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 985 毫秒
1.
All herpesviruses have a layer of protein called the tegument that lies between the virion membrane and the capsid. The tegument consists of multiple, virus-encoded protein species that together can account for nearly half the total virus protein. To clarify the structure of the tegument and its attachment to the capsid, we used electron microscopy and protein analysis to examine the tegument of herpes simplex virus type 1 (HSV-1). Electron microscopic examination of intact virions revealed that whereas the tegument was asymmetrically distributed around the capsid in extracellular virions, it was symmetrically arranged in cell-associated virus. Examination of virions after treatment with nonionic detergent demonstrated that: (i) in extracellular virus the tegument was resistant to removal with Triton X-100 (TX-100), whereas it was lost nearly completely when cell-associated virus was treated in the same way; (ii) the tegument in TX-100-treated extracellular virions was asymmetrically distributed around the capsid as it is in unextracted virus; and (iii) in some images, tegument was seen to be linked to the capsid by short, regularly spaced connectors. Further analysis was carried out with extracellular virus harvested from cells at different times after infection. It was observed that while the amount of tegument present in virions was not affected by time of harvest, the amount remaining after TX-100 treatment increased markedly as the time of harvest was increased from 24 h to 64 h postinfection. The results support the view that HSV-1 virions undergo a time-dependent change in which the tegument is transformed from a state in which it is symmetrically organized around the capsid and extractable with TX-100 to a state where it is asymmetrically arranged and resistant to extraction.All herpesviruses have a tegument, a layer of protein located between the virus membrane and the capsid. Depending on the virus species, the tegument can be 20 to 40 nm in thickness, and it may be uniformly or asymmetrically distributed about the capsid (7, 17, 24, 33). The tegument is composed predominantly of virus-encoded proteins that together can account for up to half or more of the total virion protein mass. Tegument proteins are thought to be those involved in the early stages of infection before progeny virus proteins are synthesized.The tegument has been most thoroughly studied in herpes simplex virus type 1 (HSV-1). Examination of virions by electron microscopy has demonstrated that the tegument is not highly structured. Its morphology is described as predominantly granular with fibrous elements also present (7, 19). Analysis by cryo-electron microscopy, followed by icosahedral reconstruction has shown that the tegument is not icosahedrally ordered, although a small amount of tegument density is observed close to the capsid surface at the pentons (3, 47).The HSV-1 tegument is composed of approximately 20 distinct, virus-encoded protein species whose amounts vary considerably. The predominant components are UL47, UL48, and UL49, each of which occurs in more than 800 copies per virion (8, 46). In contrast, others, such as RL2 (ICP0), RS1 (ICP4), UL36, and UL37, occur in ∼100 copies or less. Trace amounts of host cell-encoded proteins are also present (15). Many of the tegument proteins are required for virus replication (34), and functions have been defined for most (9, 12, 31, 40).Biochemical studies have demonstrated that the tegument makes noncovalent contacts with both the virus capsid and the membrane. Studies of capsid-tegument contacts have emphasized binding of UL36, a tegument protein, to UL25, a capsid protein located near the vertices and involved in DNA encapsidation (5, 20, 29). Other tegument proteins such as UL48 (VP16), UL37, and UL49 (VP22) are found to associate with UL36 and may be bound to the capsid indirectly by way of UL36 (13, 44). UL16 binds reversibly to the capsid while UL46 (VP11/12) has been shown to bind to both the membrane and the capsid (21, 22, 26). Binding of tegument proteins to the membrane has been shown to occur by way of attachment to UL11 (45) and also to the internal domains of membrane glycoproteins, including glycoprotein D (gD), gH, and gE (4, 6, 11).We describe here the results of a study in which electron microscopy and protein analysis were used to clarify the structure of the HSV-1 tegument and its attachment to the capsid. The study was designed to extend the observation that most of the HSV-1 tegument remains attached to the capsid when the membrane is removed from the virus by treatment with nonionic detergent (19). Cell-associated and extracellular virions were compared after treatment with Triton X-100 (TX-100).  相似文献   

2.
3.
Like other RNA viruses, coxsackievirus B5 (CVB5) exists as circulating heterogeneous populations of genetic variants. In this study, we present the reconstruction and characterization of a probable ancestral virion of CVB5. Phylogenetic analyses based on capsid protein-encoding regions (the VP1 gene of 41 clinical isolates and the entire P1 region of eight clinical isolates) of CVB5 revealed two major cocirculating lineages. Ancestral capsid sequences were inferred from sequences of these contemporary CVB5 isolates by using maximum likelihood methods. By using Bayesian phylodynamic analysis, the inferred VP1 ancestral sequence dated back to 1854 (1807 to 1898). In order to study the properties of the putative ancestral capsid, the entire ancestral P1 sequence was synthesized de novo and inserted into the replicative backbone of an infectious CVB5 cDNA clone. Characterization of the recombinant virus in cell culture showed that fully functional infectious virus particles were assembled and that these viruses displayed properties similar to those of modern isolates in terms of receptor preferences, plaque phenotypes, growth characteristics, and cell tropism. This is the first report describing the resurrection and characterization of a picornavirus with a putative ancestral capsid. Our approach, including a phylogenetics-based reconstruction of viral predecessors, could serve as a starting point for experimental studies of viral evolution and might also provide an alternative strategy for the development of vaccines.The group B coxsackieviruses (CVBs) (serotypes 1 to 6) were discovered in the 1950s in a search for new poliovirus-like viruses (33, 61). Infections caused by CVBs are often asymptomatic but may occasionally result in severe diseases of the heart, pancreas, and central nervous system (99). CVBs are small icosahedral RNA viruses belonging to the Human enterovirus B (HEV-B) species within the family Picornaviridae (89). In the positive single-stranded RNA genome, the capsid proteins VP1 to VP4 are encoded within the P1 region, whereas the nonstructural proteins required for virus replication are encoded within the P2 and P3 regions (4). The 30-nm capsid has an icosahedral symmetry and consists of 60 copies of each of the four structural proteins. The VP1, VP2, and VP3 proteins are surface exposed, whereas the VP4 protein lines the interior of the virus capsid (82). The coxsackievirus and adenovirus receptor (CAR), a cell adhesion molecule of the immunoglobulin superfamily, serves as the major cell surface attachment molecule for all six serotypes of CVB (5, 6, 39, 60, 98). Some strains of CVB1, CVB3 and CVB5 also interact with the decay-accelerating factor (DAF) (CD55), a member of the family of proteins that regulate the complement cascade. However, the attachment of CVBs to DAF alone does not permit the infection of cells (6, 7, 59, 85).Picornaviruses exist as genetically highly diverse populations within their hosts, referred to as quasispecies (20, 57). This genetic plasticity enables these viruses to adapt rapidly to new environments, but at the same time, it may compromise the structural integrity and enzymatic functionality of the virus. The selective constraints imposed on the picornavirus genome are reflected in the different regions used for different types of evolutionary studies. The highly conserved RNA-dependent RNA polymerase (3Dpol) gene is used to establish phylogenetic relationships between more-distantly related viruses (e.g., viruses belonging to different genera) (38), whereas the variable genomic sequence encoding the VP1 protein is used for the classification of serotypes (13, 14, 69, 71, 72).In 1963, Pauling and Zuckerkandl proposed that comparative analyses of contemporary protein sequences can be used to predict the sequences of their ancient predecessors (73). Experimental reconstruction of ancestral character states has been applied to evolutionary studies of several different proteins, e.g., galectins (49), G protein-coupled receptors (52), alcohol dehydrogenases (95), rhodopsins (15), ribonucleases (46, 88, 110), elongation factors (32), steroid receptors (10, 96, 97), and transposons (1, 45, 87). In the field of virology, reconstructed ancestral or consensus protein sequences have been used in attempts to develop vaccine candidates for human immunodeficiency virus type 1 (21, 51, 66, 81) but rarely to examine general phenotypic properties.In this study, a CVB5 virus with a probable ancestral virion (CVB5-P1anc) was constructed and characterized. We first analyzed in detail the evolutionary relationships between structural genes of modern CVB5 isolates and inferred a time scale for their evolutionary history. An ancestral virion sequence was subsequently inferred by using a maximum likelihood (ML) method. This sequence was then synthesized de novo, cloned into a replicative backbone of an infectious CVB5 cDNA clone, and transfected into HeLa cells. The hypothetical CVB5-P1anc assembled into functional virus particles that displayed phenotypic properties similar to those of contemporary clinical isolates. This is the first report describing the reconstruction and characterization of a fully functional picornavirus with a putative ancestral capsid.  相似文献   

4.
The tegument of all herpesviruses contains a high-molecular-weight protein homologous to herpes simplex virus (HSV) UL36. This large (3,164 amino acids), essential, and multifunctional polypeptide is located on the capsid surface and present at 100 to 150 copies per virion. We have been testing the idea that UL36 is important for the structural organization of the tegument. UL36 is proposed to bind directly to the capsid with other tegument proteins bound indirectly by way of UL36. Here we report the results of studies carried out with HSV type 1-derived structures containing the capsid but lacking a membrane and depleted of all tegument proteins except UL36 and a second high-molecular-weight protein, UL37. Electron microscopic analysis demonstrated that, compared to capsids lacking a tegument, these capsids (called T36 capsids) had tufts of protein located at the vertices. Projecting from the tufts were thin, variably curved strands with lengths (15 to 70 nm) in some cases sufficient to extend across the entire thickness of the tegument (∼50 nm). Strands were sensitive to removal from the capsid by brief sonication, which also removed UL36 and UL37. The findings are interpreted to indicate that UL36 and UL37 are the components of the tufts and of the thin strands that extend from them. The strand lengths support the view that they could serve as organizing features for the tegument, as they have the potential to reach all parts of the tegument. The variably curved structure of the strands suggests they may be flexible, a property that could contribute to the deformable nature of the tegument.All herpesviruses have a tegument, a layer of protein located between the virus capsid and membrane. The tegument accounts for a substantial proportion of the overall virus structure. Its thickness (30 to 50 nm), for example, may be comparable to the capsid radius, and tegument proteins can account for 40% or more of the total virion protein. Herpesvirus tegument proteins are thought to function promptly after initiation of infection, before expression of virus genes can take place (11, 13, 14, 21, 33, 37).Electron microscopic analysis of virions has demonstrated that the tegument is not highly structured (9, 22). It does not have icosahedral symmetry like the capsid, and it may be uniformly or asymmetrically arranged around the capsid (26). Tegument structure is described as fibrous or granular, and its morphology is found to change as the virus matures. Studies with herpes simplex virus type 1 (HSV-1), for example, indicate that the tegument structure is altered in cell-associated compared to extracellular virus (26).The tegument has been most thoroughly studied in HSV-1, where biochemical analyses indicate that it is composed of approximately 20 distinct, virus-encoded protein species. The predominant components are the products of the genes UL47, UL48, and UL49, with each protein present in 800 or more copies per virion (12, 40). Other tegument proteins can occur in 100 or fewer copies, and trace amounts of cell-encoded proteins are also present (17). Tegument proteins are classified as inner or outer components based on their association with the capsid after it enters the host cell cytoplasm. The inner tegument proteins (UL36, UL37, and US3) are those that remain bound to the capsid after entry, while the others (the outer tegument proteins) become detached (7, 18).The HSV-1 UL36 protein has the potential to play a central role in organizing the overall structure of the tegument. With a length of 3,164 amino acids, UL36 could span the thickness of the tegument multiple times. One hundred to 150 UL36 molecules are present in the tegument (12), and they are bound to the capsid by way of an essential C-terminal domain (2, 16). UL36 is able to bind the major tegument components by way of documented direct (UL37 and UL48) and indirect (UL46, UL47, and UL49) contacts (6, 15, 24, 38).Here we describe the results of studies designed to test the idea that UL36 serves to organize the tegument structure. Beginning with infectious virus, a novel method has been used to isolate capsids that contain UL36 and UL37 but lack the virus membrane and are depleted of all other tegument proteins. These capsids (T36 capsids) were examined by electron microscopy to clarify the structure of UL36 and UL37 molecules and their location on the capsid surface.  相似文献   

5.
6.
Incorporation of the herpes simplex virus 1 (HSV-1) portal vertex into the capsid requires interaction with a 12-amino-acid hydrophobic domain within capsid scaffold proteins. The goal of this work was to identify domains and residues in the UL6-encoded portal protein pUL6 critical to the interaction with scaffold proteins. We show that whereas the wild-type portal and scaffold proteins readily coimmunoprecipitated with one another in the absence of other viral proteins, truncation beyond the first 18 or last 36 amino acids of the portal protein precluded this coimmunoprecipitation. The coimmunoprecipitation was also precluded by mutation of conserved tryptophan (W) residues to alanine (A) at positions 27, 90, 127, 163, 241, 262, 532, and 596 of UL6. All of these W-to-A mutations precluded the rescue of a viral deletion mutant lacking UL6, except W163A, which supported replication poorly, and W596A, which fully rescued replication. A recombinant virus bearing the W596A mutation replicated and packaged DNA normally, and scaffold proteins readily coimmunoprecipitated with portal protein from lysates of infected cells. Thus, viral functions compensated for the W596A mutation''s detrimental effects on the portal-scaffold interaction seen during transient expression of portal and scaffold proteins. In contrast, the W27A mutation precluded portal-scaffold interactions in infected cell lysates, reduced the solubility of pUL6, decreased incorporation of the portal into capsids, and abrogated viral-DNA cleavage and packaging.Immature herpesvirus capsids or procapsids consist of two shells: an inner shell, or scaffold, and an outer shell that is roughly spherical and largely composed of the major capsid protein VP5 (24, 38).The capsid scaffold consists of a mixture of the UL26.5 and UL26 gene products, with the UL26.5 gene product (pUL26.5, ICP35, or VP22a) being the most abundant (1, 12, 20, 21, 32, 38). The UL26.5 open reading frame shares its coding frame and C terminus with the UL26 gene but initiates at codon 307 of UL26 (17). The extreme C termini of both VP22a and the UL26-encoded protein (pUL26) interact with the N terminus of VP5 (7, 14, 26, 40, 41). Capsid assembly likely initiates when the portal binds VP5/VP22a and/or VP5/pUL26 complexes (22, 25). The addition of more of these complexes to growing capsid shells eventually produces a closed sphere bearing a single portal. pUL26 within the scaffold contains a protease that cleaves itself between amino acids 247 and 248, separating pUL26 into an N-terminal protease domain called VP24 and a C-terminal domain termed VP21 (4, 5, 8, 9, 28, 42). The protease also cleaves 25 amino acids from pUL26 and VP22a to release VP5 (5, 8, 9). VP21 and VP22a are replaced with DNA when the DNA is packaged (12, 29).When capsids undergo maturation, the outer protein shell angularizes to become icosahedral (13). One fivefold-symmetrical vertex in the angularized outer capsid shell is biochemically distinct from the other 11 and is called the portal vertex because it serves as the channel through which DNA is inserted as it is packaged (23). In herpes simplex virus (HSV), the portal vertex is composed of 12 copies of the portal protein encoded by UL6 (2, 23, 39). We and others have shown that interactions between scaffold and portal proteins are critical for incorporation of the portal into the capsid (15, 33, 44, 45). Twelve amino acids of scaffold proteins are sufficient to interact with the portal protein, and tyrosine and proline resides within this domain are critical for the interaction with scaffold proteins and incorporation of the portal into capsids (45).One goal of the current study was to map domains and residues within the UL6-encoded portal protein that mediate interaction with scaffold proteins. We show that the portal-scaffold interaction requires all but the first 18 and last 36 amino acids of pUL6, as well as several tryptophan residues positioned throughout the portal protein.  相似文献   

7.
The African swine fever virus (ASFV) protein pE248R, encoded by the gene E248R, is a late structural component of the virus particle. The protein contains intramolecular disulfide bonds and has been previously identified as a substrate of the ASFV-encoded redox system. Its amino acid sequence contains a putative myristoylation site and a hydrophobic transmembrane region near its carboxy terminus. We show here that the protein pE248R is myristoylated during infection and associates with the membrane fraction in infected cells, behaving as an integral membrane protein. Furthermore, the protein localizes at the inner envelope of the virus particles in the cytoplasmic factories. The function of the protein pE248R in ASFV replication was investigated by using a recombinant virus that inducibly expresses the gene E248R. Under repressive conditions, the ASFV polyproteins pp220 and pp62 are normally processed and virus particles with morphology indistinguishable from that of those produced in a wild-type infection or under permissive conditions are generated. Moreover, the mutant virus particles can exit the cell as does the parental virus. However, the infectivity of the pE248R-deficient virions was reduced at least 100-fold. An investigation of the defect of the mutant virus indicated that neither virus binding nor internalization was affected by the absence of the protein pE248R, but a cytopathic effect was not induced and early and late gene expression was impaired, indicating that the protein is required for some early postentry event.African swine fever virus (ASFV) is a large enveloped deoxyvirus that causes a severe hemorrhagic disease in domestic pigs (38). The ASFV genome is a double-stranded DNA molecule of 170 to 190 kbp that encodes more than 150 polypeptides (47). The icosahedral virus particle contains more than 50 polypeptides and is composed of several concentric domains, including an internal DNA-containing nucleoid surrounded by a protein layer designated the core shell, an inner envelope, and an outer icosahedral capsid (8, 10, 20). An additional membrane acquired by budding through the plasma membrane envelops the extracellular virion (14).The complex process of virus assembly occurs at specialized cytoplasmic sites, designated viral factories, and is initiated by the recruitment and modification of endoplasmic reticulum (ER) cisternae, which collapse to form the virus inner envelope, where the viral membrane proteins p54 and p17 are localized (8, 16, 21, 32, 37). This model, however, has been recently questioned, and based on data obtained using samples prepared by high-pressure freezing, it has been suggested that the inner envelope of ASFV consists of a single lipid bilayer (28). The icosahedral capsid layer, formed by protein p72, is then progressively assembled on one side of this envelope, while on the other side, the core shell domain, mainly constituted by the processing products of the polyproteins pp220 and pp62, is simultaneously constructed (6, 7, 20, 26). Finally, the viral DNA and nucleoproteins are packaged and condensed to form the nucleoid (15).The functions of several virus proteins in the formation of the different domains of the virus particle have been investigated in recent years. Thus, the structural proteins p72 and pB438L and the nonstructural pB602L protein, described as a chaperone of p72 (22), have been shown to be required for the construction of the icosahedral capsid (24, 25, 26), while the polyprotein pp220 is essential for the formation of the inner core, constituted by the nucleoid and core shell domains (7). It has also been demonstrated that the processing of the polyproteins pp220 and pp62 by the virus-encoded protease is necessary for the assembly of a proper core (5). In addition, it is known that the transmembrane protein p54 is critical for the recruitment of envelope precursors to assembly sites (35), although the mechanisms underlying the conversion of ER cisternae into functional viral envelopes are mostly unknown. Studies of other transmembrane proteins detected as structural components of the virus particle could shed light on this matter. Some of the virion membrane proteins could also play a role in virus entry, as has been described for the proteins p12, identified as a viral attachment protein (11, 19), and p54, also involved in binding of virus to target cells (27).The ASFV protein pE248R is a late structural component of the virus particle (33) that belongs to a class of myristoylated membrane proteins related to vaccinia virus L1 (30), one of the substrates of the pathway for the formation of disulfide bonds encoded by this virus (41). The protein pE248R also contains intramolecular disulfide bridges and has been recently identified as a possible final substrate of the ASFV-encoded redox system (33). In the present study, we investigated the membrane association, the localization in the virion, and the role of the protein pE248R in ASFV replication. Our results indicate that pE248R is a myristoylated integral membrane protein localized at the inner envelope of the virus particle. By using a conditional lethal ASFV mutant, vE248Ri, with an inducible copy of the gene E248R, we showed that the protein pE248R is required for virus infectivity and an early postentry event but not virus assembly.  相似文献   

8.
9.
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   

10.
11.
The UL17 and UL25 proteins (pUL17 and pUL25, respectively) of herpes simplex virus 1 are located at the external surface of capsids and are essential for DNA packaging and DNA retention in the capsid, respectively. The current studies were undertaken to determine whether DNA packaging or capsid assembly affected the pUL17/pUL25 interaction. We found that pUL17 and pUL25 coimmunoprecipitated from cells infected with wild-type virus, whereas the major capsid protein VP5 (encoded by the UL19 gene) did not coimmunoprecipitate with these proteins under stringent conditions. In addition, pUL17 (i) coimmunoprecipitated with pUL25 in the absence of other viral proteins, (ii) coimmunoprecipitated with pUL25 from lysates of infected cells in the presence or absence of VP5, (iii) did not coimmunoprecipitate efficiently with pUL25 in the absence of the triplex protein VP23 (encoded by the UL18 gene), (iv) required pUL25 for proper solubilization and localization within the viral replication compartment, (v) was essential for the sole nuclear localization of pUL25, and (vi) required capsid proteins VP5 and VP23 for nuclear localization and normal levels of immunoreactivity in an indirect immunofluorescence assay. Proper localization of pUL25 in infected cell nuclei required pUL17, pUL32, and the major capsid proteins VP5 and VP23, but not the DNA packaging protein pUL15. The data suggest that VP23 or triplexes augment the pUL17/pUL25 interaction and that VP23 and VP5 induce conformational changes in pUL17 and pUL25, exposing epitopes that are otherwise partially masked in infected cells. These conformational changes can occur in the absence of DNA packaging. The data indicate that the pUL17/pUL25 complex requires multiple viral proteins and functions for proper localization and biochemical behavior in the infected cell.Immature herpes simplex virus (HSV) capsids, like those of all herpesviruses, consist of two protein shells. The outer shell comprises 150 hexons, each composed of six copies of VP5, and 11 pentons, each containing five copies of VP5 (23, 29, 47). One vertex of fivefold symmetry is composed of 12 copies of the protein encoded by the UL6 gene and serves as the portal through which DNA is inserted (22, 39). The pentons and hexons are linked together by 320 triplexes composed of two copies of the UL18 gene product, VP23, and one copy of the UL38 gene product, VP19C (23). Each triplex arrangement has two arms contacting neighboring VP5 subunits (47). The internal shell of the capsid consists primarily of more than 1,200 copies of the scaffold protein ICP35 (VP22a) and a smaller number of protease molecules encoded by the UL26 open reading frame, which self-cleaves to form VP24 and VP21 derived from the amino and carboxyl termini, respectively (11, 12, 19, 25; reviewed in reference 31). The outer shell is virtually identical in the three capsid types found in HSV-infected cells, termed types A, B, and C (5, 6, 7, 29, 43, 48). It is believed that all three are derived from the immature procapsid (21, 38). Type C capsids contain DNA in place of the internal shell, type B capsids contain both shells, and type A capsids consist only of the outer shell (15, 16). Cleavage of viral DNA to produce type C capsids requires not only the portal protein, but all of the major capsid proteins and the products of the UL15, UL17, UL28, UL32, and UL33 genes (2, 4, 10, 18, 26, 28, 35, 46). Only C capsids go on to become infectious virions (27).The outer capsid shell contains minor capsid proteins encoded by the UL25 and UL17 open reading frames (1, 17, 20). These proteins are located on the external surface of the viral capsid (24, 36, 44) and are believed to form a heterodimer arranged as a linear structure, termed the C capsid-specific complex (CCSC), located between pentons and hexons (41). This is consistent with the observation that levels of pUL25 are increased in C capsids as opposed to in B capsids (30). On the other hand, other studies have indicated that at least some UL17 and UL25 proteins (pUL17 and pUL25, respectively) associate with all capsid types, and pUL17 can associate with enveloped light particles, which lack capsid and capsid proteins but contain a number of viral tegument proteins (28, 36, 37). How the UL17 and UL25 proteins attach to capsids is not currently known, although the structure of the CCSC suggests extensive contact with triplexes (41). It is also unclear when pUL17 and pUL25 become incorporated into the capsid during the assembly pathway. Less pUL25 associates with pUL17(−) capsids, suggesting that the two proteins bind capsids either cooperatively or sequentially, although this could also be consequential to the fact that less pUL25 associates with capsids lacking DNA (30, 36).Both pUL25 and pUL17 are necessary for proper nucleocapsid assembly, but their respective deletion generates different phenotypes. Deletion of pUL17 precludes DNA packaging and induces capsid aggregation in the nuclei of infected cells, suggesting a critical early function (28, 34), whereas deletion of pUL25 precludes correct cleavage or retention of full-length cleaved DNA within the capsid (8, 20, 32), thus suggesting a critical function later in the assembly pathway.The current studies were undertaken to determine how pUL17 and pUL25 associate with capsids by studying their interaction and localization in the presence and absence of other capsid proteins.  相似文献   

12.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  相似文献   

13.
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   

14.
15.
16.
The process by which nonenveloped viruses cross cell membranes during host cell entry remains poorly defined; however, common themes are emerging. Here, we use correlated in vivo and in vitro studies to understand the mechanism of Flock House virus (FHV) entry and membrane penetration. We demonstrate that low endocytic pH is required for FHV infection, that exposure to acidic pH promotes FHV-mediated disruption of model membranes (liposomes), and particles exposed to low pH in vitro exhibit increased hydrophobicity. In addition, FHV particles perturbed by heating displayed a marked increase in liposome disruption, indicating that membrane-active regions of the capsid are exposed or released under these conditions. We also provide evidence that autoproteolytic cleavage, to generate the lipophilic γ peptide (4.4 kDa), is required for membrane penetration. Mutant, cleavage-defective particles failed to mediate liposome lysis, regardless of pH or heat treatment, suggesting that these particles are not able to expose or release the requisite membrane-active regions of the capsid, namely, the γ peptides. Based on these results, we propose an updated model for FHV entry in which (i) the virus enters the host cell by endocytosis, (ii) low pH within the endocytic pathway triggers the irreversible exposure or release of γ peptides from the virus particle, and (iii) the exposed/released γ peptides disrupt the endosomal membrane, facilitating translocation of viral RNA into the cytoplasm.Flock House virus (FHV), a nonenveloped, positive-sense RNA virus, has been employed as a model system in several important studies to address a wide range of biological questions (reviewed in reference 55). FHV has been instrumental in understanding virus structure and assembly (17, 19, 45), RNA replication (2, 3, 37), and specific packaging of the genome (33, 44, 53, 54). Studies of FHV infection in Drosophila melanogaster flies have provided valuable information about the antiviral innate immune response in invertebrate hosts (29, 57). FHV is also used in nanotechnology applications as an epitope-presenting platform to develop novel vaccines and medical therapies (31, 48). In this report, we use FHV as a model system to further elucidate the means by which nonenveloped viruses enter host cells and traverse cellular membranes.During cell entry enveloped and nonenveloped viral capsid proteins undergo structural rearrangements that enable the virus to breach the membrane bilayer, ultimately releasing the viral genome or nucleocapsid into the cytoplasm. These entry-related conformational changes have been well characterized for enveloped viruses, which use membrane fusion to cross membrane bilayers (reviewed in reference 59). However, the mechanisms nonenveloped viruses employ to breach cellular membranes are poorly defined. Recently, significant parallels in the mechanisms of cell entry have emerged for a diverse group of nonenveloped viruses. Specifically, programmed capsid disassembly and release of small membrane-interacting peptides appear to be a common theme (reviewed in references 4 and 50).The site of membrane penetration depends upon the route of virus entry into the cell. Viruses can enter host cells via several distinct pathways, including clathrin-mediated endocytosis, caveolae-mediated endocytosis, lipid raft-mediated endocytosis, and macropinocytosis (reviewed in reference 40). The two primary routes of virus entry are clathrin-mediated endocytosis, where viruses encounter an acidic environment, and caveolae-mediated endocytosis, which is pH neutral. Many nonenveloped viruses, including adenovirus (24, 52), parvovirus (6), and reovirus (34, 49), require acidic pH during entry. However, numerous nonenveloped viruses have acid-independent entry mechanisms, including rotavirus (28), polyomavirus (43), simian virus 40 (41, 51), and several members of the picornavirus family (7, 14, 32, 42).Upon reaching the appropriate site of membrane penetration, nonenveloped virus capsid proteins are triggered by cellular factors, such as receptor binding and/or exposure to low pH within endosomes, to undergo conformational changes necessary for membrane interactions. These tightly regulated structural rearrangements may include capsid disassembly, exposure of hydrophobic regions, and/or release of membrane-lytic factors. For example, low pH within endosomes triggers adenovirus capsid disassembly, leading to the release of the membrane lytic protein VI (24, 60). In contrast, poliovirus is activated for membrane penetration by a pH-independent mechanism. Receptor binding triggers the poliovirus capsid to undergo a conformational change, resulting in the exposure of the N terminus of VP1 and the release of VP4 (18, 23), both of which facilitate membrane interactions (20). Notably, even though some viruses, such as reovirus, enter cells via an acidic endocytic pathway, membrane penetration is not acid activated (16), indicating that exposure to low pH and membrane penetration are not always mutual events.The overall simplicity of the FHV capsid, composed of a single gene product, along with the wealth of available high-resolution structural information (reviewed in reference 45) make FHV an ideal candidate for understanding nonenveloped virus entry and infection. FHV, a member of the family Nodaviridae, is a nonenveloped insect virus with a bipartite RNA genome surrounded by an icosahedral protein capsid. The quasi-equivalent T=3 virion (∼300-Å diameter) is initially composed of 180 copies of a single coat precursor protein α (44 kDa). Following capsid assembly the α protein undergoes autocatalytic cleavage to generate two particle-associated cleavage products, a large N-terminal fragment, β (39 kDa), and a small C-terminal fragment, γ (4.4 kDa) (22), creating the infectious virion (46). Mutant FHV particles that do not undergo autocatalytic cleavage, and therefore cannot release the γ peptide, are not infectious (46). It has been hypothesized that these particles are noninfectious because they cannot mediate membrane penetration, but this has never been shown directly.The FHV X-ray structure revealed that the γ peptides were located inside the capsid shell with residues 364 to 385 forming amphipathic helices (19). Subsequent studies showed that the FHV capsid is dynamic, with γ transiently exposed to the exterior of the capsid (11). These findings led to a structure-based model of FHV membrane disruption in which the dynamic γ peptides are reversibly exposed to the surface of the capsid (11), “sampling” the environment until they encounter the appropriate cellular trigger. The virus is then activated to undergo an irreversible conformational change in which the γ helical bundles located at each fivefold axis are externalized and released from the virus particle (17, 19). Upon release, the γ pentameric helical bundles are predicted to insert into and create a local disruption of the membrane bilayer to allow the RNA to enter the cytoplasm (10).While biochemical and structural studies have provided the foundation for a model of FHV cell entry, more rigorous in vivo and in vitro studies are necessary to confirm the ideas put forth in this model. Here, we clarify the route of FHV entry and characterize the tightly regulated events required for FHV membrane penetration. We demonstrate for the first time that low endocytic pH is required for FHV infection, that acidic pH promotes FHV membrane penetration, and that particles exposed to low pH exhibit increased hydrophobicity. In addition, we provide evidence that mutant, cleavage-defective particles are blocked specifically at the membrane penetration step during cell entry. Taken together, these findings offer an experimentally supported model of FHV entry into host cells. In addition, these results add to the accumulating evidence that nonenveloped viruses employ common mechanisms to traverse cellular membranes.  相似文献   

17.
18.
Norovirus GII/4 is a leading cause of acute viral gastroenteritis in humans. We examined here how the GII/4 virus evolves to generate and sustain new epidemics in humans, using 199 near-full-length GII/4 genome sequences and 11 genome segment clones from human stool specimens collected at 19 sites in Japan between May 2006 and February 2009. Phylogenetic studies demonstrated outbreaks of 7 monophyletic GII/4 subtypes, among which a single subtype, termed 2006b, had continually predominated. Phylogenetic-tree, bootscanning-plot, and informative-site analyses revealed that 4 of the 7 GII/4 subtypes were mosaics of recently prevalent GII/4 subtypes and 1 was made up of the GII/4 and GII/12 genotypes. Notably, single putative recombination breakpoints with the highest statistical significance were constantly located around the border of open reading frame 1 (ORF1) and ORF2 (P ≤ 0.000001), suggesting outgrowth of specific recombinant viruses in the outbreaks. The GII/4 subtypes had many unique amino acids at the time of their outbreaks, especially in the N-term, 3A-like, and capsid proteins. Unique amino acids in the capsids were preferentially positioned on the outer surface loops of the protruding P2 domain and more abundant in the dominant subtypes. These findings suggest that intersubtype genome recombination at the ORF1/2 boundary region is a common mechanism that realizes independent and concurrent changes on the virion surface and in viral replication proteins for the persistence of norovirus GII/4 in human populations.Norovirus (NoV) is a nonenveloped RNA virus that belongs to the family Caliciviridae and can cause acute gastroenteritis in humans. The NoV genome is a single-stranded, positive-sense, polyadenylated RNA that encodes three open reading frames, ORF1, ORF2, and ORF3 (68). ORF1 encodes a long polypeptide (∼200 kDa) that is cleaved in the cells by the viral proteinase (3Cpro) into six proteins (4). These proteins function in NoV replication in host cells (19). ORF2 encodes a viral capsid protein, VP1. The capsid gene evolved at a rate of 4.3 × 10−3 nucleotide substitutions/site/year (7), which is comparable to the substitution rates of the envelope and capsid genes of human immunodeficiency virus (30). The capsid protein of NoV consists of a shell (S) and two protruding (P) domains: P1 and P2 (47). The S domain is relatively conserved within the same genetic lineages of NoVs (38) and is responsible for the assembly of VP1 (6). The P1 subdomain is also relatively conserved (38) and has a role in enhancing the stability of virus particles (6). The P2 domain is positioned at the most exposed surface of the virus particle (47) and forms binding clefts for putative infection receptors, such as human histo-blood group antigens (HBGA) (8, 13, 14, 60). The P2 domain also contains epitopes for neutralizing antibodies (27, 33) and is consistently highly variable even within the same genetic lineage of NoVs (38). ORF3 encodes a VP2 protein that is suggested to be a minor structural component of virus particles (18) and to be responsible for the expression and stabilization of VP1 (5).Thus far, the NoVs found in nature are classified into five genogroups (GI to GV) and multiple genotypes on the basis of the phylogeny of capsid sequences (71). Among them, genogroup II genotype 4 (GII/4), which was present in humans in the mid-1970s (7), is now the leading cause of NoV-associated acute gastroenteritis in humans (54). The GII/4 is further subclassifiable into phylogenetically distinct subtypes (32, 38, 53). Notably, the emergence and spread of a new GII/4 subtype with multiple amino acid substitutions on the capsid surface are often associated with greater magnitudes of NoV epidemics (53, 54). In 2006 and 2007, a GII/4 subtype, termed 2006b, prevailed globally over preexisting GII/4 subtypes in association with increased numbers of nonbacterial acute gastroenteritis cases in many countries, including Japan (32, 38, 53). The 2006b subtype has multiple unique amino acid substitutions that occur most preferentially in the protruding subdomain of the capsid, the P2 subdomain (32, 38, 53). Together with information on human population immunity against NoV GII/4 subtypes (12, 32), it has been postulated that the accumulation of P2 mutations gives rise to antigenic drift and plays a key role in new epidemics of NoV GII/4 in humans (32, 38, 53).Genetic recombination is common in RNA viruses (67). In NoV, recombination was first suggested by the phylogenetic analysis of an NoV genome segment clone: a discordant branching order was noted with the trees of the 3Dpol and capsid coding regions (21). Subsequently, many studies have reported the phylogenetic discordance using sequences from various epidemic sites in different study periods (1, 10, 11, 16, 17, 22, 25, 40, 41, 44-46, 49, 51, 57, 63, 64, 66). These results suggest that genome recombination frequently occurs among distinct lineages of NoV variants in vivo. However, the studies were done primarily with direct sequencing data of the short genome portion, and information on the cloned genome segment or full-length genome sequences is very limited (21, 25). Therefore, we lack an overview of the structural and temporal dynamics of viral genomes during NoV epidemics, and it remains unclear whether NoV mosaicism plays a role in these events.To clarify these issues, we collected 199 near-full-length genome sequences of GII/4 from NoV outbreaks over three recent years in Japan, divided them into monophyletic subtypes, analyzed the temporal and geographical distribution of the subtypes, collected phylogenetic evidence for the viral genome mosaicism of the subtypes, identified putative recombination breakpoints in the genomes, and isolated mosaic genome segments from the stool specimens. We also performed computer-assisted sequence and structural analyses with the identified subtypes to address the relationship between the numbers of P2 domain mutations at the times of the outbreaks and the magnitudes of the epidemics. The obtained data suggest that intersubtype genome recombination at the ORF1/2 boundary region is common in the new GII/4 outbreaks and promotes the effective acquisition of mutation sets of heterogeneous capsid surface and viral replication proteins.  相似文献   

19.
20.
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   

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