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.
Assembly and release of human immunodeficiency virus (HIV) occur at the plasma membrane of infected cells and are driven by the Gag polyprotein. Previous studies analyzed viral morphogenesis using biochemical methods and static images, while dynamic and kinetic information has been lacking until very recently. Using a combination of wide-field and total internal reflection fluorescence microscopy, we have investigated the assembly and release of fluorescently labeled HIV-1 at the plasma membrane of living cells with high time resolution. Gag assembled into discrete clusters corresponding to single virions. Formation of multiple particles from the same site was rarely observed. Using a photoconvertible fluorescent protein fused to Gag, we determined that assembly was nucleated preferentially by Gag molecules that had recently attached to the plasma membrane or arrived directly from the cytosol. Both membrane-bound and cytosol derived Gag polyproteins contributed to the growing bud. After their initial appearance, assembly sites accumulated at the plasma membrane of individual cells over 1–2 hours. Assembly kinetics were rapid: the number of Gag molecules at a budding site increased, following a saturating exponential with a rate constant of ∼5×10−3 s−1, corresponding to 8–9 min for 90% completion of assembly for a single virion. Release of extracellular particles was observed at ∼1,500±700 s after the onset of assembly. The ability of the virus to recruit components of the cellular ESCRT machinery or to undergo proteolytic maturation, or the absence of Vpu did not significantly alter the assembly kinetics.  相似文献   

2.
3.
Hepatitis C virus is a blood-borne virus that typically establishes a chronic infection in the liver, which often results in cirrhosis and hepatocellular carcinoma. Progress in understanding the complete virus life cycle has been greatly enhanced by the recent availability of a tissue culture system that produces infectious virus progeny. Thus, it is now possible to gain insight into the roles played by viral components in assembly and egress and the cellular pathways that contribute to virion formation. This minireview describes the key determining viral and host factors that are needed to produce infectious virus.  相似文献   

4.
《Cell host & microbe》2014,15(1):103-112
  1. Download : Download high-res image (185KB)
  2. Download : Download full-size image
  相似文献   

5.
Hepatitis C virus (HCV) assembly is known to occur in juxtaposition to lipid droplets, but the mechanisms of nascent virion transport and release remain poorly understood. Here we demonstrate that HCV core protein targets to early and late endosomes but not to mitochondria or peroxisomes. Further, by employing inhibitors of early and late endosome motility in HCV-infected cells, we demonstrate that the movement of core protein to the early and late endosomes and virus production require an endosome-based secretory pathway. We also observed that this way is independent of that of the internalization of endocytosed virus particles during virus entry.Hepatitis C virus (HCV) is a major causative agent of chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma. HCV usually infects host cells via receptor-mediated endocytosis (6, 21), followed by the release of genomic RNA after uncoating of the nucleocapsid in the endosome. HCV core protein constitutes the viral nucleocapsid and may possess multiple functions. Intracellular HCV core protein is localized mainly in lipid droplets (LDs) (23, 29). Recent studies have indicated that core protein promotes the accumulation of LDs to facilitate virus assembly (1, 10) and recruits viral replication complexes to LD-associated membranes, where virus assembly takes place (23). However, the precise mechanisms of HCV assembly, budding, and release remain largely unclear. Most recently, HCV virion release has been shown to require the functional endosomal sorting complex required for transport III (ESCRT-III) and Vps4 (an AAA ATPase) (13), which are required for the biogenesis of the multivesicular body (MVB), a late endosomal compartment (12). Late endosomes have been implicated in the budding of several other viruses, including retroviruses (8, 17, 24, 25, 27), rhabdoviruses (14), filoviruses (18, 20), arenaviruses (26, 32), and hepatitis B virus (35). However, little is known about the roles of late endosomes in the HCV life cycle.Since LDs are associated with the endoplasmic reticulum membrane, endosomes, peroxisomes, and mitochondria (16, 37), we investigated what subcellular compartments may be involved in HCV assembly and release. We first compared the intracellular distribution of HCV core protein with that of early endosome markers Rab5a and early endosome antigen 1 (EEA1), as well as the late endosome marker CD63 in the HCV Jc1-infected Huh7.5 cells at day 10 postinfection (p.i.). In immunofluorescence studies, we demonstrated that the core protein partially colocalized with Rab5a (Fig. (Fig.1A,1A, left panel) or EEA1 (Fig. (Fig.1A,1A, right panel). This finding was confirmed by the expression of enhanced green fluorescent protein (EGFP)-tagged Rab5a (Fig. (Fig.1A,1A, middle panel). Similarly, core protein also showed partial colocalization with CD63 (Fig. (Fig.1B).1B). In particular, core protein showed numerous vesicle-like structures of homogeneous size that partially colocalized with CD63 at the cell periphery (Fig. (Fig.1B,1B, right panel inset and drawing). This result contrasts with that of Ai et al. (2), who observed, by confocal microscopy, that core protein did not interact with markers of early and late endosomes. Ai et al. did find, however, that multimeric core complexes cofractionated with ER/late endosomal membranes in HCV-infected cells.Open in a separate windowFIG. 1.HCV core protein colocalized with early and late endosomes but not mitochondria and peroxisomes. HCV-infected cells were costained with anti-core protein (red) and anti-Rab5a (A, left panel), -EEA1 (A, right panel), or -CD63 antibodies (green) (B) or transfected with plasmids expressing enhanced green fluorescent protein (EGFP)-Rab5a (A, middle panel), enhanced yellow fluorescent protein (EYFP) (C, left panel), EYFP-mitochondria (C, middle panel), or EYFP-peroxisome (C, right panel). Cellular DNA was labeled with DAPI (4′,6-diamidine-2-phenylindole) (blue). Images shown were collected sequentially with a confocal laser scanning microscope and merged to demonstrate colocalization (yellow merge fluorescence). Enlarged views of parts of every image are shown (insets). The cartoon in panel B illustrates the core protein-containing vesicle-like structures (depicted as red circles), which partially colocalized with CD63 at the cell periphery in HCV-infected cells. PM, plasma membrane.To demonstrate the specificity of the association of core protein with endosomes, we transfected HCV-infected cells (at day 10 p.i.) with pEYFP, pEYFP-mito, and pEYFP-peroxi (Clontech) (Fig. (Fig.1C),1C), which label the cytoplasm/nucleus, mitochondria, and peroxisomes, respectively. The results showed that HCV core protein did not colocalize with mitochondria or peroxisomes. Taken together, these results indicate that core protein is partially associated with early and late endosomes.To investigate the functional involvement of the endosomes in HCV release, we employed HCV-infected cells (at day 10 p.i.). In our observation, at day 10 p.i., not all the cells were infected with HCV, as revealed by immunofluorescence staining against core protein (data not shown), suggesting that these cells are a mixture of infected and noninfected cells. We examined the effects of inhibitors of endosome movement, including 10 μM nocodazole (which induces microtubule depolymerization), 100 nM wortmannin (which inhibits early endosomes), 20 nM Baf-A1 (which blocks early endosomes from fusing with late endosomes) (Sigma), and 10 μg/ml U18666A (which arrests late endosome movement) (Biomol), on the release of HCV in the HCV-infected cells. We first determined the possible cytotoxicity of these drugs. We found that within 20 h of the drug application, no significant effect on cell viability, as revealed by 3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium salt (MTS) assay, was observed (Fig. (Fig.2C).2C). We therefore treated the cells with the various drugs for 20 h. This protocol focuses only on virus release, not on virus entry, as the reinfection of Huh7.5 cells could not account for the effects on virus release, as one round of HCV replication requires about 24 h (11).Open in a separate windowFIG. 2.HCV virion release requires endosome motility. HCV-infected cells (at day 10 p.i.) were treated with DMSO, nocodazole (10 μM), U18666A (10 μg/ml), Baf-A1 (20 nM), or wortmannin (100 nM) for 20 h, and then the levels of extracellular (A) and intracellular (B) HCV RNA and HCV core proteins (F) in cells were analyzed by RT-qPCR and Western blotting, respectively. Results of Western blotting were quantified by PhosphorImager counting. (C) Analysis of cellular proliferation and survival by MTS assay. (D) Assay of extracellular viral infectivity. The culture supernatants from the cells treated with the various drugs as indicated were used to infect naïve Huh7.5 cells. The cells were stained with anti-core protein antibody (green) and DAPI (blue). The images were analyzed by using Metamorph, and the proportion of cells (of 5,000 counted) expressing core protein was counted (E). Results are presented as percentages and are averages and standard deviations from results of triplicate experiments. (G) In parallel, the HCV-infected cells were costained with anti-Rab5a (green) and -NS5A (red) antibodies. Cellular DNA was labeled with DAPI (blue). Enlarged views of parts of every image are shown (insets). (H) Colocalization efficiency between NS5A and early endosomes was analyzed by using Zeiss LSM Zen software. Error bars represent standard deviations of the mean result from 20 cells of two experiments. (I) Assay of intracellular HCV titers. Intracellular HCV particles were prepared from the HCV-infected cells (at day 10 p.i.) treated with U18666A at concentrations varying from 2.5 to 10 μg/ml for 20 h. The titers of intracellular HCV particles were determined by immunofluorescence staining for core-positive cell foci and are reported in focus-forming units (FFU)/ml. (J) HCV-infected cells (at day 10 p.i.) were treated with U18666A at concentrations varying from 2.5 to 20 μg/ml for 8 h, and then the levels of extracellular HCV RNA were analyzed by RT-qPCR. (K) Effects of nocodazole (10 μM), U18666A (10 μg/ml), Baf-A1 (20 nM), or wortmannin (100 nM) on HCV production in single-cycle HCV growth assays. Huh7.5 cells were infected with HCV at an MOI of 1 and then incubated with DMEM containing the various drugs. At 24 h p.i., the cells and their culture supernatants were collected and used to determine the levels of intracellular and extracellular HCV RNA, which were converted to percentages of the control levels (DMSO) as 100%. Noc, nocodazole; U18, U18666A; Baf-A1, Bafilomycin A1; Wortman, wortmannin.After treatments for 20 h, the cells and their culture supernatants were collected. Intracellular RNA was isolated from cell lysates using a High Pure RNA isolation kit (Roche), and viral RNA was isolated from cell culture supernatants using a QIAamp viral RNA kit (Qiagen). Equivalent RNA volumes were subsequently analyzed on a LightCycler 1.5 real-time PCR system (Roche) for quantitative PCR (qPCR), with a primer-probe set specific for the 5′ untranslated region (UTR) sequence of HCV Jc1 and a second set specific for the glyceraldehyde-3-phosphate dehydrogenase (GAPDH) gene to quantitate the RNA amount. To calculate the percentage of HCV RNA remaining after the inhibitor treatments, the mean HCV RNA levels from triplicate wells of each sample type were standardized to the mean GAPDH RNA level in the dimethyl sulfoxide (DMSO) control wells. The relative levels of HCV RNA (percentage of control) were then analyzed with LightCycler software (version 3.53) and calculated using the relative quantification method as described in Roche Applied Science, Relative Quantification, Technical Note no. LC 13/2001 (http://www.gene-quantification.de/roche-rel-quant.pdf).As indicated in Fig. Fig.2A,2A, the extracellular HCV RNA levels were reduced to 32.1%, 21.7%, 76.2%, and 53.8% of the DMSO control after treatments with nocodazole, U18666A, Baf-A1, and wortmannin, respectively. We further determined the intracellular levels of HCV RNA to test whether these drugs have an effect on HCV RNA replication. Intracellular HCV RNA was reduced to 64% by nocodazole treatment (Fig. (Fig.2B),2B), confirming previous reports that microtubules are required for HCV RNA synthesis (9, 28). Treatments with Baf-A1 and wortmannin, however, did not affect HCV RNA replication. Interestingly, the intracellular HCV RNA level was increased to 167% by U18666A treatment (Fig. (Fig.2B),2B), suggesting that U18666A blocks HCV particle release (Fig. (Fig.2A),2A), thereby causing accumulation of HCV RNA in the late endosomes. We further determined the extracellular viral infectivity after treatments with these drugs by using the culture supernatant to infect naïve Huh7.5 cells. The infectivity was checked by counting core protein-expressing cells. Raw acquired 8-bit images (Fig. (Fig.2D)2D) were converted to 16-bit and analyzed with the Multi Wavelength Cell Scoring application module in Metamorph (Molecular Device). All of these treatments significantly inhibited the production of infectious HCV particles, as shown by the reduced infectivity in the supernatant of the infected cells (Fig. 2D and E). Determination of the amounts of core protein in the lysates showed that the levels of core protein in cells were not significantly affected by the various drug treatments (Fig. (Fig.2F).2F). Taken together, these results suggest that the inhibitors of endosome movement, including U18666A, Baf-A1, and wortmannin, reduced the secretion of HCV particles but not the HCV RNA replication. On the other hand, nocodazole-induced microtubule depolymerization reduced both HCV RNA synthesis (Fig. (Fig.2B)2B) (9, 28) and virus release (Fig. 2A and E). Since nocodazole also blocks the movement of endosomes between pericentriolar regions and the cell periphery (3, 7), it therefore should perturb particle release. Our present results show that nocodazole had a greater effect on the reduction in extracellular HCV RNA levels (to 32%) than intracellular HCV RNA levels (to 64%) (Fig. 2A and B), and this effect may be due to blocking of endosome movement and reduced HCV RNA replication. These results can be explained on the premise that nocodazole treatments may affect both HCV RNA replication and virus release; it led us to conclude that microtubules may simultaneously play a key role in both HCV RNA replication and virus egress. Thus, both microtubules and the movement of endosomes are required for HCV particle egress.Earlier reports indicated that Rab5, an early endosomal protein, interacts with NS4B (31) and is required for HCV RNA replication. This effect was demonstrated in Rab5 small interfering RNA (siRNA)-transfected replicon cells (5, 31). However, in our current studies (Fig. (Fig.2B)2B) treatments with endosome inhibitors did not reduce the levels of intracellular HCV RNA. In order to investigate the discrepancy between our results and the earlier studies, we examined the effects of endosome inhibitors on the colocalization of NS5A with Rab5a in the HCV-infected cells (at day 10 p.i.) by calculating the percentage of colocalization in the cells. Colocalization scatter diagrams were generated using the colocalization function of the Zeiss LSM Zen software. The weighted colocalization coefficient, defined as the sum of intensities of colocalizing pixels for NS5A with Rab5a in comparison to the overall sum of pixel intensities (above threshold) for NS5A, was determined. Under control conditions (DMSO), NS5A was colocalized with Rab5a throughout the cytoplasm and perinuclear region (Fig. (Fig.2G,2G, upper left panel). The proportion of NS5A that colocalized with Rab5a was 41% (Fig. (Fig.2H).2H). After treatment with U18666A, the dispersed Rab5a compartments were found only in the perinuclear region (Fig. (Fig.2G,2G, upper right panel). Importantly, the colocalization of NS5A with Rab5a was increased to 57% by U18666A treatment. Treatments with Baf-A1 and wortmannin, however, had no significant effect. In contrast, nocodazole treatment reduced the colocalization of NS5A with Rab5a to 12% (Fig. (Fig.2G2G and and2H).2H). Previous studies have reported that the expression levels of Rab5 and its colocalization with HCV NS4B (or NS5A) play a functional role in HCV RNA replication (31). In addition, Rab5 remained in an early endosome fraction and the expression levels of Rab5 showed no significant difference between wortmannin (100 nM)-treated or untreated cells (22). More importantly, the total levels of early endosome proteins, including EEA1 (Fig. (Fig.3F)3F) and Rab5a (data not shown), are not altered by the endosome inhibitors. The colocalization efficiency of NS5A with Rab5a was not reduced by U18666A, Baf-A1, or wortmannin treatments, suggesting that these drugs cannot decrease HCV RNA replication. This finding is consistent with the previous results (Fig. (Fig.2B).2B). These findings suggest that the observed discrepancy between our results in Fig. Fig.2B2B and those of the other studies (5, 31) is most likely due to differences in the expression levels of Rab5a.Open in a separate windowFIG. 3.HCV particles formed are transported from early to late endosomes. HCV-infected cells (at day 10 p.i.) were treated with the various drugs and 14 h later labeled with antibodies specific for core protein (red) and CD63 (green) (A) or EEA1 (green) (D). At the right is an enlarged area from the merged image. Nuclei were stained with DAPI (blue). (B) (E) Colocalization efficiency between core protein and early or late endosomes was analyzed by using Zeiss LSM Zen software. Error bars represent standard deviations of the mean result from 20 cells of two experiments. In parallel, the cell lysates were collected and then immunoblotted with antibodies against CD63 (C) and EEA1 (F). Results were quantified by PhosphorImager counting. The HCV-infected cells (at day 10 p.i.) were fixed either for immunofluorescence microscopy (G) or for thin-section electron microscopy (H and I). Cells were costained with anti-Rab5a (green) (G, left panel), -CD63 (green) (G, right panel) and -core protein (red). Lipid droplets (LDs) and nuclei were stained with BODYPI 493/503 (blue) and DAPI (white), respectively (G). Enlarged views of parts of every image are shown (insets). (H, left panel) Early endosome (EE) (white arrow) containing particles resembling HCV adjacent to the LDs. (I, left panel) MVB/late endosome (LE) containing particles resembling HCV and internal vesicles (white arrow). High-magnification images of the early endosome (H, right panel) and MVB/late endosome (I, right panel) harboring particles resembling HCV (black arrow).To further confirm that U18666A suppresses HCV release and/or virus assembly, we determined the titer of the accumulated infectious virus particles inside the cells. The HCV-infected cells (at day 10 p.i.) were treated with U18666A at various concentrations between 2.5 and 10 μg/ml for 20 h, and then intracellular HCV particles were isolated from the cells by repeated freezing and thawing (15). The infectivity was assayed on naïve Huh7.5 cells. The results showed that the titers of infectious intracellular HCV particles were increased in a dose-dependent manner by the U18666A treatments (Fig. (Fig.2I).2I). These data, combined with our previous results (Fig. 2A, B, D, and E), indicate that U18666A blocks HCV particle release. Further, we determined the 50% effective dose (ED50) and 50% cytotoxic concentration (CC50), which are defined as the concentration of U18666A that reduced the levels of extracellular HCV RNA by 50% and the concentration of U18666A that produced 50% cytotoxicity in an MTS assay, respectively. We observed that the HCV-infected cells (at day 10 p.i.) treated with U18666A at concentrations varying from 2.5 to 20 μg/ml for 8 h showed a dose-dependent reduction in extracellular HCV RNA levels (Fig. (Fig.2J).2J). The ED50 and CC50 were calculated by polynomial regression analysis. An ED50 of 8.18 μg/ml (19.2 μM) and a CC50 of 40.26 μg/ml (94.9 μM) were observed for U18666A for the reduction of extracellular HCV RNA and the cytotoxicity of HCV-infected cells, respectively. These results indicate that U18666A acts as a specific inhibitor of HCV release.In our previous results in Fig. 2A and D, we used a multiple-cycle virus growth assay, which could not discriminate the role of endosome movement in infection from that in virus assembly or egress. We therefore used a single-cycle HCV growth assay to further confirm that these endosome inhibitors could suppress HCV release. Previous studies have suggested that one round of HCV replication requires about 24 h (11). Therefore, Huh7.5 cells were infected with HCV JC1 at a multiplicity of infection (MOI) of 1 for 3 h. The HCV-infected cells were washed with phosphate-buffered saline (PBS) and then incubated with Dulbecco modified Eagle medium (DMEM) containing 10% fetal bovine serum (FBS) and nocodazole (10 μM), U18666A (10 μg/ml), Baf-A1 (20 nM), or wortmannin (100 nM). Therefore, the experiment will focus on virus egress and RNA replication instead of virus entry because the cells were first inoculated with HCV and then treated with the various drugs. At 24 h p.i., the cells and their culture supernatants were collected. Intracellular RNA and viral RNA were isolated from cell lysates and cell culture supernatants, respectively. The percentage of HCV RNA remaining after the inhibitor treatments was determined in the same way as for Fig. 2A and B. As indicated in Fig. Fig.2K,2K, the levels of extracellular HCV RNA (or viral particles) were reduced to 61%, 48.8%, 59.7%, and 60.4% after treatments with nocodazole, U18666A, Baf-A1, and wortmannin, respectively, compared to the control DMSO treatment. We further determined the intracellular levels of HCV RNA to test whether these drugs have an effect on HCV RNA replication. Intracellular HCV RNA was reduced to 69% by nocodazole treatment (Fig. (Fig.2K),2K), but not by U18666A, Baf-A1, and wortmannin. Overall, these results of single-cycle HCV growth assay are similar to those of multiple-cycle virus growth assay (Fig. 2A and B), again suggesting that the endosome movement inhibitors reduced the secretion of HCV particles.To further understand the roles of the early and late endosomes in the HCV life cycle, we next examined the effects of endosome inhibitors on the colocalization of core protein with CD63 or EEA1 in the HCV-infected cells (at day 10 p.i.) by calculating the percentage of their colocalization in the cells. The percentage of colocalization of core protein with CD63 or EEA1 was determined in the same way as for Fig. Fig.2H.2H. Under control conditions (DMSO), core protein was colocalized with CD63 throughout the cytoplasm, perinuclear region, and cell periphery (Fig. (Fig.3A,3A, top row). The proportion of core protein that colocalized with CD63 was 17% (Fig. (Fig.3B).3B). After treatment with U18666A, a characteristic collapse of dispersed CD63 compartments to the perinuclear region of the cells was revealed (Fig. (Fig.3A,3A, second row). Importantly, the colocalization of core protein with CD63 increased to 30% when the movement of late endosome was arrested by U18666A, whereas it was reduced to 2% and 7% by Baf-A1 and wortmannin treatments (Fig. 3A and B), respectively, suggesting that the movement of core protein was blocked by U18666A and accumulated in the juxtanuclear region. Additionally, EEA1-labeled distinct puncta, which are dispersed throughout the cytoplasm and partially colocalized with core protein in DMSO treatment (Fig. (Fig.3D,3D, top row), were found clustered in the perinuclear region following treatments with U18666A and Baf-A1 (Fig. (Fig.3D,3D, second and third rows). Correspondingly, the colocalization of core protein and EEA1 was increased after these treatments. In contrast, very little colocalization between core protein and EEA1 was seen after treatment with wortmannin (Fig. (Fig.3D,3D, bottom row). The colocalization of core protein and EEA1 was 10% in the DMSO control, in contrast with 35% and 20% after treatments with U18666A and Baf-A1, respectively, and 4% after treatment with wortmannin (Fig. (Fig.3E).3E). These data suggest that core protein was blocked by U18666A and accumulated in the juxtanuclear region. In parallel, the levels of CD63 and EEA1 protein in the lysates were determined by Western blotting. The results indicated that the total levels of CD63 and EEA1 were not altered by the various drug treatments (Fig. 3C and F, respectively). These results collectively indicate that the colocalization of core protein with CD63 or EEA1 and the release of virus particles depend on the motility of endosomes. Thus, we suggest that HCV core and/or the viral particles formed are transported from early to late endosomes.The above results prompted us to characterize the location of the early and late endosomes in relation to the site of core-LD colocalization, where HCV assembly takes place (23). In HCV-infected cells (at day 10 p.i.), early endosomes (Fig. (Fig.3G,3G, left panel) were colocalized with core protein and were located in juxtaposition to LDs, whereas the late endosomes were located far away from LDs (Fig. (Fig.3G,3G, right panel). These data suggest that following the assembly of viral particles in juxtaposition to LDs, the HCV particles are transported through early to late endosomes. To gain further insight into the trafficking patterns of HCV particles in Huh7.5 cells, we performed electron microscopy of the HCV-infected cells (at day 10 p.i.). Particles resembling HCV were present in both the early endosomes adjacent to the LDs (Fig. (Fig.3H)3H) and MVBs (late endosomes) (Fig. (Fig.3I).3I). The morphology of these endosomal compartments is similar to that reported previously (34, 36). These results again suggest that the HCV particles formed are transported from early to late endosomes.In order to rule out the possibility that the reduction in HCV particle secretion by endosome inhibitors (Fig. (Fig.2)2) was caused by the inhibition of endocytosis-mediated virus entry, we determined the percentage of cells that could be infected by HCV in the presence of endosome inhibitors. Cells were first treated with inhibitors of endosome movement, followed by HCV inoculation for 3 days. This analysis revealed that the same percentage of cells was infected and produced core protein following DMSO and U18666A treatments, 52% and 54%, respectively (Fig. 4A and B), demonstrating that U18666A did not affect HCV entry, and HCV could proceed normally to RNA replication. In contrast, Baf-A1 and wortmannin treatments yielded 7% and 20% (Fig. 4A and B), respectively, of infected cells, indicating that they blocked HCV entry, as previously reported (6, 21). Overall, these findings indicate that late endosome motility is dispensable for HCV entry and subsequent RNA replication and translation but is required for viral egress.Open in a separate windowFIG. 4.HCV entry and RNA replication are not affected by inhibiting late endosome movement, and endosomal localization of core protein is not affected by inhibiting endocytosis. Huh7.5 cells were treated with the various drugs and 14 h later washed and inoculated with HCV Jc1. At 3 days p.i., cells were stained with anti-core protein antibody (green) and DAPI (blue) (A). The images were analyzed by using Metamorph and the proportion of cells (of 5,000 counted) expressing core protein was counted (B). (C) Colocalization of HCV core protein and late endosomes is not affected by DN mutants of Esp15 or Rab5a. HCV-infected cells (at day 5 p.i.) were transfected with a control plasmid pCMV-IE (C, left panel) or with dominant negative mutants of Eps15 (pEGFP-Eps15-DN) (C, middle panel) or Rab5a (pEGFP-Rab5a-DN) (C, right panel). At day 2 posttransfection, cells were labeled with antibodies specific for core protein (blue) and CD63 (red). Cells expressed EGFP-Eps15-DN or -Rab5-DN proteins (green). Nuclei were stained with DAPI (white). Enlarged views of parts of every image (insets) are shown. Colocalization of core protein and CD63 is depicted in magenta. (D) Results from colocalization analysis are shown using Zeiss LSM Zen software. Error bars represent standard deviations of the mean result from 20 cells of two experiments.Moreover, we observed an almost complete block in virus infection after expression of either an Eps15 dominant negative (DN) mutant, EGFP-Eps15D95/295 (EGFP-Eps15-DN) (4), or a Rab5a dominant negative mutant, EGFP-Rab5a-S34N (EGFP-Rab5a-DN) (30), in naïve Huh7.5 cells (data not shown). This finding is consistent with a previous report (21). We further studied the effects of the DN mutants of Eps15 and Rab5a on the colocalization of core protein and the late endosomes to confirm that this colocalization was not due to the process of virus entry. Since the DN mutants of endocytosis will block HCV infection, we first performed virus infection followed by expression of the DN mutants. This strategy has been used to demonstrate that the trafficking of HIV-1 RNA and Gag protein to late endosome is independent of the endocytosed virus (19). EGFP-Eps15-DN and EGFP-Rab5-DN were transfected into HCV-infected cells (at day 5 p.i.). As shown in Fig. Fig.4C,4C, the localization of core protein with CD63 in these cells was not affected by the expression of these DN mutants; core protein remained well colocalized with CD63 in both the cell periphery and juxtanuclear positions in the multiple-cycle HCV growth assays. The calculated efficiency of colocalization of core protein with CD63 (19 to ∼20%; Fig. Fig.4D)4D) was not altered. These results indicate that core protein localization with late endosomes is not the result of the accumulation of endocytosed viruses but rather represents the trafficking intermediates of the core protein during the late viral replication stages. Thus, we conclude that the late endosome-based secretory pathways are not involved in virus entry but rather deliver the assembled virions to the extracellular milieu.Taken together, our results indicate that HCV egress requires the motility of early to late endosomes, which is microtubule dependent, and that this pathway is independent of the one required for virus entry. Thus, we postulate that following the assembly of virus particles in juxtaposition to LDs, the HCV particles are transported through early to late endosomes to the plasma membrane, where the membrane of late endosomes is fused with plasma membrane to release virions into the extracellular milieu. This transport appears to be important for HCV egress, but it is not clear how endosomes adapt in these processes. It is possible that HCV particles are transported into endosomes after their synthesis near LD. The endosome may facilitate transport of the virus particles to the plasma membrane or even to specialized cell surface domains, such as cell junctions. Notably, the tight junction protein claudin 1 has been reported to be required for HCV cell-to-cell transmission (33), which may help viruses to sequester away from the immune system.  相似文献   

6.
The biosynthetic pathway carries cargos from the endoplasmic reticulum (ER) to the trans Golgi network (TGN) via a typical passage through the Golgi. Interestingly, large particles such as procollagen, chylomicrons and some viruses all reach the TGN by atypical routes. Given this dichotomy, we anticipated that such cargos might rely on non-classical machineries downstream of the TGN. Using Herpes simplex virus type 1 (HSV-1) as a model and a synchronized infection protocol that focuses on TGN to plasma membrane transport, the present study revealed the surprising implication of the cellular serine-threonine protein kinase D in HSV-1 egress. These findings, confirmed by a variety of complementary means [pharmacological inhibitors, dominant negative mutant, RNA interference and electron microscopy (EM)], identify one of possibly several cellular factors that modulate the egress of viruses transiting at the TGN. Moreover, the involvement of this kinase, previously known to regulate the transport of small basolateral cargos, highlights the trafficking of both small and exceptionally large entities by a common machinery downstream of the TGN, in sharp contrast to earlier steps of transport. Conceptually, this indicates the TGN is not only a sorting station from which cargos can depart towards different destinations but also a meeting point where conventional and unconventional routes can meet along the biosynthetic pathway. Lastly, given the apical release of HSV-1 in neurons, it opens up the possibility that this kinase might regulate some apical sorting.  相似文献   

7.
Assembly of human T-cell leukemia virus type 1 (HTLV-1) particles is initiated by the trafficking of virally encoded Gag polyproteins to the inner leaflet of the plasma membrane (PM). Gag–PM interactions are mediated by the matrix (MA) domain, which contains a myristoyl group (myr) and a basic patch formed by lysine and arginine residues. For many retroviruses, Gag–PM interactions are mediated by phosphatidylinositol 4,5-bisphosphate [PI(4,5)P2]; however, previous studies suggested that HTLV-1 Gag–PM interactions and therefore virus assembly are less dependent on PI(4,5)P2. We have recently shown that PI(4,5)P2 binds directly to HTLV-1 unmyristoylated MA [myr(–)MA] and that myr(–)MA binding to membranes is significantly enhanced by inclusion of phosphatidylserine (PS) and PI(4,5)P2. Herein, we employed structural, biophysical, biochemical, mutagenesis, and cell-based assays to identify residues involved in MA–membrane interactions. Our data revealed that the lysine-rich motif (Lys47, Lys48, and Lys51) constitutes the primary PI(4,5)P2–binding site. Furthermore, we show that arginine residues 3, 7, 14 and 17 located in the unstructured N-terminus are essential for MA binding to membranes containing PS and/or PI(4,5)P2. Substitution of lysine and arginine residues severely attenuated virus-like particle production, but only the lysine residues could be clearly correlated with reduced PM binding. These results support a mechanism by which HTLV-1 Gag targeting to the PM is mediated by a trio engagement of the myr group, Arg-rich and Lys-rich motifs. These findings advance our understanding of a key step in retroviral particle assembly.  相似文献   

8.
9.
10.
In most retroviruses, plasma membrane (PM) association of the Gag structural protein is a critical step in viral assembly, relying in part on interaction between the highly basic Gag MA domain and the negatively charged inner leaflet of the PM. Assembly is thought to begin with Gag dimerization followed by multimerization, resulting in a hexameric lattice. To directly address the role of multimerization in membrane binding, we fused the MA domains of Rous sarcoma virus (RSV) and HIV-1 to the chemically inducible dimerization domain FK506-binding protein (FKBP) or to the hexameric protein CcmK4 from cyanobacteria. The cellular localization of the resulting green fluorescent protein (GFP)-tagged chimeric proteins was examined by fluorescence imaging, and the association of the proteins with liposomes was quantified by flotation in sucrose gradients, following synthesis in a reticulocyte extract or as purified proteins. Four lipid compositions were tested, representative of liposomes commonly reported in flotation experiments. By themselves, GFP-tagged RSV and HIV-1 MA proteins were largely cytoplasmic, but both hexamerized proteins were highly concentrated at the PM. Dimerization led to partial PM localization for HIV-1 MA. These in vivo effects of multimerization were reproduced in vitro. In flotation analyses, the intact RSV and HIV-1 Gag proteins were more similar to multimerized MA than to monomeric MA. RNA is reported to compete with acidic liposomes for HIV-1 Gag binding, and thus we also examined the effects of RNase treatment or tRNA addition on flotation. tRNA competed with liposomes in the case of some but not all lipid compositions and ionic strengths. Taken together, our results further underpin the model that multimerization is critical for PM association of retroviral Gag proteins. In addition, they suggest that the modulation of membrane binding by RNA, as previously reported for HIV-1, may not hold for RSV.  相似文献   

11.
In the accepted model for human immunodeficiency virus preassembly in infected host cells, the anchoring to the intracellular leaflet of the membrane of the matrix domain (MA) that lies at the N-terminus of the viral Gag protein precursor appears to be one of the crucial steps for particle assembly. In this study, we simulated the membrane anchoring of human immunodeficiency virus-1 myristoylated MA protein using a coarse-grained representation of both the protein and the membrane. Our calculations first suggest that the myristoyl group could spontaneously release from its initial hydrophobic pocket before MA protein interacts with the lipid membrane. All-atom simulations confirmed this possibility with a related energy cost estimated to be ∼5 kcal.mol−1. The phosphatidylinositol (4,5) bisphosphate (PI(4,5)P2) head binds preferentially to the MA highly basic region as described in available NMR data, but interestingly without flipping of its 2′ acyl chain into the MA protein. Moreover, MA was able to confine PI(4,5)P2 lipids all around its molecular surface after having found a stable orientation at the membrane surface. Our results suggest that this orientation is dependent on Myr anchoring and that this confinement induces a lateral segregation of PI(4,5)P2 in domains. This is consistent with a PI(4,5)P2 enrichment of the virus envelope as compared to the host cell membrane.  相似文献   

12.
In the accepted model for human immunodeficiency virus preassembly in infected host cells, the anchoring to the intracellular leaflet of the membrane of the matrix domain (MA) that lies at the N-terminus of the viral Gag protein precursor appears to be one of the crucial steps for particle assembly. In this study, we simulated the membrane anchoring of human immunodeficiency virus-1 myristoylated MA protein using a coarse-grained representation of both the protein and the membrane. Our calculations first suggest that the myristoyl group could spontaneously release from its initial hydrophobic pocket before MA protein interacts with the lipid membrane. All-atom simulations confirmed this possibility with a related energy cost estimated to be ∼5 kcal.mol−1. The phosphatidylinositol (4,5) bisphosphate (PI(4,5)P2) head binds preferentially to the MA highly basic region as described in available NMR data, but interestingly without flipping of its 2′ acyl chain into the MA protein. Moreover, MA was able to confine PI(4,5)P2 lipids all around its molecular surface after having found a stable orientation at the membrane surface. Our results suggest that this orientation is dependent on Myr anchoring and that this confinement induces a lateral segregation of PI(4,5)P2 in domains. This is consistent with a PI(4,5)P2 enrichment of the virus envelope as compared to the host cell membrane.  相似文献   

13.
Investigation of the Vpu protein of HIV-1 recently uncovered a novel aspect of the mammalian innate response to enveloped viruses: retention of progeny virions on the surface of infected cells by the interferon-induced, transmembrane and GPI-anchored protein BST-2 (CD317; tetherin). BST-2 inhibits diverse families of enveloped viruses, but how it restricts viral release is unclear. Here, immuno-electron microscopic data indicate that BST-2 is positioned to directly retain nascent HIV virions on the plasma membrane of infected cells and is incorporated into virions. Virion-incorporation was confirmed by capture of infectivity using antibody to the ectodomain of BST-2. Consistent with a direct tethering mechanism, we confirmed that proteolysis releases restricted virions and further show that this removed the ectodomain of BST-2 from the cell surface. Unexpectedly, enzymatic cleavage of GPI anchors did not release restricted virions, weighing against models in which individual BST-2 molecules span the virion and host cell membranes. Although the exact molecular topology of restriction remains unsolved, we suggest that the incorporation of BST-2 into viral envelopes underlies its broad restrictive activity, whereas its relative exclusion from virions and sites of viral assembly by proteins such as HIV-1 Vpu may provide viral antagonism of restriction.  相似文献   

14.
HIV-1 Gag can assemble and generate virions at the plasma membrane, but it is also present in endosomes where its role remains incompletely characterized. Here, we show that HIV-1 RNAs and Gag are transported on endosomal vesicles positive for TiVamp, a v-SNARE involved in fusion events with the plasma membrane. Inhibition of endosomal traffic did not prevent viral release. However, inhibiting lysosomal degradation induced an accumulation of Gag in endosomes and increased viral production 7-fold, indicating that transport of Gag to lysosomes negatively regulates budding. This also suggested that endosomal Gag-RNA complexes could access retrograde pathways to the cell surface and indeed, depleting cells of TiVamp-reduced viral production. Moreover, inhibition of endosomal transport prevented the accumulation of Gag at sites of cellular contact. HIV-1 Gag could thus generate virions using two pathways, either directly from the plasma membrane or through an endosome-dependent route. Endosomal Gag-RNA complexes may be delivered at specific sites to facilitate cell-to-cell viral transmission.The production of infectious retroviral particles is an ordered process that includes many steps (for review see Refs. 13). In particular, three major viral components, Gag, the envelope, and genomic RNAs have to traffic inside the cell to reach their assembly site. Viral biogenesis is driven by the polyprotein Gag, which is able to make viral-like particles when expressed alone (4). Upon release, HIV-14 Gag is processed by the viral protease into matrix (MA(p17)), capsid (CA(p24)), nucleocapsid (NC(p7)), p6, and smaller peptides SP1 and SP2. Gag contains several domains that are essential for viral assembly: a membrane binding domain (M) in MA; a Gag-Gag interaction domain in CA; an assembly domain (I) in NC; and a late domain (L) in p6, which recruits the cellular budding machinery. Genomic RNAs are specifically recognized by NC, and they play fundamental roles in viral biogenesis by acting as a scaffold for Gag multimerization (5).It has been demonstrated that retroviruses bud by hijacking the endosomal machinery that sorts proteins into internal vesicles of multivesicular bodies (for review, see Refs. 6, 7). Indeed, these vesicles bud with the same topology as viral particles. Proteins sorted into this pathway are usually destined for degradation in lysosomes, but some can also recycle to the plasma membrane (for review see Refs. 8, 9). They are also frequently ubiquitinated on their cytoplasmic domain (10, 11), allowing their recognition by ESCRT complexes. ESCRT-0 and ESCRT-I recognize ubiquitinated cargo present at the surface of endosomes and recruit other ESCRT complexes (1214). ESCRT-III is believed to function directly in the formation of multivesicular body intralumenal vesicles (12), even though its mechanism of action is currently not understood. Remarkably, Gag L domains interact directly with components of the multivesicular body-sorting machinery (for review see Ref. 15). HIV-1 Gag uses a PTAP motif to bind Tsg101, a component of ESCRT-I (1619), and a YPLTSL motif to interact with Alix, a protein linked to ESCRT-I and -III (2022). Finally, various ubiquitin ligases are also required directly or indirectly during HIV-1 biogenesis (23, 24; for review see Ref. 25).In many cell lines, Gag is found both at the plasma membrane and in endosomes. This has led to the hypothesis that there are several assembly sites for HIV-1 (1, 3). First, Gag can initiate and complete assembly at the plasma membrane. This is thought to occur predominantly in T lymphocytes, and this process is supported by several lines of evidences: (i) disruption of endosomal trafficking with drugs does not prevent viral production (26, 27); (ii) ESCRT complexes can be recruited at the plasma membrane, at sites where Gag accumulates (2830); (iii) Gag can be seen multimerizing and budding from the plasma membrane in live cells (31). Second, Gag could initiate assembly in endosomes, and then traffic to the cell surface to be released. This is mainly supported by the presence of Gag in endosomes in several cell lines (3234), including T cells and more strikingly macrophages (32, 35, 3639). However, we are currently lacking functional experiments addressing the role of this endosomal pool of Gag, and it is still not clear to what extent it contributes to the production of viral particles. Nevertheless, the presence of Gag in endosomes might facilitate recruitment of ESCRT complexes (34, 40), packaging of viral genomic RNAs (32, 41), and incorporation of the envelope (42). It may also be important for polarized budding (43, 44) and to create a viral reservoir in infected cells (45, 46).Despite great progress, the traffic of HIV-1 components is still not fully elucidated. In particular, the transport of the genomic RNAs is poorly understood. In this study, we have used single molecule techniques to investigate the trafficking of HIV-1 RNAs in fixed and live cells, and we show that they are transported on endosomal vesicles. We also obtained functional evidence that Gag and viral RNAs can use at least two trafficking pathways to produce virions, one going directly from the plasma membrane and another one passing through endosomes.  相似文献   

15.
We have constructed a series of human immunodeficiency virus (HIV) gag mutants by progressive truncation of the gag coding sequence from the C terminus and have combined these mutants with an assembly-competent matrix domain deletion mutation (ΔMA). By using several methods, the particle-producing capabilities of each mutant were examined. Our analysis indicated that truncated Gag precursors lacking most of C-terminal gag gene products assembled and were released from 293T cells. Additionally, a mutant with a combined deletion of the MA (ΔMA) and p6 domains even produced particles at levels comparable to that of the wild-type (wt) virus. However, most mutants derived from combination of the ΔMA and the C-terminal truncation mutations did not release particles as well as the wt. Our smallest HIV gag gene product capable of virus-like particle formation was a 28-kDa protein which consists of a few MA amino acids and the CA-p2 domain. Sucrose density gradient fractionation analysis indicated that most mutants exhibited a wt retrovirus particle density. Exceptions to this rule were mutants with an intact MA domain but deleted downstream of the p2 domains. These C-terminal truncation mutants possessed particle densities of 1.13 to 1.15 g/ml, lower than that of the wt. The N-terminal portions of the CA domain, which have been shown to be dispensable for core assembly, became critical when most of the MA domain was deleted, suggesting a requirement for an intact CA domain to assemble and release particles.  相似文献   

16.
Groups of distinct but related diseases often share common symptoms, which suggest likely overlaps in underlying pathogenic mechanisms. Identifying the shared pathways and common factors among those disorders can be expected to deepen our understanding for them and help designing new treatment strategies effected on those diseases. Neurodegeneration diseases, including Alzheimer''s disease (AD), Parkinson''s disease (PD) and Huntington''s disease (HD), were taken as a case study in this research. Reported susceptibility genes for AD, PD and HD were collected and human protein-protein interaction network (hPPIN) was used to identify biological pathways related to neurodegeneration. 81 KEGG pathways were found to be correlated with neurodegenerative disorders. 36 out of the 81 are human disease pathways, and the remaining ones are involved in miscellaneous human functional pathways. Cancers and infectious diseases are two major subclasses within the disease group. Apoptosis is one of the most significant functional pathways. Most of those pathways found here are actually consistent with prior knowledge of neurodegenerative diseases except two cell communication pathways: adherens and tight junctions. Gene expression analysis showed a high probability that the two pathways were related to neurodegenerative diseases. A combination of common susceptibility genes and hPPIN is an effective method to study shared pathways involved in a group of closely related disorders. Common modules, which might play a bridging role in linking neurodegenerative disorders and the enriched pathways, were identified by clustering analysis. The identified shared pathways and common modules can be expected to yield clues for effective target discovery efforts on neurodegeneration.  相似文献   

17.
Retroviral Gag targeting to the plasma membrane (PM) for assembly is mediated by the N-terminal matrix (MA) domain. For many retroviruses, Gag–PM interaction is dependent on phosphatidylinositol 4,5-bisphosphate (PI(4,5)P2). However, it has been shown that for human T-cell leukemia virus type 1 (HTLV-1), Gag binding to membranes is less dependent on PI(4,5)P2 than HIV-1, suggesting that other factors may modulate Gag assembly. To elucidate the mechanism by which HTLV-1 Gag binds to the PM, we employed NMR techniques to determine the structure of unmyristoylated MA (myr(–)MA) and to characterize its interactions with lipids and liposomes. The MA structure consists of four α-helices and unstructured N- and C-termini. We show that myr(–)MA binds to PI(4,5)P2 via the polar head and that binding to inositol phosphates (IPs) is significantly enhanced by increasing the number of phosphate groups on the inositol ring, indicating that the MA–IP binding is governed by charge–charge interactions. The IP binding site was mapped to a well-defined basic patch formed by lysine and arginine residues. Using an NMR-based liposome binding assay, we show that PI(4,5)P2 and phosphatidylserine enhance myr(–)MA binding in a synergistic fashion. Confocal microscopy data revealed formation of puncta on the PM of Gag expressing cells. However, G2A-Gag mutant, lacking myristoylation, is diffuse and cytoplasmic. These results suggest that although myr(–)MA binds to membranes, myristoylation appears to be key for formation of HTLV-1 Gag puncta on the PM. Altogether, these findings advance our understanding of a key mechanism in retroviral assembly.  相似文献   

18.
19.
20.
本研究目的是了解表达HIV-1六种基因的非复制型重组痘苗病毒(rNTV-C)的遗传稳定性(包括病毒载体和六种外源基因:gp160、gag、polr、evt、at和nef)。我们将rNTV-C在原代鸡胚成纤维细胞(CEF)中连续传代至25代,对第9、12、15以及25代病毒载体基因的稳定性、六种外源基因的稳定性、外源基因表达的稳定性以及外源基因的丢失率进行研究。结果显示:各代病毒均保持了非复制型天坛株痘苗病毒载体特点且传代稳定;HIV-1目的基因序列与原设计序列相符、重组位点正确且遗传稳定,连续传25代核苷酸突变率低于万分之一;目的蛋白在各代rNTV-C中均能有效表达,而且各代病毒之间表达量和分子量无明显差别;以Gag和Nef蛋白表达为标记,rNTV-C各代病毒的基因丢失率均在5%以下。本研究结果为疫苗生产提供了确保毒种稳定的关键资料。  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