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 531 毫秒
1.
Previously, a charge balance hypothesis was proposed to explain hepatitis B virus (HBV) capsid stability, assembly, RNA encapsidation, and DNA replication. This hypothesis emphasized the importance of a balanced electrostatic interaction between the positive charge from the arginine-rich domain (ARD) of the core protein (HBc) and the negative charge from the encapsidated nucleic acid. It remains unclear if any of the negative charge involved in this electrostatic interaction could come from the HBc protein per se, in addition to the encapsidated nucleic acid. HBc ARD IV mutant 173GG and ARD II mutant 173RR/R157A/R158A are arginine deficient and replication defective. Not surprisingly, the replication defect of ARD IV mutant 173GG can be rescued by restoring positively charged amino acids at the adjacent positions 174 and 175. However, most interestingly, it can be at least partially rescued by reducing negatively charged residues in the assembly domain, such as by glutamic acid-to-alanine (E-to-A) substitutions at position 46 or 117 and to a much lesser extent at position 113. Similar results were obtained for ARD II mutant 173RR/R157A/R158A. These amino acids are located on the inner surfaces of HBc icosahedral particles, and their acidic side chains point toward the capsid interior. For HBV DNA synthesis, the relative amount of positive versus negative charge in the electrostatic interactions is more important than the absolute amount of positive or negative charge. These results support the concept that balanced electrostatic interaction is important during the viral life cycle.Human hepatitis B virus (HBV) is an important human pathogen (5, 27, 34) that can replicate via an RNA intermediate (31, 33). Wild-type (WT) HBV core protein (HBc) is 183 amino acids long (adr and ayw subtypes) and consists of two distinct domains connected by a hinge region. The assembly domain spans amino acids 1 to 140, and the arginine-rich domain (ARD) spans amino acids 150 to 183. The ARD of HBc 150-183 is not required for capsid assembly in Escherichia coli (4, 8, 10, 21, 35). During nucleocapsid (capsid) formation, the HBc protein assembles into an icosahedral particle via a dimer intermediate (32). The ARD of HBc is known to be capable of binding to nucleic acids (12, 25). Serine phosphorylation at the C terminus of HBc is known to be important for RNA encapsidation, DNA synthesis, and virion secretion (2, 11, 14, 15, 17, 19, 24, 26, 39, 40). To date, there is no structural information available for the C terminus of HBc in capsids (32). The 4-helix bundle structure of HBc capsids is based on a C-terminally truncated capsid protein, HBc149 (6, 7, 36). Our research progress in the study of HBV biology has been hampered due to the lack of structural information about the HBc C-terminal tail, which plays an important regulatory role throughout the life cycle of HBV.Recently, we proposed a hypothesis that “charge balance” could be important for HBV capsid stability, assembly, RNA encapsidation, and DNA replication (17). This hypothesis postulates that many important viral activities could be influenced by the electrostatic interaction between the positive charge of the basic amino acid-rich domains of a nucleocapsid protein and the negative charge of the nucleocapsid-associated nucleic acids (17). Previously, we and others demonstrated that a mutant HBc 164, which lacks a total of eight arginine residues at the C terminus, can package both the nonspliced 3.5-kb pregenomic RNA (pgRNA) and the 2.2-kb spliced RNA (14, 17). However, pgRNA encapsidated by HBc 164 is RNase sensitive, while the encapsidated 2.2-kb spliced RNA is RNase resistant. The pgRNA and spliced RNA encapsidated by the full-length wild-type HBc 183 is also RNase resistant (14, 17). This result provided us with the first clue to invoke the previously coined “charge balance hypothesis” of RNA encapsidation and capsid stability (17). In subsequent studies, we added arginines back to the truncated HBc 164 gradually and noted that core particle-associated viral DNA also increased gradually in both size and intensity. This result provided us with the second clue to expand the previous hypothesis from RNA encapsidation to include viral DNA synthesis (17). Instead of totally nonquantitative or nonspecific binding between nucleic acids and a basic protein, our electrostatic-interaction hypothesis entails a stoichiometry-like concept between basic residues and their binding partner of nucleic acids. One of the reasons for such a demand for a more quantitative charge-charge interaction could be related to the putative intersubunit positive-charge repulsion built into the context of an icosahedral particle (22).We and others demonstrated previously that truncated HBc 173 (also denoted HBc 173RR in Fig. Fig.1)1) is sufficient for a WT-like DNA phenotype (17, 20). Mutant 173GG, containing two substitutions from arginine (R) to glycine (G) at codons 172 and 173 of HBc 173, exhibited a shorter-than-full-length HBV DNA phenotype, suggesting the importance of sufficient positive charge for viral DNA synthesis (17).Open in a separate windowFIG. 1.Arginine-deficient HBc mutants SVC173GG and SVC173/R157A/R158A exhibited a short DNA phenotype by complementation and Southern blot analysis. (A) To test the balanced electrostatic-interaction hypothesis, we engineered various HBc mutants with different arginine contents. The experimental approach is illustrated in the cartoon. (B) Amino acid sequence comparisons among the WT and HBc mutants. The hyphens represent amino acid sequences identical to that of the parental WT HBc. Roman numbers I to IV indicate the four different ARDs at the C terminus of HBc. The names SVC 173RR and SVC173 are used interchangeably in this paper. ARD IV mutant 173GG contains two arginine-to-glycine substitutions at positions 172 and 173, while ARD II mutant SVC173/R157A/R158A contains two arginine-to-alanine substitutions at positions 157 and 158. To further test the balanced electrostatic-interaction hypothesis, we restored arginines at the new positions 174 and 175 in mutant 175GGRR. (C, top) Plasmid 1903, an HBV genomic dimer containing an ablated core AUG initiation codon, was cotransfected with WT or various mutant core expression plasmids into Huh7 cells. HBV core-associated DNAs were analyzed by Southern blot analysis. The positive control mutant SVC173RR was WT-like in viral DNA synthesis (17, 20). More mature RC DNA of HBV was almost undetectable in mutants 173GG and 173/R157A/R158A, even after very long exposure to X-ray film. The replication defect of mutant 173GG could be rescued in mutant 175GGRR. The asterisk highlights the defect in synthesizing full-length viral DNA. In contrast, the black dots in lane SVC175GGRR highlight the functional rescue of full-length HBV DNA synthesis. SS, ssDNA replicative intermediate. (Bottom) Capsids collected from transfected cell lysates were measured by Western blot analysis using rabbit anti-core antibody.To test further the effect of electrostatic interaction on viral DNA synthesis, we asked if one could restore the DNA replication defect of the arginine-deficient ARD IV mutant 173GG or ARD II mutant 173RR/R157A/R158A by reducing their negative charges at specific positions. Indeed, our results showed that a single substitution from glutamic acid to alanine (E-to-A) at position 46 or 117, and to a lesser extent at position 113, could restore the WT-like DNA replication in the context of ARD IV mutant 173GG. Similarly, in the context of ARD II mutant 173RR/R157A/R158A, both E46A and E117A could rescue the full-length single-stranded DNA (ssDNA) and relaxed-circle (RC) DNA syntheses.Using an in vitro capsid assembly/disassembly assay, we demonstrated that the capsid stability of both wild-type full-length HBc 183 and truncated HBc 173 capsid particles from E. coli are dependent on the presence of encapsidated RNA (22). Upon treatment with micrococcal nuclease, encapsidated RNAs were digested and lost, leading to capsid disassembly. In this study, we demonstrated that mutant 173GG maintained capsid stability upon the loss of encapsidated RNA, probably due to the loss of intersubunit positive charge repulsion at ARD IV. Interestingly, when mutation E117A was introduced into the context of mutant 173GG, the capsid stability of mutant capsids E117A/173GG was once again dependent on the encapsidated RNA in a manner similar to that of its parental mutant, 173. Further studies also revealed that the truncated HBc mutant 172, like mutant 173, is sufficient for HBV DNA replication. Taken together, these results lend strong support to the balanced electrostatic-interaction hypothesis of HBV DNA replication.  相似文献   

2.
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   

3.
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.
7.
8.
9.
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   

10.
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   

11.
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   

12.
Retrovirus assembly is driven by polymerization of the Gag polyprotein as nascent virions bud from host cells. Gag is then processed proteolytically, releasing the capsid protein (CA) to assemble de novo inside maturing virions. CA has N-terminal and C-terminal domains (NTDs and CTDs, respectively) whose folds are conserved, although their sequences are divergent except in the 20-residue major homology region (MHR) in the CTD. The MHR is thought to play an important role in assembly, and some mutations affecting it, including the F167Y substitution, are lethal. A temperature-sensitive second-site suppressor mutation in the NTD, A38V, restores infectivity. We have used cryoelectron tomography to investigate the morphotypes of this double mutant. Virions produced at the nonpermissive temperature do not assemble capsids, although Gag is processed normally; moreover, they are more variable in size than the wild type and have fewer glycoprotein spikes. At the permissive temperature, virions are similar in size and spike content as in the wild type and capsid assembly is restored, albeit with altered polymorphisms. The mutation F167Y-A38V (referred to as FY/AV in this paper) produces fewer tubular capsids than wild type and more irregular polyhedra, which tend to be larger than in the wild type, containing ∼30% more CA subunits. It follows that FY/AV CA assembles more efficiently in situ than in the wild type and has a lower critical concentration, reflecting altered nucleation properties. However, its infectivity is lower than that of the wild type, due to a 4-fold-lower budding efficiency. We conclude that the wild-type CA protein sequence represents an evolutionary compromise between competing requirements for optimization of Gag assembly (of the immature virion) and CA assembly (in the maturing virion).In the first step of retrovirus maturation, the Gag precursor polyprotein is processed into three main components: the matrix (MA), capsid (CA), and nucleocapsid (NC). Of these, ∼1,000 to 2,000 copies of CA assemble into a capsid shell, enclosing the dimeric RNA genome and viral replication enzymes, while leaving a sizable population of unassembled CA subunits (5, 24). A properly formed capsid is thought to be essential for infectivity. The CA proteins of different retroviruses are quite uniform in size, ∼230 to 240 amino acids, but exhibit little sequence similarity except in the major homology region (MHR). Nevertheless, they share a structure with two domains (the N-terminal and C-terminal domains [NTDs and CTDs, respectively]) of conserved folds, connected by a short linker (2, 3, 8, 12, 17, 21, 22, 28). The NTDs form hexameric and pentameric rings that associate via homodimeric interactions between CTDs present in neighboring rings. Recently, structural evidence for a third interaction between the NTDs and the CTDs has been presented (10, 16, 31).The observed conservation of the MHR sequence points to its functional importance; in particular, the MHR is thought to play key roles, both in Gag assembly to produce immature virions and subsequently in the assembly of capsids within maturing virions. Consistent with this, several studies have reported adverse effects of point mutations in the MHR: certain mutations block the assembly of Gag proteins, whereas others have no evident effect on Gag assembly, genome incorporation, or budding but yield virus-like particles that are noninfectious or poorly infectious (1, 7, 13, 26, 30, 35-37). Starting with mutations of the latter kind, a series of second-site suppressors have been isolated that partially restore infectivity. Of these, two mapped in the NTD (A38V and P65Q), three in the dimerization helix in the CTD (F193L, R185W, and I190V), and one in the cleavage site between CA and the downstream peptide (S241L) (4, 7, 13, 25). The rescue of the MHR mutants by suppressing mutations was found not to be allele specific (25). The lack of allele specificity and the temperature sensitivity of some MHR mutations and their suppressors suggest that the affected residues are important for achieving a conformation(s) needed for proper assembly.In a previous study, we used cryoelectron tomography (cryo-ET) to characterize the pleiomorphic variability of wild-type Rous sarcoma virions (6, 20). Some 80% of them were found to contain capsids, which could be tubular, irregular polyhedra, or “continuous curvature” capsids, lacking angular vertices. The capsid type was found to correlate with the number of glycoprotein spikes per virion and with the efficiency of assembly, or conversely, the size of the pool of unassembled CA subunits contained within a virion. Based on these observations, we posited that in capsid assembly in situ, which is necessarily a nucleated process, different kinds of nucleation complexes initiate assembly of the various kinds of capsids. We also observed that tubular and some continuous curvature capsids had little internal material, suggesting that packaging of the viral ribonucleoprotein (RNP) had failed. Accordingly, and to accommodate the extensive polymorphism that was observed, we posited that a viable core is one with a closed capsid (of any morphology) that has successfully packaged its RNP. Subsequent in vitro studies have supported and clarified the perception of CA assembly as a variably nucleated process. Specifically, (i) it has been demonstrated that CA protein can assemble in a nucleation-driven manner into a variety of capsid-related structures (32, 33), and (ii) the mutation F167Y has been found to hamper nucleation of CA assembly in vitro, while the A38V suppressor strongly promotes assembly. Thus, CA-A38V assembles exceedingly rapidly, and in the double mutant, the A38V change overcomes the nucleation defect caused by F167Y.We have now further investigated the relationships among capsid morphology, nucleated assembly, and infectivity by performing a cryo-ET analysis of virions produced by the temperature-sensitive double mutant in which F167Y is complemented with the suppressor A38V; at the permissive temperature, the infectivity of the double mutant is restored to about 70% of wild type (25). Prior observations (32, 33) allowed the prediction that capsid assembly in vivo would be impaired in initiation for F167Y and for F167Y-A38V (abbreviated hereafter as FY/AV) at the nonpermissive temperature. Expectations for the double mutant at the permissive temperature were less clear: capsids should be produced, but this process might be altered in some way, as infectivity is lower than in the wild type. Because infectivity as measured could also be affected by Gag-related functions occurring earlier in the replication pathway, i.e., in viral budding or in proteolytic processing, we also compared the rates of virus growth, terminal Gag cleavage, and budding efficiency under these conditions for both the mutants and the wild-type virus.  相似文献   

13.
The alphaherpesvirus proteins UL31 and UL34 and their homologues in other herpesvirus subfamilies cooperate at the nuclear membrane in the export of nascent herpesvirus capsids. We studied the respective betaherpesvirus proteins M53 and M50 in mouse cytomegalovirus (MCMV). Recently, we established a random approach to identify dominant negative (DN) mutants of essential viral genes and isolated DN mutants of M50 (B. Rupp, Z. Ruzsics, C. Buser, B. Adler, P. Walther and U. H. Koszinowski, J. Virol 81:5508-5517). Here, we report the identification and phenotypic characterization of DN alleles of its partner, M53. While mutations in the middle of the M53 open reading frame (ORF) resulted in DN mutants inhibiting MCMV replication by ∼100-fold, mutations at the C terminus resulted in up to 1,000,000-fold inhibition of virus production. C-terminal DN mutants affected nuclear distribution and steady-state levels of the nuclear egress complex and completely blocked export of viral capsids. In addition, they induced a marked maturation defect of viral capsids, resulting in the accumulation of nuclear capsids with aberrant morphology. This was associated with a two-thirds reduction in the total amount of unit length genomes, indicating an accessory role for M53 in DNA packaging.Our understanding of herpesvirus morphogenesis is mainly derived from studies of Alphaherpesvirinae, such as herpes simplex virus type 1 (HSV-1) and pseudorabies virus (PrV). A faster replication cycle and a more productive infection in tissue culture aided genetic analysis of alphaherpesvirus morphogenesis. In addition, deletion mutants of key morphogenesis genes in alphaherpesviruses often maintain basic replication capacity, whereas the mutations of their homologues in Betaherpesvirinae or Gammaherpesvirinae mostly result in a lethal phenotype (for the UL31 and the UL34 family, see references 3, 6, 9-11, 16, 20, 21, and 42). These genes became amenable to comprehensive genetic analysis in betaherpesviruses only after their genomes were cloned as infectious bacterial artificial chromosomes (BACs), which obviated the need to generate replication-competent intermediates or complementing cell lines (3, 21, 23). BAC-based mutagenesis allowed viability screens mapping essential genes (8, 43) or even functional sites of essential genes in cytomegaloviruses (3, 21). However, these approaches cannot easily be applied to reveal the null phenotypes in the context of virus replication, as mutant viruses are not easily reconstituted. In addition, deletion of an essential viral gene can reveal the null phenotype of only the first of perhaps several essential functions during virus morphogenesis. This problem can be addressed to some extent by using dominant negative (DN) mutations (36). DN mutants are loss-of-function mutants that induce a null phenotype in the presence of the wild-type (wt) allele (14). Analysis of phenotypes induced by DN mutants proved to be extremely useful in genetics and cell biology, signaling, and biochemistry. Such inhibitory mutants of cellular proteins are often designed based on knowledge on the structural or functional role of a well-characterized protein domain. Unfortunately, we lack the structural information that would allow knowledge-based design of viral DN mutants for the majority of herpesvirus gene products. Thus, we established a random screen consisting of three steps to identify mutants of viral genes with DN potential (36): (i) a library of mutants is generated by random insertion of 5 amino acids (aa) or a stop codon into the open reading frame (ORF) of interest using transposon mutagenesis, (ii) nonfunctional mutants are identified by cis complementation of the respective deletion mutant mouse cytomegalovirus (MCMV) BAC, and (iii) nonfunctional mutants are tested for their inhibitory potential upon reconstitution of the wt BAC cloned genomes. In the last screen, mutants that have a specific inhibitory effect on the activity of the wt allele are selected. The specific phenotype obtained upon induction of the inhibitory mutants in the context of virus replication is then verified and further characterized using a tetracycline (Tet) regulon-based viral conditional expression system (36, 37).One intriguing aspect of herpesvirus morphogenesis is the transition of capsids from the nuclear to the cytoplasmic phase of virus morphogenesis. Two conserved nonstructural proteins, the homologues of the membrane protein pUL34 and its nuclear partner protein pUL31, form a nuclear egress complex (NEC) (18, 27, 42), which is required for primary envelopment and export of nuclear capsids to the cytoplasm (reviewed in references 24 and 25). Recent studies have revealed that the homologues of alphaherpesvirus pUL34 and pUL31, the M50 and the M53 gene products of the betaherpesvirus MCMV (pM50 and pM53, respectively) and the BFRF1 and the BFLF2 gene products of the gammaherpesvirus Epstein-Barr virus (EBV), apparently share the major functions of these two proteins. The lack of one or both proteins of the NEC generally results in the retention of viral capsids in the nucleus. This is lethal for beta- and gammaherpesvirus production (3, 9-11, 16, 18, 21, 27, 35, 42).The details of the mechanisms by which the NEC proteins mediate capsid export through the nuclear envelope are poorly understood. We (3, 21, 36, 38) and others (1, 19, 34) have started to dissect details of the NEC function using a genetic approach based on subtle mutagenesis of the respective genes. Analysis of the MCMV M50 gene by comprehensive mutagenesis localized two different functional sites. They were the M53 binding site within the N-terminal domain of M50, as well as the transmembrane region at its C terminus (3). Liang and Baines located the respective binding site in HSV-1 UL34 at aa 137 to 181 (19). Our approach, based on screens for DN mutants, identified a proline-rich sequence (aa 179 to 207) in the M50 gene product as an additional essential region (36). A recombinant virus expressing an M50 mutant lacking this site was defective in capsid egress from the nucleus despite the presence of the wt M50 protein. Consequently, the production of infectious particles after infection was reduced by more than 2 orders of magnitude. The UL34 homologues of alpha- and gammaherpesviruses lack a similar polyproline motif, but the result was confirmed by mutating the human cytomegalovirus (HCMV) homologue UL50 at the corresponding region, which is conserved within betaherpesviruses (36). The M50 mutants lacking the proline-rich motif still bind and colocalize to their respective NEC partner, pM53. Interestingly, Bjerke and coworkers also provided genetic evidence for the existence of at least one additional, yet-unknown, but essential functional entity in pUL34 of HSV-1, besides its known pUL31 binding activity, using a screen based on charged-cluster mutations (1). Further analysis of one of the noncomplementing charged-cluster mutants carrying the defect in the N-terminal domain of pUL34 also revealed a DN activity and suggested a new functional site involved in membrane curvature formation, together with the C-terminal domain of UL31 (34).The genetic analysis of M53 by Tn7-based linker scanning mutagenesis, followed by a cis complementation assay, localized the M50-binding site between aa 112 and 137 within the first of the four conserved regions (CRs) shared among the herpesvirus UL31 homologues (21). This analysis, together with a study we performed for further characterization of pM50/pM53 interaction, revealed that the large C-terminal part of pM53, comprising CR2 to -4, must carry at least one additional, yet-unknown, but essential functional site (21, 38).Here, we screened loss-of-function mutants of the MCMV M53 gene to retrieve M53 alleles with DN activity to localize this new functional domain. Mutants with a very strong inhibitory potential accumulated within CR4 of pM53 close to its C terminus. These CR4 mutants induced a block of capsid export from the nucleus. In addition, we could associate these mutations with the induction of a defect in capsid maturation and/or DNA packaging. These data suggested that pM53 is not only crucial for nuclear egress, but also involved in earlier steps of MCMV morphogenesis.  相似文献   

14.
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   

15.
16.
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   

17.
Interaction between pUL34 and pUL31 is essential for targeting both proteins to the inner nuclear membrane (INM). Sequences mediating the targeting interaction have been mapped by others with both proteins. We have previously reported identification of charge cluster mutants of herpes simplex virus type 1 UL34 that localize properly to the inner nuclear membrane, indicating interaction with UL31, but fail to complement a UL34 deletion. We have characterized one mutation (CL04) that alters a charge cluster near the N terminus of pUL34 and observed the following. (i) The CL04 mutant has a dominant-negative effect on pUL34 function, indicating disruption of some critical interaction. (ii) In infections with CL04 pUL34, capsids accumulate in close association with the INM, but no perinuclear enveloped viruses, cytoplasmic capsids, or virions or cell surface virions were observed, suggesting that CL04 UL34 does not support INM curvature around the capsid. (iii) Passage of UL34-null virus on a stable cell line that expresses CL04 resulted in selection of extragenic suppressor mutants that grew efficiently using the mutant pUL34. (iv) All extragenic suppressors contained an R229→L mutation in pUL31 that was sufficient to suppress the CL04 phenotype. (v) Immunolocalization and coimmunoprecipitation experiments with truncated forms of pUL34 and pUL31 confirm that N-terminal sequences of pUL34 and a C-terminal domain of pUL31 mediate interaction but not nuclear membrane targeting. pUL34 and pUL31 may make two essential interactions—one for the targeting of the complex to the nuclear envelope and another for nuclear membrane curvature around capsids.Egress of herpesvirus capsids from the nucleus occurs by envelopment of capsids at the inner nuclear membrane (INM) and is followed by de-envelopment at the outer nuclear membrane (ONM). This process can be broken down into a pathway of discrete steps that begin with recruitment of the viral envelopment apparatus to the INM. Herpes simplex virus type 1 (HSV-1) UL34 and UL31 and their homologs in other herpesviruses are required for efficient envelopment at the INM (7, 13, 22, 23, 29). HSV-1 pUL31 and pUL34 are targeted specifically to the INM by a mechanism that requires their interaction with each other (27, 28), and this mutual dependence is a conserved feature of herpesvirus envelopment (9, 14, 27, 28, 32, 33, 39). Localization of these two proteins at the INM results in the recruitment of other proteins, including protein kinase C delta and pUS3, to the nuclear membrane (22, 24, 30). The sequences in HSV-1 pUL34 that mediate interaction with UL31 and that lead to nuclear envelope targeting were mapped to amino acids (aa) 137 to 181 (16). The sequences in the murine cytomegalovirus (MCMV) homolog of UL31, M53, that mediate the nuclear envelope targeting interaction with the UL34 homolog, M50, were mapped to the N-terminal third of the protein in the first of four conserved regions (17), and Schnee et al. subsequently showed that this same region of pUL31 homologs from other families of herpesviruses mediates interaction with the corresponding pUL34 homologs (33).After the targeting of the pUL34/pUL31 complex to the INM, subsequent steps in nuclear egress include, it is thought, (i) local disruption of the nuclear lamina to allow capsid access to the INM, (ii) recognition and docking of capsids by the envelopment apparatus at the INM, (iii) curvature of the inner and outer nuclear membranes around the capsid, (iv) scission of the INM to create an enveloped virion in the space between the INM and ONM, (v) fusion of the virion envelope with the outer nuclear membrane, and (vi) capsid release into the cytoplasm.At least some of the viral and cellular factors critical for nuclear lamina disruption and for de-envelopment fusion have been identified. pUL34, pUL31, and pUS3 of HSV-1 have all been implicated in changes in localization, interaction, and phosphorylation of nuclear lamina components, including lamins A/C and B and the lamina-associated protein, emerin (3, 15, 19, 20, 24, 26, 34, 35). pUS3, pUL31, and glycoproteins B and H have been implicated in de-envelopment of primary virions at the ONM (8, 21, 28, 30, 38).pUL34 and pUL31 are thought to be involved in steps between lamina disruption and de-envelopment, but genetic evidence in infected cells has so far been lacking. Klupp et al. have shown that overexpression of alphaherpesvirus pUL31 and pUL34 in the absence of other viral proteins can induce formation of small vesicles derived from the INM, suggesting a role for these two proteins in membrane curvature around the capsid (12). Tight membrane curvature is an energetically unfavorable event and is thought to be accomplished by coupling curvature to energetically favorable interactions between membrane-bound proteins or protein complexes (reviewed in reference 40). The data of Klupp et al. suggest the possibility that upon recognition of a capsid, pUL31 and pUL34 may interact in a way that induces tight curvature of the INM. Here we present data in support of this hypothesis, showing that a specific point mutation in UL34 induces accumulation of docked capsids at the INM, extragenic suppression of the mutant phenotype is associated with a mutation in UL31, and pUL31 and pUL34 can interact via sequences that are not involved in their INM targeting interaction.We previously published a characterization of a library of 19 charge cluster mutants of pUL34. In each of these mutants, one charge cluster (defined as a group of five consecutive amino acids in which two or more of the residues have charged side chains) was mutated such that the charged residues were replaced by alanine. Six of the 19 charge cluster mutants tested failed to complement replication of UL34-null virus, indicating that they disrupt essential functions of pUL34. Interestingly, five of the six noncomplementing mutants were synthesized at levels comparable to that of wild-type UL34 and localized normally to the nuclear envelope, suggesting that they were unimpaired in their ability to make a nuclear envelope targeting interaction with UL31. In order to identify essential functions of pUL34 downstream of nuclear envelope targeting, we have undertaken a detailed study of the behavior and interactions of these mutants.  相似文献   

18.
19.
20.
Replicative fitness of poliovirus can be modulated systematically by replacement of preferred capsid region codons with synonymous unpreferred codons. To determine the key genetic contributors to fitness reduction, we introduced different sets of synonymous codons into the capsid coding region of an infectious clone derived from the type 2 prototype strain MEF-1. Replicative fitness in HeLa cells, measured by plaque areas and virus yields in single-step growth experiments, decreased sharply with increased frequencies of the dinucleotides CpG (suppressed in higher eukaryotes and most RNA viruses) and UpA (suppressed nearly universally). Replacement of MEF-1 capsid codons with the corresponding codons from another type 2 prototype strain (Lansing), a randomization of MEF-1 synonymous codons, increased the %G+C without increasing CpG, and reductions in the effective number of codons used had much smaller individual effects on fitness. Poliovirus fitness was reduced to the threshold of viability when CpG and UpA dinucleotides were saturated within and across synonymous codons of a capsid region interval representing only ∼9% of the total genome. Codon replacements were associated with moderate decreases in total virion production but large decreases in the specific infectivities of intact poliovirions and viral RNAs. Replication of codon replacement viruses, but not MEF-1, was temperature sensitive at 39.5°C. Synthesis and processing of viral intracellular proteins were largely unaltered in most codon replacement constructs. Replacement of natural codons with synonymous codons with increased frequencies of CpG and UpA dinucleotides may offer a general approach to the development of attenuated vaccines with well-defined antigenicities and very high genetic stabilities.Diversification of genomic sequences is constrained in all biological systems. At the level of primary sequences, the range of variability in coding regions is restricted by the codon usage bias (CUB), whereby a subset of synonymous codons are preferentially used in translation (24, 53, 69). The intensity of the CUB and the specific set of preferred codons vary widely across biological systems (39). Intertwined with the CUB is the suppression of the dinucleotides CpG and TpA (or UpA in RNA viruses) in the genomes of higher eukaryotes (4, 7, 26, 61) and many of their RNA viruses and small DNA viruses (28, 49). Variation in the primary sequences of RNA virus genomes is further constrained by requirements to maintain essential secondary and higher-order structures (42, 54, 68).We previously described the modulation of the replicative fitness of the Sabin type 2 oral poliovirus vaccine (OPV) strain (Sabin 2) by systematically changing the CUB in the capsid region, replacing the naturally occurring preferred codons with an unpreferred synonymous codon (isocodon) for each of nine amino acids (8). We called our approach “codon deoptimization” to contrast with the process of codon optimization, which is frequently used to maximize expression of foreign proteins in heterologous host systems (1, 27, 70). Apart from its potential application to development of improved poliovirus vaccines (8, 13, 38), experimental investigations of codon deoptimization directly test the relationships between replicative fitness, the extent of CUB, and the intensity of CpG and UpA suppression. As a model system for such studies, polioviruses offer several favorable properties, including (i) intrinsically high error rates for the poliovirus RNA-dependent RNA polymerase (2, 14, 16, 65), (ii) very high evolution rates (25), (iii) short generation times (8 to 10 h) and large progeny yields of prototype polioviruses, and (iv) well-developed reverse genetics (9).In this report, we extend our codon deoptimization strategy to the type 2 wild poliovirus prototype strain MEF-1. As before, we restricted our replacement of synonymous codons to the capsid coding region, which encodes two of the defining properties of polioviruses, namely, (i) the capacity to bind the CD155 poliovirus receptor (PVR) (23) and (ii) the poliovirus type-specific neutralizing antigenic sites (35). No changes were made to the flanking 5′-untranslated region and noncapsid region sequences, as they contain essential secondary structural elements (42, 54, 68) and are frequently exchanged out by recombination during circulation of poliovirus in human populations (20, 30, 32). MEF-1 was selected because of its high fitness level (hence, its use as the type 2 component of the inactivated poliovirus vaccine [IPV]) and because of its neurovirulence for humans (15), for nontransgenic mice (52), and for transgenic mice expressing the PVR (71). Type 2 polioviruses were selected first for study because the Sabin 2 OPV strain is most frequently associated with vaccine-associated paralytic poliomyelitis in contacts of OPV recipients (57, 59), with prolonged excretion of immunodeficiency-associated vaccine-derived polioviruses (VDPVs) (10, 31, 60), and with the emergence of circulating VDPVs in areas of low OPV coverage (10, 31).Consistent with our previous findings, the fitness of MEF-1 decreased in proportion to the total number of synonymous replacement codons. Fitness was reduced most efficiently by increasing the frequencies of CpG and UpA dinucleotides within and across synonymous codons. Saturation of CpG and UpA in a small capsid interval (representing only ∼9% of the genome) reduced fitness to the threshold of viability, even though the MEF-1 amino acid sequence was unaltered. The most prominent biological effect of deoptimization of codon usage and the large-scale incorporation of CpG and UpA was a sharp reduction in virus specific infectivities. In contrast, translation and processing of viral proteins and yields of intact virus particles with native antigenicities were reduced only moderately by increased CpG and UpA frequencies. Codon deoptimization with concurrent increases in the frequencies of CpG and/or UpA dinucleotides in RNA virus genomes may provide a novel general approach to the rational design of improved attenuated vaccines with predictable and stable genetic properties.  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