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 187 毫秒
1.
Adeno-associated virus (AAV) serotypes are being tailored for numerous therapeutic applications, but the parameters governing the subcellular fate of even the most highly characterized serotype, AAV2, remain unclear. To understand how cellular conditions control capsid trafficking, we have tracked the subcellular fate of recombinant AAV2 (rAAV2) vectors using confocal immunofluorescence, three-dimensional infection analysis, and subcellular fractionation. Here we report that a population of rAAV2 virions enters the nucleus and accumulates in the nucleolus after infection, whereas empty capsids are excluded from nuclear entry. Remarkably, after subcellular fractionation, virions accumulating in nucleoli were found to retain infectivity in secondary infections. Proteasome inhibitors known to enhance transduction were found to potentiate nucleolar accumulation. In contrast, hydroxyurea, which also increases transduction, mobilized virions into the nucleoplasm, suggesting that two separate pathways influence vector delivery in the nucleus. Using a small interfering RNA (siRNA) approach, we then evaluated whether nucleolar proteins B23/nucleophosmin and nucleolin, previously shown to interact with AAV2 capsids, affect trafficking and transduction efficiency. Similar to effects observed with proteasome inhibition, siRNA-mediated knockdown of nucleophosmin potentiated nucleolar accumulation and increased transduction 5- to 15-fold. Parallel to effects from hydroxyurea, knockdown of nucleolin mobilized capsids to the nucleoplasm and increased transduction 10- to 30-fold. Moreover, affecting both pathways simultaneously using drug and siRNA combinations was synergistic and increased transduction over 50-fold. Taken together, these results support the hypothesis that rAAV2 virions enter the nucleus intact and can be sequestered in the nucleolus in stable form. Mobilization from the nucleolus to nucleoplasmic sites likely permits uncoating and subsequent gene expression or genome degradation. In summary, with these studies we have refined our understanding of AAV2 trafficking dynamics and have identified cellular parameters that mobilize virions in the nucleus and significantly influence AAV infection.Adeno-associated virus (AAV) is classified as a dependovirus because it requires the presence of a helper virus, such as adenovirus or herpesvirus, in order to enter into a productive lytic cycle (6). Because of its nonpathogenicity and ability to promote sustained, long-term transgene expression in a wide variety of tissues such as the brain, liver, muscle, retina, and vasculature (51), several recombinant AAV (rAAV) serotypes are emerging as attractive vectors for gene therapy. Despite many advances in AAV vector design, barriers such as a preexisting immune response and off-target binding have necessitated administration of high viral titers to achieve efficient transduction (24, 51).Beyond the barriers of the immune response (9, 42) and cell surface targeting (52), researchers are becoming increasingly aware that subcellular processing is a significant barrier to infection (16, 29, 52). Subcellular processing may include conformational changes within the endosome or similar compartments, endosomal escape, nuclear targeting, and uncoating, but the factors that control these events are not well defined. Understanding how cellular conditions affect subcellular processing of virions will lead to improved gene delivery through exploitation of these parameters and promote better vector design.Given that the virion is an icosahedral particle only 25 nm in diameter, rAAV must contain all of the molecular components required to navigate through the subcellular environment in a remarkably small structure. Wild-type AAV is a nonenveloped parvovirus that packages a single-stranded DNA genome of approximately 4.7 kb in length. The viral genome is flanked by two inverted terminal repeats and contains two open reading frames, one that codes for replication proteins and another that codes for capsid proteins. Three capsid proteins (VP1, VP2, and VP3) are encoded in the second overlapping reading frame, each beginning with a different start codon but sharing a common C terminus and stop codon. Capsids are comprised of 60 copies of V1, VP2, and VP3 in a ratio of approximately 1:1:10, respectively (11, 43). During production, AAV capsids are known to assemble at early time points in the nucleolus (64), a subdomain of the nucleus and one of the oldest known cellular structures. Intact capsids have been shown to interact with nucleolar proteins such as nucleolin (NCL) and B23/nucleophosmin (NPM1) in the context of assembly (8, 46), but how these proteins affect infection or vector delivery is currently unknown.Initial cell surface binding of AAV capsids is mediated by expression of glycoprotein receptors and specified by residues in VP3 (45, 58, 59). After binding receptors on the host cell plasma membrane, AAV serotype 2 (AAV2) is endocytosed from the cell surface in a clathrin- and dynamin-dependent process (3, 5, 19). Following endocytosis, many AAV particles accumulate in late endosomes, lysosomes, or other compartments and do not deliver their genome to the nucleus (17). This impediment to gene delivery is exacerbated when particles lack VP1 or contain specific mutations in the unique N terminus of VP1 (23). The N terminus of VP1 is normally folded inside the capsid, harboring a phospholipase domain and putative nuclear localization signals necessary for infection (13, 23, 74). These regions of VP1 are thought to translocate to the capsid exterior during subcellular processing of the virus (10, 35, 57). Even with proper capsid composition, the vast majority of internalized particles remain clearly outside the nuclear membrane, and although recent evidence suggests that successful infection occurs when the capsid uncoats inside the nucleus (57, 61), whether AAV can enter the nucleus as an intact capsid is still vehemently debated.In general, it has proven difficult to discern whether infectious particles truly cross the nuclear membrane, due to the limitations of fluorescence microscopy (5, 67). In an in vitro setting it has been demonstrated that unmodified AAV capsids are capable of entering purified nuclei (28), yet these conditions do not accurately represent what occurs physiologically, since virus directly microinjected into cytoplasm will not enter the nucleus or efficiently transduce the cell (17, 57). In one instance, single-particle tracking of AAV has been used to follow capsids in a live-cell imaging paradigm and has found that they can be quickly and directly transported to the nucleus (54). However, another recent study has parsed confocal images of green fluorescent protein-tagged AAV2 particles during infection and has reported that few if any particles enter the nucleus during infection (38).Although it is unclear whether capsids enter the nucleus intact, it has been well established that nuclear delivery of the genome is highly inefficient and significantly limits transduction. Several studies have identified agents that surmount subcellular barriers to transduction (20, 22, 69). Two of the most well-documented agents known to improve subcellular processing are proteasome inhibitors and hydroxyurea (HU); however, their mechanisms of action remain unknown. Therefore, we set out to determine what effect, if any, these agents had on subcellular trafficking of rAAV2 in the hope of identifying specific cellular parameters that promote efficient transduction.Here we report that rAAV2 capsids accumulate in the nucleolus during infection. Proteasome inhibitors were found to potentiate nucleolar accumulation, while HU reduced nucleolar accumulation and appeared to mobilize capsids to the nucleoplasm. Acting independently, both proteasome inhibitors and HU increased transduction, and together they were cooperative, which suggests that these treatments operate through separate pathways to improve gene delivery. In addition, we found that small interfering RNA (siRNA) knockdown of nucleolar proteins NCL and NPM1 had effects similar to those of proteasome inhibition or HU and increased transduction. Based on our results, we have proposed a model wherein AAV virions initially enter the nucleus intact and can be sequestered in the nucleolus in stable form. Disruption of the nucleolus subsequently mobilizes virions from the nucleolus to nucleoplasmic sites and likely permits uncoating.  相似文献   

2.
Adeno-associated virus (AAV) serotypes differ broadly in transduction efficacies and tissue tropisms and thus hold enormous potential as vectors for human gene therapy. In reality, however, their use in patients is restricted by prevalent anti-AAV immunity or by their inadequate performance in specific targets, exemplified by the AAV type 2 (AAV-2) prototype in the liver. Here, we attempted to merge desirable qualities of multiple natural AAV isolates by an adapted DNA family shuffling technology to create a complex library of hybrid capsids from eight different wild-type viruses. Selection on primary or transformed human hepatocytes yielded pools of hybrids from five of the starting serotypes: 2, 4, 5, 8, and 9. More stringent selection with pooled human antisera (intravenous immunoglobulin [IVIG]) then led to the selection of a single type 2/type 8/type 9 chimera, AAV-DJ, distinguished from its closest natural relative (AAV-2) by 60 capsid amino acids. Recombinant AAV-DJ vectors outperformed eight standard AAV serotypes in culture and greatly surpassed AAV-2 in livers of naïve and IVIG-immunized mice. A heparin binding domain in AAV-DJ was found to limit biodistribution to the liver (and a few other tissues) and to affect vector dose response and antibody neutralization. Moreover, we report the first successful in vivo biopanning of AAV capsids by using a new AAV-DJ-derived viral peptide display library. Two peptides enriched after serial passaging in mouse lungs mediated the retargeting of AAV-DJ vectors to distinct alveolar cells. Our study validates DNA family shuffling and viral peptide display as two powerful and compatible approaches to the molecular evolution of novel AAV vectors for human gene therapy applications.A large number of inherited or acquired diseases remain promising targets for human gene therapy. One vector that has shown outstanding potential thus far in numerous preclinical and clinical evaluations is based on nonpathogenic adeno-associated virus (AAV). A unique asset among various properties that make AAV especially attractive over its competitors, such as adenoviral or lentiviral vectors, is the availability of a vast number of natural isolates which differ significantly in their properties (24). We and others have shown previously that the function of an AAV vector particle is determined mainly by the capsid protein and that viral Rep proteins and genomic packaging elements are largely interchangeable (24, 27, 85). Paradoxically, the ever-increasing repertoire of naturally occurring and synthetically generated AAV capsid sequences (>300 to date) is currently creating a dilemma for the rational selection of the optimal serotype for a given application. The importance of finding the ideal capsid for efficient and safe gene transfer has been exemplified in many preclinical studies, as well as in a clinical trial using the AAV type 2 (AAV-2) prototype in human liver tissue (36, 47). In one previous study, the treatment of patients with severe hemophilia B with recombinant AAV-2 expressing human factor IX (hFIX) resulted in mildly elevated, yet therapeutic, levels of this blood coagulation factor. However, expression was short lived, and the hFIX decline was accompanied by a transient asymptomatic increase of liver transaminases, due to a T-cell immune response against the AAV-2 capsid (47). Also, preexisting neutralizing anti-AAV-2 antibodies (frequent in humans) in these individuals likely inhibited the linear vector dose response previously observed in animals.We and others have suggested previously that the use of novel AAV serotypes, in particular, nonhuman isolates, will help to overcome some of these problems (19, 24, 63). Important examples are AAV-8 and AAV-9, which can transduce mouse liver far better than AAV-2, albeit the difference in dogs or primates is less clear (17, 52, 54, 75). The potential for the complete transduction of liver tissue and perhaps other tissues makes these two non-AAV-2 serotypes also particularly interesting for therapeutic RNA interference (RNAi) (28). We recently demonstrated the feasibility of efficiently and persistently suppressing hepatitis B virus with RNAi from a double-stranded AAV-8 vector (28). On the other hand, a potential drawback of AAV-8 and AAV-9 is their lack of specific tissue tropism (34, 52). The resulting frequent vector dissemination into all organs, including the brain, even from low peripheral doses in mice or monkeys (52, 54) is a particular concern for RNAi therapies in which control over vector biodistribution and the limitation of off-target effects will be imperative for the success of the approach (28).In order to overcome the constraints of wild-type AAV serotypes, numerous groups have recently begun to develop novel strategies to engineer “designer” AAVs tailored for the therapeutic transduction of clinically relevant organs (reviewed in detail in references 9, 12, 35, 41, 51, and 85). Briefly, the variety of strategies can be grouped into indirect or chemical approaches and direct physical modification strategies. In the indirect approaches, specific molecules (e.g., bispecific antibodies [6] or avidin-coupled ligands [4]) are allowed to react with the viral surface (biotinylated in the case of avidin [4]), as well as a cellular receptor, forming a conjugate ideally able to retarget the capsid to a refractory cell type. Yet, numerous pharmacological problems, such as concerns about in vivo complex stability and difficulties in upscaling complex manufacturing, continue to prevent the broad adaptation of these approaches. Alternative, more powerful strategies rely on the direct physical modification of the AAV capsid protein and gene. Early examples include the generation of mosaic AAV capsids via the mixing of helper plasmids carrying capsid genes from distinct serotypes, such as AAV-1 and AAV-2 (30) and pairwise combinations of AAV-1 through AAV-5 (62). Similar mosaics were generated previously via a marker rescue approach, yielding AAV-2/AAV-3 recombinants with unique properties (8). A related strategy is the rational creation of chimeric virions via domain swapping among multiple parental serotypes, involving either entire capsid loops or parts thereof or individual residues. Notable examples include AAV-1/AAV-2 chimeras with improved tropism in muscle tissue (31), with one of these chimeras presently being studied in a phase I clinical trial for the treatment of Duchenne muscular dystrophy (85). Most recently, our own group described a battery of unique chimeras comprising elements from serotypes 2 and 8, which were exploited to identify capsid subdomains responsible for efficient AAV transduction in murine liver tissue in vivo (64).A special type of chimeric capsids are those containing foreign proteins or peptides inserted into various positions of the virion shell. The methods and strategies used are widely diverse, and again, we refer to comprehensive reviews (12, 35, 41). Noteworthy here are approaches to fuse targeting ligands to the N termini of AAV capsid proteins (ideally, VP2 [45, 83]), or more powerful, to insert short peptides (up to 14 amino acids [21], typically 7) into exposed regions of the assembled virion. This strategy is referred to as viral display, in analogy to phage display, and has already been used extensively to retarget AAV-2 virions to a multitude of refractory or hard-to-infect cell types, such as vascular endothelial, smooth muscle, and pancreatic islet cells (43, 55, 77, 81, 82) and various tumor lines (22, 58, 65, 66). It has particularly benefited from comprehensive mutational analyses by various groups (e.g., references 21, 33, 56, and 83) that have resulted in the identification of prominent locations within the AAV-2 capsid tolerating peptide insertion. Most notable is the heparin binding domain (HBD), consisting of a total of four arginine (R) residues and one lysine residue, with R585 and R588 representing the most crucial components (37, 56). Numerous groups have now consistently shown that the insertion of 7-mer peptides into this region not only is frequently well tolerated and efficiently mediates virus retargeting, but also provides the extra benefit that the endogenous AAV-2 tropism can be abolished, thus enhancing target specificity (e.g., reference 21).In addition to identifying sites for vector engineering, some of the mutational AAV studies directly yielded novel capsid variants with potential benefits for clinical use. A remarkable case was a recent study by Lochrie et al. (42) in which a set of 127 AAV-2 variants with point or insertion mutations were generated and screened for multiple properties. Several capsids were isolated which differed from the wild-type AAV-2 capsid in having better in vitro transduction efficiencies (albeit being equally efficient in vivo) or, clinically most relevant, higher-level resistance to individual or pooled human antisera. Nonetheless, the limitations of the approach also became clear, most notably, the extreme effort required to generate and manually screen a large number of mutants, which in fact prevented the interesting analyses of all possible combinations of beneficial point mutations in further capsids.Indeed, the factors of time and labor are the main reasons why an increasing number of groups have recently begun to develop novel means for AAV vector evolution that no longer rely on the rational modification of the AAV-2 capsid. Instead, the new combinatorial methodologies allow for the far more efficient creation and selection of interesting candidates in a library-based high-throughput format. Thus far, two different strategies have been reported, both principally expanding on previously developed techniques. One is the use of viral display libraries, in which random 7-mer peptides are inserted into the AAV-2 HBD (at amino acid 587 or 588), yielding between 4 × 106 and 1.7 × 108 capsids potentially exposing new ligands on their surfaces (50, 58, 76). Subsequent iterative selection on diverse cell types refractory to the wild type, e.g., coronary artery endothelial cells, cardiomyoblasts, and carcinoma, leukemia, and megakaryocytic cell lines, led to enrichment with peptide mutants with increased target specificities and efficacies (48, 50, 58, 76). The second library type, independently described by two groups in 2006, relies on error-prone PCR amplification of the AAV-2 capsid gene (46, 59). Similar to the methods in earlier mutational studies, this approach resulted in the identification of AAV-2 point mutations (usually up to two per capsid) which yielded mutants that differed from the wild type in having mildly enhanced efficacies in vitro and/or improved transduction efficiencies in the presence of neutralizing anti-AAV-2 antibodies either generated in rabbits or preexisting in individual human sera.Here, for the first time, we introduce the technology of DNA family shuffling into the realm of AAV vector evolution. The basic concept of this technology is the in vitro recombination of related parental genes with >50% homology, which are first fragmented and then reassembled based on partial homology, resulting in libraries of chimeric genes. Iterative amplification under pressure can then yield hybrids not only combining parental assets, but also ideally exhibiting novel and synergistic properties (70, 71). DNA family shuffling has been used extensively in recent years to evolve and improve all types of proteins, from markers and enzymes to vaccines (e.g., references 10, 13-15, and 39). Importantly, a set of reports also suggested its power to enhance viral gene therapy vectors by creating retro- or lentiviruses with improved stability or efficacy compared to that of the parental wild types (57, 61, 69). Here, we describe the novel use of DNA family shuffling for the highly efficient molecular interbreeding of eight multispecies AAVs to create chimeric capsids and, moreover, document its compatibility and synergism with existing AAV vector evolution technology.  相似文献   

3.
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   

4.
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   

5.
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   

6.
7.
8.
The acid-dependent disassembly of foot-and-mouth disease virus (FMDV) is required for viral RNA release from endosomes to initiate replication. Although the FMDV capsid disassembles at acid pH, mutants escaping inhibition by NH4Cl of endosomal acidification were found to constitute about 10% of the viruses recovered from BHK-21 cells infected with FMDV C-S8c1. For three of these mutants, the degree of NH4Cl resistance correlated with the sensitivity of the virion to acid-induced inactivation of its infectivity. Capsid sequencing revealed the presence in each of these mutants of a different amino acid substitution (VP3 A123T, VP3 A118V, and VP2 D106G) that affected a highly conserved residue among FMDVs located close to the capsid interpentameric interfaces. These residues may be involved in the modulation of the acid-induced dissociation of the FMDV capsid. The substitution VP3 A118V present in mutant c2 was sufficient to confer full resistance to NH4Cl and concanamycin A (a V-ATPase inhibitor that blocks endosomal acidification) as well as to increase the acid sensitivity of the virion to an extent similar to that exhibited by mutant c2 relative to the sensitivity of the parental virus C-S8c1. In addition, the increased propensity to dissociation into pentameric subunits of virions bearing substitution VP3 A118V indicates that this replacement also facilitates the dissociation of the FMDV capsid.Foot-and-mouth disease virus (FMDV) is a member of the Aphthovirus genus in the family Picornaviridae. FMDV displays epithelial tropism and is responsible for a highly contagious disease of cloven-hoofed animals (23, 60). FMDV populations are quasispecies and exhibit a high potential for variation and adaptation, one consequence of which is the extensive antigenic diversity of this virus, reflected in the existence of seven serotypes and multiple antigenic variants (reviewed in references 17 and 60). Different cellular receptors, including αvβ integrins and heparan sulfate (HS) glycosaminoglycans, have been described for natural isolates and tissue culture-adapted FMDVs (3, 4, 6, 28-31, 56). However, viruses that are infectious in vivo use integrins as receptors (28). The interaction between FMDV and the integrin molecule is mediated by an Arg-Gly-Asp (RGD) triplet located at the G-H loop of capsid protein VP1 (9, 47). FMDV isolates interacting with integrins gain entry into the cell following clathrin-mediated endocytosis (8, 39, 52). On the other hand, it has been described that a genetically engineered HS-binding mutant uses caveolae to enter into cultured cells (51). After internalization, FMDV must release its genomic RNA molecule of positive polarity into the host cell cytoplasm to establish a productive infection. Early work showed that a variety of lysosomotropic agents, such as weak bases and ionophores that block acidification of endosomes, inhibit FMDV infection (5, 11-13), indicating that genome release is dependent on endosomal acidification. In addition, internalized FMDV particles colocalize with markers from early and recycling endosomes (8, 51, 52) and FMDV infection is reduced by expression of a dominant negative mutant of Rab5 (33), suggesting that FMDV may release its genome from these compartments.The FMDV capsid comprises 60 copies of each of the four structural proteins (VP1 to VP4) arranged in an icosahedral lattice of 12 pentameric subunits. FMDV particles are highly acid labile and disassemble at pH values slightly below neutrality (13). Acid lability is not a feature of the capsids of other picornaviruses, such as Enterovirus. Pentameric subunits are intermediates of FMDV assembly and disassembly (64). A high density of His residues is found close to the interpentameric interface. Protonation of these residues at the acidic pH in the endosomes has been proposed to trigger acid-induced capsid disassembly by electrostatic repulsion between the protonated His side chains (1). His 142 (H142) in VP3 of type A FMDV is involved in a His-α-helix dipole interaction, which is likely to influence the acid lability of FMDV (13). In silico predictions suggested that H142 and H145 in VP3 may have the greatest effect on this process (63). Experimental evidence of the involvement of H142 of VP3 in acid-induced disassembly of FMDV has also been reported (20). Concomitantly with capsid disassembly into pentameric intermediates, internal protein VP4 and viral RNA are released. VP4 is a highly hydrophobic and myristoylated protein (7) whose release has been suggested to mediate membrane permeabilization and ion channel formation, thus facilitating the endosomal exit of viral RNA (15, 16, 34).Besides providing information about the endosomal pH requirements for the release of virus genomes, drugs modifying endosomal acidification can reveal the molecular changes associated with viral resistance to their action. These analyses may also address whether the balance between acid lability and capsid stability required for completion of virus replication allows FMDV, which disassembles at a pH close to neutrality, to escape inhibition by drugs raising the endosomal pH. In this work, we have isolated and characterized FMDV mutants that are able to escape from the inhibition of endosomal acidification exerted by NH4Cl, a lysosomotropic weak base that raises endolysosomal pH and impairs uncoating and infection of viruses that require transit through acidic endosomal compartments for penetration (5, 26, 53). These mutants showed an increased acid lability, which is likely to allow them to uncoat at more-alkaline pH values. A single amino acid substitution close to the interpentameric interfaces in the capsid of one of these mutants was responsible for a total resistance to the elevation in endosomal pH caused by NH4Cl treatment and for the acid-labile phenotype.  相似文献   

9.
10.
After recognizing and binding to its host cell, poliovirus (like other nonenveloped viruses) faces the challenge of translocating its genome across a cellular membrane and into the cytoplasm. To avoid entanglement with the capsid, the RNA must exit via a single site on the virion surface. However, the mechanism by which a single site is selected (from among 60 equivalents) is unknown; and until now, even its location on the virion surface has been controversial. To help to elucidate the mechanism of infection, we have used single-particle cryo-electron microscopy and tomography to reconstruct conformationally altered intermediates that are formed by the poliovirion at various stages of the poliovirus infection process. Recently, we reported icosahedrally symmetric structures for two forms of the end-state 80S empty capsid particle. Surprisingly, RNA was frequently visible near the capsid; and in a subset of the virions, RNA was seen on both the inside and outside of the capsid, caught in the act of exiting. To visualize RNA exiting, we have now determined asymmetric reconstructions from that subset, using both single-particle cryo-electron microscopy and cryo-electron tomographic methods, producing independent reconstructions at ∼50-Å resolution. Contrary to predictions in the literature, the footprint of RNA on the capsid surface is located close to a viral 2-fold axis, covering a slot-shaped area of reduced density that is present in both of the symmetrized 80S reconstructions and which extends by about 20 Å away from the 2-fold axis toward each neighboring 5-fold axis.In its role as the intermediate that links one round of infection with the next, a virus particle protects the viral genome during passage from cell to cell and from host to host, it specifically recognizes and binds to target cells, and it delivers the viral genome into the appropriate compartment in the target cell. For enveloped viruses, which have their own external membranes, fusion of the viral membrane with a host membrane presents a conceptually simple mechanism for delivery of the genome or nucleoprotein into the cytoplasm. For nonenveloped viruses, the viral particle must provide the machinery necessary for either the entire virion, a nucleoprotein complex, or the viral genome to cross a membrane. This process remains poorly understood. Poliovirus provides an excellent model system for probing the mechanisms used for genome translocation. As the type member of the Picornavirus family and the etiological agent of poliomyelitis, poliovirus has been well characterized biochemically and genetically (42), its cell entry pathways have been well characterized (5, 15, 30, 52), and a number of cell entry intermediates have been identified and are accessible for structural studies (2-4, 7, 8, 18, 34, 38, 42, 55, 56).The capsid of the mature poliovirion (160S particle) consists of 60 copies of each of the four coat proteins VP1, VP2, VP3, and VP4 (which is myristolated at its amino-terminal glycine [13]) and encloses a 7.5-kbp positive-sense RNA genome. The outer surface of the capsid has a number of major features, including star-shaped mesas at its 5-fold axes, 3-fold propeller-like protrusions, canyon-like depressions surrounding each of the 5-fold mesas, and depressions at the 2-fold axes (30, 31).Poliovirus infection is initiated when the virus binds to the host-cell-surface poliovirus receptor (called Pvr or CD155) (41), triggering a conformational change of the native capsid into an altered particle called the A particle or 135S particle (18, 19). The 135S particle has been shown to be expanded by about 4% (2, 7), is infectious (16, 33), and is believed to be a productive intermediate in viral entry (30, 33). This conformational change results in the externalization of the small myristoylated capsid protein, VP4 (18), and of the amino-terminal extension of VP1 (which includes a conserved amphipathic helix) (23). Both of these externalized polypeptides then associate with membranes (17, 23). In subsequent steps, the viral genome is released from the capsid and translocated across a membrane (probably an endosomal membrane [5]) to gain access to the cytoplasm, leaving behind an end-state empty capsid shell (called the 80S particle). The trigger for RNA release and the mechanism of genome translocation are both poorly understood (30, 52).Electrophysiology and mutational experiments have shown that the externalization of VP4 and of the amino terminus of VP1 is associated with the formation of channels in membranes (17, 49, 50) and, furthermore, that point mutations in threonine 28 of VP4 can either eliminate (T28G) or alter (T28V, T28S) the ability to form channels and either eliminate (T28G) or slow (T28V, T28S) the kinetics of productive RNA release (17). These observations have led to the hypothesis that the viral polypeptides insert into host cell membranes during infection and rearrange to form channels that permit the viral genome to pass through the membrane, thereby gaining access to the cytoplasm (7, 17, 49, 50).Speculation about the sites of externalization of the viral peptides and of the viral genome began soon after the structures of mature rhinovirus and poliovirus were determined crystallographically 25 years ago (31, 44). In both structures there is a solvent-filled channel running along each 5-fold axis. This channel is closed off at the outer surface of the capsid by polypeptide loops and on the inner surface by a plug that is formed by five intertwined copies of the amino terminus of VP3, forming a parallel beta tube (31, 44). In poliovirus this tube is flanked on its inner surface by five copies of a three-stranded beta sheet in which the outermost two strands come from a beta hairpin at the amino terminus of VP4 and the innermost strand comes from residues at the extreme amino terminus of VP1 (20). The presence of this channel, together with its proximity to peptide segments that were known to be externalized upon receptor attachment, and analogies with other viruses led to a model in which both the peptides and the viral RNA are externalized via the channel at the 5-fold axis (25, 45). At that time, an alternative model for the egress of polypeptides was proposed, based on an analogy with the externalization of the amino-terminal extensions of capsid proteins in expanded states of the topologically similar T=3 plant viruses (26, 32, 43, 47) and on genetic and biochemical studies of mutations that affect cell entry and capsid stability in poliovirus (14, 39, 54). In the latter model, the peptides were proposed to exit from the base of the canyon and then proceed along the outer surface toward the 5-fold peak (43, 47). Both models suggested that five copies of each of the externalized peptides would interact in some way to form a pore in the membrane that was contiguous with one of the 5-fold channels, thus providing a way for RNA to be released from the virion at a 5-fold axis of symmetry. No data yet exist to specify what specific structural roles VP4 and the amino terminus of VP1 might play in forming pores and serving as membrane anchors. However, both the electrophysiology data (cited above) and the greater sequence conservation of VP4 suggest that its role in pore formation may be the more central (17, 49, 50).To further elucidate various steps along the infection pathway, cryo-electron microscopy (cryo-EM) reconstructions have been determined for a number of cell entry intermediates of poliovirus and rhinoviruses, and their resolutions have been improved over time (2, 3, 7, 28, 38). Structures of the complexes of polioviruses and major-group rhinoviruses with the ectodomains of their respective receptors have confirmed earlier models that suggested that the canyon is the receptor-binding site and have begun to suggest how receptor binding might lead to receptor-induced conformational rearrangements (3, 56). Cryo-EM and cryo-electron tomography structures (cryo-ET) of a poliovirus-receptor-membrane complex (using a novel receptor-decorated liposome model [51]) confirmed that initial receptor binding brings the surface of the 5-fold mesa into close proximity with the membrane and appears to produce an outward distortion of the outer leaflet of the membrane in its area of closest approach to the virus particle (4, 8).Structures have also been determined for the soluble 135S and 80S particles of poliovirus, formed by heating the virus at 50°C (135S) or 56°C (80S) in hypotonic buffers, and for the 80S particles of rhinovirus 14 and 16, formed by exposing virus to acidic pH. All of the biological and immunological evidence that is currently available indicates that the particles prepared in vitro and used for structural studies are indistinguishable from the particles that are released from the cell surface during infection (6, 53). These structures have allowed the models for peptide release and genome release to be extended and refined (7, 38) and indeed have confirmed that VP1 exits from the particle surface at the base of the canyon and climbs up the side of the 5-fold mesa. However, contrary to the assumptions of the earlier models, the 10-Å structures of the poliovirus 135S and 80S particles show that the amino end of the amino-terminal extension of VP1 does not remain associated with the mesa. Instead, it forms an alpha-helical bridge that stretches across the canyon and binds to the large EF loop of VP2, a surface projection that appears as a 3-fold propeller blade (7, 38).Until recently, the mechanism of RNA release (during the 135S-to-80S transition) has been largely a matter of conjecture. We can infer that the RNA must exit via a single site on the virion surface, to avoid entanglement with the capsid (particularly as entanglement has never been observed in electron micrographs), though the mechanism by which a single site is selected (from among 60 equivalents) is unknown. All models presented to date have assumed that the RNA is released from the channel at the 5-fold axes (2, 3, 7, 8, 25, 27, 28, 30, 42, 45). However, in the icosahedrally constrained 10-Å structures of both the poliovirus 135S and 80S particles (7, 38), the apparent intactness and stability of the 5-fold mesa argues against the 5-fold axis being the site of RNA egress, given that the diameter of the opening, as seen in those structures, would be insufficient to accommodate RNA, even if the “plug” formed by the intertwined amino termini of VP3 was displaced. Moreover, both structures revealed significant thinning between 2-fold-related pentamers in the vicinity of the 2-fold axes. Most convincingly, large holes (easily sufficient to accommodate RNA) were seen at and near the 2-fold axes in the atomic model of the late-80S structure. This coincided with an open hole in the reconstruction, when viewed at a contour level that left most of the remainder of the capsid intact. This evidence was suggestive, but not definitive, as a number of other openings were present, particularly in the interfaces between protomers. Furthermore, the behavior of the capsid structure in the immediate vicinity of the unique site of RNA exit is likely to be different from what we see in the icosahedral average, which is dominated by the remainder of the capsid.In the course of solving icosahedrally symmetric cryo-EM structures for the poliovirus end-state 80S empty capsid particle (7, 38), we were surprised to find that RNA was frequently visible near the capsid and that in a subset of about 5% of the sampled virions, RNA was seen on both the inside and outside of the capsid, apparently caught in the act of exiting. This was an exciting development, as images of viral RNA release had never previously been reported. We were able to improve the resolution to ∼10 Å by classifying the projected images into two groups: an early 80Se particle that was more prevalent in the population after a shorter heating time and a late 80Sl particle that was seen more often when the heating time was increased. The amount of RNA density remaining in the interior appears to be continuously variable in both classes, suggesting that release is gradual. Of the 5% subset of particles clearly caught in the act, almost all belonged to the 80Se class. Our interpretation was that the 80Se class may represent particles in which exiting RNA is still engaged with the capsid machinery and traversing the capsid, while the 80Sl class (in which much of the capsid resembles the 135S form more closely in structure) represents particles with the RNA disengaged, possibly after nuclease cleavage. More than two structural classes may be present, but at the current resolution, we could not distinguish them.The present report addresses the question of what we can learn about the details of RNA release from an asymmetric cryo-EM reconstruction, based on the 540-particle caught-in-the-act subset, and independently from cryo-electron tomographic reconstructions of a similarly prepared sample. In each projected particle image or subtomogram, preliminary orientation parameters are first determined from an icosahedrally symmetric calculation, and in a second stage, the symmetry is broken by choosing 1 of the 60 symmetry-equivalent orientations. Both methods have yielded similar information, at about 50-Å resolution, concerning the footprint of the RNA on the virion surface, which demonstrates that RNA is released from an asymmetric site at the base of the canyon near a particle 2-fold axis and not at the channel at the 5-fold axes, as suggested by previous models. Additionally, the demonstrated success of the methodology provides us with a blueprint for resolving the molecular details of the RNA-capsid interaction in future experiments.  相似文献   

11.
12.
Coxsackievirus B2 (CVB2), one of six human pathogens of the group B coxsackieviruses within the enterovirus genus of Picornaviridae, causes a wide spectrum of human diseases ranging from mild upper respiratory illnesses to myocarditis and meningitis. The CVB2 prototype strain Ohio-1 (CVB2O) was originally isolated from a patient with summer grippe in the 1950s. Later on, CVB2O was adapted to cytolytic replication in rhabdomyosarcoma (RD) cells. Here, we present analyses of the correlation between the adaptive mutations of this RD variant and the cytolytic infection in RD cells. Using reverse genetics, we identified a single amino acid change within the exposed region of the VP1 protein (glutamine to lysine at position 164) as the determinant for the acquired cytolytic trait. Moreover, this cytolytic virus induced apoptosis, including caspase activation and DNA degradation, in RD cells. These findings contribute to our understanding of the host cell adaptation process of CVB2O and provide a valuable tool for further studies of virus-host interactions.Virus infections depend on complex interactions between viral and cellular proteins. Consequently, the nature of these interactions has important implications for viral cell type specificity, tissue tropism, and pathogenesis. Group B coxsackieviruses (CVB1 to CVB6), members of the genus Enterovirus within the family of Picornaviridae, are human pathogens that cause a broad spectrum of diseases, ranging from mild upper respiratory illnesses to more severe infections of the central nervous system, heart, and pancreas (61). These viruses have also been associated with certain chronic muscle diseases and myocardial infarction (2, 3, 12, 13, 22).The positive single-stranded RNA genome (approximately 7,500 nucleotides in length) of CVBs is encapsidated within a small T=1, icosahedral shell (30 nm in diameter) comprised of repeating identical subunits made up of four structural proteins (VP1 to VP4). Parts of VP1, VP2, and VP3 are exposed on the outer surface of the capsid, whereas VP4 is positioned on the interior. The virion morphology is characterized by a star-shaped mesa at each 5-fold icosahedral symmetry axis, surrounded by a narrow depression referred to as the “canyon” (69). All six serotypes of CVB can use the coxsackie and adenovirus receptor (CAR) for cell attachment and entry (9, 55, 82). Some strains of CVB1, -3, and -5 also use decay accelerating factor ([DAF] CD55) for initial attachment to the host cell; however, binding to DAF alone is insufficient to permit entry into the cell (10, 54, 76).Picornaviruses are generally characterized by their cytolytic nature in cell culture. However, several in vivo and in vitro studies have shown that some picornaviruses, e.g., poliovirus, Theiler''s murine encephalomyelitis virus, foot-and-mouth disease virus, CVB3, CVB4, and CVB5, may also establish persistent, noncytolytic infections (4, 29, 35, 39, 62, 74). Recently, it has been shown that the diverse outcomes of picornaviral infections may depend on interactions between the virus and the apoptotic machinery of the infected cell (14, 30, 71). Several picornaviral proteins have been identified as inducers of an apoptotic response, including viral capsid proteins VP1, VP2, and VP3, as well as nonstructural proteins 2A and 3C (7, 20, 32, 33, 42, 50, 63). In addition, antiapoptotic activity has been assigned to the nonstructural proteins 2B and 3A (16, 59).Picornaviruses have the potential to adapt rapidly to new host environments. Virus features affecting adaptability include high mutation rates, short replication times, large populations, and frequent incidences of recombination (25-27, 53). Consequently, picornaviruses exist as genetically heterogenous populations, referred to as viral quasispecies (25, 26).Previously, the CVB2 prototype strain Ohio-1 (CVB2O) was adapted to cytolytic replication in rhabdomyosarcoma (RD) cells (66). Two amino acid changes were identified in the capsid-coding region, and one was identified in the 2C-coding region of the adapted virus. Further characterization of the virus-host interaction showed that the infection was not affected by anti-DAF antibodies, indicating the use of an alternative receptor.In this study, the amino acid substitutions associated with the adaptation of CVB2O to cytolytic infection of RD cells were evaluated. Site-directed mutagenesis studies showed that a single amino acid change in the VP1 capsid protein was responsible for the cytolytic RD phenotype. In addition, as indicated by caspase activation and DNA degradation, the apoptotic pathway was activated in RD cells infected by the cytolytic virus.  相似文献   

13.
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   

14.
15.
16.
17.
Adeno-associated virus (AAV) type 2 and 5 proteins Rep52 and Rep40 were polyubiquitinated during AAV-adenovirus type 5 (Ad5) coinfection and during transient transfection in either the presence or absence of Ad5 E4orf6 and E1b-55k. Polyubiquitination of small Rep proteins via lysine 48 (K48) linkages, normally associated with targeting of proteins for proteasomal degradation, was detected only in the presence of E4orf6. The small Rep proteins were ubiquitinated via lysine 63 (K63) following transfection in either the presence or absence of E4orf6 or following coinfection with Ad5. E4orf6/E1b-55k-dependent K48-specific polyubiquitination of small Rep proteins could be inhibited using small interfering RNA (siRNA) to cullin 5.Together, adenovirus type 5 (Ad5) early gene products E1a, E1b-55k, E2a, E4orf6, and virus-associated (VA) RNA can support efficient replication of adeno-associated virus (AAV) (4, 31). E4orf6 and E1b-55k are known to interact with cellular cullin 5 (cul5), elongins B and C, and the ring box protein Rbx1 to form an E3 ubiquitin ligase complex that specifically targets a small population of cellular proteins for degradation by the proteasome (1, 7, 21, 22, 24, 27). This property has been implicated in a number of functions presumed to be required for both Ad and AAV replication (3, 8-10, 17, 23, 24, 34, 35).Previously, only p53, Mre11, DNA ligase IV, and integrin α3 had been shown to be substrates of the Ad5 E3 ubiquitin ligase complex (1, 7, 21, 22, 24, 27); however, we have recently shown (16, 17) that the small Rep proteins and capsid proteins of AAV5 are also degraded in the presence of Ad E4orf6 and E1b-55k in a proteasome-dependent manner. These proteins were restored to levels required during infection by the action of VA RNA (17). The targeting for degradation of AAV5 protein by the E4orf6/E1b-55k E3 ubiquitin ligase complex required functional BC-box motifs in E4orf6 and could be inhibited by depletion of the scaffolding protein cullin 5 using directed small interfering RNA (siRNA) (16). In addition, the degradation of AAV5 protein was partially prevented by overexpression of pUBR7, a plasmid that generates a dominant-negative ubiquitin (16). The role this targeted degradation plays in the life cycle of AAV has not yet been clarified; however, E4orf6 mutants that cannot function in this regard do not support AAV replication as well as wild-type E4orf6 (R. Nayak and D. J. Pintel, unpublished data). Degradation of Mre11 by the Ad5 E3 ligase has also been implicated in allowing efficient Ad5 and AAV replication (24). Ubiquitination of AAV Rep proteins during viral infection, however, has not previously been reported.  相似文献   

18.
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   

19.
The parvovirus adeno-associated virus (AAV) contains a small single-stranded DNA genome with inverted terminal repeats that form hairpin structures. In order to propagate, AAV relies on the cellular replication machinery together with functions supplied by coinfecting helper viruses such as adenovirus (Ad). Here, we examined the host cell response to AAV replication in the context of Ad or Ad helper proteins. We show that AAV and Ad coinfection activates a DNA damage response (DDR) that is distinct from that seen during Ad or AAV infection alone. The DDR was also triggered when AAV replicated in the presence of minimal Ad helper proteins. We detected autophosphorylation of the kinases ataxia telangiectasia mutated (ATM) and DNA-dependent protein kinase catalytic subunit (DNA-PKcs) and signaling to downstream targets SMC1, Chk1, Chk2, H2AX, and XRCC4 and multiple sites on RPA32. The Mre11 complex was not required for activation of the DDR to AAV infection. Additionally, we found that DNA-PKcs was the primary mediator of damage signaling in response to AAV replication. Immunofluorescence revealed that some activated damage proteins were found in a pan-nuclear pattern (phosphorylated ATM, SMC1, and H2AX), while others such as DNA-PK components (DNA-PKcs, Ku70, and Ku86) and RPA32 accumulated at AAV replication centers. Although expression of the large viral Rep proteins contributed to some damage signaling, we observed that the full response required replication of the AAV genome. Our results demonstrate that AAV replication in the presence of Ad helper functions elicits a unique damage response controlled by DNA-PK.Replication of viral genomes produces a large amount of extrachromosomal DNA that may be recognized by the cellular DNA damage machinery. This is often accompanied by activation of DNA damage response (DDR) signaling pathways and recruitment of cellular repair proteins to sites of viral replication. Viruses therefore provide good model systems to study the recognition and response to DNA damage (reviewed in reference 48). The Mre11/Rad50/Nbs1 (MRN) complex functions as a sensor of chromosomal DNA double-strand breaks (DSBs) and is involved in activation of damage signaling (reviewed in reference 41). The MRN complex also localizes to DNA DSBs and is found at viral replication compartments during infection with a number of DNA viruses (6, 40, 47, 70, 75, 77, 87, 93). The phosphatidylinositol 3-kinase-like kinases (PIKKs) ataxia telangiectasia mutated (ATM), ATM and Rad3-related kinase (ATR), and the catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) are involved in the signal transduction cascades activated by DNA damage (reviewed in references 43, 51, and 71). These kinases respond to distinct types of damage and regulate DSB repair during different phases of the cell cycle (5), either through nonhomologous end-joining (NHEJ) or homologous recombination pathways (reviewed in references 63, 81, and 86). The DNA-PK holoenzyme is composed of DNA-PKcs and two regulatory subunits, the Ku70 and Ku86 heterodimer. DNA-PK functions with XRCC4/DNA ligase IV to repair breaks during NHEJ, and works with Artemis to process DNA hairpin structures during VDJ recombination and during a subset of DNA DSB events (46, 50, 86). While the kinase activity of DNA-PKcs leads to phosphorylation of a large number of substrates in vitro as well as autophosphorylation of specific residues (reviewed in references 16 and 85), it is currently unclear how DNA-PKcs contributes to signaling in cells upon different types of damage.The adeno-associated virus (AAV) genome consists of a molecule of single-stranded DNA with inverted terminal repeats (ITRs) at both ends that form double-hairpin structures due to their palindromic sequences (reviewed in reference 52). The ITRs are important for replication and packaging of the viral genome and for integration into the host genome. Four viral Rep proteins (Rep78, Rep68, Rep52, and Rep40) are also required for replication and packaging of the AAV genome into virions assembled from the Cap proteins. Although the Rep and Cap genes are replaced in recombinant AAV vectors (rAAV) that retain only the ITRs flanking the gene of interest, these vectors can be replicated by providing Rep in trans (reviewed in reference 7). Productive AAV infection requires helper functions supplied by adenovirus (Ad) or other viruses such as herpes simplex virus (HSV) (reviewed in reference 27), together with components of the host cell DNA replication machinery (54, 55, 58). In the presence of helper viruses or minimal helper proteins from Ad or HSV, AAV replicates in the nucleus at centers where the viral DNA and Rep proteins accumulate (35, 76, 84, 89). Cellular and viral proteins involved in AAV replication, including replication protein A (RPA), Ad DNA-binding protein (DBP), and HSV ICP8, localize with Rep proteins at these viral centers (29, 33, 76).A number of published reports suggest associations between AAV and the cellular DNA damage machinery. For example, transduction by rAAV vectors is increased by genotoxic agents and DNA damaging treatments (1, 62, 91) although the mechanisms involved remain unclear. Additionally, the ATM kinase negatively regulates rAAV transduction (64, 92), and we have shown that the MRN complex poses a barrier to both rAAV transduction and wild-type AAV replication (11, 67). UV-inactivated AAV particles also appear to activate a DDR involving ATM and ATR kinases that perturbs cell cycle progression (39, 60, 88). It has been suggested that this response is provoked by the AAV ITRs (60) and that UV-treated particles mimic stalled replication forks in infected cells (39). In addition to AAV genome components, the viral Rep proteins have been observed to exhibit cytotoxicity and induce S-phase arrest (3, 65).The role of cellular repair proteins in AAV genome processing has also been explored by examining the molecular fate of rAAV vectors, which are converted into circular and concatemeric forms that persist episomally (18, 19, 66). Proteins shown to regulate circularization in cell culture include ATM and the MRN complex (14, 64), while in vivo experiments using mouse models have implicated ATM and DNA-PK in this process (14, 20, 72). Additionally, DNA-PKcs and Artemis have recently been shown to cleave the ITR hairpins of rAAV vectors in vivo in a tissue-dependent manner (36). Despite these studies, it is not clear how damage response factors function together and how they impact AAV transduction and replication in human cells.In this study we examined the cellular response to AAV replication in the context of Ad infection or helper proteins. We show that coinfection with AAV and Ad activates a DDR that is distinct from that seen during infection with Ad alone. The ATM and DNA-PKcs damage kinases are activated and signal to downstream substrates, but the response does not require the MRN complex and is primarily mediated by DNA-PKcs. Although expression of the large Rep proteins induced some DDR events, full signaling appeared to require AAV replication and was accompanied by accumulation of DNA-PK at viral replication compartments. Our results demonstrate that AAV replication induces a unique DNA damage signal transduction response and provides a model system for studying DNA-PK.  相似文献   

20.
The filovirus VP40 protein is capable of budding from mammalian cells in the form of virus-like particles (VLPs) that are morphologically indistinguishable from infectious virions. Ebola virus VP40 (eVP40) contains well-characterized overlapping L domains, which play a key role in mediating efficient virus egress. L domains represent only one component required for efficient budding and, therefore, there is a need to identify and characterize additional domains important for VP40 function. We demonstrate here that the 96LPLGVA101 sequence of eVP40 and the corresponding 84LPLGIM89 sequence of Marburg virus VP40 (mVP40) are critical for efficient release of VP40 VLPs. Indeed, deletion of these motifs essentially abolished the ability of eVP40 and mVP40 to bud as VLPs. To address the mechanism by which the 96LPLGVA101 motif of eVP40 contributes to egress, a series of point mutations were introduced into this motif. These mutants were then compared to the eVP40 wild type in a VLP budding assay to assess budding competency. Confocal microscopy and gel filtration analyses were performed to assess their pattern of intracellular localization and ability to oligomerize, respectively. Our results show that mutations disrupting the 96LPLGVA101 motif resulted in both altered patterns of intracellular localization and self-assembly compared to wild-type controls. Interestingly, coexpression of either Ebola virus GP-WT or mVP40-WT with eVP40-ΔLPLGVA failed to rescue the budding defective eVP40-ΔLPLGVA mutant into VLPs; however, coexpression of eVP40-WT with mVP40-ΔLPLGIM successfully rescued budding of mVP40-ΔLPLGIM into VLPs at mVP40-WT levels. In sum, our findings implicate the LPLGVA and LPLGIM motifs of eVP40 and mVP40, respectively, as being important for VP40 structure/stability and budding.Ebola and Marburg viruses are members of the family Filoviridae. Filoviruses are filamentous, negative-sense, single-stranded RNA viruses that cause lethal hemorrhagic fevers in both humans and nonhuman primates (5). Filoviruses encode seven viral proteins including: NP (major nucleoprotein), VP35 (phosphoprotein), VP40 (matrix protein), GP (glycoprotein), VP30 (minor nucleoprotein), VP24 (secondary matrix protein), and L (RNA-dependent RNA polymerase) (2, 5, 10, 12, 45). Numerous studies have shown that expression of Ebola virus VP40 (eVP40) alone in mammalian cells leads to the production of virus-like particles (VLPs) with filamentous morphology which is indistinguishable from infectious Ebola virus particles (12, 17, 18, 25, 26, 27, 30, 31, 34, 49). Like many enveloped viruses such as rhabdovirus (11) and arenaviruses (44), Ebola virus encodes late-assembly or L domains, which are sequences required for the membrane fission event that separates viral and cellular membranes to release nascent virion particles (1, 5, 7, 10, 12, 18, 25, 27, 34). Thus far, four classes of L domains have been identified which were defined by their conserved amino acid core sequences: the Pro-Thr/Ser-Ala-Pro (PT/SAP) motif (25, 27), the Pro-Pro-x-Tyr (PPxY) motif (11, 12, 18, 19, 41, 53), the Tyr-x-x-Leu (YxxL) motif (3, 15, 27, 37), and the Phe-Pro-Ile-Val (FPIV) motif (39). Both PTAP and the PPxY motifs are essential for efficient particle release for eVP40 (25, 27, 48, 49), whereas mVP40 contains only a PPxY motif. L domains are believed to act as docking sites for the recruitment of cellular proteins involved in endocytic trafficking and multivesicular body biogenesis to facilitate virus-cell separation (8, 13, 14, 16, 28, 29, 33, 36, 43, 50, 51).In addition to L domains, oligomerization, and plasma-membrane localization of VP40 are two functions of the protein that are critical for efficient budding of VLPs and virions. Specific sequences involved in self-assembly and membrane localization have yet to be defined precisely. However, recent reports have attempted to identify regions of VP40 that are important for its overall function in assembly and budding. For example, the amino acid region 212KLR214 located at the C-terminal region was found to be important for efficient release of eVP40 VLPs, with Leu213 being the most critical (30). Mutation of the 212KLR214 region resulted in altered patterns of cellular localization and oligomerization of eVP40 compared to those of the wild-type genotype (30). In addition, the proline at position 53 was also implicated as being essential for eVP40 VLP release and plasma-membrane localization (54).In a more recent study, a YPLGVG motif within the M protein of Nipah virus (NiV) was shown to be important for stability, membrane binding, and budding of NiV VLPs (35). Whether this NiV M motif represents a new class of L domain remains to be determined. However, it is clear that this YPLGVG motif of NiV M is important for budding, perhaps involving a novel mechanism (35). Our rationale for investigating the corresponding, conserved motifs present within the Ebola and Marburg virus VP40 proteins was based primarily on these findings with NiV. In addition, Ebola virus VP40 motif maps close to the hinge region separating the N- and C-terminal domains of VP40 (4). Thus, the 96LPLGVA101 motif of eVP40 is predicted to be important for the overall stability and function of VP40 during egress. Findings presented here indicate that disruption of these filovirus VP40 motifs results in a severe defect in VLP budding, due in part to impairment in overall VP40 structure, stability and/or intracellular localization.  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