation-induced activation of FAK signaling. Moreover, by inhibiting FAK phosphorylation, we were able to reduce the radiation-induced invasiveness of the cancer cells. In this vein, we studied the effect of siRNA-mediated knockdown of uPAR on cell migration and adhesion in irradiated and non-irradiated medulloblastoma cells. Downregulation of uPAR reduced the radiation-induced adhesion, migration and invasion of the irradiated cells, primarily by inhibiting phosphorylation of FAK, Paxillin and Rac-1/Cdc42. As observed from the immunoprecipitation studies, uPAR knockdown reduced interaction among the focal adhesion molecules, such as FAK, Paxillin and p130Cas, which are known to play key roles in cancer metastasis. Pretreatment with uPAR shRNA expressing construct reduced uPAR and phospho FAK expression levels in pre-established medulloblastoma in nude mice.

Conclusion/Significance

Taken together, our results show that radiation enhances uPAR-mediated FAK signaling and by targeting uPAR we can inhibit radiation-activated cell adhesion and migration both in vitro and in vivo.  相似文献   

15.
Mitogen-activated protein kinases (MPKs) play essential roles in guard cell signaling, but whether MPK cascades participate in guard cell ethylene signaling and interact with hydrogen peroxide (H2O2), nitric oxide (NO), and ethylene-signaling components remain unclear. Here, we report that ethylene activated MPK3 and MPK6 in the leaves of wild-type Arabidopsis thaliana as well as ethylene insensitive2 (ein2), ein3, nitrate reductase1 (nia1), and nia2 mutants, but this effect was impaired in ethylene response1 (etr1), nicotinamide adenine dinucleotide phosphate oxidase AtrbohF, mpk kinase1 (mkk1), and mkk3 mutants. By contrast, the constitutive triple response1 (ctr1) mutant had constitutively active MPK3 and MPK6. Yeast two-hybrid, bimolecular fluorescence complementation, and pull-down assays indicated that MPK3 and MPK6 physically interacted with MKK1, MKK3, and the C-terminal region of EIN2 (EIN2 CEND). mkk1, mkk3, mpk3, and mpk6 mutants had typical levels of ethylene-induced H2O2 generation but impaired ethylene-induced EIN2 CEND cleavage and nuclear translocation, EIN3 protein accumulation, NO production in guard cells, and stomatal closure. These results show that the MKK1/3–MPK3/6 cascade mediates ethylene-induced stomatal closure by functioning downstream of ETR1, CTR1, and H2O2 to interact with EIN2, thereby promoting EIN3 accumulation and EIN3-dependent NO production in guard cells.  相似文献   

16.
Early human immunodeficiency virus infection is characterized by the predominance of CCR5-tropic (R5) virus. However, in many individuals CXCR4-tropic (X4) virus appears in late infection. The reasons for this phenotypic switch are unclear. The patterns of chemokine receptor expression suggest that X4 and R5 viruses have a preferential tropism for na?ve and memory T cells, respectively. Since memory cells divide approximately 10 times as often as na?ve cells in uninfected individuals, a tropism for memory cells in early infection may provide an advantage. However, with disease progression both na?ve and memory cell division frequencies increase, and at low CD4 counts, the na?ve cell division frequency approaches that of memory cells. This may provide a basis for the phenotypic switch from R5 to X4 virus observed in late infection. We show that a model of infection using observed values for cell turnover supports this mechanism. The phenotypic switch from R5 to X4 virus occurs at low CD4 counts and is accompanied by a rapid rise in viral load and drop in CD4 count. Thus, low CD4 counts are both a cause and an effect of X4 virus dominance. We also investigate the effects of different antiviral strategies. Surprisingly, these results suggest that both conventional antiretroviral regimens and CCR5 receptor-blocking drugs will promote R5 virus over X4 virus.  相似文献   

17.
Alzheimer’s disease (AD) is defined both by its progressive cognitive deterioration and hallmark increase in neuronal Aβ plaque formation. However, many of the underlying neurobiological facets of this disease are still being elucidated. Previous research has demonstrated that production of neuronal hydrogen sulfide (H2S) is significantly decreased in patients with AD. Moreover, systemic plasma H2S levels are negatively correlated with its severity. However, how a decrease in H2S production might be correlated with either the etiology or pathophysiology of AD remains unknown. To better understand the role of H2S in AD, we examined both levels of H2S and the expression and activity H2S-synthesizing enzyme (cystathionine beta synthase or CBS) in an APP/PS1 transgenic mouse line at 3, 6, 9 and 12 months. After intraperitoneal (i.p.) administration of an H2S donor (NaHS) into APP/PS1 mice, application of exogenous H2S resulted in improved spatial learning and memory acquisition in APP/PS1 mice. H2S administration also led to significant decrease in extracellular levels of Aβ40 and Aβ42, the expression of BACE1 and PS1, and a significant increase of ADAM17 expression. Similarly, an increase in non-amyloidogenic C83 fragment generation and a decrease in amyloidogenic C99 fragment generation were also observed. Thus, NaHS application resulted in a shift from the plaque-forming beta pathway to the non-plaque forming alpha pathway of APP cleavage in 6 and 12 month APP/PS1 mice. These results indicate the importance of H2S to AD severity and that administration of exogenous H2S can promote a non-amyloidogenic processing of APP.  相似文献   

18.
19.
20.
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