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 141 毫秒
1.
The p6 region of HIV-1 Gag contains two late (L) domains, PTAP and LYPXnL, that bind Tsg101 and Alix, respectively. Interactions with these two cellular proteins recruit members of the host''s fission machinery (ESCRT) to facilitate HIV-1 release. Other retroviruses gain access to the host ESCRT components by utilizing a PPXY-type L domain that interacts with cellular Nedd4-like ubiquitin ligases. Despite the absence of a PPXY motif in HIV-1 Gag, interaction with the ubiquitin ligase Nedd4-2 was recently shown to stimulate HIV-1 release. We show here that another Nedd4-like ubiquitin ligase, Nedd4-1, corrected release defects resulting from the disruption of PTAP (PTAP), suggesting that HIV-1 Gag also recruits Nedd4-1 to facilitate virus release. Notably, Nedd4-1 remediation of HIV-1 PTAP budding defects is independent of cellular Tsg101, implying that Nedd4-1''s function in HIV-1 release does not involve ESCRT-I components and is therefore distinct from that of Nedd4-2. Consistent with this finding, deletion of the p6 region decreased Nedd4-1-Gag interaction, and disruption of the LYPXnL motif eliminated Nedd4-1-mediated restoration of HIV-1 PTAP. This result indicated that both Nedd4-1 interaction with Gag and function in virus release occur through the Alix-binding LYPXnL motif. Mutations of basic residues located in the NC domain of Gag that are critical for Alix''s facilitation of HIV-1 release, also disrupted release mediated by Nedd4-1, further confirming a Nedd4-1-Alix functional interdependence. In fact we found that Nedd4-1 binds Alix in both immunoprecipitation and yeast-two-hybrid assays. In addition, Nedd4-1 requires its catalytic activity to promote virus release. Remarkably, RNAi knockdown of cellular Nedd4-1 eliminated Alix ubiquitination in the cell and impeded its ability to function in HIV-1 release. Together our data support a model in which Alix recruits Nedd4-1 to facilitate HIV-1 release mediated through the LYPXnL/Alix budding pathway via a mechanism that involves Alix ubiquitination.Retroviral Gag polyproteins bear short conserved sequences that control virus budding and release. As such, these motifs have been dubbed late or L domains (49). Three types of L domains have thus far been characterized: PT/SAP, LYPXnL, and PPPY motifs (5, 9, 32). They recruit host proteins known to function in the vacuolar protein sorting (vps) of cargo proteins and the generation of multivesicular bodies (MVB) compartments (2). It is currently accepted that budding of vesicles into MVB involves the sequential recruitment of endosomal sorting complexes required for transport (ESCRT-I, -II, and -III) and the activity of the VPS4 AAA-ATPase (22). These sorting events are believed to be triggered by recognition of ubiquitin molecules conjugated to cargo proteins (20, 24, 41). For retrovirus budding, L domain motifs are the primary signals in Gag that elicit the recruitment of ESCRT components to facilitate viral budding. Consequently, mutations in L domain motifs or dominant-negative interference with the function of ESCRT-III members or the VPS4 ATPase adversely affect virus release. This indicates that Gag interactions with the ESCRT machinery are necessary for virus budding and separation from the cell (7, 10, 15, 16, 21, 28, 44).Two late domains have been identified within the p6 region of human immunodeficiency virus type 1 (HIV-1) Gag protein: the PTAP and LYPXnL motifs. The PTAP motif binds the cellular protein Tsg101 (15, 39, 40, 47), whereas the LYPXnL motif is the docking site for Alix (44). Tsg101 functions in HIV-1 budding (15) as a member of ESCRT-I (30, 48), a soluble complex required for the generation of MVB. This process is topologically similar to HIV-1 budding and requires the recruitment of ESCRT-III members called the charged-multivesicular body proteins (3, 29, 48) and the activity of the VPS4 AAA-ATPase (4, 48). In addition to binding the LYPXnL motif, Alix also interacts with the nucleocapsid (NC) domain of HIV-1 Gag (13, 38), thus linking Gag to components of ESCRT-III that are critical for virus release (13).Other retroviruses, including the human T-cell leukemia virus (HTLV) and the Moloney murine leukemia virus (MoMLV), utilize the PPPY-type L domain to efficiently release virus (7, 26, 51). The PPPY motif binds members of the Nedd4-like ubiquitin ligase family (6, 7, 16, 19, 25, 43), whose normal cellular function is to ubiquitinate cargo proteins and target them into the MVB sorting pathway (11, 12, 20). Members of the Nedd4-like ubiquitin ligase family include Nedd4-1, Nedd4-2 (also known as Nedd4L), WWP-1/2, and Itch. They contain three distinct domains: an N-terminal membrane binding C2 domain (12), a central PPPY-interacting WW domain (43), and a C-terminal HECT domain that contains the ubiquitin ligase active site (42). The functional requirement for the binding of Nedd4-like ubiquitin ligases to the PPPY motif in virus budding has been demonstrated (7, 16, 18, 19, 25, 26, 28, 50, 51). Overexpression of dominant-negative mutants of Nedd4-like ligases, ESCRT-III components, or VPS4 cause a potent inhibition of PPPY-dependent virus release (7, 19, 29, 31, 52) and induce assembly and budding defects similar to those observed after perturbation of the PPPY motif (26, 51). These observations demonstrated that Nedd4-like ligases connect Gag encoding PPPY motif to ESCRT-III and VPS4 proteins to facilitate virus release.Whereas the role of Nedd4-like ubiquitin ligases in virus budding has been established, the protein interactions that link them to the cell''s ESCRT-III pathway are still unknown. Evidence for associations of Nedd4-like ligases with ESCRT proteins have been previously reported and include: the binding of Nedd4-like ubiquitin ligases LD1 and Nedd4-1 to ESCRT-I member Tsg101 (6, 31), the colocalization of multiple Nedd4-like ubiquitin ligases with endosomal compartments (1, 28), the requirement of the cell''s ESCRT pathway for Itch mediated L domain independent stimulation of MoMLV release (23), and the ubiquitination of ESCRT-I components with a shorter isoform, Nedd4-2s (8). Therefore, Nedd4-like ubiquitin ligase interactions with members of the cell''s ESCRT pathway may provide retroviral Gag with access to the host budding machinery required for virus release.Although HIV-1 Gag does not carry the PPPY canonical sequence known to interact with Nedd4-like ubiquitin ligases, both Nedd4-1 and Nedd4-2 were shown to restore the release of the HIV-1 PTAP mutant, albeit Nedd4-1 with less efficiency than Nedd4-2 (8, 46). These findings suggested that HIV-1 might utilize cellular Nedd4-like ubiquitin ligases to increase virus release. We present here evidence demonstrating that Nedd4-1 interacts with Gag and enhances HIV-1 PTAP virus release. Furthermore, we show that Nedd4-1''s function in HIV-1 release is distinct from that of Nedd4-2 in both its viral and cellular requirements. Notably, we found that Nedd4-1 enhancement of HIV-1 release requires the Alix-binding LYPXnL L domain motif in the p6 region and basic residues in the NC domain. In addition, Alix''s facilitation of HIV-1 release requires cellular Nedd4-1, since mutations in NC that prevented Alix-mediated HIV-1 release also eliminated release by overexpression of Nedd4-1. This suggested a Nedd4-1-Alix physical and functional interdependence. In agreement with this, we found Nedd4-1 to bind and ubiquitinate Alix in the cell. Taken together, these results support a model in which Alix recruits Nedd4-1 to facilitate late steps of HIV-1 release through the LYPXnL L domain motif via a mechanism that involves Alix ubiquitination.  相似文献   

2.
The efficient release of newly assembled retrovirus particles from the plasma membrane requires the recruitment of a network of cellular proteins (ESCRT machinery) normally involved in the biogenesis of multivesicular bodies and in cytokinesis. Retroviruses and other enveloped viruses recruit the ESCRT machinery through three classes of short amino acid consensus sequences termed late domains: PT/SAP, PPXY, and LYPXnL. The major late domain of Rous sarcoma virus (RSV) has been mapped to a PPPY motif in Gag that binds members of the Nedd4 family of ubiquitin ligases. RSV Gag also contains a second putative late domain motif, LYPSL, positioned 5 amino acids downstream of PPPY. LYPXnL motifs have been shown to support budding in other retroviruses by binding the ESCRT adaptor protein Alix. To investigate a possible role of the LYPSL motif in RSV budding, we constructed PPPY and LYPSL mutants in the context of an infectious virus and then analyzed the budding rates, spreading profiles, and budding morphology. The data imply that the LYPSL motif acts as a secondary late domain and that its role in budding is amplified in the absence of a fully functional PPPY motif. The LYPXL motif proved to be a stronger late domain when an aspartic acid was substituted for the native serine, recapitulating the properties of the LYPDL late domain of equine infectious anemia virus. The overexpression of human Alix in the absence of a fully functional PPPY late domain partially rescued both the viral budding rate and viral replication, supporting a model in which the RSV LYPSL motif mediates budding through an interaction with the ESCRT adaptor protein Alix.Retroviruses acquire their lipid envelopes from the plasma membrane as they bud from the cell. Although the structural protein Gag is both necessary and sufficient for the assembly of virus-like particles (VLPs), the membrane scission step of virus egress requires the recruitment of a network of cellular proteins normally involved in two analogous cellular membrane fission events, the budding of cargo-containing vesicles into multivesicular bodies (MVBs) (for review, see references 1, 5, 11, and 50) and the separation of two daughter cells during cytokinesis (3, 4). This cellular network of proteins, collectively called the ESCRT (endosomal sorting complex required for transport) machinery, includes four sequentially recruited high-molecular-weight protein complexes (ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III) and is essential for the transport of transmembrane cargo proteins to the lysosome for degradation via an MVB intermediate.In addition to the multiprotein ESCRT complexes, other proteins are required to promote the budding of vesicles into the MVB. Ubiquitin ligases (such as Nedd4) monoubiquitinate both ESCRT components and transmembrane cargo proteins, tagging them for the MVB pathway. Adaptor proteins connect cargo proteins to ESCRT complexes or ESCRT complexes to each other. Ultimately, the final membrane fission event of vesicle budding is mediated by an AAA ATPase (Vps4).Retroviruses as well as other enveloped viruses use three amino acid consensus sequences, PPXY, PT/SAP, and LYPXnL, as docking sites for the components of the cellular ESCRT machinery. The deletion or mutation of these sequences, termed late domains, results in the failure of the virus to recruit the budding machinery to the site of assembly and thereby results in a block at the late stage of virus release in which fully assembled but immature virus particles remain attached to the plasma membrane. The PPXY late domain interacts with the WW domains of the Nedd4 family of ubiquitin ligases. Multiple ESCRT components bind to monoubiquitin tags on both cargo and ESCRT proteins. The PT/SAP late domain binds the ESCRT-I complex component, Tsg101 (tumor susceptibility gene 101). The LYPXnL late domain interacts with an adaptor protein of the ESCRT pathway, Alix (ALG-2-interacting protein X; also called AIP1) (reviewed in reference 12). Alix interacts with both Tsg101 of the ESCRT-I complex and CMHP4 of the ESCRT-III complex. A possible fourth class of late domains for the paramyxovirus SV5 was reported previously (47). The late domain function in this case has been mapped to an FPIV sequence in the M (matrix) protein. To date, this motif has yet to be shown to be important for the budding of any other virus, and an FPIV-interacting cellular protein has yet to be identified.Often, retroviruses rely on multiple late domains for efficient budding (2, 13, 16, 29, 30). For example, in addition to its PT/SAP motif in human immunodeficiency virus type 1 (HIV-1) p6, which binds Tsg101 (6, 14, 34, 52), HIV-1 also harbors an Alix-binding LYPXnL motif that functions in budding (13, 33, 34, 48, 52). Mutation of this LYPXnL motif results in only a modest reduction in HIV-1 budding (10). However, the effects of mutations in the LYPXnL motif become more obvious in the context of a minimal Gag in which the globular domain of MA and the N-terminal domain of CA are absent (48). Furthermore, the role of this motif also seems to vary among cell types. For example, the deletion of this motif decreases HIV-1 particle production 2- to 3-fold in COS-7 cells (15) but has no consequence for HeLa cells (7). The relationship of the two viral late domains to each other is unknown. It is possible that they are partially redundant, are cooperative (since they act at slightly different steps in the ESCRT pathway), or are cell type specific. It has been observed that the mutation of one late domain has a larger effect on budding than the mutation of the other, implying a hierarchy of function. For example, in HIV-1, PTAP acts as the dominant late domain and LYPXnL acts as a secondary late domain. Equine infectious anemia virus (EIAV) seems to be an exception in that it relies only on a single LYPDL motif for late domain function.Like other retroviruses, the avian alpharetrovirus Rous sarcoma virus (RSV) requires the ESCRT pathway for release, as evidenced by the observation that a dominant-negative mutant of the ATPase Vps4, which is required for the final step of the ESCRT pathway that releases the ESCRT-III complex, inhibits RSV budding in a dose-dependent manner (37). Mutational analysis mapped the RSV late domain to the PPPY motif in the small spacer peptide p2b of Gag (41, 54, 56). This PPPY motif was previously shown to interact with chicken members of the Nedd4 family of ubiquitin ligases (21, 51). RSV Gag also harbors an LYPSL late domain consensus motif 5 amino acids downstream from PPPY in the p10 domain, which could potentially promote budding via an interaction with Alix.Alix, a 97-kDa adaptor protein with diverse functions, is composed of an N-terminal Bro1 domain, a central V domain, and a C-terminal proline-rich region (10, 22, 26, 58). The proline-rich region is assumed to be unstructured and binds Tsg101 and endophilins. The Bro1 domain, which binds CHMP4, is curved and resembles a banana shape. CHMP4 binding is functionally important for promoting HIV-1 budding (10). It was suggested previously that its convex face may allow Alix to sense negative curvatures in membranes (17, 22). At least for HIV-1, the Alix Bro1 domain also interacts with the Gag NC domain (42, 43). The central V domain of Alix, which is named for its shape, has a conserved hydrophobic pocket on the second arm near the apex of the V that is responsible for the binding of the LYPXnL late domains of HIV-1 and EIAV (10, 26, 58).In the present study, we investigated the role of the LYPSL motif in RSV budding and replication. We report here that not only the PPPY motif but also the LYPSL motif act as late domains. The contribution of the LYPSL motif to the budding rate and spreading rate is secondary to that of the PPPY motif but increases in the absence of a fully functional PPPY motif. The Alix overexpression-mediated rescue of PPPY mutants supports a model in which the LYPSL late domain functions through an interaction with Alix.  相似文献   

3.
To facilitate the release of infectious progeny virions, human immunodeficiency virus type 1 (HIV-1) exploits the Endosomal Sorting Complex Required for Transport (ESCRT) pathway by engaging Tsg101 and ALIX through late assembly (L) domains in the C-terminal p6 domain of Gag. However, the L domains in p6 are known to be dispensable for efficient particle production by certain HIV-1 Gag constructs that have the nucleocapsid (NC) domain replaced by a foreign dimerization domain to substitute for the assembly function of NC. We now show that one such L domain-independent HIV-1 Gag construct (termed ZWT) that has NC-p1-p6 replaced by a leucine zipper domain is resistant to dominant-negative inhibitors of the ESCRT pathway that block HIV-1 particle production. However, ZWT became dependent on the presence of an L domain when NC-p1-p6 was restored to its C terminus. Furthermore, when the NC domain was replaced by a leucine zipper, the p1-p6 region, but not p6 alone, conferred sensitivity to inhibition of the ESCRT pathway. In an authentic HIV-1 Gag context, the effect of an inhibitor of the ESCRT pathway on particle production could be alleviated by deleting a portion of the NC domain together with p1. Together, these results indicate that the ESCRT pathway dependence of HIV-1 budding is determined, at least in part, by the NC-p1 region of Gag.Human immunodeficiency virus type 1 (HIV-1) and other retroviruses hijack the cellular Endosomal Sorting Complex Required for Transport (ESCRT) pathway to promote the detachment of virions from the cell surface and from each other (3, 21, 42, 44, 47). The ESCRT pathway was initially identified based on its requirement for the sorting of ubiquitinated cargo into multivesicular bodies (MVB) (50, 51). During MVB biogenesis, the ESCRT pathway drives the membrane deformation and fission events required for the inward vesiculation of the limiting membrane of this organelle (26, 29, 50, 51). More recently, it emerged that the ESCRT pathway is also essential for the normal abscission of daughter cells during the final stage of cell division (10, 43). Most of the components of the ESCRT pathway are involved in the formation of four heteromeric protein complexes termed ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III. Additional components include ALIX, which interacts both with ESCRT-I and ESCRT-III, and the AAA ATPase Vps4, which mediates the disassembly of ESCRT-III (29, 42).The deformation and scission of endocytic membranes is thought to be mediated by ESCRT-III, which, together with Vps4, constitutes the most conserved element of the pathway (23, 26, 42). Indeed, it was recently shown that purified yeast ESCRT-III induces membrane deformation (52), and in another study three subunits of yeast ESCRT-III were sufficient to promote the formation of intralumenal vesicles in an in vitro assay (61). In mammals, ESCRT-III is formed by the charged MVB proteins (CHMPs), which are structurally related and tightly regulated through autoinhibition (2, 33, 46, 53, 62). The removal of an inhibitory C-terminal domain induces polymerization and association with endosomal membranes and converts CHMPs into potent inhibitors of retroviral budding (34, 46, 53, 60, 62). Alternatively, CHMPs can be converted into strong inhibitors of the ESCRT pathway and of HIV-1 budding through the addition of a bulky tag such as green fluorescent protein (GFP) or red fluorescent protein (RFP) (27, 36, 39, 54). Retroviral budding in general is also strongly inhibited by catalytically inactive Vps4 (22, 41, 55), or upon Vsp4B depletion (31), confirming the crucial role of ESCRT-III.Retroviruses engage the ESCRT pathway through the activity of so-called late assembly (L) domains in Gag. In the case of HIV-1, the primary L domain maps to a conserved PTAP motif in the C-terminal p6 domain of Gag (24, 28) and interacts with the ESCRT-I component Tsg101 (15, 22, 40, 58). HIV-1 p6 also harbors an auxiliary L domain of the LYPxnL type, which interacts with the V domain of ALIX (20, 35, 39, 54, 59, 63). Interestingly, Tsg101 binding site mutants of HIV-1 can be fully rescued through the overexpression of ALIX, and this rescue depends on the ALIX binding site in p6 (20, 56). In contrast, the overexpression of a specific splice variant of the ubiquitin ligase Nedd4-2 has been shown to rescue the release and infectivity of HIV-1 mutants lacking all known L domains in p6 (12, 57). Nedd4 family ubiquitin ligases had previously been implicated in the function of PPxY-type L domains, which also depend on an intact ESCRT pathway for function (4, 32, 38). However, HIV-1 Gag lacks PPxY motifs, and the WW domains of Nedd4-2, which mediate its interaction with PPxY motifs, are dispensable for the rescue of HIV-1 L domain mutants (57).ALIX also interacts with the nucleocapsid (NC) region of HIV-1 Gag (18, 49), which is located upstream of p6 and the p1 spacer peptide. ALIX binds HIV-1 NC via its Bro1 domain, and the capacity to interact with NC and to stimulate the release of a minimal HIV-1 Gag construct is shared among widely divergent Bro1 domain proteins (48). Based on these findings and the observation that certain mutations in NC cause a phenotype that resembles that of L domain mutants, it has been proposed that NC cooperates with p6 to recruit the machinery required for normal HIV-1 budding (18, 49).NC also plays a role in Gag polyprotein multimerization, and this function of NC depends on its RNA-binding activity (5-8). It has been proposed that the role of the NC-nucleic acid interaction during assembly is to promote the formation of Gag dimers (37), and HIV-1 assembly in the absence of NC can indeed be efficiently rescued by leucine zipper dimerization domains (65). Surprisingly, in this setting the L domains in p6 also became dispensable, since particle production remained efficient even when the entire NC-p1-p6 region of HIV-1 Gag was replaced by a leucine zipper (1, 65). These findings raised the possibility that the reliance of wild-type (WT) HIV-1 Gag on a functional ESCRT pathway is, at least in part, specified by NC-p1-p6. However, it also remained possible that the chimeric Gag constructs engaged the ESCRT pathway in an alternative manner.In the present report, we provide evidence supporting the first of those two possibilities. Particle production became independent of ESCRT when the entire NC-p1-p6 region was replaced by a leucine zipper, and reversion to ESCRT dependence was shown to occur as a result of restoration of p1-p6 but not of p6 alone. Furthermore, although the deletion of p1 alone had little effect in an authentic HIV-1 Gag context, the additional removal of a portion of NC improved particle production in the presence of an inhibitor of the ESCRT pathway. Together, these data imply that the NC-p1 region plays an important role in the ESCRT-dependence of HIV-1 particle production.  相似文献   

4.
The human immunodeficiency virus type 1 structural polyprotein Pr55Gag is necessary and sufficient for the assembly of virus-like particles on cellular membranes. Previous studies demonstrated the importance of the capsid C-terminal domain (CA-CTD), nucleocapsid (NC), and membrane association in Gag-Gag interactions, but the relationships between these factors remain unclear. In this study, we systematically altered the CA-CTD, NC, and the ability to bind membrane to determine the relative contributions of, and interplay between, these factors. To directly measure Gag-Gag interactions, we utilized chimeric Gag-fluorescent protein fusion constructs and a fluorescence resonance energy transfer (FRET) stoichiometry method. We found that the CA-CTD is essential for Gag-Gag interactions at the plasma membrane, as the disruption of the CA-CTD has severe impacts on FRET. Data from experiments in which wild-type (WT) and CA-CTD mutant Gag molecules are coexpressed support the idea that the CA-CTD dimerization interface consists of two reciprocal interactions. Mutations in NC have less-severe impacts on FRET between normally myristoylated Gag proteins than do CA-CTD mutations. Notably, when nonmyristoylated Gag interacts with WT Gag, NC is essential for FRET despite the presence of the CA-CTD. In contrast, constitutively enhanced membrane binding eliminates the need for NC to produce a WT level of FRET. These results from cell-based experiments suggest a model in which both membrane binding and NC-RNA interactions serve similar scaffolding functions so that one can functionally compensate for a defect in the other.The human immunodeficiency virus type 1 (HIV-1) structural precursor polyprotein Pr55Gag is necessary and sufficient for the assembly of virus-like particles (VLPs). Gag is composed of four major structural domains, matrix (MA), capsid (CA), nucleocapsid (NC), and p6, as well as two spacer peptides, SP1 and SP2 (3, 30, 94). Following particle assembly and release, cleavage by HIV-1 protease separates these domains. However, these domains must work together in the context of the full-length Gag polyprotein to drive particle assembly.Previous studies have mapped two major functional domains involved in the early steps of assembly: first, Gag associates with cellular membranes via basic residues and N-terminal myristoylation of the MA domain (10, 17, 20, 35, 39, 87, 91, 106); second, the Gag-Gag interaction domains that span the CA C-terminal domain (CA-CTD) and NC domain promote Gag multimerization (3, 11, 14, 16, 18, 23, 27, 29, 30, 33, 36, 46, 64, 88, 94, 102, 103). Structural and genetic studies have identified two residues (W184 and M185) within a dimerization interface in the CA-CTD that are critical to CA-CA interactions (33, 51, 74, 96). Analytical ultracentrifugation of heterodimers formed between wild-type (WT) Gag and Gag mutants with changes at these residues suggests that the dimerization interface consists of two reciprocal interactions, one of which can be disrupted to form a “half-interface” (22).In addition to the CA-CTD, NC contributes to assembly via 15 basic residues (8, 9, 11, 14, 18, 23, 25, 28, 34, 40, 43, 54, 57, 58, 74, 79, 88, 97, 104, 105), although some researchers have suggested that NC instead contributes to the stability of mature virions after assembly (75, 98, 99). It is thought that the contribution of NC to assembly is due to its ability to bind RNA, since the addition of RNA promotes the formation of particles in vitro (14-16, 37, 46), and RNase treatment disrupts Gag-Gag interactions (11) and immature viral cores (67). However, RNA is not necessary per se, since dimerization motifs can substitute for NC (1, 4, 19, 49, 105). This suggests a model in which RNA serves a structural role, such as a scaffold, to promote Gag-Gag interactions through NC. Based on in vitro studies, it has been suggested that this RNA scaffolding interaction facilitates the low-order Gag multimerization mediated by CA-CTD dimerization (4, 37, 49, 62, 63, 85). Despite a wealth of biochemical data, the relative contributions of the CA-CTD and NC to Gag multimerization leading to assembly are yet to be determined in cells.Mutations in Gag interaction domains alter membrane binding in addition to affecting Gag multimerization. In particular, mutations or truncations of CA reduce membrane binding (21, 74, 82), and others previously reported that mutations or truncations of NC affect membrane binding (13, 78, 89, 107). These findings are consistent with a myristoyl switch model of membrane binding in which Gag can switch between high- and low-membrane-affinity states (38, 71, 76, 83, 86, 87, 92, 95, 107). Many have proposed, and some have provided direct evidence (95), that Gag multimerization mediated by CA or NC interactions promotes the exposure of the myristoyl moiety to facilitate membrane associations.Gag membrane binding and multimerization appear to be interrelated steps of virus assembly, since membrane binding also facilitates Gag multimerization. Unlike betaretroviruses that fully assemble prior to membrane targeting and envelopment (type B/D), lentiviruses, such as HIV, assemble only on cellular membranes at normal Gag expression levels (type C), although non-membrane-bound Gag complexes exist (45, 58, 60, 61, 65). Consistent with this finding, mutations that reduce Gag membrane associations cause a defect in Gag multimerization (59, 74). Therefore, in addition to their primary effects on Gag-Gag interactions, mutations in Gag interaction domains cause a defect in membrane binding, which, in turn, causes a secondary multimerization defect. To determine the relative contributions of the CA-CTD and the NC domain to Gag-Gag interactions at the plasma membrane, it is essential to eliminate secondary effects due to a modulation of membrane binding.Except for studies using a His-tag-mediated membrane binding system (5, 46), biochemical studies of C-type Gag multimerization typically lack membranes. Therefore, these studies do not fully represent particle assembly, which occurs on biological membranes in cells. Furthermore, many biochemical and structural approaches are limited to isolated domains or truncated Gag constructs. Thus, some of these studies are perhaps more relevant to the behavior of protease-cleaved Gag in mature virions. With few exceptions (47, 74), cell-based studies of Gag multimerization have typically been limited to measuring how well mutant Gag is incorporated into VLPs when coexpressed or not with WT Gag. Since VLP production is a complex multistep process, effects of mutations on other steps in the process can confound this indirect measure. For example, NC contributes to VLP production by both promoting multimerization and interacting with the host factor ALIX to promote VLP release (26, 80). To directly assay Gag multimerization in cells, several groups (24, 45, 52, 56) developed microscopy assays based on fluorescence resonance energy transfer (FRET). These assays measure the transfer of energy between donor and acceptor fluorescent molecules that are brought within ∼5 nm by the association of the proteins to which they are attached (41, 48, 90). However, these microscopy-based Gag FRET assays have not been used to fully elucidate several fundamental aspects of HIV-1 Gag multimerization at the plasma membrane of cells, such as the relative contributions of the CA-CTD and NC and the effect of membrane binding on Gag-Gag interactions. In this study, we used a FRET stoichiometry method based on calibrated spectral analysis of fluorescence microscopy images (41). This algorithm determines the fractions of both donor and acceptor fluorescent protein-tagged Gag molecules participating in FRET. For cells expressing Gag molecules tagged with donor (cyan fluorescent protein [CFP]) and acceptor (yellow fluorescent protein [YFP]) molecules, this method measures the apparent FRET efficiency, which is proportional to the mole fraction of Gag constructs in complex. By measuring apparent FRET efficiencies, quantitative estimates of the mole fractions of interacting proteins can be obtained.Using this FRET-based assay, we aim to answer two questions: (i) what are the relative contributions of CA-CTD and NC domains to Gag multimerization when secondary effects via membrane binding are held constant, and (ii) what is the effect of modulating membrane binding on the ability of Gag mutants to interact with WT Gag?Our data demonstrate that the CA-CTD dimerization interface is essential for Gag multimerization at the plasma membrane, as fully disrupting the CA-CTD interaction abolishes FRET, whereas a modest level of FRET is still detected in the absence of NC. We also present evidence that the CA-CTD dimerization interface consists of two reciprocal interactions, allowing the formation of a half-interface that can still contribute to Gag multimerization. Notably, when Gag derivatives with an intact CA-CTD were coexpressed with WT Gag, either membrane binding ability or NC was required for the Gag mutants to interact with WT Gag, suggesting functional compensation between these factors.  相似文献   

5.
To promote the release of infectious virions, human immunodeficiency virus type 1 (HIV-1) exploits the endosomal sorting complex required for transport (ESCRT) pathway by engaging Tsg101 and ALIX through late assembly (L) domains in p6 Gag. An LYPxnL motif in p6 serves as docking site for the central V domain of ALIX and is required for its ability to stimulate HIV-1 budding. Additionally, the nucleocapsid (NC) domain of Gag binds to the N-terminal Bro1 domain of ALIX, which connects ALIX to the membrane-deforming ESCRT-III complex via its CHMP4 subunits. Since the isolated Bro1 domain of ALIX is sufficient to markedly stimulate virus-like particle (VLP) production in a minimal Gag rescue assay, we examined whether the Bro1 domains of other human proteins possess a similar activity. We now show that the Bro1 domain-only protein Brox and the isolated Bro1 domains of HD-PTP and rhophilin all bind to HIV-1 NC. Furthermore, all shared the capacity to stimulate VLP production by a minimal HIV-1 Gag molecule, and Brox in particular was as potent as the Bro1 domain of ALIX in this assay. Unexpectedly, Brox retained significant activity even if its CHMP4 binding site was disrupted. Thus, the ability to assist in VLP production may be an intrinsic property of the boomerang-shaped Bro1 domain.Retroviruses engage an endosomal budding machinery via so-called late assembly (L) domains in Gag to promote virus budding at the plasma membrane (4, 17, 33). In the case of human immunodeficiency virus type 1 (HIV-1), the C-terminal p6 domain of Gag harbors a conserved P(T/S)AP motif, which binds to the host protein Tsg101 and functions as the primary L domain (18, 29, 44). Additionally, HIV-1 p6 contains an auxiliary L domain of the LYPxnL type, which serves as a docking site for ALIX (28, 41, 45). Tsg101 and ALIX are both components of a protein network that is required for the biogenesis of multivesicular bodies (MVB) (22, 38). These compartments are formed through the budding of vesicles from the limiting membrane of endosomes into their lumen, a process that is topologically equivalent to virus budding at the plasma membrane. Recently, it emerged that the protein network essential for MVB formation also functions in cytokinesis, which requires a membrane fission event of similar topology (7, 32).Most of the components of the protein network that mediates these events are subunits of heteromeric endosomal sorting complexes required for transport (ESCRT) (3, 22, 38). For instance, Tsg101 is a subunit of the heterotetrameric ESCRT-I complex (22, 38). ESCRT-I and the downstream ESCRT-II are stable complexes, whereas ESCRT-III assembles only upon membrane binding (38). ESCRT-III is formed by the structurally related human CHMP proteins, which exist in an autoinhibited monomeric conformation in the cytosol (40, 46). A conformational change from a closed to an open conformation is thus likely required for the activation of CHMP proteins and the assembly of ESCRT-III. Interestingly, the uncontrolled activation of CHMP proteins through the removal of autoinhibitory C-terminal sequences results in the potent inhibition of HIV-1 budding, indicating a central role for ESCRT-III in retroviral release (46).ALIX consists of a boomerang-shaped N-terminal Bro1 domain, a central ligand binding domain that is shaped like a V, and a C-terminal proline-rich region (16). While ALIX is essential for equine anemia virus budding, its role in HIV-1 budding is less critical than that of Tsg101 (8, 16, 28, 41). However, ALIX can clearly support efficient HIV-1 budding, because its overexpression potently rescues the release defect of Tsg101 binding site mutants (16, 43). This effect of ALIX depends on the interaction between its central V domain and the LYPxnL motif in HIV-1 p6 (16, 43), confirming that this motif constitutes a functional L domain.The Bro1 domain of ALIX interacts tightly with ESCRT-III subunit CHMP4B and less avidly with CHMP4A and CHMP4C (25, 28, 41, 45). The ability of ALIX to rescue HIV-1 L domain mutants depends on the interaction between its Bro1 domain and CHMP4, indicating that CHMP4 is of particular importance in viral budding (16, 43). Interestingly, human CHMP4A assembles into membrane-attached filaments if overexpressed in mammalian cells, and these filaments can be induced to form circular arrays that drive the formation of buds and tubules with the same topology as that of a retroviral bud (21). Also, the single yeast ortholog of the mammalian CHMP4 proteins forms homo-oligomeric filaments on endosomes that appear to drive MVB sorting and biogenesis (42).By binding to membranes with its convex surface, the Bro1 domain of ALIX could also contribute directly to the generation of negative curvature required for budding away from the cytosol. In support of this notion, we recently observed that the isolated Bro1 domain of ALIX can potently enhance the formation of virus-like particles (VLP) by a minimal HIV-1 Gag construct that retains the primary L domain but lacks certain assembly domains and thus is presumably defective in its ability to deform membranes (37). We also observed that the Bro1 domain of ALIX physically interacts with the nucleocapsid (NC) region of HIV-1 Gag and that mutations in NC that interfere with the interaction induce a phenotype that resembles that of L domain mutants (37).Despite limited sequence homology between human ALIX and a yeast counterpart, the structures of their Bro1 domains are largely superimposable (16, 26), suggesting that all Bro1 domains have a shape that would be compatible with a membrane-deforming function. We therefore asked whether the ability to stimulate VLP production is unique to the Bro1 domain of ALIX or a property of Bro1 domains in general. We now show that widely divergent Bro1 domains share the ability to associate with HIV-1 Gag in an NC-dependent manner and to enhance VLP production by a minimal Gag molecule. In particular, a human Bro1 domain-only protein termed Brox (23) was as potent as the ALIX Bro1 domain in stimulating VLP production, and even forms of Brox that did not bind to CHMP4 retained significant activity. We thus propose that Bro1 domains are inherently capable of promoting budding events away from the cytosol.  相似文献   

6.
7.
The structural precursor polyprotein, Gag, encoded by all retroviruses, including the human immunodeficiency virus type 1 (HIV-1), is necessary and sufficient for the assembly and release of particles that morphologically resemble immature virus particles. Previous studies have shown that the addition of Ca2+ to cells expressing Gag enhances virus particle production. However, no specific cellular factor has been implicated as mediator of Ca2+ provision. The inositol (1,4,5)-triphosphate receptor (IP3R) gates intracellular Ca2+ stores. Following activation by binding of its ligand, IP3, it releases Ca2+ from the stores. We demonstrate here that IP3R function is required for efficient release of HIV-1 virus particles. Depletion of IP3R by small interfering RNA, sequestration of its activating ligand by expression of a mutated fragment of IP3R that binds IP3 with very high affinity, or blocking formation of the ligand by inhibiting phospholipase C-mediated hydrolysis of the precursor, phosphatidylinositol-4,5-biphosphate, inhibited Gag particle release. These disruptions, as well as interference with ligand-receptor interaction using antibody targeted to the ligand-binding site on IP3R, blocked plasma membrane accumulation of Gag. These findings identify IP3R as a new determinant in HIV-1 trafficking during Gag assembly and introduce IP3R-regulated Ca2+ signaling as a potential novel cofactor in viral particle release.Assembly of the human immunodeficiency virus (HIV) is determined by a single gene that encodes a structural polyprotein precursor, Gag (71), and may occur at the plasma membrane or within late endosomes/multivesicular bodies (LE/MVB) (7, 48, 58; reviewed in reference 9). Irrespective of where assembly occurs, the assembled particle is released from the plasma membrane of the host cell. Release of Gag as virus-like particles (VLPs) requires the C-terminal p6 region of the protein (18, 19), which contains binding sites for Alix (60, 68) and Tsg101 (17, 37, 38, 41, 67, 68). Efficient release of virus particles requires Gag interaction with Alix and Tsg101. Alix and Tsg101 normally function to sort cargo proteins to LE/MVB for lysosomal degradation (5, 15, 29, 52). Previous studies have shown that addition of ionomycin, a calcium ionophore, and CaCl2 to the culture medium of cells expressing Gag or virus enhances particle production (20, 48). This is an intriguing observation, given the well-documented positive role for Ca2+ in exocytotic events (33, 56). It is unclear which cellular factors might regulate calcium availability for the virus release process.Local and global elevations in the cytosolic Ca2+ level are achieved by ion release from intracellular stores and by influx from the extracellular milieu (reviewed in reference 3). The major intracellular Ca2+ store is the endoplasmic reticulum (ER); stores also exist in MVB and the nucleus. Ca2+ release is regulated by transmembrane channels on the Ca2+ store membrane that are formed by tetramers of inositol (1,4,5)-triphosphate receptor (IP3R) proteins (reviewed in references 39, 47, and 66). The bulk of IP3R channels mediate release of Ca2+ from the ER, the emptying of which signals Ca2+ influx (39, 51, 57, 66). The few IP3R channels on the plasma membrane have been shown to be functional as well (13). Through proteomic analysis, we identified IP3R as a cellular protein that was enriched in a previously described membrane fraction (18) which, in subsequent membrane floatation analyses, reproducibly cofractionated with Gag and was enriched in the membrane fraction only when Gag was expressed. That IP3R is a major regulator of cytosolic calcium concentration (Ca2+) is well documented (39, 47, 66). An IP3R-mediated rise in cytosolic Ca2+ requires activation of the receptor by a ligand, inositol (1,4,5)-triphosphate (IP3), which is produced when phospholipase C (PLC) hydrolyzes phosphatidylinositol-4,5-bisphosphate [PI(4,5)P2] at the plasma membrane (16, 25, 54). Paradoxically, PI(4,5)P2 binds to the matrix (MA) domain in Gag (8, 55, 59), and the interaction targets Gag to PI(4,5)P2-enriched regions on the plasma membrane; these events are required for virus release (45). We hypothesized that PI(4,5)P2 binding might serve to target Gag to plasma membrane sites of localized Ca2+ elevation resulting from PLC-mediated PI(4,5)P2 hydrolysis and IP3R activation. This idea prompted us to investigate the role of IP3R in Gag function.Here, we show that HIV-1 Gag requires steady-state levels of IP3R for its efficient release. Three isoforms of IP3R, types 1, 2, and 3, are encoded in three independent genes (39, 47). Types 1 and 3 are expressed in a variety of cells and have been studied most extensively (22, 39, 47, 73). Depletion of the major isoforms in HeLa or COS-1 cells by small interfering RNA (siRNA) inhibited viral particle release. Moreover, we show that sequestration of the IP3R activating ligand or blocking ligand formation also inhibited Gag particle release. The above perturbations, as well as interfering with receptor expression or activation, led to reduced Gag accumulation at the cell periphery. The results support the conclusion that IP3R activation is required for efficient HIV-1 viral particle release.  相似文献   

8.
9.
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   

10.
Ubiquitin conjugation to lysine residues regulates a variety of protein functions, including endosomal trafficking and degradation. While ubiquitin plays an important role in the release of many viruses, the requirement for direct ubiquitin conjugation to viral structural proteins is less well understood. Some viral structural proteins require ubiquitin ligase activity, but not ubiquitin conjugation, for efficient release. Recent evidence has shown that, like other viruses, hepatitis B virus (HBV) requires a ubiquitin ligase for release from the infected cell. The HBV core protein contains two lysine residues (K7 and K96), and K96 has been suggested to function as a potential ubiquitin acceptor site based on the fact that previous studies have shown that mutation of this amino acid to alanine blocks HBV release. We therefore reexamined the potential connection between core lysine ubiquitination and HBV replication, protein trafficking, and virion release. In contrast to alanine substitution, we found that mutation of K96 to arginine, which compared to alanine is more conserved but also cannot mediate ubiquitin conjugation, does not affect either virus replication or virion release. We also found that the core lysine mutants display wild-type sensitivity to the antiviral activity of interferon, which demonstrates that ubiquitination of core lysines does not mediate the interferon-induced disruption of HBV capsids. However, mutation of K96 to arginine alters the nuclear-cytoplasmic distribution of core, leading to an accumulation in the nucleolus. In summary, these studies demonstrate that although ubiquitin may regulate the HBV replication cycle, these mechanisms function independently of direct lysine ubiquitination of core protein.The hepatitis B virus (HBV) particle consists of an enveloped nucleocapsid that contains the viral polymerase (Pol) and an incomplete 3.2-kb double-stranded DNA genome (9). In the cytoplasm, the viral core structural proteins interact to form homodimers, which further self-assemble into capsid particles that package Pol and the viral pregenomic RNA. Encapsidated Pol subsequently reverse transcribes pregenomic RNA to give rise to mature double-stranded relaxed circular DNA-containing capsids. HBV DNA-containing capsids are released from the cell as mature virions after acquiring an envelope consisting of cellular membrane lipids and the viral small, middle, and large envelope proteins (4, 9, 41). Due to the directed insertion of the envelope proteins in the endoplasmic reticulum and Golgi membrane, and the requirement of the large envelope protein for virion release, nucleocapsids are hypothesized to bud at intracellular membranes for release through the constitutive secretory pathway (5). Although the mechanism and site of HBV nucleocapsid envelopment and release remain poorly understood, emerging evidence indicates that the cellular ubiquitin pathway may play a role in this process.Structural proteins of some enveloped RNA viruses contain highly conserved sequences [PPXY, P(T/S)AP, and YPXL] termed late (L) domains that mediate interactions with proteins of the endocytic pathway to facilitate virus budding and release (1). The P(T/S)AP motif binds Tsg101 (8, 10, 19, 27, 47), a key ESCRT (for endosomal sorting complex required for transport) component for the recognition and sorting of ubiquitinated proteins to internal vesicles of the multivesicular body (MVB), while the YPXL motif binds Alix, an ESCRT-associated protein (26, 44, 48). The PPXY motif binds proteins of the Nedd4 family ubiquitin ligases, which are responsible for ubiquitination of proteins targeted for endocytosis and sorting to the MVB (20), suggesting a link between ubiquitin and viral budding (3, 16, 17, 22, 43, 55). The observation that proteasome inhibition, which depletes free cellular ubiquitin by interfering with ubiquitin recycling, results in a viral budding defect similar to that seen in virus L domain mutants further supports the implication that ubiquitin plays a role in mediating virion release (15, 31, 40, 43). Furthermore, fusion of ubiquitin to the Rous sarcoma virus (RSV) PPPY-containing Gag protein and the equine infectious anemia virus (EIAV) Gag protein containing a heterologous PTAP or PPPY motif rescues the virus-like particle release defect induced by proteasome inhibition (18, 31). While the role of L domains in mediating virion release is relatively well established, it remains unclear whether direct ubiquitination of viral structural proteins is generally required for virion release. Mutation of ubiquitin acceptor lysine residues in the RSV Gag protein inhibits virus budding, but such mutations in human immunodeficiency virus type 1 (HIV-1) or murine leukemia virus Gag protein exert no effect on virus release (29, 42). Recently, a retroviral (i.e., prototypic foamy virus) Gag protein engineered to lack ubiquitin acceptor lysines and encoding either the PSAP or PPXY motif of the L domain displayed no defect in viruslike particle release (58). Altogether, these results suggest that recruitment of host proteins to the L domain and ubiquitination of interacting proteins, but not the viral structural proteins, is required for ubiquitin-dependent virion release, at least for some viruses.The HBV core structural protein contains two potential ubiquitin acceptor lysine residues (K7 and K96) and an L-domain-like PPAY motif (Fig. (Fig.1A).1A). Structural studies indicate that residue K96 and the PPAY motif may be exposed on the surface of HBV capsid particles, at least transiently (4, 32, 37). Studies aimed at identifying interaction factors important for HBV particle release demonstrated a number of interesting findings. First, γ2-adaptin, a cellular trafficking adaptor that contains a ubiquitin-interacting motif (UIM), interacts with both the viral large envelope protein and HBV core, and disruption of the HBV/γ2-adaptin interaction inhibits virus secretion (14, 39). Second, core protein interacts with the Nedd4 ubiquitin ligase through the PPAY motif in core (39). Mutation of the tyrosine in the PPAY motif results in disrupted binding of Nedd4, and overexpression of a catalytically inactive Nedd4 mutant inhibits HBV particle secretion (39). Third, mutation of core K96, but not K7, to alanine results in a defective release phenotype, suggesting that K96 may serve as a ubiquitin conjugation site that aids virion release (32, 39). Recently, overexpression of dominant-negative proteins of the MVB machinery, such as the Vps4 ATPases and the ESCRT-III complex-forming CHMP proteins, were also shown to disrupt HBV budding and virion release, while subviral particles comprised only of envelope proteins were released efficiently (21, 24, 49). This suggests that nucleocapsids may release from the cell by a mechanism distinct from constitutive secretion. These studies show that similar to RNA viruses, HBV utilizes components of the cellular protein trafficking machinery to mediate virion release.Open in a separate windowFIG. 1.Generation of core lysine mutants. (A) The 21-kDa HBV core structural protein contains two lysine residues at positions 7 and 96 that serve as potential ubiquitin conjugation sites. These residues are highly conserved among the four major HBV genotypes (6). Core contains a late-domain-like PPXY motif that serves as a binding site for the Nedd4 E3 ubiquitin ligase. Core additionally contains a potential noncanonical SUMOylation motif at position 96. (B) Lysine mutations were generated by site-directed mutagenesis in the core gene contained within the HBV genome under the control of a CMV promoter. K7R contains a lysine-to-arginine mutation at position 7, K96R contains a lysine-to-arginine mutation at position 96, K96A contains a lysine-to-alanine mutation at position 96, and K7R/K96R contains arginine substituted at position 7 and position 96.Although these findings imply that core ubiquitination may be necessary for HBV particle release, direct evidence of core ubiquitination has been elusive (33, 39; unpublished results). As suggested by previous Gag lysine mutagenesis studies, however, ubiquitin may instead indirectly be required through conjugation to an interacting protein that is essential for mediating HBV release (29, 58). Although core K7 and K96 have been previously assayed in the context of virion release by mutation of the lysine residues to alanine (32, 39), we expanded these studies by assaying core mutants with an arginine substitution at position K7 (K7R) and K96 (K96R), as well as a double lysine-to-arginine mutation (K7R/K96R). Compared to alanine, arginine serves as a more conserved mutation for lysine while still abolishing the potential ubiquitin conjugation site. In the present study, we utilized these mutants to comprehensively examine the role of the core lysines in HBV virus release, the formation of replication intermediates, intracellular localization of core, and the interferon (IFN)-mediated antiviral response.  相似文献   

11.
12.
VP40, the major matrix protein of Marburg virus, is the main driving force for viral budding. Additionally, cellular factors are likely to play an important role in the release of progeny virus. In the present study, we characterized the influence of the vacuolar protein sorting (VPS) pathway on the release of virus-like particles (VLPs), which are induced by Marburg virus VP40. In the supernatants of HEK 293 cells expressing VP40, different populations of VLPs with either a vesicular or a filamentous morphology were detected. While the filaments were almost completely composed of VP40, the vesicular particles additionally contained considerable amounts of cellular proteins. In contrast to that in the vesicles, the VP40 in the filaments was regularly organized, probably inducing the elimination of cellular proteins from the released VLPs. Vesicular particles were observed in the supernatants of cells even in the absence of VP40. Mutation of the late-domain motif in VP40 resulted in reduced release of filamentous particles, and likewise, inhibition of the VPS pathway by expression of a dominant-negative (DN) form of VPS4 inhibited the release of filamentous particles. In contrast, the release of vesicular particles did not respond significantly to the expression of DN VPS4. Like the budding of VLPs, the budding of Marburg virus particles was partially inhibited by the expression of DN VPS4. While the release of VLPs from VP40-expressing cells is a valuable tool with which to investigate the budding of Marburg virus particles, it is important to separate filamentous VLPs from vesicular particles, which contain many cellular proteins and use a different budding mechanism.In recent years, virus-like particles (VLPs), which are formed upon recombinant expression of the viral matrix and/or surface glycoproteins, have been recognized as representing powerful tools for developing novel vaccines and investigating certain aspects of the viral replication cycle (24, 44, 59, 63). Matrix proteins from many enveloped RNA viruses, including retroviruses, rhabdoviruses, filoviruses, paramyxoviruses, orthomyxoviruses, and arenaviruses, are able to induce VLPs (10, 14, 18, 28-30, 48, 49, 52). Increasing evidence also indicates that budding activity, and thus the release of VLPs, is often influenced by a complex interplay with components of the endosomal sorting complexes required for transport (ESCRTs), which mainly constitute the vacuolar protein sorting (VPS) pathway (16, 38, 42, 54). ESCRTs trigger the formation and budding of vesicles into the lumina of multivesicular bodies (MVBs), and the constituents of the ESCRTs are recycled by the activity of VPS4, an AAA-type ATPase. Expression of dominant-negative (DN) VPS4 mutants, which lack the ability to bind or hydrolyze ATP, blocks recycling of the ESCRTs and induces the formation of enlarged endosomes lacking internal vesicle accumulation (2, 3, 7). The inward budding of vesicles into the MVBs is topologically similar to the budding of viruses, since the vesicles bud away from the cytosol and into the lumen (reviewed in references 1, 20, and 26). Therefore, it is not entirely surprising that viruses use the cellular ESCRT machinery to organize the budding of viral progeny. Interactions between viral matrix proteins and ESCRTs occur through tetrapeptide motifs, known as late domains, which were first identified in retroviruses. Known late domains consist of the amino acid sequence P(T/S)AP, PPxY, or YxxL, where “x” represents any amino acid (19, 25, 62). The P(T/S)AP motif, for example, mediates interaction with tumor susceptibility gene 101 (Tsg101) (16, 36, 57); the PPxY motif mediates binding to WW domains of Nedd4-like ubiquitin ligases (9, 22); and the YxxL motif mediates interaction with AIP1/Alix (35, 47, 58). Recently, a novel late-domain motif, FPIV, has been identified in paramyxoviruses (46), and it is thought that additional late-domain motifs remain to be discovered (for a review, see reference 5).Inhibition of the VPS pathway has been shown to inhibit the budding of various viruses that are released with the help of ESCRTs. However, the budding of viruses and VLPs depends on the activity of ESCRTs to different degrees. Downregulation of Tsg101, a member of the ESCRT-I complex, inhibited the release of VLPs mediated by lymphocytic choriomeningitis virus Z protein and Marburg virus (MARV) VP40 (42, 54) but did not substantially inhibit the release of Gag-induced VLPs of Moloney murine leukemia virus and Rous sarcoma virus or that of matrix protein-induced VLPs of rabies virus (16, 27, 38). Expression of DN VPS4 inhibited the release of VLPs induced by the Gag proteins of Rous sarcoma virus and Moloney murine leukemia virus (16, 38) as well as that of VLPs induced by Lassa virus Z protein (55) but had no effect on the budding of rabies virus and cytomegalovirus (13, 27). These data indicate that in spite of the presence of late-domain motifs, a block in the VPS pathway may not always be critical for the budding of VLPs. In addition, the lack of known late domains in many enveloped viruses raises the question of whether they use other entry points into the VPS pathway or whether they exploit entirely different mechanisms of budding (60). To date, knowledge of how viral matrix proteins engage cellular machineries, such as the VPS pathway, to induce viral budding at the plasma membrane is very limited (8).The matrix protein VP40 of MARV contains only one known late-domain motif, PPPY, and a recent study showed that mutation of this late domain inhibited the release of VP40-induced VLPs. In addition, depletion of Tsg101 reduced the release of VP40-induced VLPs, suggesting that ESCRT-I is involved in this process (54). Whether a functional VPS pathway is important for the release of MARV VP40-induced VLPs or MARV particles remains unknown.VLPs induced by many viral matrix proteins have a morphology similar to that of cellular vesicles, which makes it difficult to separate the spherical VLPs from released cellular vesicles (4, 17, 53). In contrast, VLPs induced by the filovirus matrix protein VP40 are elongated and similar in morphology to viral particles (30, 49). Nevertheless, we observed that the supernatants of cells expressing VP40 contained various populations of particles with different morphologies. This raised the questions of whether the different particles are released by the same mechanism, whether they are all induced by VP40, and whether they are dependent on the same cellular pathways.The aim of the present study was to analyze the populations of particles released from cells expressing the MARV matrix protein VP40 and to gain further insights into the interaction between MARV and the cellular machinery involved in the budding of VLPs and MARV particles.We found that cells expressing VP40 released vesicular and filamentous particles, which could be separated by gradient centrifugation. Fractions with mainly vesicular particles represented a mixture of vesicles containing exclusively cellular proteins and vesicles also containing VP40 and few short filamentous particles. Longer filamentous particles, whose morphology resembled that of MARV particles but which displayed a much higher variability in length (400 nm to 5 μm), were found in denser gradient fractions. Filamentous VP40-induced VLPs were able to sort out cellular proteins efficiently. Release of VP40-induced filamentous VLPs was supported by the late-domain motif present in VP40, and inhibition of the cellular ESCRT machinery reduced the amount of these VLPs in the supernatant. Interestingly, the release of VLPs induced by a mutant of VP40 lacking the late domain was also reduced by inhibition of the cellular ESCRT machinery. Expression of a DN mutant of VPS4 diminished the budding of infectious MARV particles by 50%, a finding consistent with the idea that the activity of the ESCRT machinery supports viral budding but is not essential.  相似文献   

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

14.
The assembly of retroviruses is driven by oligomerization of the Gag polyprotein. We have used cryo-electron tomography together with subtomogram averaging to describe the three-dimensional structure of in vitro-assembled Gag particles from human immunodeficiency virus, Mason-Pfizer monkey virus, and Rous sarcoma virus. These represent three different retroviral genera: the lentiviruses, betaretroviruses and alpharetroviruses. Comparison of the three structures reveals the features of the supramolecular organization of Gag that are conserved between genera and therefore reflect general principles of Gag-Gag interactions and the features that are specific to certain genera. All three Gag proteins assemble to form approximately spherical hexameric lattices with irregular defects. In all three genera, the N-terminal domain of CA is arranged in hexameric rings around large holes. Where the rings meet, 2-fold densities, assigned to the C-terminal domain of CA, extend between adjacent rings, and link together at the 6-fold symmetry axis with a density, which extends toward the center of the particle into the nucleic acid layer. Although this general arrangement is conserved, differences can be seen throughout the CA and spacer peptide regions. These differences can be related to sequence differences among the genera. We conclude that the arrangement of the structural domains of CA is well conserved across genera, whereas the relationship between CA, the spacer peptide region, and the nucleic acid is more specific to each genus.Retrovirus assembly is driven by the oligomerization of Gag, a multidomain protein, including an N-terminal membrane binding domain (MA), a two-domain structural component (CA), and an RNA binding domain (NC). The Gag proteins of all orthoretroviruses, including the alpha-, beta-, and lentiretroviruses discussed here, share this conserved modular architecture (Fig. (Fig.1).1). Despite very weak sequence conservation, the tertiary structures of MA, CA, and NC are conserved among retroviruses. Outside these conserved domains the Gag proteins of different retroviruses exhibit substantial variability. Other domains may be present or absent, and the length and sequence of linker peptides may also vary (12) (Fig. (Fig.11).Open in a separate windowFIG. 1.Modular architecture of the full-length Gag proteins of HIV, M-PMV, and RSV. White rectangles illustrate Gag polyprotein cleavage products. The extent of the constructs used in the electron microscopic analysis is specified under each protein as a black rectangle. Gray triangles specify cleavage sites. Residue numbers are counted from the beginning of Gag.Oligomerization of Gag in an infected cell leads to the formation of roughly spherical immature virus particles, where Gag is arranged in a radial fashion with the N-terminal MA domain associated with a surrounding lipid bilayer, and the more C-terminal NC pointing toward the center of the particle (15, 44, 46). Subsequent multiple cleavages of Gag by the viral protease lead to a rearrangement of the virus. NC and the RNA condense in the center of the particle, CA assembles into a capsid or shell around the nucleoprotein, and MA remains associated with the viral membrane. This proteolytic maturation is required to generate an infectious virion (2). In contrast to the mature CA lattice, which has been extensively studied (11, 16, 36), the Gag lattice in immature particles is incompletely understood.Gag itself contains all of the necessary determinants for particle assembly. For example, the expression of Gag alone in an insect cell expression system is sufficient to generate viruslike particles (3, 17, 22, 38). Retroviral Gag proteins also can be assembled in vitro in the presence of nucleic acids to form spherical particles (9, 19, 39, 43, 47). The arrangement of Gag within these in vitro-assembled Gag particles is indistinguishable from that found in immature virus particles (6), and the in vitro assembly systems have proved valuable for unraveling the principles of virus assembly (18, 28, 29, 39). Multiple layers of interaction promote the assembly of Gag in vivo, including MA-membrane-MA interactions, CA-CA interactions, and NC-RNA-NC interactions. An extensive body of literature has explored which regions of Gag are required for assembly and which can be replaced or deleted without compromising assembly. MA-membrane-MA interactions contribute but are not essential. NC-RNA-NC interactions appear to function to nonspecifically link Gag molecules together and can be replaced both in vivo and in vitro by other interaction domains such as leucine zippers (4, 13, 20, 32, 48). The C-terminal domain of CA (referred to here as C-CA) and the stretch of amino acids immediately following this domain (termed the spacer peptide [SP] region) are critical for assembly and sensitive to mutation (1, 22, 27, 30).We set out to understand how the substantial sequence variation among Gag proteins in different retroviruses is manifested in structural differences in the immature Gag lattice. To do this, we studied three retroviruses from different genera: the lentivirus human immunodeficiency virus type 1 (HIV-1), the betaretrovirus Mason-Pfizer monkey virus (M-PMV), and the alpharetrovirus Rous sarcoma virus (RSV). These retroviruses are those for which in vitro assembly was first established and has been most extensively studied (6, 19, 24, 28, 29, 35, 43, 47).The domain structures of the three retroviruses differ most substantially upstream of CA. Both M-PMV and RSV have domains located between MA and CA that are absent in HIV (Fig. (Fig.1).1). In M-PMV there are 198 residues forming the pp24 and p12 domains; in RSV there are 84 residues forming the p2a, p2b, and p10 domains. The three retroviruses have different requirements for regions upstream of CA during assembly. The C-terminal 25 residues of p10 are essential for proper immature RSV assembly, both in vitro and in vivo, and these residues are inferred to interact directly with N-CA to stabilize the hexamer by forming contacts between adjacent N-CA domains (35). An equivalent assembly domain has not been described for other retroviruses. Within M-PMV p12 is the so-called internal scaffolding domain that is not essential for assembly in vitro (43) but is required for particle assembly when the precursor is expressed under the control of the M-PMV promoter (41). It is a key domain for the membrane-independent assembly of immature capsids (40).In HIV, five residues upstream of CA must be present for assembly of immature virus-like spherical particles in vitro, although larger upstream extensions, including part of MA, are required for efficient assembly of regular particles, both for HIV and RSV. For HIV, if the entire MA domain is included, in vitro assembly requires the presence of inositol penta- or hexakis phosphate (8). If no sequences upstream of CA are present, the in vitro particles in both of these viruses adopt a mature-type tubular morphology (10, 18). It has been hypothesized that cleavage at the N terminus of N-CA during maturation leads to the N-terminal residues of CA folding back into the N-CA structure to form a β-hairpin. The β-hairpin is important for assembly of the mature CA lattice, whereas its absence is important for immature assembly (23, 42). These requirements explain why, in HIV and RSV, immature Gag lattice-like structures are formed only if regions upstream of CA are present (18). In M-PMV, an immature Gag lattice can be produced when the regions upstream of CA are deleted if this is combined with mutations (such as deleting the initial proline of CA), which prevent β-hairpin formation (43).During maturation, HIV and RSV Gag proteins are cleaved twice between CA and NC to release a small peptide called SP1 or SP. In RSV the most N-terminal of these two cleavages can occur at one of two possible positions such that the released peptide is either 9 or 12 amino acids long (33). In M-PMV only one cleavage occurs between CA and NC, and no short peptide is produced. The region between the final helix of CA and the Zn fingers has been proposed to adopt a helical bundle architecture in HIV and RSV based on bioinformatic prediction, on mutational analysis, and on structural studies (1, 22, 27, 45). In all three viruses, C-CA and the residues immediately downstream are critical for assembly and are sensitive to mutation. C-CA contains the major homology region, a group of residues that are highly conserved across the retroviruses.Cryo-electron tomography (cET) studies of immature virus particles (6, 45) have resolved the electron density of the HIV Gag lattice in three dimensions at low resolution. Using these methods, we have also described the three-dimensional architecture of in vitro-assembled HIV Gag particles (6). In immature viruses and in vitro-assembled particles, Gag is seen to adopt an 8 nm hexameric lattice, as was predicted from previous Fourier analysis of two-dimensional images (7, 46). The hexameric lattice is interrupted by irregularly shaped holes and cracks in the lattice (6, 45). A similar observation has been made using AFM of in vitro-assembled particles of M-PMV Gag (26). These holes and cracks allow an otherwise planar hexameric lattice to form the surface of an approximately spherical particle.The radial positions of the MA, CA, and NC domains had been assigned previously from cryo-electron micrographs (44, 46). Based on these assignments and the shape of the density, the position and relative orientations of CA domains can be modeled into the low-resolution structure of the HIV lattice (6, 45). Density ascribed to the N-terminal domain of CA (N-CA) forms rings around large holes at the 6-fold symmetry positions in the lattice. Below this layer, at the expected radius of the C-CA, are 2-fold densities, interpreted as corresponding to dimers of C-CA. These densities are linked by rodlike densities, which descend into the NC-nucleic acid layer.HIV is the only retrovirus for which the arrangement of Gag in the immature particle has been described in three dimensions. Prior to this work, important open questions were therefore: which features of the arrangement of Gag are conserved between genera and therefore reflect general principles of Gag-Gag interactions, and which features are specific to certain genera? We have applied subtomogram averaging of cryo-electron tomograms to generate reconstructions of in vitro-assembled Gag particles from HIV, M-PMV, and RSV. These allow identification of the general and variable features of the arrangement of Gag and the architecture of immature retroviruses.  相似文献   

15.
16.
Membrane glycoproteins of alphavirus play a critical role in the assembly and budding of progeny virions. However, knowledge regarding transport of viral glycoproteins to the plasma membrane is obscure. In this study, we investigated the role of cytopathic vacuole type II (CPV-II) through in situ electron tomography of alphavirus-infected cells. The results revealed that CPV-II contains viral glycoproteins arranged in helical tubular arrays resembling the basic organization of glycoprotein trimers on the envelope of the mature virions. The location of CPV-II adjacent to the site of viral budding suggests a model for the transport of structural components to the site of budding. Thus, the structural characteristics of CPV-II can be used in evaluating the design of a packaging cell line for replicon production.Semliki Forest virus (SFV) is an enveloped alphavirus belonging to the family Togaviridae. This T=4 icosahedral virus particle is approximately 70 nm in diameter (30) and consists of 240 copies of E1/E2 glycoprotein dimers (3, 8, 24). The glycoproteins are anchored in a host-derived lipid envelope that encloses a nucleocapsid, made of a matching number of capsid proteins and a positive single-stranded RNA molecule. After entry of the virus via receptor-mediated endocytosis, a low-pH-induced fusion of the viral envelope with the endosomal membrane delivers the nucleocapsid into the cytoplasm, where the replication events of SFV occur (8, 19, 30). Replication of the viral genome and subsequent translation into structural and nonstructural proteins followed by assembly of the structural proteins and genome (7) lead to budding of progeny virions at the plasma membrane (18, 20). The synthesis of viral proteins shuts off host cell macromolecule synthesis, which allows for efficient intracellular replication of progeny virus (7). The expression of viral proteins leads to the formation of cytopathic vacuolar compartments as the result of the reorganization of cellular membrane in the cytoplasm of an infected cell (1, 7, 14).Early studies using electron microscopy (EM) have characterized the cytopathic vacuoles (CPVs) in SFV-infected cells (6, 13, 14) and identified two types of CPV, namely, CPV type I (CPV-I) and CPV-II. It was found that CPV-I is derived from modified endosomes and lysosomes (18), while CPV-II is derived from the trans-Golgi network (TGN) (10, 11). Significantly, the TGN and CPV-II vesicles are the major membrane compartments marked with E1/E2 glycoproteins (9, 11, 12). Inhibition by monensin results in the accumulation of E1/E2 glycoproteins in the TGN (12, 26), thereby indicating the origin of CPV-II. While CPV-II is identified as the predominant vacuolar structure at the late stage of SFV infection, the exact function of this particular cytopathic vacuole is less well characterized than that of CPV-I (2, 18), although previous observations have pointed to the involvement of CPV-II in budding, because an associated loss of viral budding was observed when CPV-II was absent (9, 36).In this study, we characterized the structure and composition of CPV-II in SFV-infected cells in situ with the aid of electron tomography and immuno-electron microscopy after physical fixation of SFV-infected cells by high-pressure freezing and freeze substitution (21, 22, 33). The results revealed a helical array of E1/E2 glycoproteins within CPV-II and indicate that CPV-II plays an important role in intracellular transport of glycoproteins prior to SFV budding.  相似文献   

17.
ACK (activated Cdc42-associated tyrosine kinase) (also Tnk2) is an ubiquitin-binding protein and plays an important role in ligand-induced and ubiquitination-mediated degradation of epidermal growth factor receptor (EGFR). Here we report that ACK is ubiquitinated by HECT E3 ubiquitin ligase Nedd4-1 and degraded along with EGFR in response to EGF stimulation. ACK interacts with Nedd4-1 through a conserved PPXY WW-binding motif. The WW3 domain in Nedd4-1 is critical for binding to ACK. Although ACK binds to both Nedd4-1 and Nedd4-2 (also Nedd4L), Nedd4-1 is the E3 ubiquitin ligase for ubiquitination of ACK in cells. Interestingly, deletion of the sterile alpha motif (SAM) domain at the N terminus dramatically reduced the ubiquitination of ACK by Nedd4-1, while deletion of the Uba domain dramatically enhanced the ubiquitination. Use of proteasomal and lysosomal inhibitors demonstrated that EGF-induced ACK degradation is processed by lysosomes, not proteasomes. RNA interference (RNAi) knockdown of Nedd4-1, not Nedd4-2, inhibited degradation of both EGFR and ACK, and overexpression of ACK mutants that are deficient in either binding to or ubiquitination by Nedd4-1 blocked EGF-induced degradation of EGFR. Our findings suggest an essential role of Nedd4-1 in regulation of EGFR degradation through interaction with and ubiquitination of ACK.Activated Cdc42-associated tyrosine kinase (ACK) (also Tnk2) is a member of the type VIII tyrosine kinase family. Activation of ACK, including both ACK1 and ACK2, occurs in response to signaling of epidermal growth factor receptor (EGFR), platelet-derived growth factor (PDGF) receptor, insulin receptor, Gas-6 receptor (Mer), M3 muscarinic receptor, integrins, or proteoglycan (3, 7, 11, 23, 26, 30, 44, 47). In Drosophila, D-ACK mediates the function of Cdc42 in dorsal closure during embryonic development (31). The ACK homologue, Ark-1, in Caenorhabditis elegans negatively regulates EGF signaling (15).A number of studies suggest a role for ACK in EGFR degradation. ACK1 and ACK2, two alternatively spliced isoforms, possess a highly conserved clathrin-binding motif and interact with clathrin (37, 45). Overexpression of ACK2 severely impairs transferrin receptor endocytosis, causes aberrant localization of AP-2, and induces changes in clathrin assembly. Furthermore, ACK2 interacts with sorting nexin 9 (SNX9, also named SH3PX1), a member of the sorting nexin family, via its proline-rich domain 1 and phosphorylates SNX9 to facilitate the degradation of EGF receptors (22). In C. elegans, Ark-1 genetically interacts with UNC101, the homologue of mammalian clathrin-associated protein AP47, and SLI-1, the homologue of mammalian Cbl that is an E3 ubiquitin ligase for ubiquitination of EGFR, and negatively regulates EGFR signaling (15).Our previous studies showed that ACK1 interacts with EGFR upon EGF stimulation via a region at the carboxyl terminus, designated the EGFR-binding domain (EBD), which is highly homologous to the EGFR/ErbB2-binding domain of Gene-33/Mig-6/RALT (32, 43). The interaction of ACK1 with EGFR is dependent on kinase activity and tyrosine phosphorylation of EGFR. Immunofluorescent staining using anti-EGFR and GFP-ACK1 indicates that ACK1 is colocalized with EGFR on large vacuolar structures upon EGF stimulation. Suppression of the expression of ACK1 by ACK-RNA interference (RNAi) inhibits ligand-induced degradation of EGFR, suggesting that ACK1 plays an important role in the regulation of EGFR degradation in cells. Furthermore, we identified ACK1 as an ubiquitin-binding protein. Through an ubiquitin association (Uba) domain at the carboxyl terminus, ACK1 is capable of interacting with both poly- and monoubiquitin. Overexpression of an Uba domain deletion mutant of ACK1 blocked the ligand-dependent degradation of EGFR, suggesting that ACK1 regulates EGFR degradation via its Uba domain. Thus, ACK1 senses EGF signaling and regulates degradation of EGFR.EGF-induced degradation of EGFR is mediated by ubiquitination (16). The ubiquitination of EGFR is activated upon EGF stimulation by recruiting the RING family E3 ubiquitin ligase Cbl to pY1045 (20, 21). This ubiquitination functions as a sorting signal for transporting EGFR to lysosomes for degradation (14). Nedd4, the HECT domain-containing E3 ubiquitin ligase, is also involved in the regulation of EGFR trafficking by ubiquitination of endocytic or vesicle sorting proteins (28). For example, it has been observed that Nedd4 ubiquitinates Cbl, Eps15, Tsg101, Hrs, and secretory carrier membrane proteins (SCAMPs) and participates in the processes of EGFR endocytosis and degradation (1, 18, 25, 42). However, exactly how Nedd4 engages in the EGFR degradation process in response to EGF stimulation is not known.In this report, we show that EGF stimulation induces ACK degradation. This degradation is associated with ubiquitination of ACK. Nedd4-1, but not Nedd4-2, is identified as the E3 ubiquitin ligase for ubiquitination of ACK. Furthermore, EGF-induced degradation of ACK is EGFR activation dependent and processed by lysosomes. RNAi knockdown and mutational analysis demonstrated that Nedd4-1 and Nedd4-1-catalyzed ubiquitination of ACK are required for EGF-induced degradation of EGFR and ACK. Our findings suggest a new mechanism in regulation of EGFR degradation.  相似文献   

18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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