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 546 毫秒
1.
2.
3.
4.
5.
Streptomyces development was analyzed under conditions resembling those in soil. The mycelial growth rate was much lower than that in standard laboratory cultures, and the life span of the previously named first compartmentalized mycelium was remarkably increased.Streptomycetes are gram-positive, mycelium-forming, soil bacteria that play an important role in mineralization processes in nature and are abundant producers of secondary metabolites. Since the discovery of the ability of these microorganisms to produce clinically useful antibiotics (2, 15), they have received tremendous scientific attention (12). Furthermore, its remarkably complex developmental features make Streptomyces an interesting subject to study. Our research group has extended our knowledge about the developmental cycle of streptomycetes, describing new aspects, such as the existence of young, fully compartmentalized mycelia (5-7). Laboratory culture conditions (dense inocula, rich culture media, and relatively elevated temperatures [28 to 30°C]) result in high growth rates and an orderly-death process affecting these mycelia (first death round), which is observed at early time points (5, 7).In this work, we analyzed Streptomyces development under conditions resembling those found in nature. Single colonies and soil cultures of Streptomyces antibioticus ATCC 11891 and Streptomyces coelicolor M145 were used for this analysis. For single-colony studies, suitable dilutions of spores of these species were prepared before inoculation of plates containing GYM medium (glucose, yeast extract, malt extract) (11) or GAE medium (glucose, asparagine, yeast extract) (10). Approximately 20 colonies per plate were obtained. Soil cultures were grown in petri dishes with autoclaved oak forest soil (11.5 g per plate). Plates were inoculated directly with 5 ml of a spore suspension (1.5 × 107 viable spores ml−1; two independent cultures for each species). Coverslips were inserted into the soil at an angle, and the plates were incubated at 30°C. To maintain a humid environment and facilitate spore germination, the cultures were irrigated with 3 ml of sterile liquid GAE medium each week.The development of S. coelicolor M145 single colonies growing on GYM medium is shown in Fig. Fig.1.1. Samples were collected and examined by confocal microscopy after different incubation times, as previously described (5, 6). After spore germination, a viable mycelium develops, forming clumps which progressively extend along the horizontal (Fig. 1a and b) and vertical (Fig. 1c and d) axes of a plate. This mycelium is fully compartmentalized and corresponds to the first compartmentalized hyphae previously described for confluent surface cultures (Fig. 1e, f, and j) (see below) (5); 36 h later, death occurs, affecting the compartmentalized hyphae (Fig. 1e and f) in the center of the colony (Fig. (Fig.1g)1g) and in the mycelial layers below the mycelial surface (Fig. 1d and k). This death causes the characteristic appearance of the variegated first mycelium, in which alternating live and dead segments are observed (Fig. 1f and j) (5). The live segments show a decrease in fluorescence, like the decrease in fluorescence that occurs in solid confluent cultures (Fig. (Fig.11 h and i) (5, 9). As the cycle proceeds, the intensity of the fluorescence in these segments returns, and the segments begin to enlarge asynchronously to form a new, multinucleated mycelium, consisting of islands or sectors on the colony surfaces (Fig. 1m to o). Finally, death of the deeper layers of the colony (Fig. (Fig.1q)1q) and sporulation (Fig. (Fig.1r)1r) take place. Interestingly, some of the spores formed germinate (Fig. (Fig.1s),1s), giving rise to a new round of mycelial growth, cell death, and sporulation. This process is repeated several times, and typical, morphologically heterogeneous Streptomyces colonies grow (not shown). The same process was observed for S. antibioticus ATCC 11891, with minor differences mainly in the developmental time (not shown).Open in a separate windowFIG. 1.Confocal laser scanning fluorescence microscopy analysis of the development-related cell death of S. coelicolor M145 in surface cultures containing single colonies. Developmental culture times (in hours) are indicated. The images in panels l and n were obtained in differential interference contrast mode and correspond to the same fields as in panels k and m, respectively. The others are culture sections stained with SYTO 9 and propidium iodide. Panels c, d, k, l, p, and q are cross sections; the other images are longitudinal sections (see the methods). Panels h and i are images of the same field taken with different laser intensities, showing low-fluorescence viable hyphae in the center of the colonies that develop into a multinucleated mycelium. The arrows in panels e and s indicate septa (e) and germinated spores (s). See the text for details.Figure Figure22 shows the different types of mycelia present in S. coelicolor cultures under the conditions described above, depending on the compartmentalization status. Hyphae were treated with different fluorescent stains (SYTO 9 plus propidium iodide for nucleic acids, CellMask plus FM4-64 for cell membranes, and wheat germ agglutinin [WGA] for cell walls). Samples were processed as previously described (5). The young initial mycelia are fully compartmentalized and have membranous septa (Fig. 2b to c) with little associated cell wall material that is barely visible with WGA (Fig. (Fig.2d).2d). In contrast, the second mycelium is a multinucleated structure with fewer membrane-cell wall septa (Fig. 2e to h). At the end of the developmental cycle, multinucleated hyphae begin to undergo the segmentation which precedes the formation of spore chains (Fig. 2i to m). Similar results were obtained for S. antibioticus (not shown), but there were some differences in the numbers of spores formed. Samples of young and late mycelia were freeze-substituted using the methodology described by Porta and Lopez-Iglesias (13) and were examined with a transmission electron microscope (Fig. 2n and o). The septal structure of the first mycelium (Fig. (Fig.2n)2n) lacks the complexity of the septal structure in the second mycelium, in which a membrane with a thick cell wall is clearly visible (Fig. (Fig.2o).2o). These data coincide with those previously described for solid confluent cultures (4).Open in a separate windowFIG. 2.Analysis of S. coelicolor hyphal compartmentalization with several fluorescent indicators (single colonies). Developmental culture times (in hours) are indicated. (a, e, and i) Mycelium stained with SYTO 9 and propidium iodide (viability). (b, f, and j) Hyphae stained with Cell Mask (a membrane stain). (c, g, and l) Hyphae stained with FM 4-64 (a membrane stain). (d, h, and m) Hyphae stained with WGA (cell wall stain). Septa in all the images in panels a to j, l, and m are indicated by arrows. (k) Image of the same field as panel j obtained in differential interference contrast mode. (n and o) Transmission electron micrographs of S. coelicolor hyphae at different developmental phases. The first-mycelium septa (n) are comprised of two membranes separated by a thin cell wall; in contrast, second-mycelium septa have thick cell walls (o). See the text for details. IP, propidium iodide.The main features of S. coelicolor growing in soils are shown in Fig. Fig.3.3. Under these conditions, spore germination is a very slow, nonsynchronous process that commences at about 7 days (Fig. 3c and d) and lasts for at least 21 days (Fig. 3i to l), peaking at around 14 days (Fig. 3e to h). Mycelium does not clump to form dense pellets, as it does in colonies; instead, it remains in the first-compartmentalized-mycelium phase during the time analyzed. Like the membrane septa in single colonies, the membrane septa of the hyphae are stained with FM4-64 (Fig. 3j and k), although only some of them are associated with thick cell walls (WGA staining) (Fig. (Fig.3l).3l). Similar results were obtained for S. antibioticus cultures (not shown).Open in a separate windowFIG. 3.Confocal laser scanning fluorescence microscopy analysis of the development-related cell death and hyphal compartmentalization of S. coelicolor M145 growing in soil. Developmental culture times (in days) are indicated. The images in panels b, f, and h were obtained in differential interference contrast mode and correspond to the same fields as the images in panels a, e, and g, respectively. The dark zone in panel h corresponds to a particle of soil containing hyphae. (a, c, d, e, g, i, j, and k) Hyphae stained with SYTO 9, propidium iodide (viability stain), and FM4-64 (membrane stain) simultaneously. (i) SYTO 9 and propidium iodide staining. (j) FM4-64 staining. The image in panel k is an overlay of the images in panels i and j and illustrates that first-mycelium membranous septa are not always apparent when they are stained with nucleic acid stains (SYTO 9 and propidium iodide). (l) Hyphae stained with WGA (cell wall stain), showing the few septa with thick cell walls present in the cells. Septa are indicated by arrows. IP, propidium iodide.In previous work (8), we have shown that the mycelium currently called the substrate mycelium corresponds to the early second multinucleated mycelium, according to our nomenclature, which still lacks the hydrophobic layers characteristic of the aerial mycelium. The aerial mycelium therefore corresponds to the late second mycelium which has acquired hydrophobic covers. This multinucleated mycelium as a whole should be considered the reproductive structure, since it is destined to sporulate (Fig. (Fig.4)4) (8). The time course of lysine 6-aminotransferase activity during cephamycin C biosynthesis has been analyzed by other workers using isolated colonies of Streptomyces clavuligerus and confocal microscopy with green fluorescent protein as a reporter (4). A complex medium and a temperature of 29°C were used, conditions which can be considered similar to the conditions used in our work. Interestingly, expression did not occur during the development of the early mycelium and was observed in the mycelium only after 80 h of growth. This suggests that the second mycelium is the antibiotic-producing mycelium, a hypothesis previously confirmed using submerged-growth cultures of S. coelicolor (9).Open in a separate windowFIG. 4.Cell cycle features of Streptomyces growing under natural conditions. Mycelial structures (MI, first mycelium; MII, second mycelium) and cell death are indicated. The postulated vegetative and reproductive phases are also indicated (see text).The significance of the first compartmentalized mycelium has been obscured by its short life span under typical laboratory culture conditions (5, 6, 8). In previous work (3, 7), we postulated that this structure is the vegetative phase of the bacterium, an hypothesis that has been recently corroborated by proteomic analysis (data not shown). Death in confluent cultures begins shortly after germination (4 h) and continues asynchronously for 15 h. The second multinucleated mycelium emerges after this early programmed cell death and is the predominant structure under these conditions. In contrast, as our results here show, the first mycelium lives for a long time in isolated colonies and soil cultures. As suggested in our previous work (5, 6, 8), if we assume that the compartmentalized mycelium is the Streptomyces vegetative growth phase, then this phase is the predominant phase in individual colonies (where it remains for at least 36 h), soils (21 days), and submerged cultures (around 20 h) (9). The differences in the life span of the vegetative phase could be attributable to the extremely high cell densities attained under ordinary laboratory culture conditions, which provoke massive differentiation and sporulation (5-7, 8).But just exactly what are “natural conditions”? Some authors have developed soil cultures of Streptomyces to study survival (16, 17), genetic transfer (14, 17-19), phage-bacterium interactions (3), and antibiotic production (1). Most of these studies were carried out using amended soils (supplemented with chitin and starch), conditions under which growth and sporulation were observed during the first few days (1, 17). These conditions, in fact, might resemble environments that are particularly rich in organic matter where Streptomyces could conceivably develop. However, natural growth conditions imply discontinuous growth and limited colony development (20, 21). To mimic such conditions, we chose relatively poor but more balanced carbon-nitrogen soil cultures (GAE medium-amended soil) and less dense spore inocula, conditions that allow longer mycelium growth times. Other conditions assayed, such as those obtained by irrigating the soil with water alone, did not result in spore germination and mycelial growth (not shown). We were unable to detect death, the second multinucleated mycelium described above, or sporulation, even after 1 month of incubation at 30°C. It is clear that in nature, cell death and sporulation must take place at the end of the long vegetative phase (1, 17) when the imbalance of nutrients results in bacterial differentiation.In summary, the developmental kinetics of Streptomyces under conditions resembling conditions in nature differs substantially from the developmental kinetics observed in ordinary laboratory cultures, a fact that should be born in mind when the significance of development-associated phenomena is analyzed.  相似文献   

6.
7.
8.
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   

9.
Borna disease virus (BDV) is characterized by highly neurotropic infection. BDV enters its target cells using virus surface glycoprotein (G), but the cellular molecules mediating this process remain to be elucidated. We demonstrate here that the N-terminal product of G, GP1, interacts with the 78-kDa chaperone protein BiP. BiP was found at the surface of BDV-permissive cells, and anti-BiP antibody reduced BDV infection as well as GP1 binding to the cell surface. We also reveal that BiP localizes at the synapse of neurons. These results indicate that BiP may participate in the cell surface association of BDV.Borna disease virus (BDV) belongs to the Bornaviridae family of nonsegmented, negative-strand RNA viruses and is characterized by highly neurotropic and noncytopathic infection (18, 33). BDV infects a wide variety of host species and causes central nervous system (CNS) diseases in animals, which are frequently associated with behavioral disorders (14, 19, 29, 31). BDV cell entry is mediated by endocytosis, following the attachment of viral envelope glycoprotein (G) to the cellular receptor (2, 7, 8). BDV G is translated as a precursor protein, GP, which is posttranslationally cleaved by the cellular protease furin to generate two functional subunits of the N (GP1) and C (GP2) termini (28). Recent studies revealed that GP1 is involved in virus interaction with as-yet-unidentified cell surface receptor(s) and that GP2 mediates a pH-dependent fusion event between viral and cell membranes (2, 7, 27). In addition, a previous work using a hippocampal culture system suggested that BDV G is required for viral dissemination in neurons (2); however, cellular factors involved in BDV cell entry, especially cell surface association, remain to be elucidated.To extend our understanding of the role of BDV G in the interaction with the cell plasma membrane, we transfected GP1 fused with hemagglutinin-tobacco etch virus protease cleavage site-FLAG tags (GP1-TAP) into human oligodendroglioma OL cells. GP1-TAP was purified using anti-FLAG M2 affinity gel (Sigma). To verify that GP1-TAP binds to OL cells, the cells were incubated with 4 μg/ml GP1-TAP, and binding was detected by anti-FLAG M2 antibody (Sigma). A flow cytometric analysis indicated that GP1-TAP binds to OL cells (Fig. (Fig.1A).1A). To further validate the binding of GP1-TAP, we tested whether GP1-TAP inhibits BDV infection. OL cells were pretreated with 4 μg/ml GP1-TAP for 30 min. Proteins purified from mock-transfected cells using an anti-FLAG M2 affinity gel served as a control. The cells were then mixed with cell-free BDV. After 1 h of absorption, the supernatants were removed and fresh medium was added. At 3 days postinfection, the viral antigens were stained with anti-nucleoprotein (N) monoclonal and anti-matrix (M) polyclonal antibodies. As shown in Fig. Fig.1B,1B, GP1-TAP reduced BDV infection by 40% compared to levels for mock-treated cells. This result was consistent with earlier reports showing that recombinant GP1 protein binds to the cell surface and inhibits BDV infection (6, 20).Open in a separate windowFIG. 1.BDV GP1 binds to the cell surface. (A) Binding of BDV GP1 to OL cells. OL cells were incubated with GP1-TAP (solid line), and its binding was detected using anti-FLAG M2 antibody and flow cytometry. As a control, cells incubated with proteins purified from mock-transfected cells were detected by an anti-FLAG M2 antibody (dotted line). (B) Inhibition of BDV infection by GP1. OL cells pretreated with GP1-TAP were inoculated with the BDV huP2br strain. Values are the means + standard deviations (SD) from three independent experiments. **, P < 0.01.To investigate the host factor(s) that mediates the interaction of GP1 with the cell surface, a combination of tandem affinity purification (TAP) and liquid chromatography tandem mass spectrometry analyses was designed (13). We transfected GP1-TAP into OL cells and then purified GP1 from cell homogenates using a TAP strategy. We compared the purified proteins from the whole-cell and cytosol fractions (Fig. (Fig.2A),2A), and the bands detected only in the whole-cell fraction were determined as GP1-binding proteins in the membrane and/or nuclear fractions. In addition to GP1 protein (Fig. (Fig.2A,2A, arrow), we identified a specific band around 80 kDa in the whole-cell homogenate, but not in the cytosol fraction (Fig. (Fig.2A,2A, arrowhead), and determined that the band corresponded to the BiP (immunoglobulin heavy chain-binding protein) molecular chaperone, also called glucose-regulated protein 78 (GRP78), by mass spectrometry analysis. We confirmed the specific interaction between endogenous BiP and BDV G in infected cells by immunoprecipitation analysis (Fig. (Fig.2B).2B). To map the binding domain on BiP to GP1, we constructed a series of deletion mutants of the green fluorescent protein (GFP)-tagged BiP plasmid (Fig. (Fig.2C).2C). We transfected the mutant plasmids into BDV-infected OL cells and then performed an immunoprecipitation assay using anti-GFP antibody (Invitrogen). As shown in Fig. Fig.2D,2D, BDV G was coimmunoprecipitated with truncated BiP mutants, except for BiPΔN-GFP, which lacks the ATP-binding domain of BiP (lane 3), suggesting that BiP interacts with GP1 via its N-terminal region.Open in a separate windowFIG. 2.BDV GP1 interacts with BiP molecular chaperone. (A) TAP analysis of BDV GP1. Proteins coimmunoprecipitated with GP1-TAP in OL cells were separated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and visualized by silver staining. Cyt, cytosol fraction; Wc, whole-cell homogenate. Arrow, GP1-TAP; arrowhead, BiP. (B) Coimmunoprecipitation (IP) of BDV G and endogenous BiP. BDV G was immunoprecipitated from BDV-infected OL cells by anti-BDV G polyclonal antibody. Endogenous BiP was then detected by anti-BiP monoclonal antibody (Becton Dickinson). IgG, immunoglobulin G. (C) Schematic representation of deletion mutants of recombinant BiP-GFP. The known functional regions are indicated. (D) Immunoprecipitation analysis of BiP-GFP mutants in BDV-infected OL cells. The deletion plasmids were transfected and immunoprecipitated by anti-GFP antibody. Specific binding was detected using anti-BDV G antibody. Lane 1, GFP; lane 2, BiP-GFP; lane 3, BiPΔN-GFP; lane 4, BiPΔPB-GFP; lane 5, BiPΔC-GFP.BiP is known to be resident primarily in the endoplasmic reticulum and functions as a molecular chaperone involved in the folding process of nascent proteins, mostly through interaction with its peptide-binding domain (12, 17, 21). On the other hand, BiP has been reported to serve as a coreceptor of certain viruses at the plasma membrane (15, 34). Recent studies also revealed that cell surface BiP mediates the internalization of its ligands into cells (1, 10). We first investigated whether BiP is expressed on the cell surface of BDV-permissive OL and 293T cells using an anti-BiP polyclonal antibody (H-129; Santa Cruz Biotechnology, Inc.). As shown in Fig. Fig.3A,3A, BiP expression is detected on the surface of both cell lines. This result is in agreement with recent observations that BiP is expressed on the surface of various types of cells (9, 10, 15, 23, 24, 34). We also investigated whether BiP is expressed on the cell surface of BDV-nonpermissive cell lines, such as HeLa and CHO cells. As shown in Fig. Fig.3A,3A, we detected BiP expression on the surface of HeLa, but not CHO, cells. These observations were confirmed by immunofluorescence analysis (Fig. (Fig.3B).3B). Note that BiP is clearly detected at the endoplasmic reticulum in the permeabilized CHO cells by the antibody (see Fig. S1 in the supplemental material), suggesting that BiP is expressed at a very low level, if at all, on the surface of CHO cells. We next examined whether cell surface BiP serves as a binding molecule of BDV GP1. To test this, we performed an inhibition assay using an anti-BiP polyclonal antibody (N-20; Santa Cruz Biotechnology, Inc.) which recognizes the N terminus of BiP. As shown in Fig. Fig.3C,3C, the antibody inhibited GP1 binding to the cell surface by 40%. Furthermore, BDV infection was found to decrease by 70% when cells were treated with the antibody (Fig. (Fig.3D3D).Open in a separate windowFIG. 3.Cell surface BiP mediates cell association of BDV. (A) Flow cytometric analysis was performed with anti-BiP antibody (H-129) in BDV-permissive (OL and 293T) and -nonpermissive (HeLa and CHO) cells (solid lines). Cells stained with normal rabbit immunoglobulin G were used as a control (dotted lines). (B) Immunofluorescence analysis was performed by using anti-BiP antibody (H-129) with BDV-permissive and -nonpermissive cells. Arrows indicate BiP staining at the membrane. Scale bars, 10 μm. (C) Inhibition of GP1 binding by anti-BiP antibody (N-20). OL cells were pretreated with anti-BiP antibody, followed by labeling with GP1. GP1 binding on the cell surface was detected using flow cytometry. Values are the means + SD from three independent experiments. *, P < 0.05. (D) Inhibition of BDV infection by anti-BiP antibody. OL cells were incubated with 10 μg/ml anti-BiP antibody or normal goat immunoglobulin G and then the cells were mixed with cell-free BDV. After 1 h absorption, the supernatants were replaced with fresh medium. Virus infection was measured by immunofluorescence analysis using anti-N and -M antibodies at 3 days postinfection. Values are the means + SD from three independent experiments. *, P < 0.05. IgG, immunoglobulin G.To investigate the role of cell surface BiP in the infection of BDV, the BiP expression was inhibited by short interfering RNA (siRNA) in OL cells (see Fig. S2A in the supplemental material). We selected an siRNA (Hs_HSPA5_4; Qiagen, Inc.) which could partially downregulate the cell surface expression of BiP (see Fig. S2B in the supplemental material). However, siRNA treatment of BiP did not influence the infectivity of BDV in OL cells (see Fig. S2C in the supplemental material). This may be due to an incomplete reduction of BiP expression on the cell surface. Alternatively, while BiP mediates at least in part the cell surface association of BDV particles, this result may exhibit the presence of another, as-yet-unidentified BDV G-binding protein that is involved in the binding and subsequent cell entry of BDV.Previous studies demonstrated that BDV can be traced centripetally and transsynaptically after olfactory, ophthalmic, or intraperitoneal inoculation (3, 25). Migration of BDV to the CNS after footpad infection can be prevented by sciatic nerve transection (3). These observations suggest that BDV may disseminate primarily via neural networks. Recently, it has been demonstrated that BDV G was expressed at the termini of neurites or at contact sites of neurites (2), suggesting that local assembly of BDV may take place at the presynaptic terminals of synapses, similar to assembly of other neurotropic viruses (22, 26, 32). If BiP localizes at synapse sites, BiP may efficiently participate in the transmission of BDV particles at the synapses. To evaluate this hypothesis, we examined BiP localization in primary culture of mouse hippocampal neurons. After in vitro culture for 17 days, BiP localization was determined by an immunofluorescence assay without permeabilization. As shown in Fig. Fig.4A,4A, BiP signals were clearly detected at neurites, including the contact sites between dendrites and axons, as punctate staining (arrows), suggesting that BiP is expressed at the neuronal surface, most likely at the synapses. We next examined the localization of BiP with postsynaptic density 95 (PSD-95), a marker of postsynaptic density (5). Although BiP signals were detected mainly in the perinuclear area of the hippocampal neurons, punctate staining was also found at neurites colocalized with PSD-95 (Fig. (Fig.4B,4B, arrows). Taken together, these observations suggested that BiP is distributed at the synaptic surface, including the postsynaptic membrane, of neurons, a possible site for BDV budding and entry (2).Open in a separate windowFIG. 4.BiP localizes at the synaptic surface of hippocampus neurons. (A) Localization of BiP at synaptic surface. Hippocampal neurons were immunostained with anti-BiP antibody (N-20) without permeabilization. A differential interference contrast (DIC) image is shown. Dotted lines in the Merge panel indicate the dendrite outline. Arrows indicate BiP staining at the contact sites between axons and dendrites. (B) Colocalization between BiP and a postsynaptic protein. Hippocampal neurons were immunostained with anti-BiP (N-20) and anti-PSD-95 (Millipore) antibodies. Arrows indicate colocalized signals of BiP and PSD-95 at neurites. Scale bars, 10 μm.In summary, this study demonstrates that BiP is a GP1-binding protein at the synaptic surface. This is the first report showing the BDV G-binding factor on the cell surface. The first step of BDV entry might be mediated by the interaction of GP1 with as-yet-unidentified cell surface receptors, which may form a complex with other molecules, such as BiP. We showed that treatment with anti-BiP antibody affects BDV infection as well as GP1 binding to the cell surface (Fig. (Fig.3).3). Furthermore, synaptic distribution of BiP was found in hippocampal primary neurons (Fig. (Fig.4).4). These findings strongly suggest that BiP plays critical roles in BDV association with the neuronal surface via interaction with GP1. On the other hand, a BDV-nonpermissive cell line, HeLa, appeared to express BiP on the cell surface, suggesting that the cell surface BiP may not be necessarily involved in the infectivity of BDV. A recent study by Clemente et al. (6) revealed that following initial attachment to the cell surface, BDV is recruited to the plasma membrane lipid raft (LR) prior to internalization of the particles. The study suggested that BDV may use the LR as a platform to interact with additional host cell factor(s) required for efficient BDV internalization. Because BiP does not contain transmembrane regions, BiP needs another host protein(s) with transmembrane regions on the cell surface. It has been reported that cell surface BiP interacts with diverse proteins, such as major histocompatibility complex class I molecules (34), the voltage-dependent anion channel (9), and the DnaJ-like protein MTJ-1 (4), all of which associate with LR in the plasma membrane (16, 24, 35). Once BDV has attached to the cell surface, it might utilize such BiP-associated LR proteins for efficient cell surface attachment or internalization. Previously, it has been proposed that kainate 1 (KA-1) receptor might represent the BDV receptor within the CNS (11). Because some glutamate receptors are shown to bind to BiP (30), KA-1 receptors might interact with BiP and serve as a receptor complex for BDV. Further studies are required for a full understanding of the cell association processes, especially receptor binding, of BDV.   相似文献   

10.
11.
Vector insect cells infected with Rice gall dwarf virus, a member of the family Reoviridae, contained the virus-associated microtubules adjacent to the viroplasms, as revealed by transmission electron, electron tomographic, and confocal microscopy. The viroplasms, putative sites of viral replication, contained the nonstructural viral proteins Pns7 and Pns12, as well as core protein P5, of the virus. Microtubule-depolymerizing drugs suppressed the association of viral particles with microtubules and prevented the release of viruses from cells without significantly affecting viral multiplication. Thus, microtubules appear to mediate viral transport within and release of viruses from infected vector cells.Rice gall dwarf virus (RGDV), Rice dwarf virus (RDV), and Wound tumor virus, members of the genus Phytoreovirus in the family Reoviridae, multiply both in plants and in invertebrate insect vectors. Each virus exists as icosahedral particles of approximately 65 to 70 nm in diameter, with two concentric layers (shells) of proteins that enclose a core (1, 13). The viral genome of RGDV consists of 12 segmented double-stranded RNAs that encode six structural (P1, P2, P3, P5, P6, and P8) and six nonstructural (Pns4, Pns7, Pns9, Pns10, Pns11, and Pns12) proteins (reference 21 and references therein). The core capsid is composed of P3, the major protein, which encloses P1, P5, and P6 (12). The outer layer consists of two proteins, namely, P2 and P8 (10, 12).Cytoplasmic inclusion bodies, known as viroplasms or viral factories, are assumed to be the sites of replication of viruses in the family Reoviridae. After infecting insect vector cell monolayers (VCMs) in culture with RDV, Wei et al. (19) examined the generation of RDV particles in and at the periphery of such viroplasms. VCMs are also useful for studies of RGDV, allowing detailed analysis of the synchronous replication and multiplication of this virus (14). In order to identify the viroplasms in RGDV-infected VCMs, we examined the subcellular localization of Pns7, Pns12, P5, and RGDV particles by confocal immunofluorescence microscopy. Pns7 and Pns12 of RGDV correspond to Pns6 and Pns11, respectively, which are components of the viroplasm of RDV (12, 19). RGDV P5 is a counterpart of RDV P5, a core protein that locates inside the viroplasm in RDV-infected cells. We inoculated VCMs with RGDV, purified by the method reported in reference 15, at a multiplicity of infection (MOI) of 1; fixed them 48 h postinfection (p.i.); probed the cells with Pns7-, Pns12-, P5-, and viral-antigen-specific antibodies (11, 12) that had been conjugated to fluorescein isothiocyanate (FITC) (Sigma, St. Louis, MO) or rhodamine (Sigma); and examined them by confocal microscopy, as described previously (19). In RGDV-infected cells, Pns7, Pns12, and P5 were detected as punctate inclusions (Fig. (Fig.1).1). Immunostained viral antigens formed ringlike structures around the punctate inclusions. When the images were merged, Pns7, Pns12, and P5 were colocalized in the punctate inclusions, indicating that these proteins were constituents of the viral inclusions (Fig. (Fig.1).1). Our observations revealed the similar respective localizations of the corresponding nonstructural proteins, core proteins, and viral particles of two phytoreoviruses, RGDV and RDV, in infected cells. Thus, Pns7 and Pns12 of RGDV had attributes common to their functional counterparts—Pns6 and Pns11, respectively—of RDV (19). The core protein P5 was located inside the viroplasms, and the viral antigens were distributed at the periphery of the viroplasms. The results, together, suggest that RGDV and RDV exploit similar replication strategies. Specific fluorescence was not detected in noninfected cells after incubation with Pns7-, Pns12-, P5-, and viral-antigen-specific antibodies (data not shown).Open in a separate windowFIG. 1.Subcellular localization of Pns7, Pns12, and P5 of RGDV and viral antigens in RGDV-infected VCMs 48 h p.i. Arrowheads show ringlike profiles of viral antigens that surround viral inclusions, which have been immunostained with the Pns12-specific antibodies. Arrows show the fibrillar profiles of immunostained viral antigens. Bars, 5 μm.In addition to the viral location at the periphery of the viral inclusions visualized as immunostained Pns12 (Fig. (Fig.1),1), the antigens were distributed as bundles of fibrillar structures, a form not observed in RDV-infected cells. To analyze the entity of the bundles of fibrillar structures, VCMs on coverslips were inoculated with RGDV at an MOI of 1, fixed at 48 h p.i., and examined by electron microscopy (EM), as described previously (19). We observed viral particles of approximately 70 nm in diameter in close association with the free ends, as well as along the edges, of tubules of approximately 25 nm in diameter (Fig. 2A to D). The abundant bundles of tubules with closely associated viral particles were clearly in contact with the periphery of granular, electron-dense inclusions of 800 to 1,200 nm in diameter (Fig. (Fig.2B),2B), namely, viroplasms. The dimensions and appearance of the tubular structures resembled those of microtubules (Fig. (Fig.2C)2C) (17). Transverse sections of tubules revealed arrays of closed circles of approximately 25 nm in diameter, with viral particles attached directly or via a filament to the circumference (Fig. (Fig.2D2D).Open in a separate windowFIG. 2.Association of RGDV particles with microtubules. (A) Electron micrograph showing RGDV particles associated with microtubules in virus-infected VCMs 48 h p.i. Bar, 300 nm. VP, electron-dense inclusion. (B) Virus-associated microtubules in contact with the periphery of the electron-dense inclusion indicated by a white rectangle in panel A. Bar, 300 nm. (C) Viral particles along the edges of tubules of approximately 25 nm in diameter. Bar, 150 nm. (D) Transverse sections of arrays of closed circles of approximately 25 nm in diameter with viral particles attached to their circumference directly (arrow) or via a filament (arrowhead). Bar, 150 nm. (E) Confocal micrograph showing the association of viral particles with microtubules in virus-infected VCMs 48 h p.i. Microtubules were stained with α-tubulin-specific antibodies conjugated to FITC; viral particles were stained with viral-antigen-specific antibodies conjugated to rhodamine. Arrowheads indicate the ringlike organization of viral antigens. Arrows show the colocalization of fibrillar profiles of viral antigens with microtubules. The insets show ringlike and fibrillar profiles of immunostained viral antigens. The circular areas inside the ringlike structures are viroplasms. Bar, 5 μm.Our observations suggested that RGDV particles might attach to microtubules in infected cells. To examine this possibility, we inoculated VCMs with RGDV at an MOI of 1, fixed the cells 48 h p.i., immunostained them with α-tubulin-specific antibodies conjugated to FITC and with viral-antigen-specific antibodies conjugated to rhodamine, and analyzed them by confocal microscopy, as described previously (19). Viral antigens were visualized as ringlike and fibrillar structures (Fig. (Fig.2E).2E). Double immunostaining of the infected cells revealed that a network of microtubule-based filaments colocalized with most of the fibrillar structures that represented viral antigens, confirming the association of viral particles with the microtubule-like inclusions visualized by EM (Fig. (Fig.2A).2A). Nonspecific reactions were not detected with either of the stainings (data not shown). Our results suggested that RGDV particles, which assembled at the periphery of viroplasms, might be transported along microtubules. Due to the lack of RGDV infectious clones fused with green fluorescent protein and the effective gene transfection system for VCMs, we could not observe the trafficking of RGDV particles along microtubules in living cells.We then used three-dimensional (3-D) electron tomographic microscopy (ET) to reveal a new level of morphological detail about the association of RGDV with microtubules. To produce 3-D reconstructions of RGDV-infected cells, we fixed, embedded, and sectioned infected leafhopper cells as described previously (5). We chose a representative region that showed numerous RGDV particles close to bundles of microtubules for this novel tomographic analysis. A single-axis tilt series was collected manually from −60° to 60° with 2° increments using an H9500SD EM (Hitachi, Tokyo) operated at 200 kV. These tomographic data were recorded at a defocus of 3.6 μm on the TVIPS 2k × 2k charge-coupled-device camera (TVIPS, Gauting, Germany). Microscopic magnification of ×15,000, providing 1.28 nm/pixel, was enough to view the microtubules and virus particles following tomographic reconstruction of the tilt series using IMOD (7). As shown in the 3-D tomogram in Fig. Fig.3,3, most of the RGDV particles were bound to the edges of bundles of microtubules. The RGDV particles along the edges of microtubules were arrayed in an orderly but uncrowded manner (Fig. (Fig.3).3). Our ET analysis also revealed that some viral particles were linked to filaments of approximately 10 nm in diameter (Fig. (Fig.3B).3B). Morphologically, these filaments resembled vimentin intermediate filaments (4). In many lines of cultured cells, vimentin intermediate filaments partially overlap the microtubules, and there is evidence that the two filament systems interact (3, 9, 20). Unfortunately, vimentin-specific monoclonal antibodies from mouse and rabbit did not react specifically with our leafhopper cells (data not shown), but the nature of the intermediate filaments was apparent from their dimensions, intracellular location, and organization. Thus, our ET analysis indicated that RGDV particles were able to associate directly and/or via intermediate filaments with microtubules.Open in a separate windowFIG. 3.ET analysis showing the association of RGDV particles with microtubules either directly or via intermediate filaments. (A) Translucent representation of the reconstructed viruses lining up with microtubules. (B) Slice of the reconstructed volumes from the inset of A to show the association of RGDV particles with intermediate filaments (arrows). Bars, 150 nm.To examine the role of the microtubules for RGDV activity, we added a microtubule-disrupting agent, either nocodazole (Sigma) or colchicine (Sigma), 2 hours after inoculation of VCMs with RGDV at an MOI of 1 and then continued the incubation for a further 46 h. Cells were fixed 48 h p.i. and stained with α-tubulin-specific antibodies conjugated to FITC (Sigma) and viral-particle-specific antibodies conjugated to rhodamine, with subsequent confocal fluorescence microscopy, as described previously (19). We tested a range of drug concentrations in preliminary experiments (data not shown) and determined optimal concentrations. Treatment of infected cells with 10 μM nocodazole or 5 μg/ml colchicine resulted in the complete disassembly of microtubules, with the accumulation of ringlike structures exclusively and no fibrillar structures representative of viral antigens in the cytoplasm (Fig. (Fig.4A).4A). These ringlike aggregates of viral antigens were confirmed to surround viroplasms when the latter were immunostained for Pns12, as described above and shown in Fig. Fig.1.1. Nonspecific reactions were not detected with either staining (data not shown). These results suggest that RGDV particles multiply around the viroplasm but are unable to distribute along the microtubules in the presence of the chemicals.Open in a separate windowFIG. 4.(A) Effects of microtubule-disrupting agents on the formation of microtubules and fibrillar profiles of immunostained viral antigens. Bars, 5 μm. The insets show the ringlike profiles of immunostained viral antigens after treatment with inhibitors, suggesting that viral replication occurs in the presence of each inhibitor. (B) Effects of drugs on the production of cell-associated (gray bars) and extracellular (black bars) viruses in VCMs infected with RGDV. The error bars indicate standard deviations.During the process of infection, microtubules play important roles in viral entry, intracellular trafficking, and extracellular release (2, 8, 16). We next investigated the effects of the microtubule-disrupting agents on the production in and release of viruses from virus-infected cells by the method described previously (18). Nocodazole or colchicine was added 2 h after inoculation of VCMs with RGDV at an MOI of 1, and incubation was continued for a further 46 h. The extracellular medium and the cells were collected separately. The medium was centrifuged for 30 min at 15,000 × g, and the supernatant was collected. The cells were subjected to three cycles of freezing and thawing to release viral particles. The viral titer of each sample was determined, in duplicate, by the fluorescent focus assay as described previously (6), with VCMs and a magnification of ×10. As shown in Fig. Fig.4B,4B, nocodazole (20 μM) and colchicine (10 μg/ml) caused a fivefold reduction in the number of released viruses, compared to that from untreated control infected cells. In contrast, each inhibitor at the selected dose failed to significantly reduce the titer of cell-associated viruses (less then 5% compared to that from untreated control). These results suggest that the inhibitors impeded the release of viruses into the medium without affecting viral production in infected cells. We do not yet understand why the viral titer was not elevated in drug-treated cells from which viral release was inhibited. However, our data show clearly that disruption of microtubules directly inhibited the release of mature viral particles from infected cells.In conclusion, EM, ET, immunofluorescence staining, and experiments with two inhibitors support the hypothesis that the transport of RGDV from viroplasms to the plasma membrane and into the medium is dependent on microtubules. In the case of RDV, vesicular compartment-containing viral particles that locate adjacent to the viroplasms were considered to play an important role in the transport and release of the virus from the viroplasm to the culture medium in infected VCMs (18). On the other hand, a fibrillar structure (Fig. (Fig.11 and and2),2), not observed in RDV-infected cells, was considered to function in the trafficking of RGDV from viroplasm into the culture medium (Fig. (Fig.4)4) in the present study. RGDV and RDV, both members of the Phytoreovirus genus, have some common biological and biochemical properties but are distinct from each other (13). For example, viruses are restricted to phloem-related cells in RGDV-infected plants but distributed in many types of cells in RDV-infected plants, and a P2 protein with a function to adsorb to and/or penetrate into insect vector cells is present in RGDV and absent in RDV in particles purified using carbon tetrachloride. The present molecular cytopathological study revealed one more difference between the viruses: they have different means for transporting and releasing infectious particles to the cell exterior. The presence of such a molecular mechanism may accelerate the secondary infections by the viruses in infected vector insects, and the high propagation speed would allow the viruses to complete infection cycles through insects and plants.  相似文献   

12.
13.
14.
The sheath of the Serratia entomophila antifeeding prophage, which is pathogenic to the New Zealand grass grub Costelytra zealandica, is a 3-fold helix formed by a 4-fold symmetric repeating motif disposed around a helical inner tube. This structure, determined by electron microscopy and image processing, is distinct from that of the other known morphologically similar bacteriophage sheaths.The antifeeding prophage (Afp) of Serratia entomophila and Serratia proteamaculans is a naturally occurring virus tail-like structure which delivers a putative toxin molecule that leads to starvation of the New Zealand grass grub Costelytra zealandica (5). Afp is composed of 18 different gene products (molecular masses of 6.5 to 263 kDa). The first 16 open reading frames have orthologues (Photorhabdus virulence cassettes [PVC]) in the insecticidal bacterium Photorhabdus luminescens TTO1 genome (5). Afp and PVCs morphologically resemble a typical R-type bacteriocin (6, 12, 16) However, Afp is the only known phage tail-like protein complex that is not a bacteriocin-protein complex of considerable medical relevance that targets the same or closely related bacterial strains (1, 3, 8, 12). The major component of Afp is a contractile cylindrical outer sheath encasing an inner tube speculated to house the toxin molecule (6). A dome-shaped “head” defines one extremity of the tube, while the other end is attached to a “bell-shaped” structure with a base morphologically similar to the base plate of the T4 bacteriophage tail (9).Transmission electron micrographs of two-dimensional (2D) projections of negatively stained (Fig. (Fig.11 A) or frozen-hydrated and vitrified (Fig. (Fig.1B)1B) recombinant Afp particles (see Fig. S1 in the supplemental material) were used for computational image analysis. A globally averaged image of the Afp particle in the major configuration (called E here) (Fig. (Fig.1C),1C), generated using negatively stained specimens, clearly distinguished the morphologies of the various constituent structural parts. Thus, the cylindrical sheath appears to be formed by a periodic structure harboring a distinctive, inverted-V-shaped feature. A minor population of Afp particles displays an alternate configuration (called C here) where, concomitant with contraction of the sheath (averaged axial compression of ∼52% [see Table S1 in the supplemental material]), the inner tube, shorn off the bell-shaped structure, is revealed (Fig. (Fig.1A)1A) (6). Several other bacteriocins undergo such a high degree of compression, which has been characterized in detail for the tail sheath of bacteriophage T4 (9). We also generated individual global averages for the periodic sheath structure, for the bell-shaped structure, and for the inner tube (Fig. (Fig.1C)1C) which provide more accurate dimensions of these different sections (see Table S1 in the supplemental material) than those reported earlier (6).Open in a separate windowFIG. 1.(A) Electron micrograph of a negatively stained preparation of partially purified recombinant Afp particles. The gray, white, and black arrows point to an Afp particle in the extended (E) configuration, to an Afp particle with the sheath contracted, exposing the inner tube in the contracted (C) configuration, and to an inner tube with the bell-shaped structure attached at one end, respectively. (B) Cryoelectron micrograph recorded from a preparation similar to that shown in panel A. (C) Globally averaged images of Afp particles (3,026 images) in the E state and the three distinguishable parts visualized by negative staining. Bars, 200 Å. (D) Averaged power spectrum of the sheath of vitrified Afp particles. The black and white arrows indicate reflections delineating the axial rise (1/78.5 Å) and the helical pitch (1/118 Å), respectively. In panels A and C, lighter regions represent protein and the contrast is reversed in panel B. Global averages were created using classalign2 of the EMAN suite (10) and visualized in bshow (4).For a better insight, we determined the 3D structure of the central periodic section of the Afp particle in the E state. A global power spectrum derived from the cryoimages established the structure to be helical with a clear first meridional reflection at 1/78.5 Å and the first, strongest nonmeridional reflection at 1/118 Å (Fig. (Fig.1D).1D). These reflections correspond to the axial rise (Δx) and the pitch of the helix, respectively, and reflect a turn angle (Δψ) of about ±240° (32 helix) for the repeating motif. The correct sign, i.e., the hand, of the helix remains to be determined. Computationally excised overlapping segments of this helical section from images of vitrified and negatively stained Afp particles were subjected to 3D reconstruction using the iterative helical real-space reconstruction (IHRSR) algorithm (2) using the determined helical parameters (see Fig. S2 in the supplemental material). After a few iterations, the presence of an in-plane 4-fold symmetry (C4) was apparent in the density map (see Fig. S2 in the supplemental material), which was then imposed in the subsequent reconstruction exercises. However, no stable solution was forthcoming, even after many (e.g., 30) iterative cycles. This is generally indicative of the presence of heterogeneity in the form of variations in helix translation and/or twist angle (15) in the structure. As a first step, we focused our attention on the pitch value, and following classification (see Fig. S3 in the supplemental material), we found that the majority of the image segments correspond to a pitch of 120 Å. These segments were then selected out of the full data set and led to a stable and refined 3D reconstruction. We also obtained very comparable results for the helical section when images of negatively stained Afp sheath sections were used, thus supporting our computational approach (see Fig. S4 in the supplemental material) and general conclusions about the E state described below.Figure Figure22 A is an isosurface representation of the density map of the helical Afp sheath in the E state calculated at ∼21.5-Å resolution (see Fig. S5 in the supplemental material). To the best of our knowledge, a 4-fold rotational symmetry has not been seen for any other contractile T4 bacteriophage taillike structure, which points to the unique architecture of the Afp sheath. A power spectrum generated using the 2D projection from the final density map, compared to the experimental global power spectrum (Fig. (Fig.2D),2D), showed strong agreement, confirming the fidelity of the computational image analysis. The density map displays protein layers, ∼80 Å thick, that are stacked on each other in a periodic fashion. The uneven outer surface of the sheath is perforated and decorated with ∼35-Å protrusions. When rendered with a raised threshold, a characteristic feature of the map is a contiguous, high-electron-density region having an inverted-Y-shaped structure (Fig. 2C and E; see Fig. S4 in the supplemental material). At the modest ∼21.5-Å resolution, the boundary of the repeating subunit cannot be defined. A 25 ± 3-Å-wide central lumen is seen clearly when viewed along the helix axis (Fig. (Fig.2B)2B) and likely represents the pore of the inner tube (see also below). Using scanning transmission electron microscopy (STEM) (see Fig. S6 in the supplemental material), we estimated the averaged molecular mass of the central helical section of an Afp particle to be 9.8 ± 0.4 kDa/Å (Fig. (Fig.3)3) (14) and that of only the inner tube to be ∼2.5 kDa/Å, based on a relatively small pool of such images. These values translate to a mass contribution of approximately 145 kDa of the subunit whose periodic arrangement forms the outer component of the sheath (i.e., excluding the inner tube) (see Fig. S6 in the supplemental material). This value is not very different from the cumulative mass of the different proteins, i.e., homologous afp2, afp3, and afp4, thought to be involved in Afp sheath formation (5) (see Fig. S7 in the supplemental material).Open in a separate windowFIG. 2.Orthogonal isosurface rendering, at 1 σ (standard deviation) of the computed ∼21.5-Å density map of the helical sheath of the vitrified Afp particle viewed normal (A) and parallel (B) to the helix. The images were generated using the software package CHIMERA (13). The arrow indicates a surface perforation. (C) The Afp density map rendered at 3.5 σ to highlight the largest contiguous high-electron-density regions; one circumscribed by a black ellipse is computationally extracted and shown in panel E. (D) Comparison of the experimental, averaged power spectrum of the helical sheath of Afp (left) with that computed (right) from the 2D projection of the calculated density map. (F) Global average of the inner tube of the Afp particle and a plot of the surface density variation (scaled from 0 to 1) along the helical (y) axis. The dimension along the tube is plotted on the x axis.Open in a separate windowFIG. 3.(A) Dark-field micrograph of a freeze-dried, unstained preparation of Afp particles used in STEM measurements. An Afp particle in the E state, an Afp inner tube with the attached bell-shaped structure, and a tobacco mosaic virus particle, used as a calibration standard, are marked by the arrowhead and the gray and white arrows, respectively. (B) Histogram plot of the measured distributions of mass per unit length corresponding to the uniform periodic section of the Afp particles overlaid with a fitted Gaussian curve produced by using the ORIGIN6 software package.A paucity of images of the C state (∼5% of the complete data set) precluded a full, refined 3D reconstruction, but based on the available 2,774 overlapping image segments of the isolated inner tube, a global average was calculated. A plot of contrast variation (Fig. (Fig.2F)2F) indicates that the surface is characterized by ∼40-Å spaced elevated crests and invaginated grooves, in agreement with the calculated axial rise of ∼39 Å for the subunit (see Fig. S8 in the supplemental material) comprising the tube. Based on these preliminary results, it appears that the helical symmetry of the inner tube is markedly different from that of the sheath.Our observation that the pitch of the helix in the E state can vary by as much as ∼50 Å attests to the flexible nature of the sheath, which is required for compressibility and may be facilitated by the somewhat porous nature of the sheath (Fig. (Fig.2).2). Preliminary deductions (data not shown) based on a small pool of images of the C state suggest profound rearrangement of the elements of the sheath. How that translates to extrusion of the toxin remains to be revealed.   相似文献   

15.
16.
17.
We have previously reported on the ubiquitylation and degradation of hepatitis C virus core protein. Here we demonstrate that proteasomal degradation of the core protein is mediated by two distinct mechanisms. One leads to polyubiquitylation, in which lysine residues in the N-terminal region are preferential ubiquitylation sites. The other is independent of the presence of ubiquitin. Gain- and loss-of-function analyses using lysineless mutants substantiate the hypothesis that the proteasome activator PA28γ, a binding partner of the core, is involved in the ubiquitin-independent degradation of the core protein. Our results suggest that turnover of this multifunctional viral protein can be tightly controlled via dual ubiquitin-dependent and -independent proteasomal pathways.Hepatitis C virus (HCV) core protein, whose amino acid sequence is highly conserved among different HCV strains, not only is involved in the formation of the HCV virion but also has a number of regulatory functions, including modulation of signaling pathways, cellular and viral gene expression, cell transformation, apoptosis, and lipid metabolism (reviewed in references 9 and 15). We have previously reported that the E6AP E3 ubiquitin (Ub) ligase binds to the core protein and plays an important role in polyubiquitylation and proteasomal degradation of the core protein (22). Another study from our group identified the proteasome activator PA28γ/REG-γ as an HCV core-binding partner, demonstrating degradation of the core protein via a PA28γ-dependent pathway (16, 17). In this work, we further investigated the molecular mechanisms underlying proteasomal degradation of the core protein and found that in addition to regulation by the Ub-mediated pathway, the turnover of the core protein is also regulated by PA28γ in a Ub-independent manner.Although ubiquitylation of substrates generally requires at least one Lys residue to serve as a Ub acceptor site (5), there is no consensus as to the specificity of the Lys targeted by Ub (4, 8). To determine the sites of Ub conjugation in the core protein, we used site-directed mutagenesis to replace individual Lys residues or clusters of Lys residues with Arg residues in the N-terminal 152 amino acids (aa) of the core (C152), within which is contained all seven Lys residues (Fig. (Fig.1A).1A). Plasmids expressing a variety of mutated core proteins were generated by PCR and inserted into the pCAGGS (18). Each core-expressing construct was transfected into human embryonic kidney 293T cells along with the pMT107 (25) encoding a Ub moiety tagged with six His residues (His6). Transfected cells were treated with the proteasome inhibitor MG132 for 14 h to maximize the level of Ub-conjugated core intermediates by blocking the proteasome pathway and were harvested 48 h posttransfection. His6-tagged proteins were purified from the extracts by Ni2+-chelation chromatography. Eluted protein and whole lysates of transfected cells before purification were analyzed by Western blotting using anticore antibodies (Fig. (Fig.1B).1B). Mutations replacing one or two Lys residues with Arg in the core protein did not affect the efficiency of ubiquitylation: detection of multiple Ub-conjugated core intermediates was observed in the mutant core proteins comparable to the results seen with the wild-type core protein as previously reported (23). In contrast, a substitution of four N-terminal Lys residues (C152K6-23R) caused a significant reduction in ubiquitylation (Fig. (Fig.1B,1B, lane 9). Multiple Ub-conjugated core intermediates were not detected in the Lys-less mutant (C152KR), in which all seven Lys residues were replaced with Arg (Fig. (Fig.1B,1B, lane 11). These results suggest that there is not a particular Lys residue in the core protein to act as the Ub acceptor but that more than one Lys located in its N-terminal region can serve as the preferential ubiquitylation site. In rare cases, Ub is known to be conjugated to the N terminus of proteins; however, these results indicate that this does not occur within the core protein.Open in a separate windowFIG. 1.In vivo ubiquitylation of HCV core protein. (A) The HCV core protein (N-terminal 152 aa) is represented on the top. The positions of the amino acid residues of the core protein are indicated above the bold lines. The positions of the seven Lys residues in the core are marked by vertical ticks. Substitution of Lys with Arg (R) is schematically depicted. (B) Detection of ubiquitylated forms of the core proteins. The transfected cells with core expression plasmids and pMT107 were treated with the proteasome inhibitor MG132 and harvested 48 h after transfection. His6-tagged proteins were purified and subsequently analyzed by Western blot analysis using anticore antibody (upper panel). Core proteins conjugated to a number of His6-Ub are denoted with asterisks. Whole lysates of transfected cells before purification were also analyzed (lower panel). Lanes 1 to 11, C152 to C152KR, as indicated for panel A. Lane 12; empty vector.To investigate how polyubiquitylation correlates with proteasome degradation of the core protein, we performed kinetic analysis of the wild-type and mutated core proteins by use of the Ub protein reference (UPR) technique, which can compensate for data scatter of sample-to-sample variations such as levels of expression (10, 24). Fusion proteins expressed from UPR-based constructs (Fig. (Fig.2A)2A) were cotranslationally cleaved by deubiquitylating enzymes, thereby generating equimolar quantities of the core proteins and the reference protein, dihydrofolate reductase-hemagglutinin (DHFR-HA) tag-modified Ub, in which the Lys at aa 48 was replaced by Arg to prevent its polyubiquitylation (UbR48). After 24 h of transfection with UPR constructs, cells were treated with cycloheximide and the amounts of core proteins and DHFR-HA-UbR48 at the indicated time points were determined by Western blot analysis using anticore and anti-HA antibodies. The mature form of the core protein, aa 1 to 173 (C173) (13, 20), and C152 were degraded with first-order kinetics (Fig. 2B and D). MG132 completely blocked the degradation of C173 and C152 (Fig. (Fig.2B),2B), and C152K6-23R and C152KR were markedly stabilized (Fig. (Fig.2C).2C). The half-lives of C173 and C152 were calculated to be 5 to 6 h, whereas those of C152K6-23R and C152KR were calculated to be 22 to 24 h (Fig. (Fig.2D),2D), confirming that the Ub plays an important role in regulating degradation of the core protein. Nevertheless, these results also suggest possible involvement of the Ub-independent pathway in the turnover of the core protein, as C152KR is more destabilized than the reference protein (Fig. (Fig.2C2C and and2D2D).Open in a separate windowFIG. 2.Kinetic analysis of degradation of HCV core proteins. (A) The fusion constructs used in the UPR technique. Open boxes indicate the DHFR sequence, which is extended at the C terminus by a sequence containing the HA epitope (hatched boxes). UbR48 moieties bearing the Lys-Arg substitution at aa 48 are represented by open ellipses. Bold lines indicate the regions of the core protein. The amino acid positions of the core protein are indicated above the bold lines. The arrows indicate the sites of in vivo cleavage by deubiquitylating enzymes. (B and C) Turnover of the core proteins. After a 24-h transfection with each UPR construct, cells were treated with 50 μg of cycloheximide/ml in the presence or absence of 10 μM MG132 for the different time periods indicated. Cells were lysed at the different time points indicated, followed by evaluation via sodium dodecyl sulfate-polyacrylamide gel electrophoresis and Western blot analysis using antibodies against the core protein and HA. (D) Quantitation of the data shown in panels B and C. At each time point, the ratio of band intensity of the core protein relative to the reference DHFR-HA-UbR48 was determined by densitometry and is plotted as a percentage of the ratio at time zero.We have shown that PA28γ specifically binds to the core protein and is involved in its degradation (16, 17). Recent studies demonstrated that PA28γ is responsible for Ub-independent degradation of the steroid receptor coactivator SRC-3 and cell cycle inhibitors such as p21 (3, 11, 12). Thus, we next investigated the possibility of PA28γ involvement in the degradation of either C152KR or C152. Since C152KR carries two amino acid substitutions in the PA28γ-binding region (aa 44 to 71) (17), we tested the influence of the mutations of C152KR on the interaction with PA28γ by use of a coimmunoprecipitation assay. When Flag-tagged PA28γ (F-PA28γ) was expressed in cells along with C152 or C152KR, F-PA28γ precipitated along with both C152 and C152KR, indicating that PA28γ interacts with both core proteins (Fig. (Fig.3A).3A). Figure Figure3B3B reveals the effect of exogenous expression of F-PA28γ on the steady-state levels of C152 and C152KR. Consistent with previous data (17), the expression level of C152 was decreased to a nearly undetectable level in the presence of PA28γ (Fig. (Fig.3B,3B, lanes 1 and 3). Interestingly, exogenous expression of PA28γ led to a marked reduction in the amount of C152KR expressed (Fig. (Fig.3B,3B, lanes 5 and 7). Treatment with MG132 increased the steady-state level of the C152KR in the presence of F-PA28γ as well as the level of C152 (Fig. (Fig.3B,3B, lanes 4 and 8).Open in a separate windowFIG. 3.PA28γ-dependent degradation of the core protein. (A) Interaction of the core protein with PA28γ. Cells were cotransfected with the wild-type (C152) or Lys-less (C152KR) core expression plasmid in the presence of a Flag-PA28γ (F-PA28γ) expression plasmid or an empty vector. The transfected cells were treated with MG132. After 48 h, the cell lysates were immunoprecipitated with anti-Flag antibody and visualized by Western blotting with anticore antibodies. Western blot analysis of whole cell lysates was also performed. (B) Degradation of the wild-type and Lys-less core proteins via the PA28γ-dependent pathway. Cells were transfected with the UPR construct with or without F-PA28γ. In some cases, cells were treated with 10 μM MG132 for 14 h before harvesting. Western blot analysis was performed using anticore, anti-HA, and anti-Flag antibodies. (C) After 24 h of transfection with UPR-C152KR and UPR-C191KR with or without F-PA28γ (an empty vector), cells were treated with 50 μg of cycloheximide/ml for different time periods as indicated (chase time). Western blot analysis was performed using anticore and anti-HA antibodies. The precursor core protein and the core that was processed, presumably by signal peptide peptidase, are denoted by open and closed triangles, respectively.We further investigated whether PA28γ affects the turnover of Lys-less core protein through time course experiments. C152KR was rapidly destabilized and almost completely degraded in a 3-h chase experiment using cells overexpressing F-PA28γ (Fig. (Fig.3C,3C, left panels). A similar result was obtained using an analogous Lys-less mutant of the full-length core protein C191KR (Fig. (Fig.3C,3C, right panels), thus demonstrating that the Lys-less core protein undergoes proteasomal degradation in a PA28γ-dependent manner. These results suggest that PA28γ may play a role in accelerating the turnover of the HCV core protein that is independent of ubiquitylation.Finally, we examined gain- and loss-of-function of PA28γ with respect to degradation of full-length wild-type (C191) and mutated (C191KR) core proteins in human hepatoma Huh-7 cells. As expected, exogenous expression of PA28γ or E6AP caused a decrease in the C191 steady-state levels (Fig. (Fig.4A).4A). In contrast, the C191KR level was decreased with expression of PA28γ but not of E6AP. We further used RNA interference to inhibit expression of PA28γ or E6AP. An increase in the abundance of C191KR was observed with PA28γ small interfering RNA (siRNA) but not with E6AP siRNA (Fig. (Fig.4B).4B). An increase in the C191 level caused by the activity of siRNA against PA28γ or E6AP was confirmed as well.Open in a separate windowFIG. 4.Ub-dependent and Ub-independent degradation of the full-length core protein in hepatic cells. (A) Huh-7 cells were cotransfected with plasmids for the full-length core protein (C191) or its Lys-less mutant (C191KR) in the presence of F-PA28γ or HA-tagged-E6AP expression plasmid (HA-E6AP). After 48 h, cells were lysed and Western blot analysis was performed using anticore, anti-HA, anti-Flag, or anti-GAPDH. (B) Huh-7 cells were cotransfected with core expression plasmids along with siRNA against PA28γ or E6AP or with negative control siRNA. Cells were harvested 72 h after transfection and subjected to Western blot analysis.Taking these results together, we conclude that turnover of the core protein is regulated by both Ub-dependent and Ub-independent pathways and that PA28γ is possibly involved in Ub-independent proteasomal degradation of the core protein. PA28 is known to specifically bind and activate the 20S proteasome (19). Thus, PA28γ may function by facilitating the delivery of the core protein to the proteasome in a Ub-independent manner.Accumulating evidence suggests the existence of proteasome-dependent but Ub-independent pathways for protein degradation, and several important molecules, such as p53, p73, Rb, SRC-3, and the hepatitis B virus X protein, have two distinct degradation pathways that function in a Ub-dependent and Ub-independent manner (1, 2, 6, 7, 14, 21, 27). Recently, critical roles for PA28γ in the Ub-independent pathway have been demonstrated; SRC-3 and p21 can be recognized by the 20S proteasome independently of ubiquitylation through their interaction with PA28γ (3, 11, 12). It has also been reported that phosphorylation-dependent ubiquitylation mediated by GSK3 and SCF is important for SRC-3 turnover (26). Nevertheless, the precise mechanisms underlying turnover of most of the proteasome substrates that are regulated in both Ub-dependent and Ub-independent manners are not well understood. To our knowledge, the HCV core protein is the first viral protein studied that has led to identification of key cellular factors responsible for proteasomal degradation via dual distinct mechanisms. Although the question remains whether there is a physiological significance of the Ub-dependent and Ub-independent degradation of the core protein, it is reasonable to consider that tight control over cellular levels of the core protein, which is multifunctional and essential for viral replication, maturation, and pathogenesis, may play an important role in representing the potential for its functional activity.  相似文献   

18.
19.
2-Oxobutyrate is an important intermediate in the chemical, drug, and food industries. Whole cells of Pseudomonas stutzeri SDM, containing NAD-independent lactate dehydrogenases, effectively converted 2-hydroxybutyrate into 2-oxobutyrate. Under optimal conditions, the biocatalytic process produced 2-oxobutyrate at a high concentration (44.4 g liter−1) and a high yield (91.5%).2-Oxobutyrate (2-OBA) is used as a raw material in the synthesis of chiral 2-aminobutyric acid, isoleucine, and some kinds of medicines (1, 8). There is no suitable starting material for 2-OBA production by chemical synthesis; therefore, the development of innovative biotechnology-based techniques for 2-OBA production is desirable (12).2-Hydroxybutyrate (2-HBA) is cheaper than 2-OBA and can be substituted for 2-OBA in the production of isoleucine, as reported previously (9, 10). The results of those studies also indicated that it might be possible to produce 2-OBA from 2-HBA by a suitable biocatalytic process. In the presence of NAD, NAD-dependent 2-hydroxybutyrate dehydrogenase can catalyze the oxidation of 2-HBA to 2-OBA (4). However, due to the high cost of pyridine cofactors (11), it is preferable to use a biocatalyst that directly catalyzes the formation of 2-OBA from 2-HBA without any requirement for NAD as a cofactor.In our previous report, we confirmed that NAD-independent lactate dehydrogenases (iLDHs) in the pyruvate-producing strain Pseudomonas stutzeri SDM (China Center for Type Culture Collection no. M206010) could oxidize lactate and 2-HBA (6). Therefore, in addition to pyruvate production from lactate, P. stutzeri SDM might also have a potential application in 2-OBA production.To determine the 2-OBA production capability of P. stutzeri SDM, the strain was first cultured at 30°C in a minimal salt medium (MSM) supplemented with 5.0 g liter−1 dl-lactate as the sole carbon source (5). The whole-cell catalyst was prepared by centrifuging the medium and resuspending the cell pellet, and biotransformation was then carried out under the following conditions using 2-HBA as the substrate and whole cells of P. stutzeri SDM as the biocatalyst: 2-HBA, 10 g liter−1; dry cell concentration, 6 g liter−1; buffer, 100 mM potassium phosphate (pH 7.0); temperature, 30°C; shaking speed, 300 rpm. After 4 h of reaction, the mixture was analyzed by high-performance liquid chromatography (HPLC; Agilent 1100 series; Hewlett-Packard) using a refractive index detector (3). The HPLC system was fitted with a Bio-Rad Aminex HPX-87 H column. The mobile phase consisted of 10 mM H2SO4 pumped at 0.4 ml min−1 (55°C). Biotransformation resulted in the production of a compound that had a retention time of 19.57 min, which corresponded to the peak of authentic 2-OBA (see Fig. S1 in the supplemental material).After acidification and vacuum distillation, the new compound was analyzed by negative-ion mass spectroscopy. The molecular ion ([M − H], m/z 101.1) signal of the compound was consistent with the molecular weight of 2-OBA, i.e., 102.1 (see Fig. S2 in the supplemental material). These results confirmed that 2-HBA was oxidized to 2-OBA by whole cells of P. stutzeri SDM.To investigate whether iLDHs are responsible for 2-OBA production in the above-described biocatalytic process, 2-HBA oxidation activity in P. stutzeri SDM was probed by native polyacrylamide gel electrophoresis. After electrophoresis, the gels were soaked in a substrate solution [50 mM Tris-HCl buffer (pH 8.0) containing 0.1 mM phenazine methosulfate, 0.1 mM 3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide, and 1 mM l-lactate, dl-lactate, or dl-2-HBA] and gently shaken. As shown in Fig. Fig.1,1, d- and l-iLDH migrated as two bands with distinct mobilities. The activities responsible for d- and l-2-HBA oxidation were located at the same positions as the d- and l-iLDH activities, respectively. No other bands responsible for d- and l-2-HBA oxidation were detected. Moreover, the dialysis of the crude cell extract did not lead to loss of 2-HBA oxidation activity and the addition of 10 mM NAD+ could not stimulate the reaction (see Table S1 in the supplemental material). These results implied that in the biocatalytic system, 2-HBA was oxidized to 2-OBA by iLDHs present in P. stutzeri SDM.Open in a separate windowFIG. 1.Activity staining of iLDHs after native polyacrylamide gel electrophoresis with lactate or 2-HBA as the substrate.Although the SDM strain could not use 2-HBA or 2-OBA for growth (see Fig. S3 in the supplemental material), 2-HBA might induce some of the enzymes responsible for 2-OBA production in the biocatalytic process. To exclude this possibility, the SDM strain was cultured in MSM containing dl-lactate or pyruvate as the sole carbon source. As shown in Fig. Fig.2,2, the enzyme activities that catalyzed lactate and 2-HBA oxidation were simultaneously present in the cells cultured on lactate and were absent in those cultured on pyruvate. After the lactate or pyruvate was exhausted, 5.05 g liter−1 dl-2-HBA was added to the medium. It was observed that dl-2-HBA was efficiently converted to 2-OBA in the medium containing dl-lactate (Fig. (Fig.2a).2a). No 2-OBA production was detected in the medium containing pyruvate. Because 2-HBA addition did not induce the enzymes involved in 2-HBA oxidation (Fig. 2a and b), we concluded that the iLDHs induced by dl-lactate catalyzed 2-HBA oxidation in this biocatalytic process.Open in a separate windowFIG. 2.Time course of P. stutzeri SDM growth on media containing dl-lactate (a) and pyruvate (b). 2-HBA was added to the medium after the exhaustion of lactate or pyruvate. Symbols: ▴, lactate; ▵, pyruvate; •, 2-HBA; ○, 2-OBA; ▪, cell density; ▧, iLDHs activity with dl-lactate as the substrate; ▒, iLDHs activity with dl-2-HBA as the substrate.iLDHs could catalyze the oxidation of the substrate in a flavin-dependent manner and might use membrane quinone as the electron acceptor. Unlike the oxidases, which directly use the oxygen as the electron acceptor, this substrate oxidation mechanism could prevent the formation of H2O2 (see Fig. S4 in the supplemental material). The P. stutzeri SDM strain efficiently converted dl-2-HBA to 2-OBA with high yields (4.97 g liter−1 2-OBA was produced from 5.05 g liter−1 dl-2-HBA); therefore, 2-OBA production by this strain can be a valuable and technically feasible process. To increase the efficiency of P. stutzeri SDM in the biotechnological production of 2-OBA, the conditions for biotransformation using whole cells of P. stutzeri SDM were first optimized. The influence of the reaction pH and 2-HBA concentration on 2-OBA production was determined in 100 mM phosphate buffer containing whole cells harvested from the medium containing dl-lactate as the sole carbon source. The reaction was initiated by adding the whole cells and 2-HBA at 37°C, followed by incubation for 10 min. After stopping the reaction by adding 1 M HCl, the 2-OBA concentration was determined by HPLC.As shown in Fig. Fig.3a,3a, ,2-OBA2-OBA production was highest at pH 7.0. Under acidic or alkaline conditions, the transformation of 2-HBA to 2-OBA decreased. The optimal 2-HBA concentration was found to be 0.4 M, as shown in Fig. Fig.3b.3b. 2-OBA production increased as the 2-HBA concentration increased up to about 0.4 M and decreased thereafter. The concentration of the whole-cell catalyst was then optimized using 0.4 M 2-HBA as the substrate at pH 7.0. As shown in Fig. Fig.3c,3c, the highest 2-OBA concentration was obtained with 20 g (dry cell weight [DCW]) liter−1 of P. stutzeri SDM. The 2-OBA concentration decreased with any increase beyond this cell concentration.Open in a separate windowFIG. 3.Optimization of the biocatalysis conditions. (a) Effect of pH on 2-OBA production activity. (b) Effect of 2-HBA concentrations on 2-OBA production activity. (c) Effect of the concentration of P. stutzeri SDM on biotransformation. OD, optical density.After optimizing the biocatalytic conditions, we studied the biotechnological production of 2-OBA from 2-HBA by using the whole-cell catalyst P. stutzeri SDM. As shown in Fig. Fig.4,4, when 20 g (DCW) liter−1 P. stutzeri SDM was used as the biocatalyst, 48.5 g liter−1 2-HBA was biotransformed into 44.4 g liter−1 2-OBA in 24 h.Open in a separate windowFIG. 4.Time course of production of 2-OBA from 2-HBA under the optimum conditions. Symbols: ▪, 2-OBA; •, 2-HBA.Biocatalytic production of 2-OBA was carried out using crotonic acid, propionaldehyde, 1,2-butanediol, or threonine as the substrate (2, 7, 8, 12). Resting cells of the strain Rhodococcus erpi IF0 3730 produced 15.7 g liter−1 2-OBA from 20 g liter−1 1,2-butanediol, which is the highest reported yield of 2-OBA to date (8). By using the whole-cell catalyst P. stutzeri SDM, it was possible to produce 2-OBA at a high concentration (44.4 g liter−1) and a high yield (91.5%). Due to the simple composition of the biocatalytic system (see Fig. S5 in the supplemental material), 2-HBA and 2-OBA could be easily separated on a column using a suitable resin. Separation of 2-OBA from the biocatalytic system was relatively inexpensive. The biocatalytic process presented in this report could be a promising alternative for the biotechnological production of 2-OBA.   相似文献   

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