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 46 毫秒
1.
Studies of the hepatitis C virus (HCV) life cycle have been aided by development of in vitro systems that enable replication of viral RNA and production of infectious virus. However, the functions of the individual proteins, especially those engaged in RNA replication, remain poorly understood. It is considered that NS4B, one of the replicase components, creates sites for genome synthesis, which appear as punctate foci at the endoplasmic reticulum (ER) membrane. In this study, a panel of mutations in NS4B was generated to gain deeper insight into its functions. Our analysis identified five mutants that were incapable of supporting RNA replication, three of which had defects in production of foci at the ER membrane. These mutants also influenced posttranslational modification and intracellular mobility of another replicase protein, NS5A, suggesting that such characteristics are linked to focus formation by NS4B. From previous studies, NS4B could not be trans-complemented in replication assays. Using the mutants that blocked RNA synthesis, defective NS4B expressed from two mutants could be rescued in trans-complementation replication assays by wild-type protein produced by a functional HCV replicon. Moreover, active replication could be reconstituted by combining replicons that were defective in NS4B and NS5A. The ability to restore replication from inactive replicons has implications for our understanding of the mechanisms that direct viral RNA synthesis. Finally, one of the NS4B mutations increased the yield of infectious virus by five- to sixfold. Hence, NS4B not only functions in RNA replication but also contributes to the processes engaged in virus assembly and release.Recent estimates predict that the prevalence of hepatitis C virus (HCV) infection is approximately 2.2% worldwide, equivalent to about 130 million persons (22). The virus typically establishes a chronic infection that frequently leads to serious liver disease (1), and current models indicate that both morbidity and mortality as a consequence of HCV infection will continue to rise for about the next 20 years (10, 11, 29).HCV is the only assigned species of the Hepacivirus genus within the family Flaviviridae. The virus can be classified into six genetic groups or clades (numbered 1 to 6) and then further separated into subtypes (e.g., 1a, 1b, 2a, 2b, etc.) (53, 55). HCV has a single-stranded, positive-sense RNA genome that is approximately 9.6 kb in length (reviewed in reference 46). Genomic RNA carries a single open reading frame flanked by 5′ and 3′ nontranslated regions, which are important for both replication and translation (19, 20, 34, 47, 56). Viral RNA is translated by the host ribosomal machinery, and the resultant polyprotein is co- and posttranslationally cleaved to generate the mature viral proteins. The structural proteins (core, E1, and E2) and a small hydrophobic polypeptide called p7 are produced by the cellular proteases signal peptidase and signal peptide peptidase (28, 45, 54). Two virus-encoded proteases, the NS2-3 autoprotease and the NS3 serine protease (5, 13, 26), are responsible for maturation of the nonstructural (NS) proteins (NS2, NS3, NS4A, NS4B, NS5A, and NS5B). With the exception of NS2, the NS proteins are necessary for genome replication (8, 40) and form replication complexes (RCs), which are located at the endoplasmic reticulum (ER) membrane (14, 24, 52, 57, 59). The functions of all viral constituents of RCs have not been characterized in detail. It is known that NS5B is the RNA-dependent RNA polymerase (6), while NS3 possesses helicase and nucleoside triphosphatase activities in addition to acting as a protease (32, 58). However, the precise roles of the other proteins remain to be firmly established.Expression of NS4B, one of the replicase proteins, generates rearrangements at the ER membrane that have been termed the “membranous web” (14, 24) and “membrane-associated foci” (MAFs) (25). Detection of viral RNA at such foci suggests that NS4B is involved in creating the sites where genome synthesis occurs (18, 24, 59). It is predicted that NS4B has an amphipathic α-helix within its N-terminal region, which is followed by four transmembrane domains (TMDs) in the central portion of the protein (17, 42). As a result, the majority of NS4B is likely to be tightly anchored to membranes, and experimental evidence indicates that it has characteristics consistent with an integral membrane protein (27). It is thought that after membrane association, NS4B rearranges membranes into a network, thereby generating foci which act as a “scaffold” to facilitate RNA replication. The mechanisms engaged in formation of foci are not known but include the notion that the NS4B N terminus can translocate into the ER lumen, resulting in rearrangement of cellular membranes (41, 42). Alternatively, palmitoylation, a lipid modification, might facilitate polymerization of NS4B, in turn promoting formation of RCs on the ER membrane (68).Apart from inducing membranous changes required for replication, NS4B may perform other tasks in HCV RNA synthesis. For example, studies of cell culture adaptive mutations in subgenomic replicons (SGRs) have identified amino acid changes that can stimulate RNA production (39), suggesting that NS4B may exert a regulatory role in determining replication efficiency. In support of a regulatory function, replacement of NS4B sequences in an SGR from strain H77 (a genotype 1a strain) with those from strain Con-1 (a genotype 1b strain) gave higher levels of replication than for a wild-type (wt) strain H77 SGR (7). The corresponding replacement of strain Con-1 NS4B sequences with those from strain H77 reduced the replication efficiency of a Con-1 SGR (7). Moreover, interactions of NS4B with the RC can affect the behavior of other replicase proteins. For example, NS4B is needed for hyperphosphorylation of NS5A (35, 48) and restricts its intracellular movement (30).To try to gain greater insight into the functional organization of the components that constitute RCs, trans-complementation assays using defective and helper SGRs have been established (2, 64). Such studies reveal that the only protein capable of trans-complementation is NS5A, while active replication cannot be restored for replicons harboring deleterious mutations in NS3, NS4B, and NS5B. These data led to the conclusion that functional NS5A may be able to exchange between RCs (2), whereas, by inference, such exchange would not be possible for other HCV replicase proteins. In transient-replication assays, complementation by NS5A also relied on its expression as part of a polyprotein (minimally NS3-NS5A), and production of the protein alone failed to restore replication for an inactive SGR (2). However, in a separate study, stable expression of wt NS5A was capable of complementing a defective replicon (64). Thus, different assay systems can give dissimilar results for complementation by NS5A.In this study, we have created a series of mutations in the NS4B gene of HCV strain JFH1 (31) to explore the function of the protein in the HCV life cycle. We focused our attention on the C-terminal portion of NS4B, downstream from the predicted TMD regions, since it is relatively well conserved and is predicted to lie on the cytosolic side of the ER membrane (15, 42). Our analysis examines the impact of mutations on replication efficiency and the intracellular characteristics of the mutants compared to the behavior of the wt protein. In addition, we have utilized this series of mutants to reassess trans-complementation of NS4B in replication assays. Finally, we also analyze the impact of mutations which do not affect replication on the production of infectious virus to determine whether NS4B plays a role in virus assembly and release.  相似文献   

2.
Replication of hepatitis C virus (HCV) RNA occurs on intracellular membranes, and the replication complex (RC) contains viral RNA, nonstructural proteins, and cellular cofactors. We previously demonstrated that cyclophilin A (CyPA) is an essential cofactor for HCV infection and the intracellular target of cyclosporine''s anti-HCV effect. Here we investigate the mechanism by which CyPA facilitates HCV replication. Cyclosporine treatment specifically blocked the incorporation of NS5B into the RC without affecting either the total protein level or the membrane association of the protein. Other nonstructural proteins or viral RNAs in the RC were not affected. NS5B from the cyclosporine-resistant replicon was resistant to this disruption of RC incorporation. We also isolated membrane fractions from both naïve and HCV-positive cells and found that CyPA is recruited into membrane fractions in HCV-replicating cells via an interaction with RC-associated NS5B, which is sensitive to cyclosporine treatment. Finally, we introduced point mutations in the prolyl-peptidyl isomerase (PPIase) motif of CyPA and demonstrated a critical role of this motif in HCV replication in cDNA rescue experiments. We propose a model in which the incorporation of the HCV polymerase into the RC depends on its interaction with a cellular chaperone protein and in which cyclosporine inhibits HCV replication by blocking this critical interaction and the PPIase activity of CyPA. Our results provide a mechanism of action for the cyclosporine-mediated inhibition of HCV and identify a critical role of CyPA''s PPIase activity in the proper assembly and function of the HCV RC.Hepatitis C virus (HCV), of the family Flaviviridae, is an enveloped, positive-stranded RNA virus. Spread mostly by blood-borne transmission, HCV infects more than 170 million people worldwide. The viral genome is composed of a single open reading frame (ORF) plus 5′- and 3′-nontranslated regions. The ORF encodes a large polyprotein that is cleaved by cellular and viral proteases into 10 viral proteins. The structural proteins, including the capsid protein (core), two glycoproteins (E1 and E2), and a small ion channel protein (p7), reside in the N-terminal half of the polyprotein. The rest of the ORF encodes six nonstructural (NS) proteins: NS2, NS3, NS4A, NS4B, NS5A, and NS5B. NS3 through NS5B assemble into a replication complex (RC) and are necessary and sufficient for HCV RNA replication in cell culture (8, 42). NS3 is a multifunctional protein with both a serine protease and an RNA helicase activity. The protease activity is responsible for cleavage at the NS3-NS4A, NS4A-NS4B, NS4B-NS5A, and NS5A-NS5B junctions (5), and the helicase activity is probably required to unwind the double-stranded RNA intermediates formed during replication (38). NS4A serves as an essential cofactor for the NS3 protease and anchors the NS3 protein to intracellular membranes (25, 36, 39). NS4B induces the formation of a “membranous web” that is probably the site of HCV replication (16). It also contains a GTP-binding motif that is required for replication (17). The web is derived from the endoplasmic reticulum (ER) compartment, although proteins of early-endosome origin have also been found to locate to the web (62). NS5A is a phosphoprotein and an integral component of the viral RC. The precise function of NS5A in replication is still unknown but appears to be regulated by phosphorylation and its interaction with several cellular proteins (19, 22, 24, 51, 52, 59, 63, 67). In addition, it may be involved in the transition from replication and particle formation (4, 45, 64). NS5B is the RNA-dependent RNA polymerase that is responsible for copying the RNA genome of the virus during replication. Several cellular cofactors interact with NS5B and modulate its activity in the context of the viral RC (22, 24, 35, 69, 71).Positive-stranded RNA viruses alter the intracellular membranes of host cells to form an RC in which RNA replication occurs. Modifications include the proliferation and reorganization of certain cellular membranes (1). HCV forms an RC associated with altered cellular membranes (16, 23), and crude RCs (CRCs) that maintain the replicase activity in vitro can be isolated by membrane sedimentation or flotation techniques (2, 3, 18, 27, 37).Cyclosporine is a widely used immunosuppressive and anti-inflammatory drug for organ transplant patients. It functions by forming an inhibitory complex with cyclophilins (CyPs) that inhibits the phosphatase activity of calcineurin, which is important for T-cell activation. In recent years, cyclosporine and its derivatives have been shown to be highly effective in suppressing HCV replication in vitro (44, 49, 53, 68) and in vivo (30). The mechanism of this inhibition is independent of its immunosuppressive function and distinct from that of interferon (IFN) (44, 53, 56, 68).We recently showed that HCV infection in vitro is inhibited when CyPA, a major intracellular target of cyclosporine, is downregulated by RNA interference, and mutations in NS5B that confer cyclosporine-resistant binding to CyPA contribute to the cyclosporine resistance of the replicons harboring these mutations (56, 71). Here we report that CyPA is recruited into the HCV RC together with NS5B in HCV replicon or in HCV-infected cells. Cyclosporine disrupts the association between RC-incorporated NS5B and CyPA and results in an exclusion of the polymerase from the viral RC. We also show that the prolyl-peptidyl isomerase (PPIase) motif of CyPA is essential for HCV replication.  相似文献   

3.
Nonstructural protein 4B (NS4B) plays an essential role in the formation of the hepatitis C virus (HCV) replication complex. It is a relatively poorly characterized integral membrane protein predicted to comprise four transmembrane segments in its central portion. Here, we describe a novel determinant for membrane association represented by amino acids (aa) 40 to 69 in the N-terminal portion of NS4B. This segment was sufficient to target and tightly anchor the green fluorescent protein to cellular membranes, as assessed by fluorescence microscopy as well as membrane extraction and flotation analyses. Circular dichroism and nuclear magnetic resonance structural analyses showed that this segment comprises an amphipathic α-helix extending from aa 42 to 66. Attenuated total reflection infrared spectroscopy and glycosylation acceptor site tagging revealed that this amphipathic α-helix has the potential to traverse the phospholipid bilayer as a transmembrane segment, likely upon oligomerization. Alanine substitution of the fully conserved aromatic residues on the hydrophobic helix side abrogated membrane association of the segment comprising aa 40 to 69 and disrupted the formation of a functional replication complex. These results provide the first atomic resolution structure of an essential membrane-associated determinant of HCV NS4B.With 120 to 180 million chronically infected individuals worldwide, hepatitis C virus (HCV) infection represents a major cause of chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma (38). HCV contains a 9.6-kb positive-strand RNA genome that encodes a polyprotein of about 3,000 amino acids (reviewed in references 36 and 51). The polyprotein precursor is co- and posttranslationally processed by cellular and viral proteases to yield the mature structural and nonstructural proteins. The structural proteins include the core and the envelope glycoproteins E1 and E2. The nonstructural proteins include the p7 ion channel polypeptide, the NS2-3 and NS3-4A proteases, an RNA helicase located in the C-terminal two-thirds of NS3, the NS4B and NS5A proteins, and the NS5B RNA-dependent RNA polymerase. HCV replication takes place in a membrane-associated complex, composed of viral proteins, replicating RNA, altered cellular membranes, and other host factors (7, 18, 31, 43). Determinants for membrane association of the HCV nonstructural proteins have been mapped and a likely endoplasmic reticulum (ER)-derived membrane alteration, designated the membranous web, was found to harbor the HCV replication complex (7, 18; reviewed in reference 36).NS4B is a 27-kDa integral ER membrane protein (21). The expression of NS4B alone induces the formation of the membranous web (7). Thus, an essential function of NS4B is the induction of the specific membrane alteration that serves as a scaffold for the HCV replication complex. In addition, a nucleotide-binding motif has been proposed to reside in the middle of NS4B (8), and RNA binding properties have recently been reported for NS4B (9).Both the N and the C termini of NS4B are believed to be oriented toward the cytosol, and prediction algorithms indicate the presence of four putative transmembrane segments in the central portion of the protein (21, 27, 28, 42). The cytosolic orientation of the bulk of the protein was confirmed experimentally (21), but a more refined membrane topology is so far elusive. The introduction of glycosylation acceptor sites at various positions in NS4B validated the prediction of ER luminal loops around amino acid positions 112 and 161 (27, 28). Intriguingly, the N terminus of NS4B was reported to be translocated into the ER lumen at least partially, presumably by a posttranslational mechanism (28). Interestingly, the coexpression of the other HCV proteins appears to limit this translocation (27). A recent report indicates that NS4B is palmitoylated at C-terminal residues Cys 257 and Cys 261 and forms oligomers (59).In order to define the membrane topology of NS4B we have analyzed a comprehensive panel of green fluorescent protein (GFP) fusion constructs comprising different segments of NS4B (N. Arora, V. Castet, and D. Moradpour, unpublished data). In the course of these studies, we unexpectedly found that a fusion construct comprising the N-terminal 74 amino acids (aa) of NS4B was associated with membranes, while most prediction methods located the beginning of the first transmembrane segment around aa 74 (21, 28, 42). Here, we demonstrate that an amphipathic α-helix extending from aa 42 to 66 (α-helix 42-66) in the N-terminal portion of NS4B mediates this membrane association and plays an essential role in the formation of the HCV replication complex.  相似文献   

4.
5.
Persistent infection with hepatitis C virus (HCV) is a major cause of chronic liver diseases. The aim of this study was to identify host cell factor(s) participating in the HCV replication complex (RC) and to clarify the regulatory mechanisms of viral genome replication dependent on the host-derived factor(s) identified. By comparative proteome analysis of RC-rich membrane fractions and subsequent gene silencing mediated by RNA interference, we identified several candidates for RC components involved in HCV replication. We found that one of these candidates, creatine kinase B (CKB), a key ATP-generating enzyme that regulates ATP in subcellular compartments of nonmuscle cells, is important for efficient replication of the HCV genome and propagation of infectious virus. CKB interacts with HCV NS4A protein and forms a complex with NS3-4A, which possesses multiple enzyme activities. CKB upregulates both NS3-4A-mediated unwinding of RNA and DNA in vitro and replicase activity in permeabilized HCV replicating cells. Our results support a model in which recruitment of CKB to the HCV RC compartment, which has high and fluctuating energy demands, through its interaction with NS4A is important for efficient replication of the viral genome. The CKB-NS4A association is a potential target for the development of a new type of antiviral therapeutic strategy.Hepatitis C virus (HCV) infection represents a significant global healthcare burden, and current estimates suggest that a minimum of 3% of the world''s population is chronically infected (4, 19). The virus is responsible for many cases of severe chronic liver diseases, including cirrhosis and hepatocellular carcinoma (4, 16, 19). HCV is a positive-stranded RNA virus belonging to the family Flaviviridae. Its ∼9.6-kb genome is translated into a single polypeptide of about 3,000 amino acids (aa), in which the nonstructural (NS) proteins NS2, NS3, NS4A, NS4B, NS5A, and NS5B reside in the C-terminal half region (6, 34, 44). NS4A, a small 7-kDa protein, functions as a cofactor for NS3 to enhance NS3 enzyme activities such as serine protease and helicase activities. The hydrophobic N-terminal region of NS4A, which is predicted to form a transmembrane α-helix, is responsible for membrane anchorage of the NS3-4A complex (8, 44, 50), and the central region of NS4A is important for the interaction with NS3 (10, 44). A recent study demonstrated the involvement of the C terminus of NS4A in the regulation of NS5A hyperphosphorylation and viral replication (28).The development of HCV replicon technology several years ago accelerated research on viral RNA replication (7, 44). Furthermore, a robust cell culture system for propagation of infectious HCV particles was developed using a viral genome of HCV genotype 2a, JFH-1 strain, enabling us to study every process in the viral life cycle (27, 47, 54). RNA derived from genotype 1a, HCV H77, containing cell-culture adaptive mutations, also produces infectious viruses (52). Using these systems, it has been reported that the HCV genome replicates in a distinct, subcellular replication complex (RC) compartment, which includes NS3-5B and the viral RNA (2, 14, 33). The RC forms in a distinct compartment with high concentrations of viral and cellular components located on detergent-resistant membrane (DRM) structures, possibly a lipid-raft structure (2, 41), which may protect the RC from external proteases and nucleases. Almost all processes in viral replication are dependent on the host cell''s machinery and involve intimate interaction between viral and host proteins. However, the functional roles of host factors interacting with the HCV RC in viral genome replication remain ambiguous.To gain a better understanding of cellular factors that are components of the HCV RC and that function as regulators of viral replication, a comparative proteomic analysis of DRM fractions from HCV replicon and parental cells and subsequent RNA interference (RNAi) silencing of selected genes were performed. We identified creatine kinase B (CKB) as a key factor for the HCV genome replication. CKB catalyzes the reversible transfer of the phosphate group of phosphocreatine (pCr) to ADP to yield ATP and creatine and is known to play important roles in local delivery and cellular compartmentalization of ATP (48, 51). The findings obtained here suggest that recruitment of CKB to the HCV RC, through CKB interaction with NS4A, is essential for maintenance or enhancement of viral replicase activity.  相似文献   

6.
7.
For Bovine viral diarrhea virus (BVDV), the type species of the genus Pestivirus in the family Flaviviridae, cytopathogenic (cp) and noncytopathogenic (ncp) viruses are distinguished according to their effect on cultured cells. It has been established that cytopathogenicity of BVDV correlates with efficient production of viral nonstructural protein NS3 and with enhanced viral RNA synthesis. Here, we describe generation and characterization of a temperature-sensitive (ts) mutant of cp BVDV strain CP7, termed TS2.7. Infection of bovine cells with TS2.7 and the parent CP7 at 33°C resulted in efficient viral replication and a cytopathic effect. In contrast, the ability of TS2.7 to cause cytopathogenicity at 39.5°C was drastically reduced despite production of high titers of infectious virus. Further experiments, including nucleotide sequencing of the TS2.7 genome and reverse genetics, showed that a Y1338H substitution at residue 193 of NS2 resulted in the temperature-dependent attenuation of cytopathogenicity despite high levels of infectious virus production. Interestingly, TS2.7 and the reconstructed mutant CP7-Y1338H produced NS3 in addition to NS2-3 throughout infection. Compared to the parent CP7, NS2-3 processing was slightly decreased at both temperatures. Quantification of viral RNAs that were accumulated at 10 h postinfection demonstrated that attenuation of the cytopathogenicity of the ts mutants at 39.5°C correlated with reduced amounts of viral RNA, while the efficiency of viral RNA synthesis at 33°C was not affected. Taken together, the results of this study show that a mutation in BVDV NS2 attenuates viral RNA replication and suppresses viral cytopathogenicity at high temperature without altering NS3 expression and infectious virus production in a temperature-dependent manner.The pestiviruses Bovine viral diarrhea virus-1 (BVDV-1), BVDV-2, Classical swine fever virus (CSFV), and Border disease virus (BDV) are causative agents of economically important livestock diseases. Together with the genera Flavivirus, including several important human pathogens like Dengue fever virus, West Nile virus, Yellow fever virus, and Tick-borne encephalitis virus, and Hepacivirus (human Hepatitis C virus [HCV]), the genus Pestivirus constitutes the family Flaviviridae (8, 20). All members of this family are enveloped viruses with a single-stranded positive-sense RNA genome encompassing one large open reading frame (ORF) flanked by 5′ and 3′ nontranslated regions (NTR) (see references 8 and 28 for reviews). The ORF encodes a polyprotein which is co- and posttranslationally processed into the mature viral proteins by viral and cellular proteases. For BVDV, the RNA genome is about 12.3 kb in length and encodes a polyprotein of about 3,900 amino acids. The first third of the ORF encodes a nonstructural (NS) autoprotease and four structural proteins, while the remaining part of the genome encodes NS proteins which share many common characteristics and functions with the corresponding NS proteins encoded by the HCV genome (8, 28). NS2 of BVDV represents a cysteine autoprotease which is distantly related to the HCV NS2-3 protease (26). NS3, NS4A, NS4B, NS5A, and NS5B are essential components of the pestivirus replicase (7, 10, 49). NS3 possesses multiple enzymatic activities, namely serine protease (48, 52, 53), NTPase (46), and helicase activity (51). NS4A acts as an essential cofactor for the NS3 proteinase. NS5B represents the RNA-dependent RNA polymerase (RdRp) (22, 56). The functions of NS4B and NS5A remain to be determined. NS5A has been shown to be a phosphorylated protein that is associated with cellular serine/threonine kinases (44).According to their effects in tissue culture, two biotypes of pestiviruses are distinguished: cytopathogenic (cp) and noncytopathogenic (ncp) viruses (17, 27). The occurrence of cp BVDV in cattle persistently infected with ncp BVDV is directly linked to the induction of lethal mucosal disease in cattle (12, 13). Previous studies have shown that cp BVDV strains evolved from ncp BVDV strains by different kinds of mutations. These include RNA recombination with various cellular mRNAs, resulting in insertions of cellular protein-coding sequences into the viral genome, as well as insertions, duplications, and deletions of viral sequences, and point mutations (1, 2, 9, 24, 33, 36, 37, 42). A common consequence of all these genetic changes in cp BVDV genomes is the efficient production of NS3 at early and late phases of infection. In contrast, NS3 cannot be detected in cells at late time points after infection with ncp BVDV. An additional major difference is that the cp viruses produce amounts of viral RNA significantly larger than those of their ncp counterparts (7, 32, 50). While there is clear evidence that cell death induced by cp BVDV is mediated by apoptosis, the molecular mechanisms involved in pestiviral cytopathogenicity are poorly understood. In particular, the role of NS3 in triggering apoptosis remains unclear. It has been hypothesized that the NS3 serine proteinase might be involved in activation of the apoptotic proteolytic cascade (21, 55). Furthermore, it has been suggested that the NS3-mediated, enhanced viral RNA synthesis of cp BVDV and subsequently larger amounts of viral double-stranded RNAs may play a crucial role in triggering apoptosis (31, 54).In this study, we describe generation and characterization of a temperature-sensitive (ts) cp BVDV mutant whose ability to cause viral cytopathogenicity at high temperature is strongly attenuated. Our results demonstrate that a single amino acid substitution in NS2 attenuates BVDV cytopathogenicity at high temperature without affecting production of infectious viruses and expression of NS3 in a temperature-dependent manner.  相似文献   

8.
The hepatitis C virus NS2 protein has been recently implicated in virus particle assembly. To further understand the role of NS2 in this process, we conducted a reverse genetic analysis of NS2 in the context of a chimeric genotype 2a infectious cell culture system. Of 32 mutants tested, all were capable of RNA replication and 25 had moderate-to-severe defects in virus assembly. Through forward genetic selection for variants capable of virus spread, we identified second-site mutations in E1, E2, NS2, NS3, and NS4A that suppressed NS2 defects in assembly. Two suppressor mutations, E1 A78T and NS3 Q221L, were further characterized by additional genetic and biochemical experiments. Both mutations were shown to suppress other NS2 defects, often with mutual exclusivity. Thus, several NS2 mutants were enhanced by NS3 Q221L and inhibited by E1 A78T, while others were enhanced by E1 A78T and inhibited by NS3 Q221L. Furthermore, we show that the NS3 Q221L mutation lowers the affinity of native, full-length NS3-NS4A for functional RNA binding. These data reveal a complex network of interactions involving NS2 and other viral structural and nonstructural proteins during virus assembly.Hepatitis C virus (HCV) is a major cause of acute and chronic liver disease and contributes to the development of hepatocellular carcinoma. HCV is an enveloped, positive-strand RNA virus, the type member of the Hepacivirus genus in the family Flaviviridae (43). HCV exhibits high levels of sequence diversity that cluster into seven major genotypes and numerous subtypes (21).HCV genomes are 9.6 kb and encode a single long open reading frame of ∼3,011 codons (43). Translation of this genome produces a large polyprotein that is co- and posttranslationally processed by viral and host proteases into 10 distinct products. The N-terminal one-third of the polyprotein encodes the structural proteins, which are thought to compose the virus particle. These include an RNA-binding nucleocapsid protein, core (C), and two viral envelope glycoproteins, E1 and E2. E1 and E2 are type I membrane proteins that coordinately fold into a heterodimer complex (36). The remainder of the genome encodes the nonstructural (NS) proteins NS2, NS3, NS4A, NS4B, NS5A, and NS5B, which mediate the intracellular aspects of the viral life cycle. In addition, a small viroporin-like protein, p7, resides between the structural and NS genes.HCV encodes two proteases, the NS2-NS3 cysteine autoprotease and the NS3-NS4A serine protease. The only known substrate of the NS2-NS3 autoprotease is the NS2/3 junction. This enzyme is encoded by the C-terminal 121 amino acids (aa) of NS2, which forms a homodimer with twin composite active sites composed of two residues from one chain and one residue from the other (45). In addition, the serine protease domain of NS3 plays a noncatalytic role in stimulating NS2/3 cleavage (69). Upstream of the cysteine protease domain, the N-terminal hydrophobic region of NS2 mediates interaction with cellular membranes. While the membrane topology of NS2 is not yet fully known (67, 80), N-terminal cleavage by endoplasmic reticulum-resident signal peptidase and C-terminal cleavage by the cytosolic NS2-NS3 cysteine protease indicate that NS2 likely contains one or three transmembrane (TM) domains.The NS3-NS4A serine protease is encoded by the N-terminal domain of NS3 and is responsible for downstream cleavages at the NS3/4A, NS4A/B, NS4B/5A, and NS5A/B junctions. NS4A, a small (54-aa), membrane-anchored protein, acts as a cofactor for the serine protease activity by helping to complete the chymotrypsin-like fold of NS3 (14, 46). In addition to polyprotein processing, NS3-NS4A serine protease helps to dampen the innate antiviral response by cleaving cellular proteins involved in signal transduction (65).The C-terminal region of NS3 encodes an RNA helicase/NTPase activity that is essential for viral replication, although it is not yet clear which specific step(s) of the replication cycle requires this activity (29, 33). Interestingly, the NS3 serine protease and RNA helicase domains enhance each other''s activities, suggesting that proteolysis and RNA replication may be functionally coordinated (5, 6). In addition, NS4A helps to promote RNA-stimulated ATP hydrolysis by the NS3 helicase domain (4).In addition to their role in polyprotein processing, emerging evidence indicates that NS2 and NS3-NS4A participate in virus particle assembly (52). Prior work showed that NS2 is not essential for RNA replication of subgenomic replicons engineered to express NS3 through NS5B (44). The first evidence for an additional function of NS2 came from the construction of improved chimeric genotype 2a cDNA clones that replicated to high titers in cell culture (HCVcc). Pietschmann and colleagues showed that the Jc1 chimera containing a J6-JFH1 junction between the first and second putative TM domains of NS2 yielded higher-titer viruses than the original infectious J6/JFH chimera (41, 58). Furthermore, a number of adaptive mutations that improve virus production have been mapped to NS2 and NS3 (22, 23, 27, 53, 64, 68, 82). By using bicistronic constructs to express NS2 and NS3 independently of NS2/3 cleavage, two groups showed that full-length NS2, but not uncleaved NS2-NS3 or the NS2 cysteine protease active sites, was required for virus production (24, 25). Moreover, a limited number of mutations in NS2 were shown to inhibit virus assembly (24, 79, 83).Despite these observations, the role of NS2 in virus assembly remains unclear. We have therefore undertaken a genetic analysis to target conserved residues in NS2 for site-directed mutagenesis and identified a number of key residues that are important for virus assembly. Further analysis revealed that a network of genetic interactions among NS2, E1-E2, and NS3-NS4A helps to direct virus assembly. Finally, a suppressor mutation in NS3 was shown to influence functional RNA binding by the RNA helicase/ATPase.  相似文献   

9.
10.
The flavivirus genome comprises a single strand of positive-sense RNA, which is translated into a polyprotein and cleaved by a combination of viral and host proteases to yield functional proteins. One of these, nonstructural protein 3 (NS3), is an enzyme with both serine protease and NTPase/helicase activities. NS3 plays a central role in the flavivirus life cycle: the NS3 N-terminal serine protease together with its essential cofactor NS2B is involved in the processing of the polyprotein, whereas the NS3 C-terminal NTPase/helicase is responsible for ATP-dependent RNA strand separation during replication. An unresolved question remains regarding why NS3 appears to encode two apparently disconnected functionalities within one protein. Here we report the 2.75-Å-resolution crystal structure of full-length Murray Valley encephalitis virus NS3 fused with the protease activation peptide of NS2B. The biochemical characterization of this construct suggests that the protease has little influence on the helicase activity and vice versa. This finding is in agreement with the structural data, revealing a single protein with two essentially segregated globular domains. Comparison of the structure with that of dengue virus type 4 NS2B-NS3 reveals a relative orientation of the two domains that is radically different between the two structures. Our analysis suggests that the relative domain-domain orientation in NS3 is highly variable and dictated by a flexible interdomain linker. The possible implications of this conformational flexibility for the function of NS3 are discussed.Flaviviruses such as dengue virus (DENV), yellow fever virus (YFV), West Nile virus (WNV), and Japanese encephalitis virus (JEV) belong to the family Flaviviridae and are the causative agents of a range of serious human diseases including hemorrhagic fever, meningitis, and encephalitis (37). They remain a global health priority, as many viruses are endemic in large parts of the Americas, Africa, Australia, and Asia, and vaccines remain unavailable for most members (31, 46, 57).Flaviviruses have a positive-sense single-stranded RNA (ssRNA) genome (approximately 11 kb) that encodes one large open reading frame containing a 5′ type 1 cap and conserved RNA structures at both the 5′ and 3′ untranslated regions that are important for viral genome translation and replication. The genomic RNA is translated into a single polyprotein precursor (11) consisting of three structural (C [capsid], prM [membrane], and E [envelope]) and seven nonstructural (NS1, NS2a, NS2b, NS3, NS4a, NS4b, and NS5) proteins arranged in the order C-prM-E-NS1-NS2a-NS2b-NS3-NS4a-NS4b-NS5 (reviewed in reference 33) (Fig. (Fig.1).1). Only the structural proteins become part of the mature, infectious virion, whereas the nonstructural proteins are involved in polyprotein processing, viral RNA synthesis, and virus morphogenesis (33, 43). The precursor protein is directed by signal sequences into the host endoplasmic reticulum (ER), where NS1 and the exogenous domains of prM and E face the lumen, while C, NS3, and NS5 are cytoplasmic. NS2A, NS2B, NS4A, and NS4B are largely hydrophobic transmembrane proteins with small hydrophilic segments (Fig. (Fig.1).1). The post- and cotranslational cleavage of the polyprotein is performed by NS3 in the cytoplasm and by host proteases in the ER lumen to yield the mature proteins (Fig. (Fig.1)1) (33, 43). Of the nonstructural proteins, NS3 and NS5 are the best characterized, and both are essential for viral replication (23, 27, 41). Both proteins are multifunctional. The N-terminal one-third of NS3 contains the viral protease (NS3pro), which requires a portion of NS2B for its activity, while the remaining portion codes for the RNA helicase/NTPase/RTPase domain (NS3hel) (21, 22, 32, 55). NS5, however, contains both an N-terminal methyltransferase and a C-terminal RNA-dependent RNA polymerase (16, 51). The functions of NS1, NS2A, NS4A, and NS4B are not well understood, but they appear to play important roles in replication and virus assembly/maturation and have been found to bind to NS3 and NS5, possibly modulating their activity (33, 36).Open in a separate windowFIG. 1.Schematic diagram of flavivirus polyprotein organization and processing. (Top) Linear organization of the structural and nonstructural proteins within the polyprotein. (Middle) Putative membrane topology of the polyprotein predicted from biochemical and cellular analyses, which is then processed by cellular and viral proteases (indicated by arrows). (Bottom) Different complexes that are thought to arise in different cellular compartments during and following polyprotein processing.Because of its enzymatic activities and its critical role in viral replication and polyprotein processing, NS3 constitutes a promising drug target for antiviral therapy (31). NS3pro (residues 1 to 169) is a trypsin-like serine protease with the characteristic catalytic triad (Asp-His-Ser) and a highly specific substrate recognition sequence, conserved in all flaviviruses, consisting of two basic residues in P2 and P1 followed by a small unbranched amino acid in P1′ (11). NS3pro has an aberrant fold compared to the canonical trypsin structure, and its folding and protease activity are dependent on a noncovalent association with a central 47-amino-acid hydrophilic domain of NS2B (19, 21). The remainder of NS2B contains three transmembrane helices involved in membrane associations. NS3 mediates cleavages at the C-terminal side of the highly conserved dibasic residue located at the coding junctions NS2A/NS2B, NS2B/NS3, NS3/NS4A, and NS4B/NS5 and also between the C terminus of C and NS4A (11, 33) (Fig. (Fig.11).The C-terminal portion of NS3 (NS3hel, residues 170 to 619) performs several catalytically related activities, namely, RNA strand separation and (poly)nucleotide hydrolysis (5, 22, 32, 55) at a common, RecA-like NTPase catalytic center that couples the energy released from the hydrolysis of the triphosphate moieties of nucleotides to RNA unwinding. Although the precise role of NS3 in replication has not been established, its helicase activity is thought to separate nascent RNA strands from the template strands and to assist replication initiation by unwinding RNA secondary structure in the 3′ untranslated region (11, 13, 15, 33). NS3 is a member of the DEAH/D box family within helicase superfamily 2 (SF2) and is characterized by seven conserved sequence motifs involved in nucleic acid binding and hydrolysis (45). In addition, its RNA triphosphatase activity is thought to be involved in the capping of the viral RNA. In the process of replication, NS3 interacts, most likely via its C-terminal domain, with NS5 (13, 15, 24, 26, 58, 62). The NS3 5′ triphosphatase and NS5 methyltransferase activities probably cooperate in cap formation by removing the terminal γ-phosphate and performing sequential N7 and 2′ O methylations, respectively (16, 28, 46, 56). The guanylyltransferase activity required for cap formation remains elusive at present, although recent evidence suggests that it may be present in NS5 (8, 17). In addition, the interaction between NS3 and NS5 can stimulate NS3 helicase/NTPase activity (15, 62).The atomic structures of NS3pro in the presence and absence of ligands and/or the NS2B activating domain (2, 19, 47) and NS3hel (35, 38, 39, 49, 58-60) are known, and recently, the structure of full-length DENV4 (one of four dengue virus serotypes) NS3 fused to an 18-residue NS2B cofactor (NS2B18NS3) was reported (34). This structure revealed an elongated conformation, with the protease domain interfacing with the NTP binding pocket and being separated from NS3hel by a relatively flexible linker, which suggested that the protease domain may have a positive effect on the activity of the NTPase/helicase domain. However, other reports suggested that NS3pro has no or a very limited effect on the activity of NS3hel (32, 62). In addition, since current evidence suggests that NS2B is not part of the replication complex (Fig. (Fig.1)1) (36), and it is known that in the absence of the NS2B cofactor, NS3pro is unfolded and inactive, it becomes hard to envisage what effect the NS3 protease domain may have on the helicase domain in a biologically relevant context. Equally, it is still not clear what role the helicase domain plays during polyprotein processing by NS3pro and, in general, why these two apparently distinct and unrelated catalytic activities are harbored within a single polypeptide.In order to gain further insights into these questions, we report the biochemical analysis and crystallographic structure at a 2.75-Å resolution of full-length NS3 from Murray Valley encephalitis virus (MVEV), a member of the JEV group of flaviviruses, fused to the entire protease activation peptide of the NS2B cofactor (NS2B45NS3). The structure reveals the protease and helicase domains to be structurally independent and differs dramatically from the structure observed for DENV4 NS2B18NS3. We discuss the implications of this unexpectedly different configuration of the NS3 protein and argue that the structural flexibility observed is likely to be crucial for its multifunctional nature.  相似文献   

11.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

12.
Flavivirus NS1 is a versatile nonstructural glycoprotein, with intracellular NS1 functioning as an essential cofactor for viral replication and cell surface and secreted NS1 antagonizing complement activation. Even though NS1 has multiple functions that contribute to virulence, the genetic determinants that regulate the spatial distribution of NS1 in cells among different flaviviruses remain uncharacterized. Here, by creating a panel of West Nile virus-dengue virus (WNV-DENV) NS1 chimeras and site-specific mutants, we identified a novel, short peptide motif immediately C-terminal to the signal sequence cleavage position that regulates its transit time through the endoplasmic reticulum and differentially directs NS1 for secretion or plasma membrane expression. Exchange of two amino acids within this motif reciprocally changed the cellular targeting pattern of DENV or WNV NS1. For WNV, this substitution also modulated infectivity and antibody-induced phagocytosis of infected cells. Analysis of a mutant lacking all three conserved N-linked glycosylation sites revealed an independent requirement of N-linked glycans for secretion but not for plasma membrane expression of WNV NS1. Collectively, our experiments define the requirements for cellular targeting of NS1, with implications for the protective host responses, immune antagonism, and association with the host cell sorting machinery. These studies also suggest a link between the effects of NS1 on viral replication and the levels of secreted or cell surface NS1.West Nile virus (WNV) is a single-stranded, positive-sense enveloped RNA Flavivirus that cycles in nature between birds and Culex mosquitoes. It is endemic in parts of Africa, Europe, the Middle East, and Asia, and outbreaks occur annually in North America. More than 29,000 human cases of severe WNV infection have been diagnosed in the United States since its entry in 1999, and millions have been infected and remain undiagnosed (9). Humans can develop a febrile illness that progresses to a flaccid paralysis, meningitis, or encephalitis syndrome (59). Dengue virus (DENV) is a genetically related flavivirus that is transmitted by Aedes aegypti and Aedes albopictus mosquitoes and causes clinical syndromes in humans, ranging from an acute self-limited febrile illness (dengue fever [DF]) to a severe and life-threatening vascular leakage and bleeding diathesis (dengue hemorrhagic fever/dengue shock syndrome [DHF/DSS]). Globally, DENV causes an estimated 50 million infections annually, resulting in 500,000 hospitalizations and ∼22,000 deaths (45).The ∼10.7-kb Flavivirus RNA genome is translated as a single polyprotein, which is then cleaved into three structural proteins (C, prM/M, and E) and seven nonstructural (NS) proteins (NS1, NS2A, NS2B, NS3, NS4A, NS4B, and NS5) by virus- and host-encoded proteases (39). The multifunctional NS proteins include an RNA-dependent RNA polymerase and methyltransferase (NS5), a helicase and protease (NS3), accessory proteins that form part of the viral replication complex, and immune evasion molecules (33, 34). Flavivirus NS1 is a 48-kDa nonstructural glycoprotein with two or three N-linked glycans, depending on the flavivirus, and is absent from the virion. The Japanese encephalitis virus (JEV) serogroup (West Nile, Japanese, Murray Valley, and St. Louis encephalitis viruses) generate NS1 and NS1′ proteins, the latter of which is a product of a ribosomal frameshift event that occurs at a heptanucleotide motif located at the beginning of the NS2A gene (25, 47).NS1 is an essential gene as it is required for efficient viral RNA replication (34, 41, 44). In infected mammalian cells, NS1 is synthesized as a soluble monomer, dimerizes after posttranslational modification in the lumen of the endoplasmic reticulum (ER), and accumulates extracellularly as higher-order oligomers, including hexamers (16, 26, 64, 65). Soluble NS1 binds back to the plasma membrane of uninfected cells through interactions with sulfated glycosaminoglycans (5). In infected cells, NS1 is also directly transported to and expressed on the plasma membrane although it lacks a transmembrane domain or canonical targeting motif. The mechanism of cell surface expression of flavivirus NS1 in infected cells remains uncertain although some fraction may be linked through an atypical glycosyl-phosphatidylinositol anchor (30, 50) or lipid rafts (49).NS1 has been implicated in having pathogenic consequences in flavivirus infection. The high levels of NS1 in the serum of DENV-infected patients correlate with severe disease (4, 37). NS1 has been proposed to facilitate immune complex formation (4), elicit auto-antibodies that react with host matrix proteins (21), damage endothelial cells via antibody-dependent complement-mediated cytolysis (38), or directly enhance infection (1). Flavivirus NS1 also has direct immune evasion functions and antagonizes complement activation on cell surfaces and in solution. WNV NS1 attenuates the alternative pathway of complement activation by binding the complement-regulatory protein factor H (11, 36), and DENV, WNV, and YFV NS1 proteins bind C1s and C4 in a complex to promote efficient degradation of C4 to C4b (3).Although NS1 is absent from the virion, antibodies against it can protect against infection in vivo. Immunization with purified NS1 or passive administration of some anti-WNV, anti-yellow fever virus (YFV), and anti-DENV NS1 monoclonal antibodies (MAbs) protect mice against lethal virus challenge (12, 13, 17, 22, 27, 29, 31, 32, 56-58). Initial studies with isotype switch variants and F(ab′)2 fragments of anti-YFV NS1 MAbs suggested that the Fc region of anti-NS1 MAbs was required for protection (58). Subsequent mechanistic studies with WNV NS1 indicated that only MAbs recognizing cell surface-associated NS1 trigger Fc-γ receptor I- and/or IV-mediated phagocytosis and clearance of infected cells (13).In this study, we identify a reciprocal relationship between the secretion and cell surface expression patterns of WNV and DENV NS1s. Using WNV-DENV NS1 chimeras and point mutants, we identified a novel short peptide motif immediately C-terminal to the signal sequence cleavage position that directs NS1 for secretion or to the plasma membrane. These studies begin to explain how NS1 regulates its localization to several cellular compartments (ER, cell surface, and extracellular space) and have implications for viral infectivity, association with the host cell sorting machinery, and protective immune responses.  相似文献   

13.
14.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

15.
The first morphological evidence of African swine fever virus (ASFV) assembly is the appearance of precursor viral membranes, thought to derive from the endoplasmic reticulum, within the assembly sites. We have shown previously that protein p54, a viral structural integral membrane protein, is essential for the generation of the viral precursor membranes. In this report, we study the role of protein p17, an abundant transmembrane protein localized at the viral internal envelope, in these processes. Using an inducible virus for this protein, we show that p17 is essential for virus viability and that its repression blocks the proteolytic processing of polyproteins pp220 and pp62. Electron microscopy analyses demonstrate that when the infection occurs under restrictive conditions, viral morphogenesis is blocked at an early stage, immediately posterior to the formation of the viral precursor membranes, indicating that protein p17 is required to allow their progression toward icosahedral particles. Thus, the absence of this protein leads to an accumulation of these precursors and to the delocalization of the major components of the capsid and core shell domains. The study of ultrathin serial sections from cells infected with BA71V or the inducible virus under permissive conditions revealed the presence of large helicoidal structures from which immature particles are produced, suggesting that these helicoidal structures represent a previously undetected viral intermediate.African swine fever virus (ASFV) (61, 72) is the only known DNA-containing arbovirus and the sole member of the Asfarviridae family (24). Infection by this virus of its natural hosts, the wild swine warthogs and bushpigs and the argasid ticks of the genus Ornithodoros, results in a mild disease, often asymptomatic, with low viremia titers, that in many cases develops into a persistent infection (3, 43, 71). In contrast, infection of domestic pigs leads to a lethal hemorrhagic fever for which the only available methods of disease control are the quarantine of the affected area and the elimination of the infected animals (51).The ASFV genome is a lineal molecule of double-stranded DNA of 170 to 190 kbp in length with convalently closed ends and terminal inverted repeats. The genome encodes more than 150 open reading frames, half of which lack any known or predictable function (16, 75).The virus particle, with an overall icosahedral shape and an average diameter of 200 nm (11), is organized in several concentric layers (6, 11, 15) containing more than 50 structural proteins (29). Intracellular particles are formed by an inner viral core, which contains the central nucleoid surrounded by a thick protein coat, referred to as core shell. This core is enwrapped by an inner lipid envelope (7, 34) on top of which the icosahedral capsid is assembled (26, 27, 31). Extracellular virions possess an additional membrane acquired during the budding from the plasma membrane (11). Both forms of the virus, intracellular and extracellular, are infective (8).The assembly of ASFV particles occurs in the cytoplasm of the infected cell, in viral factories located close to the cell nucleus (6, 13, 49). ASFV factories possess several characteristics similar to those of the cellular aggresomes (35), which are accumulations of aggregates of cellular proteins that form perinuclear inclusions (44).Current models propose that ASFV assembly begins with the modification of endoplasmic reticulum (ER) membranes, which are subsequently recruited to the viral factories and transformed into viral precursor membranes. These ER-derived viral membranes represent the precursors of the inner viral envelope and are the first morphological evidence of viral assembly (7, 60). ASFV viral membrane precursors evolve into icosahedral intermediates and icosahedral particles by the progressive assembly of the outer capsid layer at the convex face of the precursor membranes (5, 26, 27, 31) through an ATP- and calcium-dependent process (19). At the same time, the core shell is formed underneath the concave face of the viral envelope, and the viral DNA and nucleoproteins are packaged and condensed to form the innermost electron-dense nucleoid (6, 9, 12, 69). However, the assembly of the capsid and the internal envelope appears to be largely independent of the components of the core of the particle, since the absence of the viral polyprotein pp220 during assembly produces empty virus-like particles that do not contain the core (9).Comparative genome analysis suggests that ASFV shares a common origin with the members of the proposed nucleocytoplasmic large DNA viruses (NCLDVs) (40, 41). The reconstructed phylogeny of NCLDVs as well as the similitude in the structures and organizations of the genomes indicates that ASFV is more closely related to poxviruses than to other members of the NCLDVs. A consensus about the origin and nature of the envelope of the immature form of vaccinia virus (VV), the prototypical poxvirus, seems to be emerging (10, 17, 20, 54). VV assembly starts with the appearance of crescent-shaped structures within specialized regions of the cytoplasm also known as viral factories (21, 23). The crescent membranes originate from preexisting membranes derived from some specialized compartment of the ER (32, 37, 52, 53, 67), and an operative pathway from the ER to the crescent membrane has recently been described (38, 39). VV crescents apparently grow in length while maintaining the same curvature until they become closed circles, spheres in three dimensions, called immature virions (IV) (22). The uniform curvature is produced by a honeycomb lattice of protein D13L (36, 70), which attaches rapidly to the membranes so that nascent viral membranes always appear to be coated over their entirety. The D13L protein is evolutionarily related to the capsid proteins of the other members of the NCLDV group, including ASFV, but lacks the C-terminal jelly roll motif (40). This structural difference is probably related to the fact that poxviruses are the only member of this group without an icosahedral capsid; instead, the spherical D13L coat acts as a scaffold during the IV stage but is discarded in subsequent steps of morphogenesis (10, 28, 46, 66). Thus, although crescents in VV and precursors of the inner envelope in ASFV are the first morphogenetic stages discernible in the viral factories of these viruses, they seem to be different in nature. Crescents are covered by the D13L protein and are more akin to the icosahedral intermediates of ASFV assembly, whereas ASFV viral membrane precursors are more similar to the naked membranes seen when VV morphogenesis is arrested by rifampin treatment (33, 47, 48, 50) or when the expression of the D13L and A17L proteins are repressed during infection with lethal conditional VV viruses (45, 55, 56, 68, 74, 76).Although available evidence strongly supports the reticular origin of the ASFV inner envelope (7, 60), the mechanism of acquisition remains unknown, and the number of membranes present in the inner envelope is controversial. The traditional view of the inner envelope as formed by two tightly opposed membranes derived from ER collapsed cisternae (7, 59, 60) has recently been challenged by the careful examination of the width of the internal membrane of viral particles and the single outer mitochondrial membrane, carried out using chemical fixation, cryosectioning, and high-pressure freezing (34). The results suggest that the inner envelope of ASFV is a single lipid bilayer, which raises the question of how such a structure can be generated and stabilized in the precursors of the ASFV internal envelope. In the case of VV, the coat of the D13L protein has been suggested to play a key role in the stabilization of the single membrane structure of the crescent (10, 17, 36), but the ASFV capsid protein p72 is not a component of the viral membrane precursors. The identification and functional characterization of the proteins involved in the generation of these structures are essential for the understanding of the mechanisms involved in these early stages of viral assembly. For this reason, we are focusing our interest on the study of abundant structural membrane proteins that reside at the inner envelope of the viral particle. We have shown previously that one of these proteins, p54, is essential for the recruitment of ER membranes to the viral factory (59). Repression of protein p54 expression has a profound impact on virus production and leads to an early arrest in virion morphogenesis, resulting in the virtual absence of membranes in the viral factory.Protein p17, encoded by the late gene D117L in the BA71V strain, is an abundant structural protein (60, 65). Its sequence, which is highly conserved among ASFV isolates (16), does not show any significant similarity with the sequences present in the databases. Protein p17 is an integral membrane protein (18) that is predicted to insert in membranes with a Singer type I topology and has been localized in the envelope precursors as well as in both intracellular and extracellular mature particles (60), suggesting that it resides at the internal envelope, the only membranous structure of the intracellular particles.In this work, we analyze the role of protein p17 in viral assembly by means of an IPTG (isopropyl-β-d-thiogalactopyranoside)-dependent lethal conditional virus. The data presented indicate that protein p17 is essential for viral morphogenesis. The repression of this protein appears to block assembly at the level of viral precursor membranes, resulting in their accumulation at the viral factory.From the electron microscopy analysis of serial sections of viral factories at very early times during morphogenesis, we present experimental evidence that suggests that, during assembly, viral precursor membranes and core material organize into large helicoidal intermediates from which icosahedral particles emerge. The possible role of these structures during ASFV morphogenesis is discussed.  相似文献   

16.
West Nile virus capsid protein (WNVCp) displays pathogenic toxicity via the apoptotic pathway. However, a cellular mechanism protective against this toxic effect has not been observed so far. Here, we identified Makorin ring finger protein 1 (MKRN1) as a novel E3 ubiquitin ligase for WNVCp. The cytotoxic effects of WNVCp as well as its expression levels were inhibited in U2OS cells that stably expressed MKRN1. Immunoprecipitation analyses revealed an interaction between MKRN1 and WNVCp. Domain analysis indicated that the C terminus of MKRN1 and the N terminus of WNVCp were required for the interaction. MKRN1 could induce WNVCp ubiquitination and degradation in a proteasome-dependent manner. Interestingly, the WNVCp mutant with amino acids 1 to 105 deleted WNVCp was degraded by MKRN1, whereas the mutant with amino acids 1 to 90 deleted was not. When three lysine sites at positions 101, 103, and 104 of WNVCp were replaced with alanine, MKRN1-mediated ubiquitination and degradation of the mutant were significantly inhibited, suggesting that these sites are required for the ubiquitination. Finally, U2OS cell lines stably expressing MKRN1 were resistant to cytotoxic effects of WNV. In contrast, cells depleted of MKRN1 were more susceptible to WNVCp cytotoxicity. Confirming this, overexpression of MKRN1 significantly reduced, but depletion of MKRN1 increased, WNV proliferation in 293T cells. Taken together, our results suggest that MKRN1 can protect cells from WNV by inducing WNVCp degradation.West Nile virus (WNV) is an arthropod-borne virus that is a member of the Flaviviridae family, which includes St. Louis encephalitis virus, Kunjin virus, yellow fever virus, dengue virus, and Murray Valley encephalitis virus (2). Since its first identification in the West Nile province of Uganda in 1937, WNV has spread quickly through Asia, Europe, and the United States and has caused a serious global health problem (34). The clinical manifestations of WNV usually entail neurological diseases such as meningitis and encephalitis. This might be caused by WNV genome replication after inoculation and its subsequent spread to lymph nodes and blood, followed by its entrance into the central nervous system through Toll-like receptor and tumor necrosis factor receptor (40).WNV has the genome of a single positive-sense RNA containing one open reading frame. The encoded polypeptide is processed further by viral and cellular proteases into several nonstructural and structural proteins (2). Nonstructural (NS) proteins include NS1, NS2A, NS2B, NS3, NS4A, NS4B, and NS5. NS1 is involved in synthesis of viral RNA, and NS3 mediates the cleavage of nonstructural proteins (22, 25, 30, 48). NS5 functions as an RNA polymerase and methyltransferase, which are required for viral replication (14, 17, 18). NS2A, NS2B, NS4A, and NS4B promote the organization of viral replication factors and membrane permeabilization (3, 5, 6, 13, 37). The capsid, envelope (E), and premembrane (prM) proteins are the structural proteins, which are involved in virus assembly (43). E protein is a virion surface protein that regulates binding and fusion to the cell membrane (1, 11, 32). The prM protein is a precursor of the M protein, which is translocated to the endoplasmic reticulum (ER) by capsid (2, 21). Viral assembly occurs mainly in the ER membrane following release of viral particles (23).The capsid of WNV (WNVCp) localizes and is involved in nucleocapsid assembly on the ER membrane (15). However, extra roles of the flavivirus capsid in the nucleus has been reported. For example, capsid proteins of Japanese encephalitis virus (JEV) and hepatitis C virus (HCV), which are also members of the Flaviviridae family, participate in pathogenesis by localizing to the nucleus (33). Nucleolar and nuclear WNVCp is involved in pathogenesis via induction of the apoptotic process in cells through interaction with Hdm2, which results in the activation of the potent tumor suppressor p53 (47). It also induces apoptotic death of neuron cells via mitochondrial dysfunction and activation of caspase pathways when introduced into the brains of mice (46).The Makorin ring finger protein 1 (MKRN1) gene was first reported as the source gene of introns for the intronless imprinted MKRN gene family (10). The protein is an ancient protein conserved from invertebrates to vertebrates, and it contains several zinc finger motifs, including C3H, C3HC4, and unique Cys-His motifs (10). Furthermore, this gene is constitutively expressed in most human tissues, including neurons (10). The role of MKRN1 as an E3 ligase was first identified by its ability to degrade hTERT (16). Interestingly, MKRN1 functions as a coregulator of androgen and retinoic acid receptor (27), suggesting possible diverse roles of MKRN1 in human cells.In this study, we report on an ubiquitin (Ub) E3-ligase for WNVCp. MKRN1 was able to ubiquitinate and degrade WNVCp in a proteasome-dependent manner. Furthermore, degradation of WNVCp resulted in a reduction of WNV-induced cell death. Cells stably overexpressing MKRN1 were resistant to WNV-induced cell death. In contrast, ablation of MKRN1 by small interfering RNA (siRNA) renders cells more susceptible to the cytotoxicity of WNVCp. Furthermore, WNV proliferation was suppressed in 293T cells overexpressing MKRN1 but increased in MKRN1-depleted 293T cells. Based on these data, we suggest that MKRN1 might play a role in protection of cells against WNV infection.  相似文献   

17.
Current agricultural practices involve inclusion of antimicrobials in animal feed and result in manure containing antimicrobials and antimicrobial-resistant microorganisms. This work evaluated the effects of land application of swine manure on the levels of tetracycline, macrolide, and lincosamide antimicrobials and on macrolide, lincosamide, and streptogramin B (MLSB) resistance in field soil samples and laboratory soil batch tests. MLSB and tetracycline antimicrobials were quantified after solid-phase extraction using liquid chromatography-tandem mass spectrometry. The prevalence of the ribosomal modification responsible for MLSB resistance in the same samples was quantified using fluorescence in situ hybridization. Macrolide antimicrobials were not detected in soil samples, while tetracyclines were detected, suggesting that the latter compounds persist in soil. No significant differences in ribosomal methylation or presumed MLSB resistance were observed when amended and unamended field soils were compared, although a transient (<20-day) increase was observed in most batch tests. Clostridium cluster XIVa accounted for the largest fraction of resistant bacteria identified in amended soils. Overall, this study did not detect a persistent increase in the prevalence of MLSB resistance due to land application of treated swine manure.Treated swine manure contains substantial levels of both antimicrobial-resistant microorganisms (10, 26) and antimicrobials (7, 18, 33). Land application of manure could therefore contribute to public health risks associated with the increasing prevalence of antimicrobial resistance in pathogens both directly, through the dissemination of antimicrobial-resistant pathogens, and indirectly, through the introduction of and selection for antimicrobial resistance genes. Because limited data are available, this connection is largely a theoretical connection, particularly for the indirect effects. However, a recent retrospective study of antimicrobial resistance in soil did support the hypothesis that there is an environmental connection by documenting that there was an increase in the abundance of antibiotic resistance genes in samples collected from 1940 to 2008, during which time antimicrobial production increased dramatically (12).The fate of antimicrobials in amended soils is a function of their sorptive properties, the soil characteristics, and the potential for abiotic and biotic degradation of the antimicrobials. Tetracyclines tend to adsorb to soil (21, 23), which leads to persistence in amended soils (3, 7, 11), although they are also susceptible to degradation (3, 4). The macrolide tylosin frequently is not detected (3, 4, 7, 11, 33) and is likely rapidly degraded in manure and soils (8, 16, 24). However, persistence of tylosin for several months in amended soil has also been reported (6). The differences in degradation rates may be caused by differences in soil characteristics, manure-to-soil ratios, and/or microbial communities (15, 16, 21).Addition of both antimicrobials and antimicrobial-resistant microorganisms might be expected to result in an increase in the levels of resistance. However, most studies have not shown that there is a long-term increase in antimicrobial resistance due to land application of manure at agronomically prescribed rates (5, 9, 26). Transient (i.e., <45-day) increases have been reported (9, 26), as have elevated levels of resistance at sites near manure piles (5). In contrast, another report showed that there were significantly higher levels of tylosin resistance in soils that received animal manure from operations that used subtherapeutic levels of antimicrobials than in soils at sites where there was no use of subtherapeutic levels of antimicrobials (19). One limitation of these studies was their use of culture-based methods to quantify resistance; the results may not be representative of the entire microbial community. The molecular methods that have been used to quantify resistance also have limitations, and the most serious limitation is the inability of these methods to examine the full diversity of known and unknown resistance genes. The previous molecular studies of the impact of land application on resistance were largely restricted to qualitative analyses (10, 25), although quantitative PCR methods for analysis of tetracycline resistance genes have recently been used for cattle and swine lagoons (14, 20). In a retrospective soil study, Knapp et al. (12), who also used quantitative PCR, found multiple site differences, which made it difficult to evaluate the impact of manure application. However, the site with the highest manure application rate did not show the highest levels of antimicrobial resistance, suggesting that there are other factors that have a greater influence on the prevalence of resistance.In the present study, a variation of the fluorescence in situ hybridization (FISH) technique was used to assess the impact of land application of swine manure on the levels of macrolide-lincosamide-streptogramin B (MLSB) resistance. Although the MLSB antimicrobials are chemically distinct, methylation or mutation of a single base of the 23S rRNA prevents binding and results in cross-resistance to all three classes (29). The prevalence of MLSB antimicrobial resistance in the microbial community can therefore be quantified indirectly by hybridization of an oligonucleotide probe to unmethylated, MLSB-sensitive ribosomes, using either membrane hybridization (1, 10) or FISH (31). These methods do not require culturing or a comprehensive knowledge of the diversity of resistance gene sequences, but they do not detect resistance to specific antimicrobials that results from other mechanisms, such as macrolide efflux.This study focused on evaluating the impact of land application of swine manure on the levels of antimicrobials and the prevalence of antimicrobial resistance in the soil environment. The concentrations of tetracycline, macrolide, and lincosamide antimicrobials and the prevalence of MLSB resistance were compared for field soils that received no manure, swine manure from farms that did not use antimicrobials (referred to below as organic farms), and swine manure from conventional farms to determine whether land application affects the levels of antimicrobials and MLSB resistance. The effects of addition of manure, antimicrobials (lincomycin and chlortetracycline), and MLSB-resistant microorganisms on the prevalence of MLSB resistance were also compared using soil batch tests.  相似文献   

18.
19.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

20.
Hepatitis C virus (HCV) is an important human pathogen affecting 170 million chronically infected individuals. In search for cellular proteins involved in HCV replication, we have developed a purification strategy for viral replication complexes and identified annexin A2 (ANXA2) as an associated host factor. ANXA2 colocalized with viral nonstructural proteins in cells harboring genotype 1 or 2 replicons as well as in infected cells. In contrast, we found no obvious colocalization of ANXA2 with replication sites of other positive-strand RNA viruses. The silencing of ANXA2 expression showed no effect on viral RNA replication but resulted in a significant reduction of extra- and intracellular virus titers. Therefore, it seems likely that ANXA2 plays a role in HCV assembly rather than in genome replication or virion release. Colocalization studies with individually expressed HCV nonstructural proteins indicated that NS5A specifically recruits ANXA2, probably by an indirect mechanism. By the deletion of individual NS5A subdomains, we identified domain III (DIII) as being responsible for ANXA2 recruitment. These data identify ANXA2 as a novel host factor contributing, with NS5A, to the formation of infectious HCV particles.Hepatitis C virus (HCV) infections are characterized by a mostly unapparent acute phase leading to persistence in ca. 70% of all infected individuals. Currently, 170 million people suffer from chronic hepatitis C, and they have a high risk to develop severe liver disease. It has been estimated that HCV accounts for 27% of cirrhosis and 25% of hepatocellular carcinoma cases worldwide (2).HCV is an enveloped positive-strand RNA virus belonging to the genus Hepacivirus in the family Flaviviridae. The genome of HCV encompasses a single ∼9,600-nucleotide (nt)-long RNA molecule containing one large open reading frame (ORF) that is flanked by nontranslated regions (NTRs), which are important for viral translation and replication. HCV proteins generated from the polyprotein precursor are cleaved by cellular and viral proteases into at least 10 different products (for a review of polyprotein cleavage and the function of the individual proteins, see reference 4). The structural proteins Core, E1, and E2 are located in the amino-terminal portion of the polyprotein, followed by p7, a hydrophobic peptide that is supposed to be a viroporin, and the nonstructural proteins (NS) NS2, NS3, NS4A, NS4B, NS5A, and NS5B. Only the nonstructural proteins NS3 to NS5B are involved in viral RNA replication. NS3 is a multifunctional protein, consisting of an amino-terminal protease domain required for the processing of the NS3 to NS5B region and a carboxyterminal helicase/nucleoside triphosphatase domain. NS4A is a cofactor that activates the NS3 protease function by forming a heterodimer. The hydrophobic protein NS4B induces vesicular membrane alterations involved in RNA replication. NS5A is a phosphoprotein that seems to play an important role in viral replication and assembly (3, 35, 58). NS5B is the RNA-dependent RNA polymerase of HCV.Positive-strand RNA viruses replicate their RNA in vesicular structures originating from different cellular organelles (36). In the case of HCV, particular membrane alterations have been identified by electron microscopy, designated the membranous web, consisting of accumulations of vesicles primarily derived from the endoplasmic reticulum (17). Important insights into the organization of HCV replication complexes were obtained by the in vitro analysis of viral RNA synthesis in membrane preparations of cells harboring subgenomic HCV replicons, so-called crude replication complexes (CRCs) (1, 20). A current model based on a stoichiometric analysis of CRCs suggests that each vesicular structure contains multiple copies of viral nonstructural proteins and has a connection to the cytoplasm, allowing the constant supply of nucleotides for RNA synthesis (45), presumably analogously to the replication complex of the closely related dengue virus (DV) (64). Viral RNA synthesis in CRCs is highly resistant to proteinases and nucleases (39), and the membranes are detergent resistant at 4°C, resembling features of lipid rafts (54).Several purification techniques have been established to identify relevant HCV host factors by proteomics, based on either the extraction of detergent-resistant membranes (19, 34) or the immunoprecipitation of vesicles (24), revealing different sets of cellular proteins potentially involved in viral replication. In most of these studies, cell lines harboring persistent subgenomic replicons were utilized (33); however, with the availability of a fully permissive cell culture system supporting the complete HCV replication cycle (31, 63, 66), it became evident that viral RNA replication and assembly are closely linked. Recent work revealed an intimate connection of viral replication complexes and assembly sites in close proximity to cytoplasmic lipid droplets (38), with Core and especially NS5A functioning as central regulators by a poorly defined mechanism. NS5A is phosphorylated at multiple serine and threonine residues, binds RNA, and is composed of three domains, which are separated by trypsin-sensitive low-complexity regions (LCS I and II) (59). An N-terminal amphipathic alpha helix tightly associates NS5A with intracellular membranes. Domain I and LCS1 most likely are involved in viral RNA replication, since replication-enhancing mutations primarily mapped to this region (8, 32). The role of domain II is unknown, while domain III recently has been shown to be dispensable for RNA replication but essential for viral particle assembly (3, 35, 58). One of the proposed mechanisms points to a critical interaction with the Core protein, for which phosphorylation in the C-terminal part of domain III of NS5A appears to be required (35). The interaction of Core and NS5A has been proposed to be important for the recruitment of the replication complexes to lipid droplets (3), thereby allowing a coordinated packaging of the newly synthesized RNA.In this study, we identified annexin A2 (also called annexin II, calpactin 1, and ANXA2) as an HCV host factor by a proteomic analysis. ANXA2 belongs to a family of proteins characterized by their Ca2+-dependent binding to negatively charged phospholipids. The annexin proteins consist of two principle domains, a variable N-terminal and a conserved C-terminal domain, which harbors the Ca2+ and membrane binding sites (for a review, see references 14 and 15). All annexins show cytosolic and membrane localizations. Membrane recruitment probably is regulated by intracellular Ca2+ fluctuations, and target membrane selection differs for different annexins.In addition to showing a cytosolic distribution, ANXA2 can associate with the plasma membrane and the membrane of early endosomes. Plasma membrane-associated ANXA2 typically is found in a tight heterotetrameric complex with the S100 protein S100A10 (p11). ANXA2 specifically interacts with phosphatidylinositol(4,5)bisphosphate (PIP2) (22, 48) and binds to membranes enriched in cholesterol, supporting a role in the organization of lipid raft-like membrane microdomains. Due to the direct binding of ANXA2 to F-actin, the protein has been proposed to provide a direct link between cytoskeletal elements and PIP2/cholesterol-rich membrane domains (47).ANXA2 has been implicated in several cellular transport processes, including the internalization and transport of cholesteryl esters, the biogenesis of multivesicular bodies, the recycling of plasma membrane receptors, and the Ca2+-induced exocytosis of certain secretory granules (14). Here, we show that ANXA2 is present at HCV replication sites within the membranous web. The recruitment of ANXA2 is mediated by domain III of NS5A and probably is required for efficient virus assembly.  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