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 312 毫秒
1.
The Merkel cell polyomavirus (MCPyV) was identified recently in human Merkel cell carcinomas, an aggressive neuroendocrine skin cancer. Here, we identify a putative host cell receptor for MCPyV. We found that recombinant MCPyV VP1 pentameric capsomeres both hemagglutinated sheep red blood cells and interacted with ganglioside GT1b in a sucrose gradient flotation assay. Structural differences between the analyzed gangliosides suggest that MCPyV VP1 likely interacts with sialic acids on both branches of the GT1b carbohydrate chain. Identification of a potential host cell receptor for MCPyV will aid in the elucidation of its entry mechanism and pathophysiology.Members of the polyomavirus (PyV) family, including simian virus 40 (SV40), murine PyV (mPyV), and BK virus (BKV), bind cell surface gangliosides to initiate infection (2, 8, 11, 15). PyV capsids are assembled from 72 pentamers (capsomeres) of the major coat protein VP1, with the internal proteins VP2 and VP3 buried within the capsids (7, 12). The VP1 pentamer makes direct contact with the carbohydrate portion of the ganglioside (10, 12, 13) and dictates the specificity of virus interaction with the cell. Gangliosides are glycolipids that contain a ceramide domain inserted into the plasma membrane and a carbohydrate domain that directly binds the virus. Specifically, SV40 binds to ganglioside GM1 (2, 10, 15), mPyV binds to gangliosides GD1a and GT1b (11, 15), and BKV binds to gangliosides GD1b and GT1b (8).Recently, a new human PyV designated Merkel cell PyV (MCPyV) was identified in Merkel cell carcinomas, a rare but aggressive skin cancer of neuroendocrine origin (3). It is as yet unclear whether MCPyV is the causative agent of Merkel cell carcinomas (17). A key to understanding the infectious and transforming properties of MCPyV is the elucidation of its cellular entry pathway. In this study, we identify a putative host cell receptor for MCPyV.Because an intact infectious MCPyV has not yet been isolated, we generated recombinant MCPyV VP1 pentamers in order to characterize cellular factors that bind to MCPyV. VP1 capsomeres have been previously shown to be equivalent to virus with respect to hemagglutination properties (4, 16), and the atomic structure of VP1 bound to sialyllactose has demonstrated that the capsomere is sufficient for this interaction (12, 13). The MCPyV VP1 protein (strain w162) was expressed and purified as described previously (1, 6). Briefly, a glutathione S-transferase-MCPyV VP1 fusion protein was expressed in Escherichia coli and purified using glutathione-Sepharose affinity chromatography. The fusion protein was eluted using glutathione and cleaved in solution with thrombin. The thrombin-cleaved sample was then rechromatographed on a second glutathione-Sepharose column to remove glutathione transferase and any uncleaved protein. The unbound VP1 was then chromatographed on a P-11 phosphocellulose column, and peak fractions eluting between 400 and 450 mM NaCl were collected. The purified protein was analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), followed by Coomassie blue staining (Fig. (Fig.1A,1A, left) and immunoblotting using an antibody (I58) that generally recognizes PyV VP1 proteins (Fig. (Fig.1A,1A, right) (9). Transmission electron microscopy (Philips CM10) analysis confirmed that the purified recombinant MCPyV VP1 formed pentamers (capsomeres), which did not assemble further into virus-like particles (Fig. (Fig.1B).1B). In an initial screening of its cell binding properties, we tested whether the MCPyV VP1 pentamers hemagglutinated red blood cells (RBCs). The MCPyV VP1 pentamers were incubated with sheep RBCs and assayed as previously described (5). SV40 and mPyV recombinant VP1 pentamers served as negative and positive controls, respectively. We found that MCPyV VP1 hemagglutinated the RBCs with the same efficiency as mPyV VP1 (protein concentration/hemagglutination unit) (Fig. (Fig.1C,1C, compare rows B and C from wells 1 to 11), suggesting that MCPyV VP1 engages a plasma membrane receptor on the RBCs. The recombinant murine VP1 protein used for comparison was from the RA strain, a small plaque virus (4). Thus, MCPyV VP1 has the hemagglutination characteristics of a small plaque mPyV (12, 13).Open in a separate windowFIG. 1.Characterization of MCPyV VP1. Recombinant MCPyV VP1 forms pentamers and hemagglutinates sheep RBCs. (A) Coomassie blue-stained SDS-PAGE and an immunoblot of the purified recombinant MCPyV VP1 protein are shown. Molecular mass markers are indicated. (B) Electron micrograph of the purified MCPyV VP1. MCPyV VP1 (shown in panel A) was diluted to 100 μg/ml and absorbed onto Formvar/carbon-coated copper grids. Samples were washed with phosphate-buffered saline, stained with 1% uranyl acetate, and visualized by transmission electron microscopy at 80 kV. Bar = 20 nm. (C) Sheep RBCs (0.5%) were incubated with decreasing concentrations of purified recombinant SV40 VP1 (row A), mPyV VP1 (row B), and MCPyV VP1 (row C). Wells 1 to 11 contain twofold serial dilutions of protein, starting at 2 μg/ml (well 1). Well 12 contains buffer only and serves as a negative control. Well 7 (rows B and C) corresponds to 128 hemagglutination units per 2 μg/ml VP1 protein.To characterize the chemical nature of the putative receptor for MCPyV, total membranes from RBCs were purified as described previously (15). The plasma membranes (30 μg) were incubated with MCPyV VP1 (0.5 μg) and floated on a discontinuous sucrose gradient (15). After fractionation, the samples were analyzed by SDS-PAGE, followed by immunoblotting with I58. VP1 was found in the bottom of the gradient in the absence of the plasma membranes (Fig. (Fig.2A,2A, first panel). In the presence of plasma membranes, a fraction of the VP1 floated to the middle of the gradient (Fig. (Fig.2A,2A, second panel), supporting the hemagglutination results that suggested that MCPyV VP1 binds to a receptor on the plasma membrane.Open in a separate windowFIG. 2.MCPyV VP1 binds to a protease-resistant, sialic acid-containing receptor on the plasma membrane. (A) Purified recombinant MCPyV VP1 was incubated with or without the indicated plasma membranes. The samples were floated in a discontinuous sucrose gradient, and the fractions were collected from the top of the gradient, subjected to SDS-PAGE, and immunoblotted with the anti-VP1 antibody I58. (B) Control and proteinase K-treated plasma membranes were subjected to SDS-PAGE, followed by Coomassie blue staining. (C) HeLa cells treated with proteinase K (4 μg/ml) were incubated with MCPyV at 4°C, and the resulting cell lysate was probed for the presence of MCPyV VP1. (D) As described in the legend to panel C, except 293T cells were used. (E) Purified MCPyV VP1 was incubated with plasma membranes pretreated with or without α2-3,6,8 neuraminidase and analyzed as described in the legend to panel A.To determine whether the receptor is a protein or a lipid, plasma membrane preparations (30 μg) were incubated with proteinase K (Sigma), followed by analysis with SDS-PAGE and Coomassie blue staining. Under these conditions, the majority of the proteins in the plasma membranes were degraded by the protease (Fig. (Fig.2B,2B, compare lanes 1 and 2). Despite the lack of proteins, the proteinase K-treated plasma membranes bound MCPyV VP1 as efficiently as control plasma membranes (Fig. (Fig.2A,2A, compare the second and third panels), demonstrating that MCPyV VP1 interacts with a protease-resistant receptor. The absence of VP1 in the bottom fraction in Fig. Fig.2A2A (third panel) is consistent with the fact that the buoyant density of the membranes is lowered by proteolysis. Of note, a similar result was seen with binding of the mPyV to the plasma membrane (15). Binding of MCPyV to the cell surface of two human tissue culture cells (i.e., HeLa and 293T) was also largely unaffected by pretreatment of the cells with proteinase K (Fig. 2C and D, compare lanes 1 and 2), further indicating that a nonproteinaceous molecule on the plasma membrane engages the virus.We next determined whether the protease-resistant receptor contains a sialic acid modification. Plasma membranes (10 μg) were incubated with a neuraminidase (α2-3,6,8 neuraminidase; Calbiochem) to remove the sialic acid groups. In contrast to the control plasma membranes, the neuraminidase-treated membranes did not bind MCPyV VP1 (Fig. (Fig.2E,2E, compare first and second panels), indicating that the MCPyV receptor includes a sialic acid modification.Gangliosides are lipids that contain sialic acid modifications. We asked if MCPyV VP1 binds to gangliosides similar to other PyV family members. The structures of the gangliosides used in this analysis (gangliosides GM1, GD1a, GD1b, and GT1b) are depicted in Fig. Fig.3A.3A. To assess a possible ganglioside-VP1 interaction, we employed a liposome flotation assay established previously (15). When liposomes (consisting of phosphatidyl-choline [19 μl of 10 mg/ml], -ethanolamine [5 μl of 10 mg/ml], -serine [1 μl of 10 mg/ml], and -inositol [3 μl of 10 mg/ml]) were incubated with MCPyV VP1 and subjected to the sucrose flotation assay, the VP1 remained in the bottom fraction (Fig. (Fig.3B,3B, first panel), indicating that VP1 does not interact with these phospholipids. However, when liposomes containing GT1b (1 μl of 1 mM), but not GM1 (1 μl of 1 mM) or GD1a (1 μl of 1 mM), were incubated with MCPyV VP1, the vesicles bound this VP1 (Fig. (Fig.3B).3B). A low level of virus floated partially when incubated with liposomes containing GD1b (Fig. (Fig.3B),3B), perhaps reflecting a weak affinity between MCPyV and GD1b. Importantly, MCPyV binds less efficiently to neuraminidase-treated GT1b-containing liposomes than to GT1b-containing liposomes (Fig. (Fig.3B,3B, sixth panel), suggesting that the GT1b sialic acids are involved in virus binding. This finding is consistent with the ability of neuraminidase to block MCPyV binding to the plasma membrane (Fig. (Fig.2E).2E). The level of virus flotation observed in the neuraminidase-treated GT1b-containing liposomes is likely due to the inefficiency of the neuraminidase reaction with a high concentration of GT1b used to prepare the vesicles.Open in a separate windowFIG. 3.MCPyV VP1 binds to ganglioside GT1b. (A) Structures of gangliosides GM1, GD1a, GD1b, and GT1b. The nature of the glycosidic linkages is indicated. (B) Purified MCPyV VP1 protein was incubated with liposomes only or with liposomes containing the indicated gangliosides. The samples were analyzed as described in the legend to Fig. Fig.2A.2A. Where indicated, GT1b-containing liposomes were pretreated with α2-3,6,8 neuraminidase and analyzed subsequently for virus binding. (C to E) The indicated viruses were incubated with liposomes and analyzed as described in the legend to panel B.As controls, GM1-containing liposomes bound SV40 (Fig. (Fig.3C),3C), GD1a-containing liposomes bound mPyV (Fig. (Fig.3D),3D), and GD1b-containing liposomes bound BKV (Fig. (Fig.3E),3E), demonstrating that the liposomes were functionally intact. We note that, while all of the MCPyV VP1 floated when incubated with liposomes containing GT1b (Fig. (Fig.3B,3B, sixth panel), a significant fraction of SV40, mPyV, and BKV VP1 remained in the bottom fraction despite being incubated with liposomes containing their respective ganglioside receptors (Fig. 3C to E, second panels). This result is likely due to the fact that in contrast to MCPyV, which are assembled as pentamers (Fig. (Fig.1B),1B), the SV40, mPyV, and BKV used in these experiments are fully assembled particles: their larger and denser nature prevents efficient flotation. Nonetheless, we conclude that MCPyV VP1 binds to ganglioside GT1b efficiently.The observation that GD1a does not bind to MCPyV VP1 suggests that the monosialic acid modification on the right branch of GT1b (Fig. (Fig.3A)3A) is insufficient for binding. Similarly, the failure of GD1b to bind MCPyV VP1 suggests that the sialic acid on the left arm of GT1b is necessary for binding. Together, these observations suggest that MCPyV VP1 interacts with sialic acids on both branches of GT1b (Fig. (Fig.4).4). A recent structure of SV40 VP1 in complex with the sugar portion of GM1 (10) demonstrated that although SV40 VP1 binds both branches of GM1 (Fig. (Fig.4),4), only a single sialic acid in GM1 is involved in this interaction. In the case of mPyV, structures of mPyV VP1 in complex with different carbohydrates (12, 13) revealed that the sialic acid-galactose moiety on the left branch of GD1a (and GT1b) is sufficient for mPyV VP1 binding (Fig. (Fig.4).4). Although no structure of BKV in complex with the sugar portion of GD1b (or GT1b) is available, in vitro binding studies (8) have suggested that the disialic acid modification on the right branch of GD1b (and GT1b) is responsible for binding BKV VP1 (Fig. (Fig.4).4). Thus, it appears that the unique feature of the MCPyV VP1-GT1b interaction is that the sialic acids on both branches of this ganglioside are likely involved in capsid binding.Open in a separate windowFIG. 4.A potential model of the different VP1-ganglioside interactions (see the text for discussion).The identification of a potential cellular receptor for MCPyV will facilitate the study of its entry mechanism. An important issue for further study is to determine whether MCPyV targets Merkel cells preferentially, and if so, whether GT1b is found in higher levels in these cells to increase susceptibility.  相似文献   

2.
3.
4.
5.
6.
7.
A pathway toward isobutanol production previously constructed in Escherichia coli involves 2-ketoacid decarboxylase (Kdc) from Lactococcus lactis that decarboxylates 2-ketoisovalerate (KIV) to isobutyraldehyde. Here, we showed that a strain lacking Kdc is still capable of producing isobutanol. We found that acetolactate synthase from Bacillus subtilis (AlsS), which originally catalyzes the condensation of two molecules of pyruvate to form 2-acetolactate, is able to catalyze the decarboxylation of KIV like Kdc both in vivo and in vitro. Mutational studies revealed that the replacement of Q487 with amino acids with small side chains (Ala, Ser, and Gly) diminished only the decarboxylase activity but maintained the synthase activity.We have previously shown that 2-keto acids generated from amino acid biosynthesis can serve as precursors for the Ehrlich degradation pathway (15) to higher alcohols (3). In order to produce isobutanol, the valine biosynthesis pathway was used to generate 2-ketoisovalerate (KIV), the precursor to valine, which was then converted to isobutanol via a decarboxylation and reduction step (Fig. (Fig.1A).1A). The entire pathway to isobutanol from glucose is shown in Fig. Fig.1A.1A. To produce isobutanol, we overexpressed five genes, alsS (Bacillus subtilis), ilvC (Escherichia coli), ilvD (E. coli), kdc (Lactococcus lactis), and ADH2 (Saccharomyces cerevisiae) (Fig. (Fig.1A).1A). This E. coli strain produced 6.8 g/liter isobutanol in 24 h (Fig. (Fig.1B)1B) and more than 20 g/liter in 112 h (3). More recently, we have found that an alcohol dehydrogenase (Adh) encoded by yqhD on the E. coli genome can convert isobutyraldehyde to isobutanol efficiently (5) (Fig. (Fig.1B1B).Open in a separate windowFIG. 1.Schematic representation of the pathway for isobutanol production. (A) The Kdc-dependent synthetic pathway for isobutanol production. (B) Isobutanol production with the Kdc-dependent and -independent synthetic pathways. IlvC, acetohydroxy acid isomeroreductase; IlvD, dihydroxy acid dehydratase. (C) Enzymatic reaction of Als, Ahbs, and Kdc activities.One key reaction in the production of isobutanol is the conversion of KIV to isobutyraldehyde catalyzed by 2-ketoacid decarboxylase (Kdc) (Fig. (Fig.1C).1C). Since E. coli does not have Kdc, kdc from L. lactis was overexpressed. Kdc is a nonoxidative thiamine PPi (TPP)-dependent enzyme and is relatively rare in bacteria, being more frequently found in plants, yeasts, and fungi (8, 19). Several enzymes with Kdc activity have been found, including pyruvate decarboxylase, phenylpyruvate decarboxylase (18), branched-chain Kdc (8, 19), 2-ketoglutarate decarboxylase (10, 17, 20), and indole-3-pyruvate decarboxylase (13).In this work, unexpectedly, we find that Kdc is nonessential for E. coli to produce isobutanol (Fig. (Fig.1).1). An E. coli strain overexpressing only alsS (from B. subtilis), ilvC, and ilvD (both from E. coli) is still able to produce isobutanol. Since E. coli is not a natural producer of isobutanol, it cannot be detected from the culture media in any unmodified strain. We identify that AlsS from B. subtilis, which was introduced in E. coli for acetolactate synthesis (Als), catalyzes the decarboxylation of 2-ketoisovalerate like Kdc both in vivo and in vitro. AlsS is part of the acetoin synthesis pathway and catalyzes the aldo condensation of two molecules of pyruvate to 2-acetolactate (Als activity) (Fig. (Fig.1C)1C) (11). The overall reaction catalyzed by AlsS is irreversible because of CO2 evolution. The first step in catalysis is the ionized thiazolium ring of TPP reacting with the first pyruvate, followed by decarboxylation. This intermediate then reacts with the second pyruvate. Deprotonation followed by C-C bond breakage produces 2-acetolactate. In this work, mutational approaches were used to assess the importance of Q487 in the Kdc activity of AlsS.  相似文献   

8.
9.
10.
11.
12.
13.
14.
15.
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   

16.
17.
18.
The entry of enveloped viruses into host cells is preceded by membrane fusion, which in paramyxoviruses is triggered by the fusion (F) protein. Refolding of the F protein from a metastable conformation to a highly stable postfusion form is critical for the promotion of fusion, although the mechanism is still not well understood. Here we examined the effects of mutations of individual residues of the F protein of Newcastle disease virus, located at critical regions of the protein, such as the C terminus of the N-terminal heptad repeat (HRA) and the N terminus of the C-terminal heptad repeat (HRB). Seven of the mutants were expressed at the cell surface, showing differences in antibody reactivity in comparison with the F wild type. The N211A, L461A, I463A, and I463F mutants showed a hyperfusogenic phenotype both in syncytium and in dye transfer assays. The four mutants promoted fusion more efficiently at lower temperatures than the wild type did, meaning they probably had lower energy requirements for activation. Moreover, the N211A, I463A, and I463F mutants exhibited hemagglutinin-neuraminidase (HN)-independent activity when influenza virus hemagglutinin (HA) was coexpressed as an attachment protein. The data are discussed in terms of alterations of the refolding pathway and/or the stability of the prefusion and fusion conformations.Newcastle disease virus (NDV) is an avian enveloped virus belonging to the family Paramyxoviridae. Two viral membrane-associated proteins are responsible for the entry of the virus into the host cell: they are hemagglutinin-neuraminidase (HN), a receptor-binding protein that interacts with sialoglycoconjugates at the cell surface, and F, a trimeric class I fusion protein that, upon activation, triggers the fusion of the viral and target membranes. F protein is activated after the attachment of its homotypic HN protein to the proper receptor; however, how HN activates F is not well understood. F protein is synthesized as an inactive precursor, F0, that is activated by proteolytic cleavage to the disulfide-linked F1-F2 fusion-competent form (Fig. (Fig.1)1) (10). The crystal structures of several paramyxoviral fusion proteins, in both the prefusion and postfusion conformations (3, 26, 27), have revealed that these proteins undergo major conformational changes, from a metastable conformation to a highly stable, postfusion form. Several regions in the ectodomain of class I viral fusion proteins are involved in these conformational conversions, including a hydrophobic fusion peptide at the N terminus of the F1 protein and two hydrophobic heptad repeat motifs, HRA and HRB, located at its N and C termini, respectively (Fig. (Fig.1).1). In the prefusion form, HRB shows a triple-stranded coiled-coil conformation forming the stalk of the mushroom-like protein (3, 19, 27). Its globular head contains three domains, DI to DIII (Fig. (Fig.1),1), with the base of the head being formed by the DI and DII domains, with residues predominantly located between HRA and HRB. The top of the head is formed by DIII, consisting mainly of HRA and the fusion peptide, located on the side of the head sequestered between adjacent subunits. In this prefusion state, HRA is folded as two antiparallel β-strands and four (h1 to h4) helices (27) (see Fig. Fig.6).6). The DIII domain undergoes major structural changes from the prefusion to the final postfusion conformation. HRA refolds as an α-helix, propelling the fusion peptide into the target membrane and generating a prehairpin intermediate (see Fig. Fig.6).6). The final, stable conformation consists of a six-helical bundle (6HB), comprising a dimer of trimers in which the trimeric HRA coiled coil forms the core, packed along the outside by three antiparallel HRB α-helices (1, 3, 19, 27).Open in a separate windowFIG. 1.Schematic representation of the structure of the NDV fusion protein. (A) Domain structure of F protein (27). (B) Locations of the fusion peptide, HR regions, and sequences studied. Mutated residues are indicated in bold.Open in a separate windowFIG. 6.Scheme of conformational changes in HRA from prefusion to postfusion state. (A) Ribbon model of PIV5 F protein in its metastable prefusion conformation (PDB accession number 2b9b) (27), showing some residues (named in white) from the A subunit and the corresponding residues in the NDV F protein (named in yellow). Subunits B and C are depicted in gray for clarity. (B) In the metastable, prefusion conformation, HRA is folded as a spring-loaded mixture of α-helices, turns, and β-strands, comprising 11 segments in the DIII head domain of the trimer (27). (C) After fusion, HRA is presented as a single long helix that allows the fusion peptide to be buried in the target membrane. The approximate positions of HRC and the core β-sheet are shown as dashed lines for both conformations.The refolding mechanism that triggers F protein activation is still not well understood. Mutational analysis of the HRA and HRB domains of paramyxovirus F proteins (3, 13, 18, 19, 22, 23), as well as the use of HRA- and HRB-derived peptides (6, 17), has led to the proposal of a series of discrete refolding intermediates of the F protein, from the metastable native conformation, through the prehairpin intermediate, and to the final postfusion hairpin structure (6HB) (17, 19, 27). To gain further insight into the individual residues critical for this mechanism, in this work we mutated several residues of the head and stalk of the NDV F protein (Fig. (Fig.1).1). The mutations disrupted F protein antibody reactivity, fusogenicity, and HN dependence in different ways. Interestingly, a mutant of the C-terminal h4 α-helix of HRA (N211A mutant) and two mutants of a residue located at the most N-terminal position of HRB (I463A and I463F mutants) exhibited a hyperfusogenic phenotype and HN-independent activity when influenza virus hemagglutinin (HA) was coexpressed as an attachment protein. The data are discussed in terms of alterations of the refolding pathway and/or the stability of the prefusion and fusion conformations.  相似文献   

19.
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   

20.
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   

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