首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
3.
Like other positive-strand RNA viruses, alphaviruses replicate their genomes in association with modified intracellular membranes. Alphavirus replication sites consist of numerous bulb-shaped membrane invaginations (spherules), which contain the double-stranded replication intermediates. Time course studies with Semliki Forest virus (SFV)-infected cells were combined with live-cell imaging and electron microscopy to reveal that the replication complex spherules of SFV undergo an unprecedented large-scale movement between cellular compartments. The spherules first accumulated at the plasma membrane and were then internalized using an endocytic process that required a functional actin-myosin network, as shown by blebbistatin treatment. Wortmannin and other inhibitors indicated that the internalization of spherules also required the activity of phosphatidylinositol 3-kinase. The spherules therefore represent an unusual type of endocytic cargo. After endocytosis, spherule-containing vesicles were highly dynamic and had a neutral pH. These primary carriers fused with acidic endosomes and moved long distances on microtubules, in a manner prevented by nocodazole. The result of the large-scale migration was the formation of a very stable compartment, where the spherules were accumulated on the outer surfaces of unusually large and static acidic vacuoles localized in the pericentriolar region. Our work highlights both fundamental similarities and important differences in the processes that lead to the modified membrane compartments in cells infected by distinct groups of positive-sense RNA viruses.All positive-strand RNA viruses replicate their genomes in association with cellular membranes. The formation and activity of the membrane-bound replication complexes (RCs) can result in extensive alteration of membrane structures (11, 40, 48). Different viruses use different cytoplasmic membrane compartments as platforms for replication. Currently, there is only a limited understanding of how the virus-encoded and cellular proteins coordinate the formation of the replication-induced membrane structures. We address the mechanisms of membrane-bound replication with alphaviruses, particularly Semliki Forest virus (SFV). The alphaviruses comprise several human and animal pathogens, including the encephalitogenic alphaviruses (e.g., Western, Eastern, and Venezuelan equine encephalitis viruses) as well as the recently reemerging chikungunya virus, which belongs to the SFV clade of alphaviruses. During the past 5 years, chikungunya virus has caused more than 2 million infections and 500 deaths, and a new strain has spread throughout the areas surrounding the Indian Ocean (50). The alphaviruses use mosquitoes as intermediate hosts and transmission vectors, and at present no vaccines or antivirals are available to control these infections.The cytoplasmic replication of alphaviruses depends on the four viral nonstructural (ns) proteins, nsP1 to nsP4, which are all essential and act as a membrane-bound replication complex. The nsPs are translated from the viral positive-sense RNA genome as one large polyprotein. Cleavages catalyzed by the nsP2 moiety result in the release of the individual proteins. A large fraction of the synthesized nsPs is involved in genome replication and associates with membranes, but a sizable fraction dissociates and is distributed in different cellular compartments: nsP1 binds to the inner surface of the plasma membrane (PM); nsP2 is translocated into the nucleus; nsP3 seems to form aggregates in the cytoplasm; and most of the extra nsP4, the core RNA polymerase, is degraded by the proteasome. While the major enzymatic functions of the individual nsPs have been elucidated (21), little is known of how they function together in the replication machinery.As in other positive-strand RNA viruses, the RCs of alphaviruses are associated with altered intracellular membranes, which were first described in the late 1960s and early 1970s (13, 14, 18). In these early studies, it was shown that virus replication induces bulb-shaped membrane invaginations with a diameter of ∼50 nm, which were called spherules. The spherules were found on the limiting membranes of large cytoplasmic vacuoles, which were named virus-induced cytopathic vacuoles of type I (CPV-I). On rare occasions, the spherules were seen also at the PM. By electron microscopic (EM) autoradiography, it was also shown that the spherules both at the CPV-I and at the PM could be sites of RNA synthesis (18). Subsequently, Froshauer et al. (15) showed that CPV-I are positive for endosomal and lysosomal markers. Moreover, using EM, they showed that the inside of the spherule is connected to the cytoplasm by a pore from which electron-dense material (which the authors suggest to be the newly synthesized RNA) seems to diffuse into the cytoplasm.During the past decade, our group has addressed the biogenesis of the CPV-I. We demonstrated that the formation of the spherules did not require structural proteins (44) and, more recently, that all four nsPs were associated with the spherules together with newly formed RNA (labeled by bromouridine), strongly suggesting that they were the actual units of RNA replication (RCs) (28). We also suggested as one possibility that the spherules could first arise at the PM; subsequent endocytosis of the spherules could account for the formation of the CPV-I (28, 44). Of the four nsPs, only nsP1 has affinity for membranes, and when expressed alone, it is specifically targeted to the inner surface of the PM (45). NsP1 is a monotopic membrane protein; its affinity for membranes is dictated by an amphipathic alpha helix, located in the central region of the protein (4, 31). NsP1 has a specific affinity for negatively charged phospholipids, which could potentially account for its prevalent localization to the PM, where such lipids are enriched. Later we showed that the membrane binding of nsP1 through the amphipathic helix is essential for alphavirus replication (56).Several groups of positive-sense RNA viruses make spherules, which appear very similar to those made by the alphaviruses. However, for these virus groups, the spherules arise in different locations. For the well-characterized brome mosaic virus (BMV), a plant virus very distantly related to the alphaviruses, the spherules are seen in the endoplasmic reticulum (ER) adjacent to the nucleus (51). For the unrelated nodaviruses, the spherules are localized on mitochondrial surfaces (25). Recent models of the RCs of flaviviruses suggest that their replication complexes also resemble spherules (62). For the Flaviviridae, the RCs are found on the membranes of the secretory pathway.The aim of this study was to clarify the role of different membranes in the formation and maturation of alphavirus RCs, and particularly to test our hypothesis that the RCs (spherules) are formed at the PM and are internalized thereafter. By using confocal microscopy, live-cell imaging, and novel electron microscopic techniques, we demonstrate that the RCs of SFV undergo an unprecedented, highly dynamic trafficking between different cellular compartments. They are first detected at the PM, which serves as the major platform for spherule formation. A specific endocytic event results in the transfer of spherules to the limiting membrane of small cytoplasmic vesicles. Using pharmacological inhibitors, we have been able to block the internalization process, and we found that the exit of spherules from the PM is dependent on the activity of phosphatidylinositol3- kinase (PI3K). Following the intracellular dynamics associated with spherules in live cells, we show the contribution of actin and microtubule-based transport, as well as that of fusion events with preexisting acidic organelles, providing the first complete model for the biogenesis of the large static CPV-I, where spherules are found at later stages of infection.  相似文献   

4.
Chikungunya virus (CHIKV; genus Alphavirus, family Togaviridae) has recently caused several major outbreaks affecting millions of people. There are no licensed vaccines or antivirals, and the knowledge of the molecular biology of CHIKV, crucial for development of efficient antiviral strategies, remains fragmentary. CHIKV has a 12 kb positive-strand RNA genome, which is translated to yield a nonstructural (ns) or replicase polyprotein. CHIKV structural proteins are expressed from a subgenomic RNA synthesized in infected cells. Here we have developed CHIKV trans-replication systems, where replicase expression and RNA replication are uncoupled. Bacteriophage T7 RNA polymerase or cellular RNA polymerase II were used for production of mRNAs for CHIKV ns polyprotein and template RNAs, which are recognized by CHIKV replicase and encode for reporter proteins. CHIKV replicase efficiently amplified such RNA templates and synthesized large amounts of subgenomic RNA in several cell lines. This system was used to create tagged versions of ns proteins including nsP1 fused with enhanced green fluorescent protein and nsP4 with an immunological tag. Analysis of these constructs and a matching set of replicon vectors revealed that the replicases containing tagged ns proteins were functional and maintained their subcellular localizations. When cells were co-transfected with constructs expressing template RNA and wild type or tagged versions of CHIKV replicases, formation of characteristic replicase complexes (spherules) was observed. Analysis of mutations associated with noncytotoxic phenotype in CHIKV replicons showed that a low level of RNA replication is not a pre-requisite for reduced cytotoxicity. The CHIKV trans-replicase does not suffer from genetic instability and represents an efficient, sensitive and reliable tool for studies of different aspects of CHIKV RNA replication process.  相似文献   

5.
For positive-strand RNA viruses, the viral genomic RNA also acts as an mRNA directing the translation of the replicase proteins of the virus. Replication takes place in association with cytoplasmic membranes, which are heavily modified to create specific replication compartments. Here we have expressed by plasmid DNA transfection the large replicase polyprotein of Semliki Forest virus (SFV) in mammalian cells from a nonreplicating mRNA and provided a separate RNA containing the replication signals. The replicase proteins were able to efficiently and specifically replicate the template in trans, leading to accumulation of RNA and marker gene products expressed from the template RNA. The replicase proteins and double-stranded RNA replication intermediates localized to structures similar to those seen in SFV-infected cells. Using correlative light electron microscopy (CLEM) with fluorescent marker proteins to relocate those transfected cells, in which active replication was ongoing, abundant membrane modifications, representing the replication complex spherules, were observed both at the plasma membrane and in intracellular endolysosomes. Thus, replication complexes are faithfully assembled and localized in the trans-replication system. We demonstrated, using CLEM, that the replication proteins alone or a polymerase-negative polyprotein mutant together with the template did not give rise to spherule formation. Thus, the trans-replication system is suitable for cell biological dissection and examination in a mammalian cell environment, and similar systems may be possible for other positive-strand RNA viruses.  相似文献   

6.
Alphavirus replication complexes that are located in the mitochondrial fraction of infected cells which pellets at 15,000 x g (P15 fraction) were used for the in vitro synthesis of viral 49S genome RNA, subgenomic 26S mRNA, and replicative intermediates (RIs). Comparison of the polymerase activity in P15 fractions from Sindbis virus (SIN)- and Semliki Forest virus (SFV)-infected cells indicated that both had similar kinetics of viral RNA synthesis in vitro but the SFV fraction was twice as active and produced more labeled RIs than SIN. When assayed in vitro under conditions of high specific activity, which limits incorporation into RIs, at least 70% of the polymerase activity was recovered after detergent treatment. Treatment with Triton X-100 or with Triton X-100 plus deoxycholate (DOC) solubilized some prelabeled SFV RIs but little if any SFV or SIN RNA polymerase activity from large structures that also contained cytoskeletal components. Treatment with concentrations of DOC greater than 0.25% or with 1% Triton X-100-0.5% DOC in the presence of 0.5 M NaCl released the polymerase activity in a soluble form, i.e., it no longer pelleted at 15,000 x g. The DOC-solubilized replication complexes, identified by their polymerase activity in vitro and by the presence of prelabeled RI RNA, had a density of 1.25 g/ml, were 20S to 100S in size, and contained viral nsP1, nsP2, phosphorylated nsP3, nsP4, and possibly nsP34 proteins. Immunoprecipitation of the solubilized structures indicated that the nonstructural proteins were complexed together and that a presumed cellular protein of approximately 120 kDa may be part of the complex. Antibodies specific for nsP3, and to a lesser extent antibodies to nsP1, precipitated native replication complexes that retained prelabeled RIs and were active in vitro in viral RNA synthesis. Thus, antibodies to nsP3 bound but did not disrupt or inhibit the polymerase activity of replication complexes in vitro.  相似文献   

7.
Alphavirus replicase protein nsP1 has multiple functions during viral RNA synthesis. It catalyzes methyltransferase and guanylyltransferase activities needed in viral mRNA capping, attaches the viral replication complex to cytoplasmic membranes, and is required for minus-strand RNA synthesis. Two temperature-sensitive (ts) mutations in Semliki Forest virus (SFV) were previously identified within nsP1: ts10 (E529D) and ts14 (D119N). Recombinant viruses containing these individual mutations reproduced the features of the original ts strains. We now find that the capping-associated enzymatic activities of recombinant nsP1, containing ts10 or ts14 lesions, were not ts. The mutant proteins and polyproteins also were membrane bound, mutant nsP1 interacted normally with the other nonstructural proteins, and there was no major defect in nonstructural polyprotein processing in the mutants, although ts14 surprisingly displayed slightly retarded processing. The two mutant viruses were specifically defective in minus-strand RNA synthesis at the restrictive temperature. Integrating data from SFV and Sindbis virus, we discuss the domain structure of nsP1 and the relative positioning of and interactions between the replicase proteins. nsP1 is suggested to contain a specific subdomain involved in minus-strand synthesis and interaction with the polymerase nsP4 and the protease nsP2.  相似文献   

8.
Direct ferritin immunoelectron microscopy was used to visualize the asialoglycoprotein receptor in various rat liver subcellular fractions. The cytoplasmic surfaces of cytoplasmic organelles such as the rough and smooth microsomes, Golgi cisternae and lysosomes showed hardly any ferritin label exception for the slight labeling of secretory granules found mainly in the light Golgi fraction (GF1). Occasionally, however, open membrane sheet structures, smooth vesicular or tubular structures heavily labeled with ferritin, were present in all these subcellular fractions. These structures probably correspond to fragmented sinusoidal or lateral hepatocyte plasma membranes recovered to these subcellular fractions. When the limiting membranes of the secretion granules were partially broken by mechanical force, a number of ferritin particles frequently were seen attached in large clusters to the luminal surface of the membrane, the cytoplasmic surface of the corresponding domain being slightly labeled. These observations are strong evidence that the receptor protein is never translocated vertically throughout the intracellular transport from ER to plasma membrane via Golgi apparatus and from plasma membrane back to trans-Golgi elements and also in lysosomes, always exposing the major antigenic sites to the luminal or extracellular surface and the minor counterparts to the cytoplasmic surface of the membranes. The receptor protein also is suggested to be concentrated in clusters on the luminal surface of secretion granules when they form on the trans-side of the Golgi apparatus.  相似文献   

9.
10.
One of the distinguishing features of the alphaviruses is a sequential processing of the nonstructural polyproteins P1234 and P123. In the early stages of the infection, the complex of P123+nsP4 forms the primary replication complexes (RCs) that function in negative-strand RNA synthesis. The following processing steps make nsP1+P23+nsP4, and later nsP1+nsP2+nsP3+nsP4. The latter mature complex is active in positive-strand RNA synthesis but can no longer produce negative strands. However, the regulation of negative- and positive-strand RNA synthesis apparently is not the only function of ns polyprotein processing. In this study, we developed Sindbis virus mutants that were incapable of either P23 or P123 cleavage. Both mutants replicated in BHK-21 cells to levels comparable to those of the cleavage-competent virus. They continuously produced negative-strand RNA, but its synthesis was blocked by the translation inhibitor cycloheximide. Thus, after negative-strand synthesis, the ns proteins appeared to irreversibly change conformation and formed mature RCs, in spite of the lack of ns polyprotein cleavage. However, in the cells having no defects in alpha/beta interferon (IFN-alpha/beta) production and signaling, the cleavage-deficient viruses induced a high level of type I IFN and were incapable of causing the spread of infection. Moreover, the P123-cleavage-deficient virus was readily eliminated, even from the already infected cells. We speculate that this inability of the viruses with unprocessed polyprotein to productively replicate in the IFN-competent cells and in the cells of mosquito origin was an additional, important factor in ns polyprotein cleavage development. In the case of the Old World alphaviruses, it leads to the release of nsP2 protein, which plays a critical role in inhibiting the cellular antiviral response.  相似文献   

11.
The subcellular distribution of the alpha subunit(s) of Gi has an obvious bearing on the ability of this protein to interact with receptors and targets and on its potential to serve in still unexplored capacities. In this study, we have examined the distribution of Gi alpha by means of light and electron microscopy. The cells employed were mouse 3T3 fibroblasts, normal rat kidney fibroblasts, rat C6 glioma cells, human umbilical vein endothelial cells, and human 293 kidney fibroblasts. By indirect immunofluorescence, two patterns of Gi alpha were evident. The more prominent was that associated with phase-dense, cytoplasmic structures exhibiting a tubule-like morphology. A similar distribution was noted for mitochondria, indicating attachment to a subset of microtubules. The second pattern appeared as a diffuse, particulate fluorescence associated with the plasma membrane. By immunogold labeling and electron microscopy, two populations of Gi alpha were again evident. In this instance, labeling of the plasma membrane was the more prominent. Gold particles were most often evenly distributed along the plasma membrane and were concentrated along microspikes. The second, less abundant population of Gi alpha represented the subunit (or fragments) within lysosomes. Specificity in immunolabeling was confirmed in all instances by immunotransfer blotting, the use of antibodies differing in specificities for epitopes within Gi alpha, the absence of labeling with preimmune sera, and the decrease in labeling after preincubation of antisera with appropriate peptides. These results support the proposal that several populations of Gi alpha exist: those evident within the cytoplasm by immunofluorescence, those present at the plasma membrane, and those evident within lysosomes by immunogold labeling.  相似文献   

12.
Influenza A virus RNA genome exists as eight-segmented ribonucleoprotein complexes containing viral RNA polymerase and nucleoprotein (vRNPs). Packaging of vRNPs and virus budding take place at the apical plasma membrane (APM). However, little is known about the molecular mechanisms of apical transport of newly synthesized vRNP. Transfection of fluorescent-labeled antibody and subsequent live cell imaging revealed that punctate vRNP signals moved along microtubules rapidly but intermittently in both directions, suggestive of vesicle trafficking. Using a series of Rab family protein, we demonstrated that progeny vRNP localized to recycling endosome (RE) in an active/GTP-bound Rab11-dependent manner. The vRNP interacted with Rab11 through viral RNA polymerase. The localization of vRNP to RE and subsequent accumulation to the APM were impaired by overexpression of Rab binding domains (RBD) of Rab11 family interacting proteins (Rab11-FIPs). Similarly, no APM accumulation was observed by overexpression of class II Rab11-FIP mutants lacking RBD. These results suggest that the progeny vRNP makes use of Rab11-dependent RE machinery for APM trafficking.  相似文献   

13.
The limiting membrane of the lysosome contains a group of transmembrane glycoproteins named lysosome-associated membrane proteins (Lamps). These proteins are targeted to lysosomes by virtue of tyrosine-based sorting signals in their cytosolic tails. Four adaptor protein (AP) complexes, AP-1, AP-2, AP-3, and AP-4, interact with such signals and are therefore candidates for mediating sorting of the Lamps to lysosomes. However, the role of these complexes and of the coat protein, clathrin, in sorting of the Lamps in vivo has either not been addressed or remains controversial. We have used RNA interference to show that AP-2 and clathrin-and to a lesser extent the other AP complexes-are required for efficient delivery of the Lamps to lysosomes. Because AP-2 is exclusively associated with plasma membrane clathrin coats, our observations imply that a significant population of Lamps traffic via the plasma membrane en route to lysosomes.  相似文献   

14.
Human astrovirus nonstructural C-terminal nsP1a protein (nsP1a/4) colocalizes with the endoplasmic reticulum and viral RNA. It has been suggested that nsP1a/4 protein is involved in the RNA replication process in endoplasmic reticulum-derived intracellular membranes. A hypervariable region (HVR) is contained in the nsP1a/4 protein, and different replicative patterns can be distinguished depending on its variability. In the present work, both the astrovirus RNA-dependent RNA polymerase and four types (IV, V, VI, and XII) of nsP1a/4 proteins have been cloned and expressed in the baculovirus system to analyze their interactions. Different isoforms of each of the nsP1a/4 proteins exist: a nonphosphorylated isoform and different phosphorylated isoforms. While the polymerase accumulates as a monomer, the nsP1a/4 proteins accumulate as oligomers. The oligomerization domain of nsP1a/4-V is mapped between residues 176 and 209. For all studied genotypes, oligomers mainly contain the nonphosphorylated isoform. When RNA polymerase is coexpressed with nsP1a/4 proteins, they interact, likely forming heterodimers. The polymerase binding region has been mapped in the nsP1a/4-V protein between residues 88 and 176. Phosphorylated isoforms of nsP1a/4 type VI show a stronger interactive pattern with the polymerase than the nonphosphorylated isoform. This difference is not observed in genotypes IV and V, suggesting a role of the HVR in modulating the interaction of the nsP1a/4 protein with the polymerase through phosphorylation/dephosphorylation of some critical residues.  相似文献   

15.
The aim of the present study was to define the site of replication of the coronavirus mouse hepatitis virus (MHV). Antibodies directed against several proteins derived from the gene 1 polyprotein, including the 3C-like protease (3CLpro), the putative polymerase (POL), helicase, and a recently described protein (p22) derived from the C terminus of the open reading frame 1a protein (CT1a), were used to probe MHV-infected cells by indirect immunofluorescence (IF) and electron microscopy (EM). At early times of infection, all of these proteins showed a distinct punctate labeling by IF. Antibodies to the nucleocapsid protein also displayed a punctate labeling that largely colocalized with the replicase proteins. When infected cells were metabolically labeled with 5-bromouridine 5'-triphosphate (BrUTP), the site of viral RNA synthesis was shown by IF to colocalize with CT1a and the 3CLpro. As shown by EM, CT1a localized to LAMP-1 positive late endosomes/lysosomes while POL accumulated predominantly in multilayered structures with the appearance of endocytic carrier vesicles. These latter structures were also labeled to some extent with both anti-CT1a and LAMP-1 antibodies and could be filled with fluid phase endocytic tracers. When EM was used to determine sites of BrUTP incorporation into viral RNA at early times of infection, the viral RNA localized to late endosomal membranes as well. These results demonstrate that MHV replication occurs on late endosomal membranes and that several nonstructural proteins derived from the gene 1 polyprotein may participate in the formation and function of the viral replication complexes.  相似文献   

16.
Viral infections cause profound alterations in host cells. Here, we explore the interactions between proteins of the Alphavirus Sindbis and host factors during the course of mammalian cell infection. Using a mutant virus expressing the viral nsP3 protein tagged with green fluorescent protein (GFP) we directly observed nsP3 localization and isolated nsP3-interacting proteins at various times after infection. These results revealed that host factor recruitment to nsP3-containing complexes was time dependent, with a specific early and persistent recruitment of G3BP and a later recruitment of 14-3-3 proteins. Expression of GFP-tagged G3BP allowed reciprocal isolation of nsP3 in Sindbis infected cells, as well as the identification of novel G3BP-interacting proteins in both uninfected and infected cells. Note-worthy interactions include nuclear pore complex components whose interactions with G3BP were reduced upon Sindbis infection. This suggests that G3BP is a nuclear transport factor, as hypothesized previously, and that viral infection may alter RNA transport. Immunoelectron microscopy showed that a portion of Sindbis nsP3 is localized at the nuclear envelope, suggesting a possible site of G3BP recruitment to nsP3-containing complexes. Our results demonstrate the utility of using a standard GFP tag to both track viral protein localization and elucidate specific viral-host interactions over time in infected mammalian cells.  相似文献   

17.
Positive-strand RNA [(+)RNA] viruses invariably replicate their RNA genomes on modified intracellular membranes. In infected Drosophila cells, Flock House nodavirus (FHV) RNA replication complexes form on outer mitochondrial membranes inside ~50-nm, virus-induced spherular invaginations similar to RNA replication-linked spherules induced by many (+)RNA viruses at various membranes. To better understand replication complex assembly, we studied the mechanisms of FHV spherule formation. FHV has two genomic RNAs; RNA1 encodes multifunctional RNA replication protein A and RNA interference suppressor protein B2, while RNA2 encodes the capsid proteins. Expressing genomic RNA1 without RNA2 induced mitochondrial spherules indistinguishable from those in FHV infection. RNA1 mutation showed that protein B2 was dispensable and that protein A was the only FHV protein required for spherule formation. However, expressing protein A alone only "zippered" together the surfaces of adjacent mitochondria, without inducing spherules. Thus, protein A is necessary but not sufficient for spherule formation. Coexpressing protein A plus a replication-competent FHV RNA template induced RNA replication in trans and membrane spherules. Moreover, spherules were not formed when replicatable FHV RNA templates were expressed with protein A bearing a single, polymerase-inactivating amino acid change or when wild-type protein A was expressed with a nonreplicatable FHV RNA template. Thus, unlike many (+)RNA viruses, the membrane-bounded compartments in which FHV RNA replication occurs are not induced solely by viral protein(s) but require viral RNA synthesis. In addition to replication complex assembly, the results have implications for nodavirus interaction with cell RNA silencing pathways and other aspects of virus control.  相似文献   

18.
The replication of most positive-strand RNA viruses and retroviruses is regulated by proteolytic processing. Alphavirus replicase proteins are synthesized as a polyprotein, called P1234, which is cleaved into nsP1, nsP2, nsP3, and nsP4 by the carboxyl-terminal protease domain of nsP2. The cleavage intermediate P123+nsP4 synthesizes minus-strand copies of the viral RNA genome, whereas the completely processed complex is required for plus-strand synthesis. To understand the mechanisms responsible for this sequential proteolysis, we analyzed in vitro translated Semliki Forest virus polyproteins containing noncleavable processing sites or various deletions. Processing of each of the three sites in vitro required a different type of activity. Site 3/4 was cleaved in trans by nsP2, its carboxyl-terminal fragment Pro39, and by all polyprotein proteases. Site 1/2 was cleaved in cis with a half-life of about 20-30 min. Site 2/3 was cleaved rapidly in trans but only after release of nsP1 from the polyprotein exposing an "activator" sequence present in the amino terminus of nsP2. Deletion of amino-terminal amino acids of nsP2 or addition of extra amino acid residues to its amino terminus specifically inhibited the protease activity that processes the 2/3 site. This sequence of delayed processing of P1234 would explain the accumulation of P123 plus nsP4, the early short-lived minus-strand replicase. The polyprotein stage would allow correct assembly and membrane association of the RNA-polymerase complex. Late in infection free nsP2 would cleave at site 2/3 yielding P12 and P34, the products of which, nsP1-4, are distributed to the plasma membrane, nucleus, cytoplasmic aggregates, and proteasomes, respectively.  相似文献   

19.
Recent insights into the early events in Sindbis virus RNA replication suggest a requirement for either the P123 or P23 polyprotein, as well as mature nsP4, the RNA-dependent RNA polymerase, for initiation of minus-strand RNA synthesis. Based on this observation, we have succeeded in reconstituting an in vitro system for template-dependent initiation of SIN RNA replication. Extracts were isolated from cells infected with vaccinia virus recombinants expressing various SIN proteins and assayed by the addition of exogenous template RNAs. Extracts from cells expressing P123C>S, a protease-defective P123 polyprotein, and nsP4 synthesized a genome-length minus-sense RNA product. Replicase activity was dependent upon addition of exogenous RNA and was specific for alphavirus plus-strand RNA templates. RNA synthesis was also obtained by coexpression of nsP1, P23C>S, and nsP4. However, extracts from cells expressing nsP4 and P123, a cleavage-competent P123 polyprotein, had much less replicase activity. In addition, a P123 polyprotein containing a mutation in the nsP2 protease which increased the efficiency of processing exhibited very little, if any, replicase activity. These results provide further evidence that processing of the polyprotein inactivates the minus-strand initiation complex. Finally, RNA synthesis was detected when soluble nsP4 was added to a membrane fraction containing P123C>S, thus providing a functional assay for purification of the nsP4 RNA polymerase.  相似文献   

20.
The membrane-associated alphavirus RNA replication complex contains four virus-encoded subunits, the nonstructural proteins nsP1 to nsP4. Semliki Forest virus (SFV) nsP1 is hydrophobically modified by palmitoylation of cysteines 418 to 420. Here we show that Sindbis virus nsP1 is also palmitoylated on the same site (cysteine 420). When mutations preventing nsP1 palmitoylation were introduced into the genomes of these two alphaviruses, the mutant viruses remained viable and replicated to high titers, although their growth was slightly delayed. The subcellular distribution of palmitoylation-defective nsP1 was altered in the mutant: it no longer localized to filopodial extensions, and a fraction of it was soluble. The ultrastructure of the alphavirus replication sites appeared normal, and the localization of the other nonstructural proteins was unaltered in the mutants. In both wild-type- and mutant-virus-infected cells, SFV nsP3 and nsP4 could be extracted from membranes only by alkaline solutions whereas the nsP2-membrane association was looser. Thus, the membrane binding properties of the alphavirus RNA replication complex were not determined by the palmitoylation of nsP1. The nsP1 palmitoylation-defective alphaviruses produced normal plaques in several cell types, but failed to give rise to plaques in HeLa cells, although they induced normal apoptosis of these cells. The SFV mutant was apathogenic in mice: it caused blood viremia, but no infectious virus was detected in the brain.  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