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 780 毫秒
1.
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   

2.
3.
Canine parvovirus (CPV) and feline panleukopenia virus (FPV) are closely related parvoviruses that differ in their host ranges for cats and dogs. Both viruses bind their host transferrin receptor (TfR), enter cells by clathrin-mediated endocytosis, and traffic with that receptor through endosomal pathways. Infection by these viruses appears to be inefficient and slow, with low numbers of virions infecting the cell after a number of hours. Species-specific binding to TfR controls viral host range, and in this study FPV and strains of CPV differed in the levels of cell attachment, uptake, and infection in canine and feline cells. During infection, CPV particles initially bound and trafficked passively on the filopodia of canine cells while they bound to the cell body of feline cells. That binding was associated with the TfR as it was disrupted by anti-TfR antibodies. Capsids were taken up from the cell surface with different kinetics in canine and feline cells but, unlike transferrin, most did not recycle. Capsids labeled with fluorescent markers were seen in Rab5-, Rab7-, or Rab11-positive endosomal compartments within minutes of uptake, but reached the nucleus. Constitutively active or dominant negative Rab mutants changed the intracellular distribution of capsids and affected the infectivity of virus in cells.Cell infection by animal viruses involves a specific sequence of steps that deliver the virus and its genome from the cell surface to the compartment where replication can occur. For nonenveloped viruses, infection initiates with binding to a specific cell receptor and uptake into the cell by receptor-mediated endocytosis. Various factors can control the process of viral uptake, including the characteristics of the receptor(s) bound by the virus and its signaling and endocytic properties, the affinity of the virus for the receptor, and the structural features of the interaction in different environments (36, 61). Receptors may be located on the cell body or may also be displayed on the extended lamellipodia or filopodia with greater surface areas. Viruses binding to filopodia can be either passively delivered to the cell body for endocytosis by dynamic movement of the entire structure or actively trafficked by retrograde actin transport as well as the action of myosin-2 motors on the actin (32, 57). Cross-linking and clustering of receptors by viral particles can influence the rate and pathways of uptake from the cell surface (23), and many viral receptors activate signaling pathways that alter the structure of the underlying cytoskeleton to enhance uptake (see, e.g., references 12, 30, and 51). Receptor-bound viruses then enter one or more endosomal pathways that results in the capsid being enclosed in vesicles and trafficked within the endosomal pathways of the cell, where clustered virus and receptors (23) may undergo structural alterations upon exposure to conditions such as low pH or proteases (36, 61). The specific receptor-mediated binding and entry pathways often provide signals for viruses that allow endosomal escape and establish infection. A variety of markers of the endosomal compartments have been used in studies of viral entry. Rab proteins are monomeric small GTPases which regulate endosomal membrane trafficking, and specific Rab proteins are associated with different endosomal compartments. Among the many Rab proteins in the cells, Rab5 is primarily associated with the early endosome and regulates trafficking through that compartment, Rab7 is associated with the late endosome, and Rab11 is associated with the recycling endosome (14, 58). Tracking viral particles within the endosomal pathways during cell entry has been used to define the steps in the entry and infection processes of a variety of different viruses and has revealed many of the common features and variant processes that are used (7-9, 33, 71).Here, we examine the uptake and infection of cells by parvovirus capsids and compare some of the steps followed by capsids that differ in their receptor binding properties and host ranges. Feline panleukopenia virus (FPV) infects cats (50, 66), binds the transferrin receptor-1 (TfR) on feline cells, and uses that receptor for uptake and infection (27, 44). FPV does not bind the canine TfR or infect dogs or cultured canine cells. Canine parvovirus (CPV) is a natural variant of FPV which emerged in 1978 after acquiring a small number of mutations that allow its capsid to bind the canine TfR (27). The original strain of CPV (designated CPV type 2 [CPV-2]) spread worldwide in dogs during 1978, but some of the same mutations that gave it the canine host range rendered it unable to infect cats (66, 67). CPV-2 was replaced worldwide during 1979 and 1980 by a natural variant, CPV type 2a (CPV-2a), which contained an additional four to five changes in its coat protein gene (48, 49). Subsequently, the canine viruses have continued to evolve, and additional single mutations have been selected that alter antigenic epitopes. Strains altered at VP2 residue 426 are designated CPV-2b (Asn426Asp) and CPV-2c (Asp426Glu) (13, 48). CPV-2a and its variants are able to infect both dogs and cats but show reduced binding to the feline TfR on cells and in vitro (27, 42). In addition, the affinity of binding to the canine TfR is much lower than that seen for the feline TfR (42).The TfR is a type II membrane protein expressed in nonlipid raft regions of the plasma membrane, and it binds iron-loaded (holo) transferrin (Tf) at neutral pH (2). TfR expression is tightly regulated, and it is more highly expressed on dividing cells with high iron needs, which would favor binding of these viruses. The TfRs of mice and humans are used as receptors for cell infection by the mouse mammary tumor virus and the New World hemorrhagic fever arenaviruses (52, 56).The TfR is assembled as a homodimer, and each monomer of the ectodomain is composed of protease-like, apical and helical domains, as well as a 30-Å membrane-proximal stalk (5, 20, 31). The transmembrane domain mediates membrane insertion and influences some aspects of trafficking within the cell, while the cytoplasmic domain contains a tyrosine-threonine-arginine-phenylalanine (YTRF) sequence that engages the clathrin-mediated endocytic machinery through AP-2 (adaptor protein-2) (53, 55). The TfR sequence also includes one or two cysteines adjacent to the inner leaflet of the membrane that may be palmitoylated to influence the rate of receptor recycling, and it also contains sequences that control basolateral localization in polarized cells (41). In the normal pathway of TfR-mediated entry, the TfR-holo-Tf complex is transported into the endosomal system, where low pH results in conformational changes and iron release. The TfR-Tf complex enters the early endosome, from which some of the complex is rapidly recycled to the cell surface while most passes to the perinuclear recycling endosome. From there it recycles to the cell surface where the iron-free apo-Tf is released at neutral pH (21, 22, 24, 37, 38, 69, 70). The rate of uptake and the efficiency of TfR recycling depend on the form of the ligand, and more than 97% of monomeric Tf recycles to the cell surface within 10 to 30 min. However, cross-linking TfRs with oligomeric Tf or antibodies causes the complexes to be retained within endosomes for longer times, and a higher proportion is trafficked to late endosomes and lysosomes for degradation (35).Holo-Tf binds the membrane-proximal side of the feline and canine TfR ectodomain (11), while virus binding involves the apical domain of the receptor as mutations in that structure affect the ability to bind FPV and CPV capsids (43) The feline and canine TfRs differ in ∼10% of their sequences, but a major difference controlling the CPV-specific binding is a unique glycosylation site in the apical domain of the canine TfR (43). Alteration of the glycosylated Asn to Lys (the feline TfR residue) allowed the canine TfR to bind FPV and also greatly increased the affinity of binding to CPV-2 and CPV-2a-related capsids (42).CPV and FPV have small (25 nm) nonenveloped capsids that package a single-stranded DNA genome of ∼5,120 bases (68). The particles are made up of two overlapping proteins, VP1 and VP2, with 90% of the capsid protein being VP2. VP1 contains a 143-residue amino (N)-terminal sequence that encodes a phospholipase A2 enzymatic activity, as well as basic amino acid motifs that play a role in nuclear localization (72). The VP1 unique region becomes exposed during cell entry without capsid disintegration, and the phospholipase A2 modifies the endosomal membrane to enhance endosomal escape (19, 75).Previous studies of cell entry by CPV, minute virus of mice, and various adeno-associated viruses (AAVs) show that viral uptake primarily occurs through clathrin-mediated endocytosis. However, when the AP-2-interacting sequences in the cytoplasmic tail of the feline TfR were mutated or deleted, the altered receptor still allowed CPV infection at a similar efficiency to that of wild-type TfR (26). The intracellular pathways of viral entry and trafficking have been examined by using cells fixed at various times after uptake and then antibody stained for virus and cellular markers or by expressing green fluorescent protein (GFP)-labeled markers. Time courses examined were between 1 and 6 h, and sequential steps of trafficking were suggested, with the virus passed from the early endosomes to the recycling endosome, followed by localization in late endosomes and lysosomes after uptake (65). By fluorescent antibody staining, VP1 release occurred only hours after uptake, possibly in a low-pH degradative compartment (64, 72). In addition, CPV capsids appear to remain associated with the receptor for 1 to 2 h after virus uptake as antibodies against the TfR cytoplasmic tail microinjected into feline CRFK cells block infection in this time period (44). Infection is also blocked by neutralizing the low pH of the endosomal system with ammonium chloride or bafilomycin A1, although it is not clear whether this is due to direct effects on the capsid or to indirect alterations in endosomal trafficking. When the X-ray crystal structures of capsids of CPV and FPV were determined at low pH or in the presence of EDTA or when capsids were examined for changes in protease susceptibility, only small changes in surface loops of the viral structure were present (40, 60).Here, we used microscopy to examine dynamic steps in the binding, uptake, and early trafficking of parvovirus capsids in live canine and feline cells. Labeled capsids were seen to undergo rapid movement into multiple endosomal compartments shortly after entry. Initial binding of CPV to canine cells involved filopodia while in feline cells the virus bound primarily to receptors on the cell body. In cells expressing GFP-conjugated Rab proteins, particles rapidly localized to multiple endosomal compartments in the cytoplasm after uptake, which gradually accumulated near the microtubule-organizing center. The distribution of intracellular viruses and the viral infectivity in feline cells were altered by expression of either constitutively active (CA) or dominant negative (DN) mutants of the Rab proteins.  相似文献   

4.
The human papillomavirus type 16 E5 oncoprotein (16E5) enhances acute, ligand-dependent activation of the epidermal growth factor receptor (EGFR) and concomitantly alkalinizes endosomes, presumably by binding to the 16-kDa “c” subunit of the V-ATPase proton pump (16K) and inhibiting V-ATPase function. However, the relationship between 16K binding, endosome alkalinization, and altered EGFR signaling remains unclear. Using an antibody that we generated against 16K, we found that 16E5 associated with only a small fraction of endogenous 16K in keratinocytes, suggesting that it was unlikely that E5 could significantly affect V-ATPase function by direct inhibition. Nevertheless, E5 inhibited the acidification of endosomes, as determined by a new assay using a biologically active, pH-sensitive fluorescent EGF conjugate. Since we also found that 16E5 did not alter cell surface EGF binding, the number of EGFRs on the cell surface, or the endocytosis of prebound EGF, we postulated that it might be blocking the fusion of early endosomes with acidified vesicles. Our studies with pH-sensitive and -insensitive fluorescent EGF conjugates and fluorescent dextran confirmed that E5 prevented endosome maturation (acidification and enlargement) by inhibiting endosome fusion. The E5-dependent defect in vesicle fusion was not due to detectable disruption of actin, tubulin, vimentin, or cytokeratin filaments, suggesting that membrane fusion was being directly affected rather than vesicle transport. Perhaps most importantly, while bafilomycin A1 (like E5) binds to 16K and inhibits endosome acidification, it did not mimic the ability of E5 to inhibit endosome enlargement or the trafficking of EGF. Thus, 16E5 alters EGF endocytic trafficking via a pH-independent inhibition of vesicle fusion.High-risk human papillomaviruses (HPVs) are the causative agent of cervical cancer (63) and HPV type 16 (HPV-16) is associated with a majority of cervical malignancies worldwide (13). HPV-16 encodes three oncoproteins: E5, E6, and E7. While the contributions of E6 and E7 to cellular immortalization and transformation have been characterized in detail (20), the role of HPV-16 E5 (16E5) is poorly understood (53). Nevertheless, a number of studies suggest that 16E5 does contribute to the development of cervical cancer. Most high-risk HPV types encode an E5 protein (48), and targeted expression of the three HPV-16 oncogenes in basal epithelial cells of transgenic mice (4) leads to a higher incidence of cervical cancer than does the expression of E6 and E7 alone (44). In addition, targeted epithelial expression of 16E5 (without E6 and E7) in transgenic mice induces skin tumors (21). It may be noteworthy that unlike high-risk HPV-18, which integrates into the host DNA and potentially disrupts E5 gene expression (20, 64), the HPV-16 genome often persists in episomal form in malignant lesions (12, 16, 24, 36, 42).Biological activities of 16E5 that may facilitate carcinogenesis include evading host immune detection by interfering with the transport of antigen-presenting major histocompatibility complex (MHC) class I molecules to the cell surface (6), promoting anchorage-independent growth (33, 41, 52) and disrupting gap junctions responsible for cell-cell communication (37, 58). The 16E5 phenotype most frequently linked to the development of cancer is enhanced ligand-dependent activation of the epidermal growth factor receptor (EGFR) (15, 41, 46, 52). 16E5 stimulates EGF-dependent cell proliferation in vitro (7, 33, 40, 41, 52, 60) and in vivo (21), which might expand the population of basal or stemlike keratinocytes and thereby increase the probability that some of these cells would undergo malignant transformation. A number of studies indicate that 16E5 may enhance ligand-dependent EGFR activation by interfering with the acidification of early endosomes containing EGF bound to activated EGFRs (17, 51, 57). It has been hypothesized that 16E5 inhibits the H+ V-ATPase responsible for maintaining an acidic luminal pH in late endosomes and lysosomes (28) by associating with the V-ATPase 16-kDa “c” subunit (16K) (1, 5, 14, 22, 46) and disrupting assembly of the V-ATPase integral (Vo) and peripheral (Vi) subcomplexes (10). In contrast, Thomsen et al. (57) reported that 16E5 inhibits early endosome trafficking in fibroblasts by completely depolymerizing actin microfilaments.Due to the unavailability of antibodies that recognize native 16E5 and 16K, direct association of 16E5 with 16K has only been observed by overexpressing epitope-tagged forms of both proteins in vitro (5, 46) or in vivo (1, 14, 22). It is uncertain, therefore, whether these associations occur when the proteins are expressed at “physiological” levels. In yeast, both wild-type 16E5 (10) and several 16E5 mutants that associate with 16K in COS cells (1) inhibit vacuolar acidification, although another study in yeast concludes the opposite (5). 16K is a component of the V-ATPase Vo subcomplex, which is assembled in the endoplasmic reticulum (ER) (28), and 16E5 localizes to the ER and nuclear envelope in epithelial cells (32, 54). Thus, the export of Vo from the ER could potentially be inhibited by a significant level of 16K binding to 16E5, although the differential alkalinization of endosomes rather than the Golgi apparatus (17) would require specificity for those proton pumps directed to those sites.In the present study, we generated an antibody against native 16K and used it to determine whether 16K/16E5 complexes formed in primary keratinocytes. We also synthesized a new pH-sensitive fluorescent EGF conjugate to evaluate whether there was a correlation between E5-induced EGFR activation, trafficking and endosome alkalinization. Finally, we simultaneously monitored EGFR endocytic trafficking (using pH-insensitive fluorescent EGF), endosome fusion (using fluorescent EGF and dextran), and the status of cellular filaments and microtubules to evaluate whether E5 might disrupt some of these structures that mediate vesicle transport.  相似文献   

5.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

6.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

7.
8.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

9.
The arenavirus envelope glycoprotein (GPC) initiates infection in the host cell through pH-induced fusion of the viral and endosomal membranes. As in other class I viral fusion proteins, this process proceeds through a structural reorganization in GPC in which the ectodomain of the transmembrane fusion subunit (G2) engages the host cell membrane and subsequently refolds to form a highly stable six-helix bundle structure that brings the two membranes into apposition for fusion. Here, we describe a G2-directed monoclonal antibody, F100G5, that prevents membrane fusion by binding to an intermediate form of the protein on the fusion pathway. Inhibition of syncytium formation requires that F100G5 be present concomitant with exposure of GPC to acidic pH. We show that F100G5 recognizes neither the six-helix bundle nor the larger trimer-of-hairpins structure in the postfusion form of G2. Rather, Western blot analysis using recombinant proteins and a panel of alanine-scanning GPC mutants revealed that F100G5 binding is dependent on an invariant lysine residue (K283) near the N terminus of G2, in the so-called fusion peptide that inserts into the host cell membrane during the fusion process. The F100G5 epitope is located in the internal segment of the bipartite GPC fusion peptide, which also contains four conserved cysteine residues, raising the possibility that this fusion peptide may be highly structured. Collectively, our studies indicate that F100G5 identifies an on-path intermediate form of GPC. Binding to the transiently exposed fusion peptide may interfere with G2 insertion into the host cell membrane. Strategies to effectively target fusion peptide function in the endosome may lead to novel classes of antiviral agents.Enveloped viruses enter their target cells through fusion of the virus and cell membranes, in a process promoted by the viral envelope glycoprotein. For some viruses, such as human immunodeficiency virus (HIV), entry is initiated by interaction of the envelope glycoprotein with cell surface receptor proteins. Other viruses, such as influenza virus, are endocytosed and membrane fusion is triggered by exposure to acidic pH in the maturing endosome. The subsequent merger of the viral and cell membranes is accomplished through a major structural reorganization of the envelope glycoprotein. Antiviral strategies that target virus entry by using neutralizing antibodies or small-molecule fusion inhibitors can, in many cases, prevent virus infection and disease.The Arenaviridae comprise a diverse group of rodent-borne viruses, some of which are responsible for severe hemorrhagic fevers in humans. Lassa fever virus (LASV) is endemic in western Africa (59), and at least five New World species are recognized to cause fatal disease in the Americas, including the Argentine hemorrhagic fever virus Junín (JUNV) (63). New pathogenic arenavirus species continue to emerge from their distinct animal reservoirs (1, 11, 24). At present, there are no licensed vaccines or effective therapies to address the threat of arenavirus infection.Arenaviruses are enveloped, negative-strand RNA viruses whose bipartite genome encodes ambisense expression of four viral proteins (12, 22). The arenavirus envelope glycoprotein, GPC, is a member of the class I virus fusion proteins (33, 40, 75), a group that includes HIV Env, influenza virus hemagglutinin (HA), and paramyxovirus F protein. These envelope glycoproteins share several salient features. The precursor glycoproteins assemble as trimeric complexes and are subsequently rendered competent for membrane fusion by a proteolytic cleavage that results in the formation of the mature receptor-binding and transmembrane fusion subunits. The GPC precursor glycoprotein is cleaved by the cellular SKI-1/S1P protease (6, 51, 54) to generate the respective G1 and G2 subunits, which remain noncovalently associated. The ectodomain of the class I fusion subunit is distinguished by the presence of two 4-3 heptad repeat (HR1 and HR2) sequences that, in the course of membrane fusion, refold to form the now-classical six-helix bundle structure, which defines this class of envelope glycoproteins. Unlike other class I fusion proteins, GPC also contains a cleaved and stable signal peptide (SSP) as a third and essential subunit in the mature complex (2, 32, 69, 77, 81).Arenavirus infection is initiated by G1 binding to a cell surface receptor. The pathogenic clade B New World arenaviruses utilize transferrin receptor 1 (TfR1) for entry (1, 64, 65), whereas those in clades A and C, as well as the Old World viruses, bind α-dystroglycan and/or an unknown receptor (15, 34, 71). The virion particle is subsequently endocytosed (9), and membrane fusion is initiated by acidification in the maturing endosome (17, 28, 29). pH-dependent activation of GPC is modulated through a unique interaction between SSP and G2 (79, 80) and can be targeted by small-molecule inhibitors that block membrane fusion (76) and protect against arenavirus infection (8, 52).A generally accepted model for membrane fusion by the class I envelope glycoproteins (reviewed in references 45 and 73) posits that the native complex exists in a metastable state that is established on proteolytic maturation of the biosynthetic precursor. Upon activation, whether by acidic pH in the endosome or receptor binding at the plasma membrane, the fusion subunit that was sequestered in the prefusion state is exposed and undergoes a series of dramatic conformational changes leading to membrane fusion. In this process, a hydrophobic region at or near the N terminus of the fusion subunit (the fusion peptide) inserts into the host cell membrane, thus allowing the protein to bridge the two membranes. This so-called prehairpin intermediate subsequently collapses upon itself to form the highly stable six-helix bundle structure, in which the three HR2 helices pack into hydrophobic grooves on the trimeric HR1 coiled-coil in an antiparallel manner, bringing the virus and cell membranes into apposition. Free energy made available in the formation of this stable structure is thought to drive fusion of the lipid bilayers. Peptides that correspond in sequence to HR2 (C-peptides) bind to the putative prehairpin intermediate and interfere with its refolding, thereby preventing membrane fusion (18, 57, 74). While the structure of the six-helix bundle core has been elucidated in atomic detail (45, 73), information regarding the molecular pathway leading to this postfusion state is largely indirect. Indeed, the prehairpin intermediate is conceptualized through the activity of C-peptide fusion inhibitors (57, 74).In this report, we describe a G2-directed monoclonal antibody (MAb), F100G5, that recognizes a pH-induced intermediate of JUNV GPC and prevents GPC-mediated membrane fusion. This MAb binds at or near the internal fusion peptide of G2 and may act by interfering with its penetration into the host cell membrane. These studies highlight the feasibility of targeting short-lived GPC intermediates for inhibition of membrane fusion.  相似文献   

10.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

11.
12.
13.
Deleting individual genes for outer surface c-type cytochromes in Geobacter sulfurreducens partially inhibited the reduction of humic substances and anthraquinone-2,6,-disulfonate. Complete inhibition was obtained only when five of these genes were simultaneously deleted, suggesting that diverse outer surface cytochromes can contribute to the reduction of humic substances and other extracellular quinones.Humic substances can play an important role in the reduction of Fe(III), and possibly other metals, in sedimentary environments (6, 34). Diverse dissimilatory Fe(III)-reducing microorganisms (3, 5, 7, 9, 11, 19-22, 25) can transfer electrons onto the quinone moieties of humic substances (38) or the model compound anthraquinone-2,6-disulfonate (AQDS). Reduced humic substances or AQDS abiotically reduces Fe(III) to Fe(II), regenerating the quinone. Electron shuttling in this manner can greatly increase the rate of electron transfer to insoluble Fe(III) oxides, presumably because soluble quinone-containing molecules are more accessible for microbial reduction than insoluble Fe(III) oxides (19, 22). Thus, catalytic amounts of humic substances have the potential to dramatically influence rates of Fe(III) reduction in soils and sediments and can promote more rapid degradation of organic contaminants coupled to Fe(III) reduction (1, 2, 4, 10, 24).To our knowledge, the mechanisms by which Fe(III)-reducing microorganisms transfer electrons to humic substances have not been investigated previously for any microorganism. However, reduction of AQDS has been studied using Shewanella oneidensis (17, 40). Disruption of the gene for MtrB, an outer membrane protein required for proper localization of outer membrane cytochromes (31), inhibited reduction of AQDS, as did disruption of the gene for the outer membrane c-type cytochrome, MtrC (17). However, in each case inhibition was incomplete, and it was suggested that there was a possibility of some periplasmic reduction (17), which would be consistent with the ability of AQDS to enter the cell (40).The mechanisms for electron transfer to humic substances in Geobacter species are of interest because molecular studies have frequently demonstrated that Geobacter species are the predominant Fe(III)-reducing microorganisms in sedimentary environments in which Fe(III) reduction is an important process (references 20, 32, and 42 and references therein). Geobacter sulfurreducens has routinely been used for investigations of the physiology of Geobacter species because of the availability of its genome sequence (29), a genetic system (8), and a genome-scale metabolic model (26) has made it possible to take a systems biology approach to understanding the growth of this organism in sedimentary environments (23).  相似文献   

14.
15.
16.
Several mycoplasma species feature a membrane protrusion at a cell pole, and unknown mechanisms provide gliding motility in the direction of the pole defined by the protrusion. Mycoplasma gallisepticum, an avian pathogen, is known to form a membrane protrusion composed of bleb and infrableb and to glide. Here, we analyzed the gliding motility of M. gallisepticum cells in detail. They glided in the direction of the bleb at an average speed of 0.4 μm/s and remained attached around the bleb to a glass surface, suggesting that the gliding mechanism is similar to that of a related species, Mycoplasma pneumoniae. Next, to elucidate the cytoskeletal structure of M. gallisepticum, we stripped the envelopes by treatment with Triton X-100 under various conditions and observed the remaining structure by negative-staining transmission electron microscopy. A unique cytoskeletal structure, about 300 nm long and 100 nm wide, was found in the bleb and infrableb. The structure, resembling an asymmetrical dumbbell, is composed of five major parts from the distal end: a cap, a small oval, a rod, a large oval, and a bowl. Sonication likely divided the asymmetrical dumbbell into a core and other structures. The cytoskeletal structures of M. gallisepticum were compared with those of M. pneumoniae in detail, and the possible protein components of these structures were considered.Mycoplasmas are commensal and occasionally pathogenic bacteria that lack a peptidoglycan layer (50). Several species feature a membrane protrusion at a pole; for Mycoplasma mobile, this protrusion is called the head, and for Mycoplasma pneumoniae, it is called the attachment organelle (25, 34-37, 52, 54, 58). These species bind to solid surfaces, such as glass and animal cell surfaces, and exhibit gliding motility in the direction of the protrusion (34-37). This motility is believed to be essential for the mycoplasmas'' pathogenicity (4, 22, 27, 36). Recently, the proteins directly involved in the gliding mechanisms of mycoplasmas were identified and were found to have no similarities to those of known motility systems, including bacterial flagellum, pilus, and slime motility systems (25, 34-37).Mycoplasma gallisepticum is an avian pathogen that causes serious damage to the production of eggs for human consumption (50). The cells are pear-shaped and have a membrane protrusion, consisting of the so-called bleb and infrableb (29), and gliding motility (8, 14, 22). Their putative cytoskeletal structures may maintain this characteristic morphology because M. gallisepticum, like other mycoplasma species, does not have a cell wall (50). In sectioning electron microscopy (EM) studies of M. gallisepticum, an intracellular electron-dense structure in the bleb and infrableb was observed, suggesting the existence of a cytoskeletal structure (7, 24, 29, 37, 58). Recently, the existence of such a structure has been confirmed by scanning EM of the structure remaining after Triton X-100 extraction (13), although the details are still unclear.A human pathogen, M. pneumoniae, has a rod-shaped cytoskeletal structure in the attachment organelle (9, 15, 16, 31, 37, 57). M. gallisepticum is related to M. pneumoniae (63, 64), as represented by 90.3% identity between the 16S rRNA sequences, and it has some open reading frames (ORFs) homologous to the component proteins of the cytoskeletal structures of M. pneumoniae (6, 17, 48). Therefore, the cytoskeletal structures of M. gallisepticum are expected to be similar to those of M. pneumoniae, as scanning EM images also suggest (13).The fastest-gliding species, M. mobile, is more distantly related to M. gallisepticum; it has novel cytoskeletal structures that have been analyzed through negative-staining transmission EM after extraction by Triton X-100 with image averaging (45). This method of transmission EM following Triton X-100 extraction clearly showed a cytoskeletal “jellyfish” structure. In this structure, a solid oval “bell,” about 235 nm wide and 155 nm long, is filled with a 12-nm hexagonal lattice. Connected to this bell structure are dozens of flexible “tentacles” that are covered with particles 20 nm in diameter at intervals of about 30 nm. The particles appear to have 180° rotational symmetry and a dimple at the center. The involvement of this cytoskeletal structure in the gliding mechanism was suggested by its cellular localization and by analyses of mutants lacking proteins essential for gliding.In the present study, we applied this method to M. gallisepticum and analyzed its unique cytoskeletal structure, and we then compared it with that of M. pneumoniae.  相似文献   

17.
18.
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   

19.
An intracellular multiplication F (IcmF) family protein is a conserved component of a newly identified type VI secretion system (T6SS) encoded in many animal and plant-associated Proteobacteria. We have previously identified ImpLM, an IcmF family protein that is required for the secretion of the T6SS substrate hemolysin-coregulated protein (Hcp) from the plant-pathogenic bacterium Agrobacterium tumefaciens. In this study, we characterized the topology of ImpLM and the importance of its nucleotide-binding Walker A motif involved in Hcp secretion from A. tumefaciens. A combination of β-lactamase-green fluorescent protein fusion and biochemical fractionation analyses revealed that ImpLM is an integral polytopic inner membrane protein comprising three transmembrane domains bordered by an N-terminal domain facing the cytoplasm and a C-terminal domain exposed to the periplasm. impLM mutants with substitutions or deletions in the Walker A motif failed to complement the impLM deletion mutant for Hcp secretion, which provided evidence that ImpLM may bind and/or hydrolyze nucleoside triphosphates to mediate T6SS machine assembly and/or substrate secretion. Protein-protein interaction and protein stability analyses indicated that there is a physical interaction between ImpLM and another essential T6SS component, ImpKL. Topology and biochemical fractionation analyses suggested that ImpKL is an integral bitopic inner membrane protein with an N-terminal domain facing the cytoplasm and a C-terminal OmpA-like domain exposed to the periplasm. Further comprehensive yeast two-hybrid assays dissecting ImpLM-ImpKL interaction domains suggested that ImpLM interacts with ImpKL via the N-terminal cytoplasmic domains of the proteins. In conclusion, ImpLM interacts with ImpKL, and its Walker A motif is required for its function in mediation of Hcp secretion from A. tumefaciens.Many pathogenic gram-negative bacteria employ protein secretion systems formed by macromolecular complexes to deliver proteins or protein-DNA complexes across the bacterial membrane. In addition to the general secretory (Sec) pathway (18, 52) and twin-arginine translocation (Tat) pathway (7, 34), which transport proteins across the inner membrane into the periplasm, at least six distinct protein secretion systems occur in gram-negative bacteria (28, 46, 66). These systems are able to secrete proteins from the cytoplasm or periplasm to the external environment or the host cell and include the well-documented type I to type V secretion systems (T1SS to T5SS) (10, 15, 23, 26, 30) and a recently discovered type VI secretion system (T6SS) (4, 8, 22, 41, 48, 49). These systems use ATPase or a proton motive force to energize assembly of the protein secretion machinery and/or substrate translocation (2, 6, 41, 44, 60).Agrobacterium tumefaciens is a soilborne pathogenic gram-negative bacterium that causes crown gall disease in a wide range of plants. Using an archetypal T4SS (9), A. tumefaciens translocates oncogenic transferred DNA and effector proteins to the host and ultimately integrates transferred DNA into the host genome. Because of its unique interkingdom DNA transfer, this bacterium has been extensively studied and used to transform foreign DNA into plants and fungi (11, 24, 40, 67). In addition to the T4SS, A. tumefaciens encodes several other secretion systems, including the Sec pathway, the Tat pathway, T1SS, T5SS, and the recently identified T6SS (72). T6SS is highly conserved and widely distributed in animal- and plant-associated Proteobacteria and plays an important role in the virulence of several human and animal pathogens (14, 19, 41, 48, 56, 63, 74). However, T6SS seems to play only a minor role or even a negative role in infection or virulence of the plant-associated pathogens or symbionts studied to date (5, 37-39, 72).T6SS was initially designated IAHP (IcmF-associated homologous protein) clusters (13). Before T6SS was documented by Pukatzki et al. in Vibrio cholerae (48), mutations in this gene cluster in the plant symbiont Rhizobium leguminosarum (5) and the fish pathogen Edwardsiella tarda (51) caused defects in protein secretion. In V. cholerae, T6SS was responsible for the loss of cytotoxicity for amoebae and for secretion of two proteins lacking a signal peptide, hemolysin-coregulated protein (Hcp) and valine-glycine repeat protein (VgrG). Secretion of Hcp is the hallmark of T6SS. Interestingly, mutation of hcp blocks the secretion of VgrG proteins (VgrG-1, VgrG-2, and VgrG-3), and, conversely, vgrG-1 and vgrG-2 are both required for secretion of the Hcp and VgrG proteins from V. cholerae (47, 48). Similarly, a requirement of Hcp for VgrG secretion and a requirement of VgrG for Hcp secretion have also been shown for E. tarda (74). Because Hcp forms a hexameric ring (41) stacked in a tube-like structure in vitro (3, 35) and VgrG has a predicted trimeric phage tail spike-like structure similar to that of the T4 phage gp5-gp27 complex (47), Hcp and VgrG have been postulated to form an extracellular translocon. This model is further supported by two recent crystallography studies showing that Hcp, VgrG, and a T4 phage gp25-like protein resembled membrane penetration tails of bacteriophages (35, 45).Little is known about the topology and structure of T6SS machinery subunits and the distinction between genes encoding machinery subunits and genes encoding regulatory proteins. Posttranslational regulation via the phosphorylation of Fha1 by a serine-threonine kinase (PpkA) is required for Hcp secretion from Pseudomonas aeruginosa (42). Genetic evidence for P. aeruginosa suggested that the T6SS may utilize a ClpV-like AAA+ ATPase to provide the energy for machinery assembly or substrate translocation (41). A recent study of V. cholerae suggested that ClpV ATPase activity is responsible for remodeling the VipA/VipB tubules which are crucial for type VI substrate secretion (6). An outer membrane lipoprotein, SciN, is an essential T6SS component for mediating Hcp secretion from enteroaggregative Escherichia coli (1). A systematic study of the T6SS machinery in E. tarda revealed that 13 of 16 genes in the evp gene cluster are essential for secretion of T6S substrates (74), which suggests the core components of the T6SS. Interestingly, most of the core components conserved in T6SS are predicted soluble proteins without recognizable signal peptide and transmembrane (TM) domains.The intracellular multiplication F (IcmF) and H (IcmH) proteins are among the few core components with obvious TM domains (8). In Legionella pneumophila Dot/Icm T4SSb, IcmF and IcmH are both membrane localized and partially required for L. pneumophila replication in macrophages (58, 70, 75). IcmF and IcmH are thought to interact with each other in stabilizing the T4SS complex in L. pneumophila (58). In T6SS, IcmF is one of the essential components required for secretion of Hcp from several animal pathogens, including V. cholerae (48), Aeromonas hydrophila (63), E. tarda (74), and P. aeruginosa (41), as well as the plant pathogens A. tumefaciens (72) and Pectobacterium atrosepticum (39). In E. tarda, IcmF (EvpO) interacted with IcmH (EvpN), EvpL, and EvpA in a yeast two-hybrid assay, and its putative nucleotide-binding site (Walker A motif) was not essential for secretion of T6SS substrates (74).In this study, we characterized the topology and interactions of the IcmF and IcmH family proteins ImpLM and ImpKL, which are two essential components of the T6SS of A. tumefaciens. We adapted the nomenclature proposed by Cascales (8), using the annotated gene designation followed by the letter indicated by Shalom et al. (59). Our data indicate that ImpLM and ImpKL are both integral inner membrane proteins and interact with each other via their N-terminal domains residing in the cytoplasm. We also provide genetic evidence showing that ImpLM may function as a nucleoside triphosphate (NTP)-binding protein or nucleoside triphosphatase to mediate T6S machinery assembly and/or substrate secretion.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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