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 31 毫秒
1.
In animal models of infection, glycoprotein E (gE) is required for efficient herpes simplex virus type 1 (HSV-1) spread from the inoculation site to the cell bodies of innervating neurons (retrograde direction). Retrograde spread in vivo is a multistep process, in that HSV-1 first spreads between epithelial cells at the inoculation site, then infects neurites, and finally travels by retrograde axonal transport to the neuron cell body. To better understand the role of gE in retrograde spread, we used a compartmentalized neuron culture system, in which neurons were infected in the presence or absence of epithelial cells. We found that gE-deleted HSV-1 (NS-gEnull) retained retrograde axonal transport activity when added directly to neurites, in contrast to the retrograde spread defect of this virus in animals. To better mimic the in vivo milieu, we overlaid neurites with epithelial cells prior to infection. In this modified system, virus infects epithelial cells and then spreads to neurites, revealing a 100-fold retrograde spread defect for NS-gEnull. We measured the retrograde spread defect of NS-gEnull from a variety of epithelial cell lines and found that the magnitude of the spread defect from epithelial cells to neurons correlated with epithelial cell plaque size defect, indicating that gE plays a similar role in both types of spread. Therefore, gE-mediated spread between epithelial cells and neurites likely explains the retrograde spread defect of gE-deleted HSV-1 in vivo.Herpes simplex virus type 1 (HSV-1) is an alphaherpesvirus that characteristically infects skin and mucosal surfaces before spreading to sensory neurons, where it establishes a lifelong persistent infection. The virus periodically returns to the periphery via sensory axons and causes recurrent lesions as well as asymptomatic shedding. This life cycle requires viral transport along axons in two directions: toward the neuron cell body (retrograde direction) and away from the neuron cell body (anterograde direction).Many studies of alphaherpesvirus neuronal spread have focused on pseudorabies virus (PRV), a virus whose natural host is the pig. Three PRV proteins, glycoprotein E (gE), gI, and Us9, have been shown to mediate anterograde neuronal spread both in animal models of infection and in cultured neurons. However, these three proteins are dispensable for retrograde spread (3, 8, 11, 12, 31, 46). In contrast, numerous animal models of infection have shown that HSV-1 gE is required for retrograde spread from the inoculation site to the cell bodies of innervating neurons (4, 9, 44, 56). In the murine flank model, wild-type (WT) virus replicates in the skin and then infects sensory neurons and spreads in a retrograde direction to the dorsal root ganglia (DRG). In this model, gE-deleted HSV-1 replicates in the skin but is not detected in the DRG (9, 44). This phenotype differs from gE-deleted PRV, which is able to reach the DRG at WT levels (8). Thus, unlike PRV, gE-deleted HSV-1 viruses have a retrograde spread defect in vivo.HSV-1 gE is a 552-amino-acid type I membrane protein found in the virion membrane as well as in the trans-Golgi and plasma membranes of infected cells (1). gE forms a heterodimer with another viral glycoprotein, gI. The gE/gI complex is important for HSV-1 immune evasion through its Fc receptor activity. gE/gI binds to the Fc domain of antibodies directed against other viral proteins, sequestering these antibodies and blocking antibody effector functions (27, 32, 40). Additionally, gE/gI promotes spread between epithelial cells. Viruses lacking either gE or gI form characteristically small plaques in cell culture and small inoculation site lesions in mice (4, 9, 18, 40, 58). In animal models, gE and gI also mediate viral spread in both anterograde and retrograde directions (4, 19, 44, 56).In order to better understand the role of gE in HSV-1 retrograde neuronal spread, we employed a compartmentalized neuron culture system that has been used to study directional neuronal spread of PRV and West Nile virus (12, 14, 45). In the Campenot chamber system, neurites are contained in a compartment that is separate from their corresponding cell bodies. Therefore, spread in an exclusively retrograde direction can be measured by infecting neurites and detecting spread to neuron cell bodies.HSV-1 replication requires retrograde transport of incoming viral genomes to the nucleus. In neurites, fusion between viral and cellular membranes occurs at the plasma membrane (43, 48). Upon membrane fusion, the capsid and a subset of tegument proteins (the inner tegument) dissociate from glycoproteins and outer tegument proteins, which remain at the plasma membrane (28, 38). Unenveloped capsids and the associated inner tegument proteins are then transported in the retrograde direction to the nucleus (7, 48, 49).For both neurons and epithelial cells, retrograde transport is dependent upon microtubules, ATP, the retrograde microtubule motor dynein, and the dynein cofactor dynactin (22, 34, 49, 52). Several viral proteins interact with components of the dynein motor complex (23, 39, 60). However, none of these proteins suggest a completely satisfactory mechanism by which viral retrograde transport occurs, either because they are not components of the complex that is transported to the nucleus (UL34, UL9, VP11/12) or because capsids lacking that protein retain retrograde transport activity (VP26) (2, 17, 21, 28, 37). This implies that additional viral proteins are involved in retrograde trafficking.We sought to better characterize the role of gE in retrograde spread and found that gE is dispensable for retrograde axonal transport; however, it promotes HSV-1 spread from epithelial cells to neurites. This epithelial cell-to-neuron spread defect provides a plausible explanation for the retrograde spread defect of gE-deleted HSV-1 in animal models of infection.  相似文献   

2.
Borna disease virus (BDV), the prototypic member of the family Bornaviridae within the order Mononegavirales, provides an important model for the investigation of viral persistence within the central nervous system (CNS) and of associated brain disorders. BDV is highly neurotropic and enters its target cell via receptor-mediated endocytosis, a process mediated by the virus surface glycoprotein (G), but the cellular factors and pathways determining BDV cell tropism within the CNS remain mostly unknown. Cholesterol has been shown to influence viral infections via its effects on different viral processes, including replication, budding, and cell entry. In this work, we show that cell entry, but not replication and gene expression, of BDV was drastically inhibited by depletion of cellular cholesterol levels. BDV G-mediated attachment to BDV-susceptible cells was cholesterol independent, but G localized to lipid rafts (LR) at the plasma membrane. LR structure and function critically depend on cholesterol, and hence, compromised structural integrity and function of LR caused by cholesterol depletion likely inhibited the initial stages of BDV cell internalization. Furthermore, we also show that viral-envelope cholesterol is required for BDV infectivity.Borna disease virus (BDV) is an enveloped virus with a nonsegmented negative-strand RNA genome whose organization (3′-N-p10/P-M-G-L-5′) is characteristic of mononegaviruses (6, 28, 46, 48). However, based on its unique genetics and biological features, BDV is considered to be the prototypic member of a new virus family, Bornaviridae, within the order Mononegavirales (8, 28, 46, 49).BDV can infect a variety of cell types in cell culture but in vivo exhibits exquisite neurotropism and causes central nervous system (CNS) disease in different vertebrate species, which is frequently manifested in behavioral abnormalities (19, 33, 44, 53). Both host and viral factors contribute to a variable period of incubation and heterogeneity in the symptoms and pathology associated with BDV infection (14, 16, 29, 42, 44). BDV provides an important model for the investigation of both immune-mediated pathological events associated with virus-induced neurological disease and mechanisms whereby noncytolytic viruses induce neurodevelopmental and behavioral disturbances in the absence of inflammation (15, 18, 41). Moreover, serological data and molecular epidemiological studies suggest that BDV, or a BDV-like virus, can infect humans and that it might be associated with certain neuropsychiatric disorders (17, 24), which further underscores the interest in understanding the mechanisms underlying BDV persistence in the CNS and its effect on brain cell functions. The achievement of these goals will require the elucidation of the determinants of BDV cell tropism within the CNS.BDV enters its target cell via receptor-mediated endocytosis, a process in which the BDV G protein plays a central role (1, 5, 13, 14, 39). Cleavage of BDV G by the cellular protease furin generates two functional subunits: GP1 (GPN), involved in virus interaction with a yet-unidentified cell surface receptor (1, 39), and GP2 (GPC), which mediates a pH-dependent fusion event between viral and cellular membranes (13). However, a detailed characterization of cellular factors and pathways involved in BDV cell entry remains to be done.Besides cell surface molecules that serve as viral receptors, many other cell factors, including nonproteinaceous molecules, can influence cell entry by virus (52). In this regard, cholesterol, which plays a critical role in cellular homeostasis (55), has also been identified as a key factor required for productive infection by different viruses. Accordingly, cholesterol participates in a variety of processes in virus-infected cells, including fusion events between viral and cellular membranes (3), viral replication (23), and budding (35, 37), as well as maintenance of lipid rafts (LR) (12) as scaffold structures where the viral receptor and coreceptor associate (11, 26, 32, 36). LR are specialized microdomains within cellular membranes constituted principally of proteins, sphingolipids, and cholesterol. LR facilitate the close proximity and interaction of specific sets of proteins and contribute to different processes associated with virus multiplication (38). Cholesterol can also influence virus infection by contributing to the maintenance of the properties of the viral envelope required for virus particle infectivity (21, 54). Here, we show for the first time that cholesterol plays a critical role in BDV infection. Depletion of cellular cholesterol prior to, but not after, BDV cell entry prevented productive BDV infection, likely due to disruption of plasma membrane LR that appear to be the cell entry point for BDV. In addition, we document that cholesterol also plays an essential role in the properties of the BDV envelope required for virus particle infectivity.  相似文献   

3.
A major obstacle to gene transduction by viral vectors is inactivation by human complement in vivo. One way to overcome this is to incorporate complement regulatory proteins, such as CD55/decay accelerating factor (DAF), into viral particles. Lentivirus vectors pseudotyped with the baculovirus envelope protein GP64 have been shown to acquire more potent resistance to serum inactivation and longer transgene expression than those pseudotyped with the vesicular stomatitis virus (VSV) envelope protein G. However, the molecular mechanisms underlying resistance to serum inactivation in pseudotype particles bearing the GP64 have not been precisely elucidated. In this study, we generated pseudotype and recombinant VSVs bearing the GP64. Recombinant VSVs generated in human cell lines exhibited the incorporation of human DAF in viral particles and were resistant to serum inactivation, whereas those generated in insect cells exhibited no incorporation of human DAF and were sensitive to complement inactivation. The GP64 and human DAF were detected on the detergent-resistant membrane and were coprecipitated by immunoprecipitation analysis. A pseudotype VSV bearing GP64 produced in human DAF knockdown cells reduced resistance to serum inactivation. In contrast, recombinant baculoviruses generated in insect cells expressing human DAF or carrying the human DAF gene exhibited resistance to complement inactivation. These results suggest that the incorporation of human DAF into viral particles by interacting with baculovirus GP64 is involved in the acquisition of resistance to serum inactivation.Gene therapy is a potential treatment option for genetic diseases, malignant diseases, and other acquired diseases. To this end, safe and efficient gene transfer into specific target cells is a central requirement, and a variety of nonviral and viral vector systems have been developed (6, 44). Recombinant viruses can be used for efficient gene transfer. Retroviruses, adeno-associated viruses, and lentiviruses are able to integrate foreign genes into host genomes and are suitable for gene therapeutics by virtue of their permanent expression of the therapeutic genes, whereas adenoviruses, herpesviruses, and baculoviruses can transiently express foreign genes (6, 12, 44). Pseudotype particles bearing other viral envelope proteins have been developed to improve transduction efficiency and the safety of viral vectors, including retrovirus (4, 7), lentivirus (25), vesicular stomatitis virus (VSV) (29), and baculovirus (17, 42). Pseudotype retroviruses and lentiviruses bearing the baculovirus envelope protein GP64 of Autographa californica nucleopolyhedrosis virus (AcNPV) have been shown to exhibit efficient gene transduction into a wide variety of cells with a lower cytotoxicity compared to those bearing the VSV envelope protein G (VSVG), which is commonly used for pseudotyping (18, 32, 35, 36).However, a drawback of gene transduction by viral vectors is that human sera inactivate the vectors (11, 40). Complement is a major element of the innate immune response and serves to link innate and adaptive immunity (8). Complement activation can occur via classical, lectin, and alternative pathways (2, 8). All pathways invoke several responses, such as virus opsonization, virolysis, anaphylatoxin, and chemotaxin production, as well as others (2, 8). VSV and baculovirus are inactivated by human sera via the classical pathway (1, 11). Because complement activation also induces potential damage to host cells, the complement system is tightly regulated by the complement regulatory proteins (CRPs), including CD55/decay-accelerating factor (DAF), CD46/membrane cofactor protein (MCP), and CD59 (2, 8, 15). DAF and CD46 inhibit activation of C3/C5-converting enzymes, which regulate the activation of classical and alternative pathways, whereas CD59 regulates the assembly of the membrane attack complex (2, 8, 15).Viral vectors can be manipulated to confer resistance to the complement inactivation. Human immunodeficiency virus (HIV) is known to develop resistance to human complement through the incorporation of DAF, CD46, and CD59 to the viral particles (22, 30, 31, 38). Moloney murine leukemia virus vectors produced in HT1080 cells are resistant to complement inactivation (5). Baculovirus and lentivirus vectors bearing DAF or the fusion protein between the functional domains of human DAF and the GP64 were resistant to complement inactivation (9, 13). It has been shown that lentivirus vectors pseudotyped with the GP64 are more resistant to inactivation in the sera of mice and rats (14, 32) and are capable of executing longer expression of the transgenes in nasal epithelia compared to those pseudotyped with the VSVG (35, 36). However, the precise mechanisms underlying the resistance to complement inactivation by pseudotyping of the GP64 is not known.To clarify the molecular mechanisms underlying the resistance of the viral vectors pseudotyped with the GP64 to the complement inactivation, we produced pseudotype and recombinant VSVs bearing the GP64. The recombinant VSVs carrying the gp64 gene generated in human cells but not in insect cells exhibited incorporation of human DAF on the viral particles and were resistant to the complement inactivation. Furthermore, production of the gp64 pseudotype VSV in the DAF knockdown human cells impaired serum resistance, whereas production of the gp64 recombinant VSV in the CHO cell lines stably expressing human DAF and the recombinant baculoviruses in the insect cells stably expressing human DAF or encoding the DAF gene in the genome conferred resistance to the complement inactivation. These results suggest that DAF incorporation into viral particles bearing baculovirus GP64 confers resistance to serum inactivation.  相似文献   

4.
Highly active antiretroviral therapy (HAART) can reduce human immunodeficiency virus type 1 (HIV-1) viremia to clinically undetectable levels. Despite this dramatic reduction, some virus is present in the blood. In addition, a long-lived latent reservoir for HIV-1 exists in resting memory CD4+ T cells. This reservoir is believed to be a source of the residual viremia and is the focus of eradication efforts. Here, we use two measures of population structure—analysis of molecular variance and the Slatkin-Maddison test—to demonstrate that the residual viremia is genetically distinct from proviruses in resting CD4+ T cells but that proviruses in resting and activated CD4+ T cells belong to a single population. Residual viremia is genetically distinct from proviruses in activated CD4+ T cells, monocytes, and unfractionated peripheral blood mononuclear cells. The finding that some of the residual viremia in patients on HAART stems from an unidentified cellular source other than CD4+ T cells has implications for eradication efforts.Successful treatment of human immunodeficiency virus type 1 (HIV-1) infection with highly active antiretroviral therapy (HAART) reduces free virus in the blood to levels undetectable by the most sensitive clinical assays (18, 36). However, HIV-1 persists as a latent provirus in resting, memory CD4+ T lymphocytes (6, 9, 12, 16, 48) and perhaps in other cell types (45, 52). The latent reservoir in resting CD4+ T cells represents a barrier to eradication because of its long half-life (15, 37, 40-42) and because specifically targeting and purging this reservoir is inherently difficult (8, 25, 27).In addition to the latent reservoir in resting CD4+ T cells, patients on HAART also have a low amount of free virus in the plasma, typically at levels below the limit of detection of current clinical assays (13, 19, 35, 37). Because free virus has a short half-life (20, 47), residual viremia is indicative of active virus production. The continued presence of free virus in the plasma of patients on HAART indicates either ongoing replication (10, 13, 17, 19), release of virus after reactivation of latently infected CD4+ T cells (22, 24, 31, 50), release from other cellular reservoirs (7, 45, 52), or some combination of these mechanisms. Finding the cellular source of residual viremia is important because it will identify the cells that are still capable of producing virus in patients on HAART, cells that must be targeted in any eradication effort.Detailed analysis of this residual viremia has been hindered by technical challenges involved in working with very low concentrations of virus (13, 19, 35). Recently, new insights into the nature of residual viremia have been obtained through intensive patient sampling and enhanced ultrasensitive sequencing methods (1). In a subset of patients, most of the residual viremia consisted of a small number of viral clones (1, 46) produced by a cell type severely underrepresented in the peripheral circulation (1). These unique viral clones, termed predominant plasma clones (PPCs), persist unchanged for extended periods of time (1). The persistence of PPCs indicates that in some patients there may be another major cellular source of residual viremia (1). However, PPCs were observed in a small group of patients who started HAART with very low CD4 counts, and it has been unclear whether the PPC phenomenon extends beyond this group of patients. More importantly, it has been unclear whether the residual viremia generally consists of distinct virus populations produced by different cell types.Since the HIV-1 infection in most patients is initially established by a single viral clone (23, 51), with subsequent diversification (29), the presence of genetically distinct populations of virus in a single individual can reflect entry of viruses into compartments where replication occurs with limited subsequent intercompartmental mixing (32). Sophisticated genetic tests can detect such population structure in a sample of viral sequences (4, 39, 49). Using two complementary tests of population structure (14, 43), we analyzed viral sequences from multiple sources within individual patients in order to determine whether a source other than circulating resting CD4+ T cells contributes to residual viremia and viral persistence. Our results have important clinical implications for understanding HIV-1 persistence and treatment failure and for improving eradication strategies, which are currently focusing only on the latent CD4+ T-cell reservoir.  相似文献   

5.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

6.
7.
8.
9.
10.
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   

11.
12.
The cellular immune response to primary influenza virus infection is complex, involving multiple cell types and anatomical compartments, and is difficult to measure directly. Here we develop a two-compartment model that quantifies the interplay between viral replication and adaptive immunity. The fidelity of the model is demonstrated by accurately confirming the role of CD4 help for antibody persistence and the consequences of immune depletion experiments. The model predicts that drugs to limit viral infection and/or production must be administered within 2 days of infection, with a benefit of combination therapy when administered early, and cytotoxic CD8 T cells in the lung are as effective for viral clearance as neutralizing antibodies when present at the time of challenge. The model can be used to investigate explicit biological scenarios and generate experimentally testable hypotheses. For example, when the adaptive response depends on cellular immune cell priming, regulation of antigen presentation has greater influence on the kinetics of viral clearance than the efficiency of virus neutralization or cellular cytotoxicity. These findings suggest that the modulation of antigen presentation or the number of lung resident cytotoxic cells and the combination drug intervention are strategies to combat highly virulent influenza viruses. We further compared alternative model structures, for example, B-cell activation directly by the virus versus that through professional antigen-presenting cells or dendritic cell licensing of CD8 T cells.Understanding how the immune system combats influenza virus infection and how the virus can affect the immune system is crucial to predicting and designing prophylactic and therapeutic strategies against the infection (58). Antigenic shift and antigenic drift alter the degree to which preexisting immunity can control the virus. These factors also influence whether different arms of the adaptive immune system can cross-react against new strains of the virus. For example, shifts of the hemagglutinin (HA) and neuraminidase (NA) protein sequences limit the ability of antibodies to neutralize new variants of the virus and may make cross-reactive T-cell responses to conserved viral proteins more important. Other viral proteins, such as NS1, affect both the induction of type I interferon as well as the susceptibility of infected cells to interferon-mediated inhibition of viral gene expression (43). The efficiencies of viral replication and cell-to-cell viral spread are altered by mutations in the viral matrix and polymerase genes, while the survival of infected cells can be altered by the viral PB1-F2 protein. These attributes are influenced by mutations in the viral matrix (50, 51) and polymerase (30, 69) genes, while the survival of infected cells can be altered by the viral PB1-F2 protein (17). The multigenic aspect of influenza virus pathogenesis makes experimental prediction difficult and time-consuming. Computer simulation tools would be useful to independently dissect the potential contribution and relative importance of each factor or to investigate unexpected scenarios that are difficult to replicate experimentally.Mathematical models and computer simulations have been widely used to study viral dynamics and immune responses to viral infections, such as human immunodeficiency virus type 1 (HIV-1) and simian immunodeficiency viruses (SIV), lymphocytic choriomeningitis virus (19, 55, 60, 61), and influenza A virus (3, 7, 8, 13, 34, 35, 52). More complex compartmental models of the immune system (4, 23) and models incorporating differential delay equations (21, 48, 68) have been used to better reflect the time that cells reside in a particular compartment or the duration of transit between compartments. In this study, we sought to develop a two-compartment mathematical model to assess the individual contributions of antigen presentation and activation of naïve T and B cells by antigen-presenting cells (APC), CD4 T-cell help, CD8 T-cell-mediated cytotoxicity, B cells, and antibody to control influenza A virus (IAV) infection and to explore the influence of anatomical location. We developed a model which represented published experimental findings on primary influenza virus infection. More importantly, the model was used to explore alternative structures for interactions between virus and immune cells, for example, comparing virus kinetics when antigen delivery and immune cell priming occurred through direct interaction of virus and immune cells or through a cellular intermediate. The model predicts that, under some circumstances, changes affecting antigen presentation more strongly impacted viral kinetics than other viral or immune factors (28, 73, 75, 78). This model highlights the importance of the assumptions used to synthesize a model and gaps in our understanding of the immune response regulating primary influenza virus infection. We discuss the implications of these findings for future influenza virus research and theories of influenza virus virulence based on influenza virus-immune system interactions.  相似文献   

13.
Influenza A virus buds through the apical plasma membrane, forming enveloped virus particles that can take the shape of pleomorphic spheres or vastly elongated filaments. For either type of virion, the factors responsible for separation of viral and cell membranes are not known. We find that cellular Rab11 (a small GTP-binding protein involved in endocytic recycling) and Rab11-family interacting protein 3 ([FIP3] which plays a role in membrane trafficking and regulation of actin dynamics) are both required to support the formation of filamentous virions, while Rab11 is additionally involved in the final budding step of spherical particles. Cells transfected with Rab11 GTP-cycling mutants or depleted of Rab11 or FIP3 content by small interfering RNA treatment lost the ability to form virus filaments. Depletion of Rab11 resulted in up to a 100-fold decrease in titer of spherical virus released from cells. Scanning electron microscopy of Rab11-depleted cells showed high densities of virus particles apparently stalled in the process of budding. Transmission electron microscopy of thin sections confirmed that Rab11 depletion resulted in significant numbers of abnormally formed virus particles that had failed to pinch off from the plasma membrane. Based on these findings, we see a clear role for a Rab11-mediated pathway in influenza virus morphogenesis and budding.Influenza A virus is a highly infectious respiratory pathogen, causing 3 to 5 million severe cases yearly while the recent H1N1 pandemic has spread to over 200 countries and resulted in over 15,000 WHO-confirmed deaths since its emergence in March 2009 (57). Influenza virus particles are enveloped structures that contain nine identified viral polypeptides. The lipid envelope is derived by budding from the apical plasma membrane and contains the viral integral membrane proteins hemagglutinin (HA) and neuraminidase (NA) as well as the M2 ion channel. Internally, virus particles contain a matrix protein (M1), small quantities of the NS2/NEP polypeptide, and eight genomic segments of negative-sense RNA that are separately encapsidated into ribonucleoprotein (RNP) particles by the viral nucleoprotein (NP) and tripartite polymerase complex (PB1, PB2, and PA). M1 is thought to form a link between the RNPs and the cytoplasmic tails of the viral membrane proteins though M2 may also play a role (39). The minimal viral protein requirements for budding are disputed; while initial studies suggested that M1 was the main driver of budding (21, 34), more recent work proposes that the glycoproteins HA and NA are responsible (8).Further complicating the analysis of influenza A virus budding is the observation that most strains of the virus form two distinct types of virions: spherical particles approximately 100 nm in diameter and much longer filamentous particles up to 30 μm in length (38). Of the viral proteins, M1 is the primary determinant of particle shape (3, 17) although other virus genes also play a role. It is also likely that host factors are involved in the process as cells with fully differentiated apical and basolateral membranes produce more filaments than nonpolarized cell types (42). While it is tempting to speculate that virus morphology and budding are regulated by the same cellular process, the fact that spherical budding occurs in the absence of an intact actin cytoskeleton while filament formation does not (42, 48) indicates some level of divergence in the mechanisms responsible for spherical and filamentous virion morphogenesis.The means by which viral and cellular membranes are separated are also unclear. Unlike many other enveloped viruses, including retroviruses (19, 36, 52) and herpes simplex virus (12), influenza A virus does not utilize the cellular endosomal sorting complex required for transport (ESCRT) pathway (5, 8). However, recent reports indicate that some viruses, including human cytomegalovirus (HCMV) (32), the hantavirus Andes virus (44), and respiratory syncytial virus (RSV) may employ a Rab11-mediated pathway during assembly and/or budding (4, 51). The Rab family of small GTPases is involved in targeting vesicle trafficking, mediating a wide range of downstream processes including endosomal trafficking and membrane fusion/fission events (reviewed in references 53 and 58). Rab11 is involved in trafficking proteins and vesicles between the trans-Golgi network (TGN), recycling endosome, and the plasma membrane (9, 49, 50) as well as playing a role in actin remodeling, cytokinesis, and abscission (27, 41, 55). Apical recycling endosome (ARE) trafficking is of particular interest in the context of viral infection as other negative-sense RNA viruses have been shown to assemble and/or traffic virion components through the ARE prior to final assembly and budding at the plasma membrane (4, 44, 51). Rab11 function is modulated and targeted through interactions with Rab11 family interacting proteins (Rab11-FIPs) that direct it to specific subcellular locations (23, 25, 26) by binding to actin or microtubule-based motor proteins (24, 26, 47). While Rab11-FIPs recognize both isoforms of Rab11 (a and b [Rab11a/b]) through a conserved amphipathic α-helical motif, they differ in their ability to bind either the GTP-bound form of Rab11 (FIP1, FIP3, FIP4, and Rip11) or both the GTP and GDP-bound forms (FIP2) (23, 30). FIP1 and FIP2 have been implicated in RSV budding (4, 51) while FIP4 is important for trafficking of HCMV components (32). FIP3 has not previously been linked with virus budding but plays an important role in both cell motility and cytokinesis, regulating actin dynamics and endosomal membrane trafficking (29, 55).In light of the normal cellular functions of Rab11 and its effectors and of their reported involvement in the budding of other viruses, we examined the role of this cellular pathway in influenza virus budding. We find that Rab11-FIP3 is essential for filamentous but not spherical virion formation while Rab11 is required for both forms of virus budding.  相似文献   

14.
BST-2/tetherin is an interferon-inducible protein that restricts the release of enveloped viruses from the surface of infected cells by physically linking viral and cellular membranes. It is present at both the cell surface and in a perinuclear region, and viral anti-tetherin factors including HIV-1 Vpu and HIV-2 Env have been shown to decrease the cell surface population. To map the domains of human tetherin necessary for both virus restriction and sensitivity to viral anti-tetherin factors, we constructed a series of tetherin derivatives and assayed their activity. We found that the cytoplasmic tail (CT) and transmembrane (TM) domains of tetherin alone produced its characteristic cellular distribution, while the ectodomain of the protein, which includes a glycosylphosphatidylinositol (GPI) anchor, was sufficient to restrict virus release when presented by the CT/TM regions of a different type II membrane protein. To counteract tetherin restriction and remove it from the cell surface, HIV-1 Vpu required the specific sequence present in the TM domain of human tetherin. In contrast, the HIV-2 Env required only the ectodomain of the protein and was sensitive to a point mutation in this region. Strikingly, the anti-tetherin factor, Ebola virus GP, was able to overcome restriction conferred by both tetherin and a series of functional tetherin derivatives, including a wholly artificial tetherin molecule. Moreover, GP overcame restriction without significantly removing tetherin from the cell surface. These findings suggest that Ebola virus GP uses a novel mechanism to circumvent tetherin restriction.Pathogenic viruses often have evolved mechanisms to neutralize host defenses that act at the cellular level to interfere with the virus life cycle. Such cellular restriction factors have been most extensively characterized for HIV-1 (38) and include the interferon-inducible membrane protein BST-2/HM1.24/CD317/tetherin (28, 40). If unchecked, tetherin blocks the release of newly formed HIV-1 particles from cells by physically tethering them at the cell surface (7, 28, 32, 40). In addition, tetherin has been shown to act against a broad range of enveloped viral particles, including retroviruses, filoviruses, arenaviruses, and herpesviruses (17, 18, 23, 35). In turn, certain viruses that are targeted by tetherin appear to have evolved counteracting activities, and anti-tetherin factors so far identified include HIV-1 Vpu; HIV-2 Env; simian immunodeficiency virus (SIV) Nef, Vpu, and Env proteins; Ebola virus GP; and Kaposi''s sarcoma-associated herpesvirus (KSHV) K5 (11, 16, 18, 20, 23, 28, 36, 40, 44, 45).Tetherin is a homodimeric type II integral membrane protein containing an N-terminal cytoplasmic tail (CT), a single-pass transmembrane domain (TM), an ectodomain-containing predicted coiled-coil regions, two glycoslyation sites, three conserved cysteines, and a C-terminal glycosylphosphatidylinositol (GPI) anchor (2, 19, 31). This unusual topology, with two independent membrane anchors, has led to the suggestion that the retention of virions at the cell surface arises from tetherin''s ability to be inserted simultaneously in both host and viral membranes (28, 32, 41) or, alternatively, that dimers or higher-order complexes of tetherin conferred by the ectodomain mediate this effect (39). Interestingly, an artificial tetherin containing the same structural features as the native protein but constructed from unrelated sequences was able to restrict both HIV-1 and Ebola virus particles (32). This suggests that the viral lipid envelope is the target of tetherin and provides an explanation for tetherin''s broad activity against diverse enveloped viruses.A fraction of tetherin is present at the plasma membrane of cells (9, 14), and it has been proposed that viral anti-tetherin factors function by removing this cell surface fraction (40). This now has been shown to occur in the presence of HIV-1 Vpu (5, 7, 15, 26, 34, 40, 44), HIV-2 Env (5, 20), SIV Env (11), SIV Nef (15), and KSHV K5 (3, 23). In addition, certain anti-tetherin factors also may promote the degradation of tetherin, as has been observed for both HIV-1 Vpu (3, 5, 7, 10, 22, 26, 27) and KSHV K5 (3, 23), although Vpu also appears able to block tetherin restriction in the absence of degradation (8), and no effects on tetherin steady-state levels have been observed in the presence of either the HIV-2 or SIVtan Env (11, 20). Simply keeping tetherin away from the cell surface, or targeting it for degradation, may not be the only mechanism used by anti-tetherin factors, since it also has been reported that Vpu does not affect the levels of surface tetherin or its total cellular levels in certain T-cell lines (27).The interactions between tetherin and viral anti-tetherin factors show evidence of species specificity, suggesting ongoing evolution between viruses and their hosts. HIV-1 Vpu is active against human and chimpanzee tetherin but not other primate tetherins (10, 25, 34, 36, 44, 45), while SIV Nef proteins are active against primate but not human tetherins (16, 36, 44, 45). This suggests that, unlike tetherin restriction, the action of the anti-tetherin factors may involve specific sequence interactions. Indeed, the TM domain has been recognized as a target for HIV-1 Vpu (10, 15, 16, 25, 34), while a single point mutation introduced into the extracellular domain of human tetherin can block its antagonism by the SIVtan Env (11).In the present study, we investigated the roles of the different domains of tetherin in both promoting virus restriction and conferring susceptibility to the anti-tetherin factors encoded by HIV-1, HIV-2, and Ebola virus. We confirmed that tetherin restriction can be conferred by proteins that retain the two distinct membrane anchors, while signals for the cellular localization of the protein reside in the CT/TM domains of the protein. We found that the Vpu protein targets the TM domain of tetherin, while the HIV-2 Env targets the ectodomain of the protein. In contrast, the Ebola virus GP appears to use a non-sequence-specific mechanism to counteract tetherin restriction, since even an artificial tetherin could be successfully overcome by GP expression. Interestingly, Ebola virus GP counteracted tetherin restriction without removing the protein from the cell surface, suggesting that it is possible to overcome this restriction by mechanisms other than blocking tetherin''s cell surface expression.  相似文献   

15.
Human immunodeficiency virus type 1 (HIV-1) envelope protein (Env) is subject to both neutralizing antibody (NAb) and CD8 T-cell (cytotoxic T-lymphocyte [CTL]) immune pressure. We studied the reversion of the Env CTL escape mutant virus to the wild type and the relationship between the reversion of CTL mutations with N-linked glycosylation site (NLGS)-driven NAb escape in pigtailed macaques. Env CTL mutations either did not revert to the wild type or only transiently reverted 5 to 7 weeks after infection. The CTL escape mutant reversion was coincident, for the same viral clones, with the loss of NLGS mutations. At one site studied, both CTL and NLGS mutations were needed to confer NAb escape. We conclude that CTL and NAb escape within Env can be tightly linked, suggesting opportunities to induce effective multicomponent anti-Env immunity.CD8 T-cell responses against human immunodeficiency virus (HIV) have long been observed to select for viral variants that avoid cytotoxic T-lymphocyte (CTL) recognition (2, 5, 15, 18, 27). These immune escape mutations may, however, result in reduced replication competence (“fitness cost”) (11, 20, 26). CTL escape variants have been shown to revert to the wild type (WT) upon passage to major histocompatibility complex-mismatched hosts, both in macaques with simian immunodeficiency virus (SIV) or chimeric SIV/HIV (SHIV) infection (11, 12) and in humans with HIV type 1 (HIV-1) infection (1, 19).Most analyses of CTL escape and reversion have studied Gag CTL epitopes known to facilitate control of viremia (7, 14, 21, 30). Fewer analyses have studied Env-specific CTL epitopes. Recent sequencing studies suggest the potential for mutations within predicted HIV-1 Env-specific CTL epitopes to undergo reversion to the WT (16, 23). Env-specific CTL responses may, however, have less impact on viral control of both HIV-1 and SIV/SHIV than do Gag CTL responses (17, 24, 25), presumably reflecting either less-potent inhibition of viral replication or minimal fitness cost of escape (9).Serial viral escape from antibody pressure also occurs in both macaques and humans (3, 13, 28). Env is extensively glycosylated, and this “evolving glycan shield” can sterically block antibody binding without mutation at the antibody-binding site (8, 16, 31). Mutations at glycosylation sites, as well as other mutations, are associated with escape from neutralizing antibody (NAb) responses (4, 13, 29). Mutations in the amino acid sequences of N-linked glycosylation sites (NLGS) can alter the packing of the glycan cloud that surrounds the virion, by a loss, gain, or shift of an NLGS (32), thus facilitating NAb escape.Env is the only viral protein targeted by both CTL and NAb responses. The serial viral escape from both Env-specific CTL and NAb responses could have implications for viral fitness and the reversion of multiple mutations upon transmission to naïve hosts.We previously identified three common HIV-1 Env-specific CD8 T cell epitopes, RY8788-795, SP9110-118, and NL9671-679, and their immune escape patterns in pigtail macaques (Macaca nemestrina) infected with SHIVmn229 (25). SHIVmn229 is a chimeric virus constructed from an SIVmac239 backbone and an HIV-1HXB2 env fragment that was passaged through macaques to become pathogenic (11). This earlier work provided an opportunity for detailed studies of how viruses with Env-specific CTL escape mutations, as well as mutations in adjacent NLGS, evolve when transmitted to naïve pigtail macaques.  相似文献   

16.
HIV-1 possesses an exquisite ability to infect cells independently from their cycling status by undergoing an active phase of nuclear import through the nuclear pore. This property has been ascribed to the presence of karyophilic elements present in viral nucleoprotein complexes, such as the matrix protein (MA); Vpr; the integrase (IN); and a cis-acting structure present in the newly synthesized DNA, the DNA flap. However, their role in nuclear import remains controversial at best. In the present study, we carried out a comprehensive analysis of the role of these elements in nuclear import in a comparison between several primary cell types, including stimulated lymphocytes, macrophages, and dendritic cells. We show that despite the fact that none of these elements is absolutely required for nuclear import, disruption of the central polypurine tract-central termination sequence (cPPT-CTS) clearly affects the kinetics of viral DNA entry into the nucleus. This effect is independent of the cell cycle status of the target cells and is observed in cycling as well as in nondividing primary cells, suggesting that nuclear import of viral DNA may occur similarly under both conditions. Nonetheless, this study indicates that other components are utilized along with the cPPT-CTS for an efficient entry of viral DNA into the nucleus.Lentiviruses display an exquisite ability to infect dividing and nondividing cells alike that is unequalled among Retroviridae. This property is thought to be due to the particular behavior or composition of the viral nucleoprotein complexes (NPCs) that are liberated into the cytoplasm of target cells upon virus-to-cell membrane fusion and that allow lentiviruses to traverse an intact nuclear membrane (17, 28, 29, 39, 52, 55, 67, 79). In the case of the human immunodeficiency type I virus (HIV-1), several studies over the years identified viral components of such structures with intrinsic karyophilic properties and thus perfect candidates for mediation of the passage of viral DNA (vDNA) through the nuclear pore: the matrix protein (MA); Vpr; the integrase (IN); and a three-stranded DNA flap, a structure present in neo-synthesized viral DNA, specified by the central polypurine tract-central termination sequence (cPPT-CTS). It is clear that these elements may mediate nuclear import directly or via the recruitment of the host''s proteins, and indeed, several cellular proteins have been found to influence HIV-1 infection during nuclear import, like the karyopherin α2 Rch1 (38); importin 7 (3, 30, 93); the transportin SR-2 (13, 20); or the nucleoporins Nup98 (27), Nup358/RANBP2, and Nup153 (13, 56).More recently, the capsid protein (CA), the main structural component of viral nucleoprotein complexes at least upon their cytoplasmic entry, has also been suggested to be involved in nuclear import or in postnuclear entry steps (14, 25, 74, 90, 92). Whether this is due to a role for CA in the shaping of viral nucleoprotein complexes or to a direct interaction between CA and proteins involved in nuclear import remains at present unknown.Despite a large number of reports, no single viral or cellular element has been described as absolutely necessary or sufficient to mediate lentiviral nuclear import, and important controversies as to the experimental evidences linking these elements to this step exist. For example, MA was among the first viral protein of HIV-1 described to be involved in nuclear import, and 2 transferable nuclear localization signals (NLSs) have been described to occur at its N and C termini (40). However, despite the fact that early studies indicated that the mutation of these NLSs perturbed HIV-1 nuclear import and infection specifically in nondividing cells, such as macrophages (86), these findings failed to be confirmed in more-recent studies (23, 33, 34, 57, 65, 75).Similarly, Vpr has been implicated by several studies of the nuclear import of HIV-1 DNA (1, 10, 21, 43, 45, 47, 64, 69, 72, 73, 85). Vpr does not possess classical NLSs, yet it displays a transferable nucleophilic activity when fused to heterologous proteins (49-51, 53, 77, 81) and has been shown to line onto the nuclear envelope (32, 36, 47, 51, 58), where it can truly facilitate the passage of the viral genome into the nucleus. However, the role of Vpr in this step remains controversial, as in some instances Vpr is not even required for viral replication in nondividing cells (1, 59).Conflicting results concerning the role of IN during HIV-1 nuclear import also exist. Indeed, several transferable NLSs have been described to occur in the catalytic core and the C-terminal DNA binding domains of IN, but for some of these, initial reports of nuclear entry defects (2, 9, 22, 46, 71) were later shown to result from defects at steps other than nuclear import (60, 62, 70, 83). These reports do not exclude a role for the remaining NLSs in IN during nuclear import, and they do not exclude the possibility that IN may mediate this step by associating with components of the cellular nuclear import machinery, such as importin alpha and beta (41), importin 7 (3, 30, 93, 98), and, more recently, transportin-SR2 (20).The central DNA flap, a structure present in lentiviruses and in at least 1 yeast retroelement (44), but not in other orthoretroviruses, has also been involved in the nuclear import of viral DNA (4, 6, 7, 31, 78, 84, 95, 96), and more recently, it has been proposed to provide a signal for viral nucleoprotein complexes uncoating in the proximity of the nuclear pore, with the consequence of providing a signal for import (8). However, various studies showed an absence or weakness of nuclear entry defects in viruses devoid of the DNA flap (24, 26, 44, 61).Overall, the importance of viral factors in HIV-1 nuclear import is still unclear. The discrepancies concerning the role of MA, IN, Vpr, and cPPT-CTS in HIV-1 nuclear import could in part be explained by their possible redundancy. To date, only one comprehensive study analyzed the role of these four viral potentially karyophilic elements together (91). This study showed that an HIV-1 chimera where these elements were either deleted or replaced by their murine leukemia virus (MLV) counterparts was, in spite of an important infectivity defect, still able to infect cycling and cell cycle-arrested cell lines to similar efficiencies. If this result indicated that the examined viral elements of HIV-1 were dispensable for the cell cycle independence of HIV, as infections proceeded equally in cycling and arrested cells, they did not prove that they were not required in nuclear import, because chimeras displayed a severe infectivity defect that precluded their comparison with the wild type (WT).Nuclear import and cell cycle independence may not be as simply linked as previously thought. On the one hand, there has been no formal demonstration that the passage through the nuclear pore, and thus nuclear import, is restricted to nondividing cells, and for what we know, this passage may be an obligatory step in HIV infection in all cells, irrespective of their cycling status. In support of this possibility, certain mutations in viral elements of HIV affect nuclear import in dividing as well as in nondividing cells (4, 6, 7, 31, 84, 95). On the other hand, cell cycle-independent infection may be a complex phenomenon that is made possible not only by the ability of viral DNA to traverse the nuclear membrane but also by its ability to cope with pre- and postnuclear entry events, as suggested by the phenotypes of certain CA mutants (74, 92).Given that the cellular environment plays an important role during the early steps of viral infection, we chose to analyze the role of the four karyophilic viral elements of HIV-1 during infection either alone or combined in a wide comparison between cells highly susceptible to infection and more-restrictive primary cell targets of HIV-1 in vivo, such as primary blood lymphocytes (PBLs), monocyte-derived macrophages (MDM), and dendritic cells (DCs).In this study, we show that an HIV-1-derived virus in which the 2 NLSs of MA are mutated and the IN, Vpr, and cPPT-CTS elements are removed displays no detectable nuclear import defect in HeLa cells independently of their cycling status. However, this mutant virus is partially impaired for nuclear entry in primary cells and more specifically in DCs and PBLs. We found that this partial defect is specified by the cPPT-CTS, while the 3 remaining elements seem to play no role in nuclear import. Thus, our study indicates that the central DNA flap specifies the most important role among the viral elements involved thus far in nuclear import. However, it also clearly indicates that the role played by the central DNA flap is not absolute and that its importance varies depending on the cell type, independently from the dividing status of the cell.  相似文献   

17.
Membrane glycoproteins of alphavirus play a critical role in the assembly and budding of progeny virions. However, knowledge regarding transport of viral glycoproteins to the plasma membrane is obscure. In this study, we investigated the role of cytopathic vacuole type II (CPV-II) through in situ electron tomography of alphavirus-infected cells. The results revealed that CPV-II contains viral glycoproteins arranged in helical tubular arrays resembling the basic organization of glycoprotein trimers on the envelope of the mature virions. The location of CPV-II adjacent to the site of viral budding suggests a model for the transport of structural components to the site of budding. Thus, the structural characteristics of CPV-II can be used in evaluating the design of a packaging cell line for replicon production.Semliki Forest virus (SFV) is an enveloped alphavirus belonging to the family Togaviridae. This T=4 icosahedral virus particle is approximately 70 nm in diameter (30) and consists of 240 copies of E1/E2 glycoprotein dimers (3, 8, 24). The glycoproteins are anchored in a host-derived lipid envelope that encloses a nucleocapsid, made of a matching number of capsid proteins and a positive single-stranded RNA molecule. After entry of the virus via receptor-mediated endocytosis, a low-pH-induced fusion of the viral envelope with the endosomal membrane delivers the nucleocapsid into the cytoplasm, where the replication events of SFV occur (8, 19, 30). Replication of the viral genome and subsequent translation into structural and nonstructural proteins followed by assembly of the structural proteins and genome (7) lead to budding of progeny virions at the plasma membrane (18, 20). The synthesis of viral proteins shuts off host cell macromolecule synthesis, which allows for efficient intracellular replication of progeny virus (7). The expression of viral proteins leads to the formation of cytopathic vacuolar compartments as the result of the reorganization of cellular membrane in the cytoplasm of an infected cell (1, 7, 14).Early studies using electron microscopy (EM) have characterized the cytopathic vacuoles (CPVs) in SFV-infected cells (6, 13, 14) and identified two types of CPV, namely, CPV type I (CPV-I) and CPV-II. It was found that CPV-I is derived from modified endosomes and lysosomes (18), while CPV-II is derived from the trans-Golgi network (TGN) (10, 11). Significantly, the TGN and CPV-II vesicles are the major membrane compartments marked with E1/E2 glycoproteins (9, 11, 12). Inhibition by monensin results in the accumulation of E1/E2 glycoproteins in the TGN (12, 26), thereby indicating the origin of CPV-II. While CPV-II is identified as the predominant vacuolar structure at the late stage of SFV infection, the exact function of this particular cytopathic vacuole is less well characterized than that of CPV-I (2, 18), although previous observations have pointed to the involvement of CPV-II in budding, because an associated loss of viral budding was observed when CPV-II was absent (9, 36).In this study, we characterized the structure and composition of CPV-II in SFV-infected cells in situ with the aid of electron tomography and immuno-electron microscopy after physical fixation of SFV-infected cells by high-pressure freezing and freeze substitution (21, 22, 33). The results revealed a helical array of E1/E2 glycoproteins within CPV-II and indicate that CPV-II plays an important role in intracellular transport of glycoproteins prior to SFV budding.  相似文献   

18.
Poxviruses produce complement regulatory proteins to subvert the host''s immune response. Similar to the human pathogen variola virus, ectromelia virus has a limited host range and provides a mouse model where the virus and the host''s immune response have coevolved. We previously demonstrated that multiple components (C3, C4, and factor B) of the classical and alternative pathways are required to survive ectromelia virus infection. Complement''s role in the innate and adaptive immune responses likely drove the evolution of a virus-encoded virulence factor that regulates complement activation. In this study, we characterized the ectromelia virus inhibitor of complement enzymes (EMICE). Recombinant EMICE regulated complement activation on the surface of CHO cells, and it protected complement-sensitive intracellular mature virions (IMV) from neutralization in vitro. It accomplished this by serving as a cofactor for the inactivation of C3b and C4b and by dissociating the catalytic domain of the classical pathway C3 convertase. Infected murine cells initiated synthesis of EMICE within 4 to 6 h postinoculation. The levels were sufficient in the supernatant to protect the IMV, upon release, from complement-mediated neutralization. EMICE on the surface of infected murine cells also reduced complement activation by the alternative pathway. In contrast, classical pathway activation by high-titer antibody overwhelmed EMICE''s regulatory capacity. These results suggest that EMICE''s role is early during infection when it counteracts the innate immune response. In summary, ectromelia virus produced EMICE within a few hours of an infection, and EMICE in turn decreased complement activation on IMV and infected cells.Poxviruses encode in their large double-stranded DNA genomes many factors that modify the immune system (30, 56). The analysis of these molecules has revealed a delicate balance between viral pathogenesis and the host''s immune response (2, 21, 31, 61). Variola, vaccinia, monkeypox, cowpox, and ectromelia (ECTV) viruses each produce an orthologous complement regulatory protein (poxviral inhibitor of complement enzymes [PICE]) that has structural and functional homology to host proteins (14, 29, 34, 38, 41, 45, 54). The loss of the regulatory protein resulted in smaller local lesions with vaccinia virus lacking the vaccinia virus complement control protein (VCP) (29) and in a greater local inflammatory response in the case of cowpox lacking the inflammation-modulatory protein (IMP; the cowpox virus PICE) (35, 45, 46). Additionally, the complete loss of the monkeypox virus inhibitor of complement enzymes (MOPICE) may account for part of the reduced mortality observed in the West African compared to Congo basin strains of monkeypox virus (12).The complement system consists of proteins on the cell surface and in blood that recognize and destroy invading pathogens and infected host cells (36, 52). Viruses protect themselves from the antiviral effects of complement activation in a variety of ways, including hijacking the host''s complement regulatory proteins or producing their own inhibitors (7, 8, 15, 20, 23). Another effective strategy is to incorporate the host''s complement regulators in the outermost viral membrane, which then protects the virus from complement attack (62). The extracellular enveloped virus (EEV) produced by poxviruses acquires a unique outer membrane derived from the Golgi complex or early endosomes that contain the protective host complement regulators (58, 62). Poxviruses have multiple infectious forms, and the most abundant, intracellular mature virions (IMV), are released when infected cells lyse (58). The IMV lacks the outermost membrane found on EEV and is sensitive to complement-mediated neutralization. The multiple strategies viruses have evolved to evade the complement system underscore its importance to innate and adaptive immunity (15, 36).The most well-characterized PICE is VCP (24-29, 34, 49, 50, 53, 55, 59, 60). Originally described as a secreted complement inhibitor (34), VCP also attaches to the surface of infected cells through an interaction with the viral membrane protein A56 that requires an unpaired N-terminal cysteine (26). This extra cysteine also adds to the potency of the inhibitor by forming function-enhancing dimers (41). VCP and the smallpox virus inhibitor of complement enzymes (SPICE) bind heparin in vitro, and this may facilitate cell surface interactions (24, 38, 50, 59). The coevolution of variola virus with its only natural host, humans, likely explains the enhanced activity against human complement observed with SPICE compared to the other PICEs (54, 64).Our recent work with ECTV, the causative agent of mousepox infection, demonstrated that the classical and alternative pathways of the complement system are required for host survival (48). The mouse-specific pathogen ECTV causes severe disease in most strains and has coevolved with its natural host, analogous to variola virus in humans (9). This close host-virus relationship is particularly important for evaluating the role of the complement system, given the species specificity of many complement proteins, receptors, and regulators (10, 47, 62). Additionally, the availability of complement-deficient mice permits dissection of the complement activation pathways involved. Naïve C57BL/6 mouse serum neutralizes the IMV of ECTV in vitro, predominately through opsonization (48). Maximal neutralization requires natural antibody, classical-pathway activation, and amplification by the alternative pathway. C3 deficiency in the normally resistant C57BL/6 strain results in acute mortality, similar to immunodeficiencies in important elements of the antiviral immune response, including CD8+ T cells (19, 32), natural killer cells (18, 51), and gamma interferon (33). During ECTV infection, the complement system acts in the first few hours and days to delay the spread of infection, resulting in lower levels of viremia and viral burden in tissues (48).This study characterized the PICE produced by ECTV, ectromelia virus inhibitor of complement enzymes (EMICE), and assessed its complement regulatory activity. Recombinant EMICE (rEMICE) decreased activation of both human and mouse complement. Murine cells produced EMICE at 4 to 6 h postinfection prior to the release of the majority of the complement-sensitive IMV from infected cells. rEMICE protected ECTV IMV from complement-mediated neutralization. Further, EMICE produced during natural infection inhibited complement deposition on infected cells by the alternative pathway. ECTV likely produces this abundance of EMICE to protect both the IMV and infected cells.  相似文献   

19.
The endosomal sorting complex required for transport (ESCRT) machinery controls the incorporation of cargo into intraluminal vesicles of multivesicular bodies. This machinery is used during envelopment of many RNA viruses and some DNA viruses, including herpes simplex virus type 1. Other viruses mature independent of ESCRT components, instead relying on the intrinsic behavior of viral matrix and envelope proteins to drive envelopment. Human cytomegalovirus (HCMV) maturation has been reported to proceed independent of ESCRT components (A. Fraile-Ramos et al. Cell. Microbiol. 9:2955-2967, 2007). A virus complementation assay was used to evaluate the role of dominant-negative (DN) form of a key ESCRT ATPase, vacuolar protein sorting-4 (Vps4DN) in HCMV replication. Vps4DN specifically inhibited viral replication, whereas wild-type-Vps4 had no effect. In addition, a DN form of charged multivesicular body protein 1 (CHMP1DN) was found to inhibit HCMV. In contrast, DN tumor susceptibility gene-101 (Tsg101DN) did not impact viral replication despite the presence of a PTAP motif within pp150/ppUL32, an essential tegument protein involved in the last steps of viral maturation and release. Either Vps4DN or CHMP1DN blocked viral replication at a step after the accumulation of late viral proteins, suggesting that both are involved in maturation. Both Vps4A and CHMP1A localized in the vicinity of viral cytoplasmic assembly compartments, sites of viral maturation that develop in CMV-infected cells. Thus, ESCRT machinery is involved in the final steps of HCMV replication.Cellular endosomal sorting complex required for transport (ESCRT) machinery controls the evolutionarily conserved process (33) of membrane budding that is normally a component of cytokinesis (6, 46), endosome sorting and multivesicular body (MVB) formation (28). After the initial characterization in retroviruses, many enveloped viruses have been shown to rely on this machinery during envelopment and release from cells (1, 18, 35, 40, 47, 69). Other viruses, such as influenza virus, mature independent of ESCRT machinery and are believed to use an alternative virus-intrinsic pathway (7). The core of the ESCRT machinery consists of five multiprotein complexes (ESCRT-0, -I, -II, and -III and Vps4-Vta1) (27). Vacuolar protein sorting-4 (Vps4) is a critical ATPase that functions downstream of most ESCRT components. Based on sensitivity to dominant-negative (DN) inhibitors of protein function, replication of several RNA viruses, as well as of the DNA virus herpes simplex virus type 1 (HSV-1) (5, 10), have been shown to rely on Vps4 in a manner that is analogous to the formation of MVBs (endosomal compartments containing intraluminal vesicles) (10, 45). Evidence based exclusively on small interfering RNA (siRNA) methods suggested cytomegalovirus (CMV) maturation was independent of ESCRT components, although the maturation of this virus remained MVB associated (16).ESCRT machinery facilitates envelopment and release at cytoplasmic membranes and recruits cargo for sorting via any of three alternative pathways that converge on a Vps4-dependent downstream step: (i) a tumor susceptibility gene-101 (Tsg101)-dependent pathway, (ii) an apoptosis linked gene-2 interacting protein X (ALIX)-dependent pathway, and (iii) a pathway that relies on a subset of Nedd4-like HECT E3 ubiquitin ligases (35). The involvement of ESCRT in viral envelopment and egress was first observed in human immunodeficiency virus (HIV) (18, 19, 40, 60) and has been extended to equine infectious anemia virus (34, 40, 52, 60), Rous sarcoma virus (29, 70, 71), Mason-Pfizer monkey virus (20, 72), rabies virus (24), Ebola virus (23), hepatitis B virus (68), vaccinia virus (25), HSV-1 (5, 10), and several other RNA and DNA viruses (7). Structural proteins in most of these viruses carry late (L) domains characterized by conserved amino acid motifs (PTAP, PPXY, and YXXL) that mediate protein-protein interactions and facilitate recruitment of ESCRT components to facilitate virus budding. The introduction of mutations in these motifs leads to defects in viral maturation and release from cells (40).Vps4 controls the release of ESCRT complexes from membranes (18, 40). Inhibition of Vps4A and Vps4B using Vps4ADN reduces levels of viral maturation mediated by L domains (47). For this reason, inhibition by a Vps4DN is considered the gold standard test to establish the role of ESCRT machinery in maturation of any virus (7). Tsg101, a component of ESCRT-I, normally functions to deliver ubiquitinated transmembrane proteins to MVBs (35). HIV-1 p6 Gag PTAP domain interacts with Tsg101 (18) and directs viral cores (capsids) to sites of viral envelopment (39). Upon disruption of HIV-1 PTAP domain, particle release becomes dependent on auxiliary factors, including an ALIX-binding YXXL domain within p6 Gag (60). A minimal amino-terminal L domain of Tsg101 functions as a DN inhibitor of PTAP-mediated viral budding without inhibiting Tsg101-independent PPXY- or YXXL-dependent pathways (40). The murine leukemia virus PPXY domain recruits a subset of Nedd4-like HECT E3 ubiquitin ligases (WWP1, WWP2, and Itch) (36) that in turn recruit ESCRT-III components (35). The YXXL L domain binds to the cellular protein ALIX (60). ALIX binds to Tsg101 (38) and also with ESCRT-III protein CHMP-4B (60), thus linking ESCRT-I and ESCRT-III. Green fluorescent protein (GFP)-, red fluorescent protein, or yellow fluorescent protein (YFP)-fused CHMPs are general DN inhibitors of all natural CHMP-associated activities and cause the formation of aberrant endosomal compartments that sequester ESCRT complexes (26, 31, 60). Through the use of these DN constructs, the recruitment and assembly of ESCRT components can be inhibited to specifically disrupt different steps of the ESCRT pathway.The best evidence supporting involvement of ESCRT machinery in the life cycle of herpesviruses comes from the inhibition of HSV-1 envelopment by Vps4DN (10), as well as by CHMP3DN (5), together with the association of HSV-1 maturation with MVB. It was recently reported that HHV-6 also induces MVB formation that controls viral egress via an exosomal release pathway (45). After losing primary envelope acquired at the nuclear membrane, Human CMV (HCMV) undergoes a secondary, or final, envelopment step within a cytoplasmic assembly compartments (AC) (59). Secondary envelopment is thought to occur within early endosomal compartments based on diverse observations: (i) purified virions and dense bodies have a lipid composition that is similar to this compartment (64); (ii) the AC of HCMV-infected fibroblasts contain endosomal markers (11); and (iii) a number of HCMV envelope proteins, including US28 (14), UL33, US27 (15), and gB (9), colocalize with endosomal markers in infected cells. A model of HCMV egress via early endosomes has been proposed (11).The approach that we have used here employed human foreskin fibroblasts (HFs) and restricted viral replication to cells that expressed the DN or wild-type (WT) component of the ESCRT pathway by including a requirement that transfected cells complement replication of virus. Confirming expression of both DN and complementing protein in transfected cells by epifluorescence microscopy ensured that an overwhelming majority of cells coexpressed these proteins. The results were scored as inhibition of viral spread to adjacent cells as well as demonstration of late gene expression in the transfected and/or infected cell. Viral progeny is released within 48 to 72 h from CMV-infected cells (44), reducing the likelihood that nonspecific or long-term toxicity of DN-ESCRT proteins would impact our analysis. This assay has been effectively used earlier for both immediate-early gene (54) and late gene (2, 62) mutants, and similar complementation assay results have been reported in diverse systems (8, 49, 73). This assay further provided an opportunity to determine when inhibition occurred relative to the viral replication cycle. Our data implicate ESCRT machinery late during HCMV maturation, which is consistent with a role in secondary envelopment and release.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号