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 31 毫秒
1.
The fusion of enveloped viruses with the host cell is driven by specialized fusion proteins to initiate infection. The “class I” fusion proteins harbor two regions, typically two heptad repeat (HR) domains, which are central to the complex conformational changes leading to fusion: the first heptad repeat (HRN) is adjacent to the fusion peptide, while the second (HRC) immediately precedes the transmembrane domain. Peptides derived from the HR regions can inhibit fusion, and one HR peptide, T20 (enfuvirtide), is in clinical use for HIV-1. For paramyxoviruses, the activities of two membrane proteins, the receptor-binding protein (hemagglutinin-neuraminidase [HN] or G) and the fusion protein (F), initiate viral entry. The binding of HN or G to its receptor on a target cell triggers the activation of F, which then inserts into the target cell and mediates the membrane fusion that initiates infection. We have shown that for paramyxoviruses, the inhibitory efficacy of HR peptides is inversely proportional to the rate of F activation. For HIV-1, the antiviral potency of an HRC-derived peptide can be dramatically increased by targeting it to the membrane microdomains where fusion occurs, via the addition of a cholesterol group. We report here that for three paramyxoviruses—human parainfluenza virus type 3 (HPIV3), a major cause of lower respiratory tract diseases in infants, and the emerging zoonotic viruses Hendra virus (HeV) and Nipah virus (NiV), which cause lethal central nervous system diseases—the addition of cholesterol to a paramyxovirus HRC-derived peptide increased antiviral potency by 2 log units. Our data suggest that this enhanced activity is indeed the result of the targeting of the peptide to the plasma membrane, where fusion occurs. The cholesterol-tagged peptides on the cell surface create a protective antiviral shield, target the F protein directly at its site of action, and expand the potential utility of inhibitory peptides for paramyxoviruses.Fusion of enveloped viruses with the host cell is a key step in viral infectivity, and interference with this process can lead to highly effective antivirals. Viral fusion is driven by specialized proteins that undergo an ordered series of conformational changes. These changes facilitate the initial, close apposition of the viral and host membranes, and they ultimately result in the formation of a fusion pore (reviewed in reference 12). The “class I” fusion proteins harbor two regions, typically two heptad repeat (HR) domains: the first one (HRN) adjacent to the fusion peptide and the second one (HRC) immediately preceding the transmembrane domain. Peptides derived from the HR regions can inhibit fusion, and one of them, T20 (enfuvirtide), is in clinical use for HIV-1 (19). Peptides derived from the HRN and HRC regions of paramyxovirus fusion (F) proteins can interact with fusion intermediates of F (3, 20, 22, 37, 46, 49) and provide a promising antiviral strategy.The current model for class I-driven fusion postulates the existence of a so-called prehairpin intermediate, a high-energy structure that bridges the viral and cell membranes, where the HRN and the HRC are separated. The prehairpin intermediate spontaneously collapses into the postfusion structure—a six-helical bundle (6HB), with an inner trimeric coiled-coil formed by the HRN onto which the HRC folds (12, 14, 30, 40). The key to these events is the initial activation step, whereby HN triggers F to initiate the process. Structural and biophysical analyses of the paramyxovirus 6HB (30, 50, 51) suggest that inhibitors bind to the prehairpin intermediate and prevent its transition to the 6HB, thus inhibiting viral entry. The peptides bind to their complementary HR region and thereby prevent HRN and HRC from refolding into the stable 6HB structure required for fusion (3, 10, 40). The efficiency of F triggering by HN critically influences the degree of fusion mediated by F and thus the extent of viral entry (35). In addition, differences in the efficiency of triggering of the fusion process impact the efficacy of potential antiviral molecules that target intermediate states of the fusion protein (36).Paramyxoviruses cause important human illnesses, significantly contributing to global disease and mortality, ranging from lower-respiratory-tract diseases in infants caused by human parainfluenza virus types 1, 2, and 3 (HPIV1, -2, and -3) (9, 48), to highly lethal central nervous system diseases caused by the emerging paramyxoviruses HeV and NiV. No antiviral therapies or vaccines yet exist for these paramyxoviruses, and vaccines would be unlikely to protect the youngest infants. Antiviral agents, therefore, would be particularly beneficial. All paramyxoviruses possess two envelope glycoproteins directly involved in viral entry and pathogenesis: a fusion protein (F) and a receptor-binding protein (HN, H, or G). The paramyxovirus F proteins belong to the group of “class I” fusion proteins (44, 45), which also include the influenza virus hemagglutinin protein and the HIV-1 fusion protein gp120. The F protein is synthesized as a precursor protein (F0) that is proteolytically processed posttranslationally to form a trimer of disulfide-linked heterodimers (F1-F2). This cleavage event places the fusion peptide at the F1 terminus in the mature F protein and is essential for membrane fusion activity. The exact triggers that initiate a series of conformational changes in F leading to membrane fusion differ depending on the pathway the virus uses to enter the cell. In the case of HPIV, HeV, and NiV, the receptor-binding protein, hemagglutinin-neuraminidase (HN) (in HPIV3) or G (in HeV and NiV), binds to cellular surface receptors, brings the viral envelope into proximity with the plasma membrane, and activates the viral F protein. This receptor-ligand interaction is required for the F protein to mediate the fusion of the viral envelope with the host cell membrane (23, 33, 35).The HRC peptide regions of a number of paramyxoviruses, including Sendai virus, measles virus, Newcastle disease virus (NDV), respiratory syncytial virus (RSV), simian virus 5 (SV5), Hendra virus (HeV), and Nipah virus (NiV), can inhibit the infectivity of the homologous virus (17, 20, 31, 37, 47, 49, 52, 53). Recently, we showed that peptides derived from the HRC region of the F protein of HPIV3 are effective inhibitors of both HPIV and HeV/NiV fusion (31) and that, for HeV, the strength of HRC peptide binding to the corresponding HRN region correlates with the potency of fusion and infection inhibition (30). However, peptides derived from the HPIV3 F protein HRC region are more effective at inhibiting HeV/NiV fusion than HPIV3 fusion, despite a stronger homotypic HRN-HRC interaction for HPIV3 (30, 31). We showed (36) that the kinetics of fusion (kinetics of F activation) impacts sensitivity to inhibition by peptides, as is the case for HIV (39). Alterations in HPIV3 HN′s property of F activation affect the kinetics of F''s progression through its conformational changes, thus altering inhibitor efficacy. Once the extended intermediate stage of F has passed, and fusion proceeds, peptide inhibitors are ineffective. We have proposed that the design of effective inhibitors may require either targeting an earlier stage of F activation or increasing the concentration of inhibitor at the location of receptor binding, in order to enhance the access and association of the inhibitor with the intermediate-stage fusion protein (36).A substantial body of evidence supports the notion that viral fusion occurs in confined areas of the interacting viral and host membranes (26). For HIV-1, the lipid composition of the viral membrane is strikingly different from that of the host cell membrane; the former is particularly enriched in cholesterol and sphingomyelin (4, 5, 7, 8). Cholesterol and sphingolipids are often laterally segregated in membrane microdomains or “lipid rafts” (7, 11). In fact, the antiviral potency of the HIV-inhibitory HRC peptide C34 is dramatically increased by targeting it to the “lipid rafts” via the addition of a cholesterol group (16).We applied the targeting strategy based on cholesterol derivatization to paramyxoviruses, and we show here that by adding a cholesterol tag to HPIV3-derived HRC E459V (30) inhibitory peptides, we increased antiviral potency by 2 log units (50% inhibitory concentrations [IC50], <2 nM). We chose to use the HPIV3-derived peptides for HeV/NiV, because we have previously shown that they are far more effective inhibitors of HeV and NiV than the homotypic peptides (30, 31). We propose that the enhanced activity resulting from the addition of a cholesterol tag is a result of the targeting of the peptide to the plasma membrane, where fusion occurs.  相似文献   

2.
3.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

4.
Of the four required herpes simplex virus (HSV) entry glycoproteins, the precise role of gH-gL in fusion remains the most elusive. The heterodimer gH-gL has been proposed to mediate hemifusion after the interaction of another required glycoprotein, gD, with a receptor. To identify functional domains of HSV-1 gH, we generated 22 randomized linker-insertion mutants. Analyses of 22 gH mutants revealed that gH is relatively tolerant of insertion mutations, as 15 of 22 mutants permitted normal processing and transport of gH-gL to the cell surface. gH mutants that were not expressed well at the cell surface did not function in fusion or viral entry. The screening of gH mutants for function revealed the following: (i) for wild-type gH and some gH mutants, fusion with nectin-1-expressing target cells occurred more rapidly than with herpesvirus entry mediator (HVEM)-expressing target cells; (ii) some gH mutants reduced the rate of cell fusion without abrogating fusion completely, indicating that gH may play a role in governing the kinetics of fusion and may be responsible for a rate-limiting first stage in HSV-1 fusion; and (iii) only one gH mutant, located within the short cytoplasmic tail, completely abrogated function, indicating that the gH cytoplasmic tail is crucial for cell fusion and viral infectivity.Herpes simplex virus (HSV), an enveloped neurotropic virus, infects target cells via membrane fusion, a process executed by viral fusion proteins capable of inserting into target membranes. Unlike many enveloped viruses that induce fusion through the activity of a single viral fusion protein, HSV requires four glycoproteins, glycoprotein B (gB), glycoprotein D (gD), glycoprotein H (gH), and glycoprotein L (gL), to execute fusion (6, 40, 42). The focus of this study, gH, is expressed as a heterodimer with gL (gH-gL). HSV gH and gL rely on one another for proper folding, posttranslational processing, and transport to the cell and virion surface (5, 23, 35).A sequential model of entry is the prevailing working hypothesis of HSV entry (1-3, 28, 32, 41). Viral attachment is mediated by the binding of glycoprotein C (gC) or gB to cell surface glycosaminoglycans such as heparan sulfate (38). The subsequent fusion between the virion envelope and host cell membrane is thought to result from a series of concerted events. First, gD binds to one of its host cell receptors. These receptors include herpesvirus entry mediator (HVEM), a member of the tumor necrosis factor (TNF) receptor family; nectin-1 and nectin-2, cell adhesion molecules of the Ig superfamily; and heparan sulfate modified by specific 3-O-sulfotransferases (39).It was previously proposed that gD binding a receptor induces a conformational change that allows for interactions between gD, gB, and/or gH-gL (1, 2, 8, 10, 16, 25, 32). It is thought that while gD functions primarily in receptor binding, gB and gH-gL function as the core fusion machinery of HSV.Based on its crystal structure, gB has structural features typical of viral fusion proteins in general and is structurally similar to vesicular stomatitis virus (VSV) glycoprotein G, the fusion protein of VSV (22, 34). In addition to its resemblance to other viral fusogens, gB also binds its own receptor, paired immunoglobulin-like receptor (PILRalpha) (36, 37). Importantly, HSV gB does not successfully execute fusion in the absence of gD or gH-gL (41). Compared to the other required HSV entry glycoproteins, relatively little is known about the specific roles of gH-gL during fusion. The structure of gH-gL is unknown, although in silico analyses and studies of synthetic gH peptides suggested that gH also has fusogenic properties (12, 13, 17-20).gD, a gD receptor, and gH-gL have been shown to be sufficient for inducing hemifusion, the mixing of the proximal leaflets of the viral and host cell bilayers (41). Several lines of research suggest that the subsequent step in fusion is an interaction between gH-gL and gB, with the latter glycoprotein being required for a committed and expanding fusion pore (1-3, 16, 28, 41). However, it is still unclear whether the gB and gH-gL interaction requires that gD first bind a receptor (1, 3), indicating that another viable model of HSV entry may be nonsequential gD-gB-gH-gL complex formation.Several domains important for fusion within HSV gH have been discerned. The only function associated with the N-terminal domain of HSV gH, to date, is gL binding. Residues 377 to 397 within a predicted alpha-helix in the gH ectodomain are required for cell-cell fusion and complementation of a gH-null virus (18). The mutation of a predicted heptad repeat region spanning residues 443 to 471 abrogated cell-cell fusion (17). Insertion mutations within what has been termed the pretransmembrane region of gH have also been shown to abrogate fusion and viral entry (11). The glycine residue at position 812 within the predicted gH transmembrane domain was shown previously to be important for fusion (21). Finally, although the deletion of the final six residues of gH (residues 832 to 838), which are within its short cytoplasmic tail, has no effect on fusion, further deletions were shown to decrease polykaryocyte formation by a syncytial HSV strain (4, 43).We used a transposon-based comprehensive random linker-insertion mutagenesis strategy to generate a library of mutants spanning the entire length of HSV-1 gH, an 838-amino-acid type I membrane protein. A panel of 22 insertion mutants was generated, 15 of which were expressed at near-normal levels on the cell surface. Interestingly, some insertions reduced the rate of cell fusion rather than abrogating cell fusion activity altogether, suggesting that gH may have a role in governing the kinetics of fusion and may be responsible for a rate-limiting first stage in HSV-1 fusion. Additionally, one insertion mutation that completely abrogated cell fusion and viral infectivity is located within the gH cytoplasmic tail, indicating that the short C-terminal tail of gH is critical for cell fusion and entry mediated by HSV-1.  相似文献   

5.
The spatial organization of metastable paramyxovirus fusion (F) and attachment glycoprotein hetero-oligomers is largely unknown. To further elucidate the organization of functional fusion complexes of measles virus (MeV), an archetype of the paramyxovirus family, we subjected central predictions of alternative docking models to experimental testing using three distinct approaches. Carbohydrate shielding through engineered N-glycans indicates close proximity of a membrane-distal, but not membrane-proximal, section of the MeV attachment (H) protein stalk domain to F. Directed mutagenesis of this section identified residues 111, 114, and 118 as modulators of avidity of glycoprotein interactions and determinants of F triggering. Stalk-length variation through deletion or insertion of HR elements at positions flanking this section demonstrates that the location of the stalk segment containing these residues cannot be altered in functional fusion complexes. In contrast, increasing the distance between the H head domains harboring the receptor binding sites and this section through insertion of structurally rigid α-helical domains with a pitch of up to approximately 75 Å downstream of stalk position 118 partially maintains functionality in transient expression assays and supports efficient growth of recombinant virions. In aggregate, these findings argue against specific protein-protein contacts between the H head and F head domains but instead support a docking model that is characterized by short-range contacts between the prefusion F head and the attachment protein stalk, possibly involving H residues 111, 114, and 118, and extension of the head domain of the attachment protein above prefusion F.Paramyxoviruses infect cells through fusion of the viral envelope with target cell membranes. For all members of the Paramyxovirinae subfamily, this involves the concerted action of two envelope glycoproteins, the fusion (F) and attachment (H, HN, or G, depending on the Paramyxovirinae genus) proteins. Both proteins feature short lumenal tails, a single transmembrane domain, and large ectodomains. The F protein, in type I orientation, forms homotrimers, while homodimers or homotetramers have been suggested as functional units for attachment proteins of different Paramyxovirinae subfamily members (7, 14, 28, 41, 49, 50, 66). For entry, upon receptor binding, the attachment protein is considered to initiate a series of conformational rearrangements in the metastable prefusion F protein (15, 77), which ultimately brings together transmembrane domains and fusion peptides and, thus, donor and target membranes (3, 32, 45, 53, 80).Multiple studies have demonstrated that specific interactions between compatible F and attachment proteins of paramyxovirinae are imperative for the formation of functional fusion complexes (6, 29, 36, 42, 43, 56, 75). However, the molecular nature of these interactions and the spatial organization of functional glycoprotein hetero-oligomers remain largely unknown. Individual ectodomain and partial ectodomain crystal structures have been obtained for different paramyxovirus F (13, 76, 77) and attachment (8, 14, 17, 28, 35, 79) proteins, respectively. For F, a stabilized human parainfluenza virus type 5 (HPIV5) ectodomain that is believed to represent a prefusion conformation folds into a globular head structure that is attached to the transmembrane domains through a helical stalk consisting of the membrane-proximal heptad repeat B (HR-B) domains (77). For the attachment protein, a globular head that harbors the receptor binding sites is considered to be connected to the transmembrane region through extended stalk domains (34, 78). Crystal structures of isolated head domains have been solved for several paramyxovirus attachment proteins, including measles virus (MeV) H, and reveal the six-blade propeller fold typical of sialidase structures (8, 14, 17, 28, 79). However, morbilliviruses recognize proteinaceous receptors (for MeV, the regulator of complement activation [CD46] and/or signaling lymphocytic activation molecule [SLAM], depending on the virus strain) (21, 40, 46, 51, 64, 65). X-ray data do not extend to the stalk domains, but circular dichroism analysis (78) and structure predictions (36, 78) support an α-helical coiled-coil configuration of the stalk.The nature of individual residues that engage in specific intermolecular interactions between glycoproteins of paramyxovirinae prior to refolding has been studied most extensively for the attachment protein. The stalk domains of several paramyxovirus HN proteins have been implicated in mediating specificity for their homotypic F proteins (18, 20, 43, 63, 70, 72). We have found that this extends to MeV and canine distemper virus H and, thus, to paramyxovirinae recognizing proteinaceous receptors (36), supporting the general hypothesis that F-interacting residues may reside in the stalk region of the attachment protein (30, 78).Considerably less information concerning the nature of F microdomains that mediate attachment protein specificity is available. Among the few exceptions are peptides derived from Newcastle disease virus (NDV) and Sendai virus F HR-B domains, which interact with soluble variants of the respective HN proteins in vitro (25, 67). Multiple domains have been suggested to mediate specificity of HPIV2 F for its HN (69). However, a conclusive N-glycan shielding study (43) and structural information (77, 78) argue against direct contacts between NDV F HR-B domains and HN in native glycoprotein complexes. Thus, the role of individual HPIV2 F residues in HN binding is unclear (25, 43).Building on the observation that MeV H is able to engage in productive heterotypic interactions with F proteins derived from some but not all isolates of closely related canine distemper virus, we have recently identified residues in morbillivirus F (MeV F residue 121) and H (H stalk residues 110 to 114) that interdependently contribute to physical MeV glycoprotein interaction and F triggering for fusion (36). While these residues could mediate reciprocal glycoprotein specificity through long-range effects, molecular modeling of the MeV H stalk in an α-helical conformation has posited F residue 121 at the same level above the viral envelope as H residues 110 to 114, making direct contacts structurally conceivable (36). This spatial organization of functional fusion complexes furthermore provides a comprehensive explanation for previous demonstrations of a specific role for attachment protein stalk domains of paramyxovirinae in functional and physical interactions with F (18, 43, 63, 70, 72). However, this “staggered-head” model mandates positioning the globular head of the attachment protein above the prefusion F trimer (36), as opposed to a suggested “parallel-head” alignment of the glycoproteins (31, 47). The latter is mostly based on transmission electron microscopy micrographs of viral particles apparently showing glycoprotein spikes of equal length (33). Unfortunately, these images lack the resolution for an identification of the molecular nature of the spikes (attachment or F protein) or the distinguishing between densely packaged H and F head domains of different heights and laterally aligned head domains. Indeed, a recent single-particle reconstruction based on cryo-electron microscopy images of HPIV5 particles revealed that defined spikes correspond to F in a postfusion conformation, which was interpreted as a product of possible premature F refolding (38). These two-dimensional images of heavy-metal-stained particles did not reveal F spikes in a prefusion conformation. Rather, a dense surface layer was considered to correspond to prefusion glycoprotein hetero-oligomers (38). In addition to further-advanced image reconstructions, biochemical assessment of alternative docking modes is imperative for the elucidation of the organization of functional fusion complexes of paramyxovirinae.In this study, we subjected central predictions of the hypothetical alignment models to experimental analysis. By employing carbohydrate shielding, directed mutagenesis, and variation of the length of the H stalk domain, we examined the proximity of different regions of the H stalk to F, probed a role of individual residues around the previously identified H stalk section from positions 110 to 114 in the formation of functional fusion complexes, tested the effect of varying the length of the H stalk membrane proximal and membrane distal to this section, and explored the general possibility of whether specific contacts between the prefusion F and H head domains are required for F triggering. Experimental data were interpreted in the light of a working model of MeV glycoprotein hetero-oligomers prior to receptor binding.  相似文献   

6.
The highly pathogenic Nipah virus (NiV) is aerially transmitted and causes a systemic infection after entering the respiratory tract. Airway epithelia are thus important targets in primary infection. Furthermore, virus replication in the mucosal surfaces of the respiratory or urinary tract in later phases of infection is essential for virus shedding and transmission. So far, the mechanisms of NiV replication in epithelial cells are poorly elucidated. In the present study, we provide evidence that bipolar targeting of the two NiV surface glycoproteins G and F is of biological importance for fusion in polarized epithelia. We demonstrate that infection of polarized cells induces focus formation, with both glycoproteins located at lateral membranes of infected cells adjacent to uninfected cells. Supporting the idea of a direct spread of infection via lateral cell-to-cell fusion, we could identify basolateral targeting signals in the cytoplasmic domains of both NiV glycoproteins. Tyrosine 525 in the F protein is part of an endocytosis signal and is also responsible for basolateral sorting. Surprisingly, we identified a dityrosine motif at position 28/29 in the G protein, which mediates polarized targeting. A dileucine motif predicted to function as sorting signal is not involved. Mutation of the targeting signal in one of the NiV glycoproteins prevented the fusion of polarized cells, suggesting that basolateral or bipolar F and G expression facilitates the spread of NiV within epithelial cell monolayers, thereby contributing to efficient virus spread in mucosal surfaces in early and late phases of infection.Nipah virus (NiV) is a zoonotic and highly pathogenic member of the genus Henipavirus within the family Paramyxoviridae. NiV emerged for the first time in 1998 and caused an outbreak of respiratory disease in pigs and fatal encephalitis in humans in Malaysia and Singapore (9). Fruit bats of the genus Pteropus have been identified as the major natural reservoir host (50). Due to the lack of therapeutic or prophylactic options and the high mortality rates associated with human infections, work with live NiV requires biosafety level 4 (BSL-4) containment.As a typical member of the family Paramyxoviridae, NiV possesses two viral surface glycoproteins that are required for virus entry and spread. Glycoprotein G is responsible for the binding of the virus to cellular ephrin-B2 and -B3 receptors (3, 35, 36). After receptor binding, the viral fusion protein F mediates pH-independent fusion of viral and cellular membranes (virus entry) or fusion of cellular membranes (cell-to-cell fusion). To be fusion active, the precursor F0 must be proteolytically cleaved into the subunits F1 and F2 by host cell proteases. This proteolytic activation requires clathrin-mediated endocytosis, due to a tyrosine-based signal in the cytoplasmic tail of the F protein, and subsequent cleavage by endosomal cathepsin L (12, 37, 45). Only after recycling from endosomes to the cell surface is fusion-active F protein available for incorporation into budding virions or for the initiation of cell-to-cell fusion (13).The respiratory tract is the most common route of virus entry into the human body. Following respiratory invasion, some viruses, such as influenza virus and severe acute respiratory syndrome (SARS) coronavirus, remain localized, and virions are disseminated only locally by transport in mucus or inflammatory exudates, which permit access to new target cells in the lung. Even if these viruses can cause severe diseases, they fail to penetrate beyond the mucosal surface. Other viruses, such as measles, mumps, rubella, and varicella viruses, also infect via the respiratory tract but then enter the blood circulation from the airways without causing major local symptoms. NiV enters via the airways and subsequently spreads systemically, with extensive endothelial involvement leading to vasculitis, which is mostly responsible for the clinical disease. In addition, NiV often causes symptomatic respiratory infections. Respiratory illness is generally observed in pigs and in about half of human infections (9, 24, 30, 38). A retrospective analysis of NiV outbreaks in Bangladesh strongly suggested that the patients with symptomatic respiratory tract infections were responsible for human-to-human transmission (24). Thus, NiV infection of the airway mucosa is relevant not only for primary NiV infection, serving as a portal of virus entry, but also for virus shedding and transmission to other hosts. Beside respiratory epithelia, epithelial cells in the kidney and bladder have been shown to be infected in vivo and are suggested to be important sites of release of progeny virions into the urine (8, 29, 34, 49; for a review, see reference 26).The major characteristics of polarized epithelial cells are structurally and functionally discrete apical and basolateral plasma membrane domains. To maintain the distinct protein compositions of these domains, newly synthesized membrane proteins must be sorted to their sites of ultimate function and residence (28). Due to the polarized nature of epithelia, virus receptors or viral proteins can be selectively expressed at either apical or basolateral cell surfaces. This can restrict virus entry, budding, or cell-to-cell fusion, with significant implications for virus spread and thus for pathogenesis. The aim of this study was to elucidate the molecular mechanisms of NiV spread within epithelial cells, focusing on the roles of the two surface glycoproteins G and F. We could show that infection of polarized MDCK cells leads to the formation of viral foci. The finding that both NiV glycoproteins not only were expressed apically but also were present at lateral membranes in infected cells adjacent to noninfected cells suggested that the infection spreads by cell-to-cell fusion. When we analyzed the distributions of F and G proteins upon single expression, we observed that the presence of the glycoproteins at (baso)lateral membranes is signal mediated. We could demonstrate that both proteins possess tyrosine-based targeting motifs in their cytoplasmic tails (Y525 in the F protein and Y28/29 in the G protein), which mediate sorting to the basolateral membranes of polarized epithelia. Fusion of polarized cells was observed only when the basolateral sorting signals of both glycoproteins were intact. These observations support the notion that basolateral or bipolar expression of F and G proteins is required and responsible for the spread of infection across the lateral junctions via glycoprotein-mediated cell-to-cell fusion.  相似文献   

7.
8.
Fusion of enveloped viruses with host cells is triggered by either receptor binding or low pH but rarely requires both except for avian sarcoma leukosis virus (ASLV). We recently reported that membrane fusion mediated by an oncogenic Jaagsiekte sheep retrovirus (JSRV) envelope (Env) requires an acidic pH, yet receptor overexpression is required for this process to occur. Here we show that a soluble form of the JSRV receptor, sHyal2, promoted JSRV Env-mediated fusion at a low pH in normally fusion-negative cells and that this effect was blocked by a synthetic peptide analogous to the C-terminal heptad repeat of JSRV Env. In contrast to the receptor of ASLV, sHyal2 induced pronounced shedding of the JSRV surface subunit, as well as unstable conformational rearrangement of its transmembrane (TM) subunit, yet full activation of JSRV Env fusogenicity, associated with strong TM oligomerization, required both sHyal2 and low pH. Consistently, sHyal2 enabled transduction of nonpermissive cells by JSRV Env pseudovirions, with low efficiency, but substantially blocked viral entry into permissive cells at both binding and postbinding steps, indicating that sHyal2 prematurely activates JSRV Env-mediated fusion. Altogether, our study supports a model that receptor priming promotes fusion activation of JSRV Env at a low pH, and that the underlying mechanism is likely to be different from that of ASLV. Thus, JSRV may provide a useful alternate model for the better understanding of virus fusion and cell entry.Fusion is a fundamental event in the life cycle of enveloped viruses and is essential for viral replication. While viral fusion proteins are highly divergent in primary sequence, their structures and modes of activation share striking similarities, permitting their classification into two major groups (41). Class I fusion proteins, as exemplified by the retrovirus envelope (Env) and influenza virus hemagglutinin (HA), are composed mainly of alpha-helices, and they are present as metastable trimers on the viral surface (11). Class II fusion proteins, represented by alphavirus E1 and flavivirus E, contain predominantly beta-sheets and exist as dimers in the prefusion state (16). Of note, the vesicular stomatitis virus G (VSV-G) and herpesvirus gB proteins were recently assigned to a newly established class III, for fusion proteins combining properties of both class I and class II (13, 30). Despite these differences, one common and intriguing characteristic of all viral fusion proteins is their ability to undergo dramatic conformational rearrangements upon activation, i.e., the formation of trimers of hairpins, which drive fusion between viral and cellular membranes (11, 17).Retrovirus Env is a typical type I transmembrane protein composed of surface (SU) and transmembrane (TM) subunits and belongs to the class I fusion proteins. SU is responsible for binding to cognate cellular receptors or cofactors, while TM directly mediates membrane fusion (6). Most retroviruses use a pH-independent pathway for entry, during which receptor binding relieves the ability of SU to restrain TM, resulting in conformational changes in TM and subsequent fusion with the cell membrane (11). Interestingly, increasing numbers of retroviruses have recently been shown to require a low pH (3, 15, 24, 28, 31) or pH-dependent protease activities to trigger fusion (18); the latter property has also been demonstrated for some other enveloped viruses (2, 14, 18, 26, 27, 33, 34). Among these, avian sarcoma leukosis virus (ASLV) is unique in that it uses a two-step mechanism for fusion, in which receptor binding primes the second trigger of low pH (24).Jaagsiekte sheep retrovirus (JSRV) is a simple betaretrovirus etiologically responsible for contagious lung tumors in sheep (12). The native Env protein of JSRV functions as a potent oncogene that induces cell transformation in vitro and in animals (4, 9, 21, 29, 42). The cell entry receptor for JSRV has been identified as hyaluronidase 2 (Hyal2), a glycosylphosphatidylinositol (GPI)-anchored protein belonging to the hyaluronidase family (29); Hyal2 itself has low hyaluronidase activity, and this activity is not associated with JSRV entry and infection (38). Intrigued by the oncogenic nature of JSRV Env, we recently examined the mechanism of JSRV entry and found that JSRV Env-mediated fusion and cell entry require a low pH (1, 8). These observations led us to hypothesize that the pH-dependent fusion activation of JSRV Env may be advantageous for its oncogenesis, given that extreme cell-cell fusion of the plasma membrane at a neutral pH would result in syncytium formation and often cell death. Curiously, we noticed that overexpression of Hyal2 is necessary for JSRV Env to induce membrane fusion at a low pH in vitro, suggesting that Hyal2 may play an active role in the pH-dependent fusion process. Here we provide direct evidence that Hyal2 functions in cooperation with a low pH to trigger the JSRV Env-mediated fusion activation yet exhibits some striking differences from the mechanism of ASLV fusion. The multistep pathway for JSRV Env-mediated fusion activation might be important for its replication fitness and oncogenesis.  相似文献   

9.
Human respiratory syncytial virus (HRSV) fusion (F) protein is an essential component of the virus envelope that mediates fusion of the viral and cell membranes, and, therefore, it is an attractive target for drug and vaccine development. Our aim was to analyze the neutralizing mechanism of anti-F antibodies in comparison with other low-molecular-weight compounds targeted against the F molecule. It was found that neutralization by anti-F antibodies is related to epitope specificity. Thus, neutralizing and nonneutralizing antibodies could bind equally well to virions and remained bound after ultracentrifugation of the virus, but only the former inhibited virus infectivity. Neutralization by antibodies correlated with inhibition of cell-cell fusion in a syncytium formation assay, but not with inhibition of virus binding to cells. In contrast, a peptide (residues 478 to 516 of F protein [F478-516]) derived from the F protein heptad repeat B (HRB) or the organic compound BMS-433771 did not interfere with virus infectivity if incubated with virus before ultracentrifugation or during adsorption of virus to cells at 4°C. These inhibitors must be present during virus entry to effect HRSV neutralization. These results are best interpreted by asserting that neutralizing antibodies bind to the F protein in virions interfering with its activation for fusion. Binding of nonneutralizing antibodies is not enough to block this step. In contrast, the peptide F478-516 or BMS-433771 must bind to F protein intermediates generated during virus-cell membrane fusion, blocking further development of this process.Human respiratory syncytial virus (HRSV), a member of the Pneumovirus genus of the Paramyxoviridae family, is the main cause of severe lower respiratory tract infections in very young children (36), and it is a pathogen of considerable importance in the elderly (24, 26) and in immunocompromised adults (22). Currently, there is no effective vaccine against the virus although it is known that passive administration of neutralizing antibodies to individuals at high risk is an effective immunoprophylaxis (37, 38).The HRSV genome is a single-stranded negative-sense RNA molecule of approximately 15 kb that encodes 11 proteins (16, 53). Two of these proteins are the main surface glycoproteins of the virion. These are (i) the attachment (G) protein, which mediates virus binding to cells (44), and (ii) the fusion (F) protein, which promotes both fusion of the viral and cell membranes at the initial stages of the infectious cycle and fusion of the membrane of infected cells with those of adjacent cells to form characteristic syncytia (72). These two glycoproteins are the only targets of neutralizing antibodies either induced in animal models (19, 63, 65, 70) or present in human sera (62).The G protein is a highly variable type II glycoprotein that shares neither sequence identity nor structural features with the attachment protein of other paramyxoviruses (75). It is synthesized as a precursor of about 300 amino acids (depending on the strain) that is modified posttranslationally by the addition of a large number of N- and O-linked oligosaccharides and is also palmitoylated (17). The G protein is oligomeric (probably a homotetramer) (23) and promotes binding of HRSV to cell surface proteoglycans (35, 40, 49, 67). Whether this is the only interaction of G with cell surface components is presently unknown.The F protein is a type I glycoprotein that is synthesized as an inactive precursor of 574 amino acids (F0) which is cleaved by furin during transport to the cell surface to yield two disulfide-linked polypeptides, F2 from the N terminus and F1 from the C terminus (18). Like other viral type I fusion proteins, the mature F protein is a homotrimer which is in a prefusion, metastable, conformation in the virus particle. After fusion, the F protein adopts a highly stable postfusion conformation. Stability of the postfusion conformation is determined to great extent by two heptad repeat (HR) sequences, HRA and HRB, present in the F1 chain. Mixtures of HRA and HRB peptides form spontaneously heterotrimeric complexes (43, 51) that assemble in six-helix bundles (6HB), consisting of an internal core of three HRA helices surrounded by three antiparallel HRB helices, as determined by X-ray crystallography (79).The three-dimensional (3D) structure of the HRSV F protein has not been solved yet. Nevertheless, the structures of the pre- and postfusion forms of two paramyxovirus F proteins have revealed substantial conformational differences between the pre- and postfusion conformations (77, 78). The present hypothesis about the mechanism of membrane fusion mediated by paramyxovirus F proteins proposes that, following binding of the virus to the cell surface, the prefusion form of the F glycoprotein is activated, and membrane fusion is triggered. The F protein experiences then a series of conformational changes which include the exposure of a hydrophobic region, called the fusion peptide, and its insertion into the target membrane. Subsequent refolding of this intermediate leads to formation of the HRA and HRB six-helix bundle, concomitant with approximation of the viral and cell membranes that finally fuse, placing the fusion peptide and the transmembrane domain in the same membrane (4, 20). The formation of the 6HB and the associated free energy change are tightly linked to the merger of the viral and cellular membranes (60).Antibodies play a major role in protection against HRSV. Animal studies have demonstrated that immunization with either F or G glycoproteins induces neutralizing antibodies and protects against a viral challenge (19, 63, 70). Furthermore, transfer of these antibodies (31, 56) or of anti-F or anti-G monoclonal antibodies (MAbs) protects mice, cotton rats, or calves against either a human or bovine RSV challenge, respectively (65, 68, 73). Likewise, infants at high risk of severe HRSV disease are protected by the prophylactic administration of immunoglobulins with high anti-HRSV neutralizing titers (33). Finally, a positive correlation was found between high titers of serum neutralizing antibodies and protection in adult volunteers challenged with HRSV (34, 74), while an inverse correlation was found between high titers of neutralizing antibodies and risk of infection in children (29) and in the elderly (25).Whereas all the anti-G monoclonal antibodies reported to date are poorly neutralizing (1, 28, 48, 71), some anti-F monoclonal antibodies have strong neutralization activity (1, 3, 5, 28, 46). It is believed that HRSV neutralization by anti-G antibodies requires simultaneous binding of several antibodies to different epitopes, leading to steric hindrance for interaction of the G glycoprotein with the cell surface. Indeed, it has been shown that neutralization is enhanced by mixtures of anti-G monoclonal antibodies (1, 50), mimicking the effect of polyclonal anti-G antibodies. In contrast, highly neutralizing anti-F monoclonal antibodies do not require cooperation by other antibodies to block HRSV infectivity efficiently (1).In addition to neutralizing antibodies, other low-molecular-weight compounds directed against the F protein are potent inhibitors of HRSV infectivity. Synthetic peptides that reproduce sequences of heptad repeat B inhibit both membrane fusion promoted by the F protein and HRSV infectivity (42). Also, other small molecules obtained by chemical synthesis have been shown to interact with F protein and inhibit HRSV infectivity. These HRSV entry inhibitors have been the topic of intense research in recent years (55).This study explores the mechanisms of HRSV neutralization by different inhibitors of membrane fusion, including anti-F monoclonal antibodies, an HRB peptide, and the synthetic compound BMS-433771 (13-15). The results obtained indicate that antibodies and low-molecular-weight compounds block membrane fusion at different stages during virus entry.  相似文献   

10.
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   

11.
12.
13.
Many viral fusion proteins are primed by proteolytic cleavage near their fusion peptides. While the coronavirus (CoV) spike (S) protein is known to be cleaved at the S1/S2 boundary, this cleavage site is not closely linked to a fusion peptide. However, a second cleavage site has been identified in the severe acute respiratory syndrome CoV (SARS-CoV) S2 domain (R797). Here, we investigated whether this internal cleavage of S2 exposes a viral fusion peptide. We show that the residues immediately C-terminal to the SARS-CoV S2 cleavage site SFIEDLLFNKVTLADAGF are very highly conserved across all CoVs. Mutagenesis studies of these residues in SARS-CoV S, followed by cell-cell fusion and pseudotyped virion infectivity assays, showed a critical role for residues L803, L804, and F805 in membrane fusion. Mutation of the most N-terminal residue (S798) had little or no effect on membrane fusion. Biochemical analyses of synthetic peptides corresponding to the proposed S2 fusion peptide also showed an important role for this region in membrane fusion and indicated the presence of α-helical structure. We propose that proteolytic cleavage within S2 exposes a novel internal fusion peptide for SARS-CoV S, which may be conserved across the Coronaviridae.The severe acute respiratory syndrome coronavirus (SARS-CoV) emerged in 2003 as a significant threat to human health, and CoVs still represent a leading source of novel viruses for emergence into the human population. The CoV spike (S) protein mediates both receptor binding (via the S1 domain) and membrane fusion (via the S2 domain) and shows many features of a class I fusion protein, including the presence of distinct heptad repeats within the fusion domain (37). A critical feature of any viral fusion protein is the so-called “fusion peptide,” which is a relatively apolar region of 15 to 25 amino acids that interacts with membranes and drives the fusion reaction (9, 34, 38). Fusion peptides can be classified as N-terminal or internal, depending on their location relative to the cleavage site of the virus fusion protein (23). One key feature of viral fusion peptides is that within a particular virus family, there is high conservation of amino acid residues; however, there is little similarity between fusion peptides of different virus families (26). Despite these differences, some common themes do emerge, including a high level of glycine and/or alanine residues, as well as critical bulky hydrophobic amino acids. In several cases, the fusion peptide is known to contain a central “kink.” In the case of influenza virus hemagglutinin (HA), which is a classic example of an N-terminal fusion peptide, the N- and C-terminal parts of the fusion peptide (which are α-helical) penetrate the outer leaflet of the target membrane, with the kink at the phospholipid surface. The inside of the kink contains hydrophobic amino acids, with charged residues on the outer face (18). Internal fusion peptides (such as Ebola virus [EBOV] GP) often contain a conserved proline near their centers but also require a mixture of hydrophobic and flexible residues similar to N-terminal fusion peptides (9, 11). It is believed that the kinked fusion peptide sits in the outer leaflet of the target membrane and possibly induces positive curvature to drive the fusion reaction (22). It is important to note that, despite the presence of key hydrophobic residues, viral fusion peptides often do not display extensive stretches of hydrophobicity and can contain one or more charged residues (8). Ultimately, fusion peptide identification must rely on an often complex set of criteria, including structures of the fusion protein in different conformations, biophysical measurements of peptide function in model membranes, and biological activity in the context of virus particles.To date, the exact location and sequence of the CoV fusion peptide are not known (4); however, by analogy with other class I viral fusion proteins, it is predicted to be in the S2 domain. Overall, three membranotropic regions in SARS-CoV S2 have been suggested as potential fusion peptides (14, 17). Based on sequence analysis and a hydrophobicity analysis of the S protein using the Wimley-White (WW) interfacial hydrophobic interface scale, initial indications were that the SARS-CoV fusion peptide resided in the N-terminal part of HR1 (heptad repeat 1) (5, 6), which is conserved across the Coronaviridae. Mutagenesis of this predicted fusion peptide inhibited fusion in syncytia assays of S-expressing cells (28). This region of SARS-CoV has also been analyzed by other groups in biochemical assays (16, 17, 29) and defined as the WW II region although Sainz et al. (29) actually identified another, less conserved and less hydrophobic, region (WW I) as being more important for fusion. Peptides corresponding to this region have also been studied in biochemical assays by other groups (13). In addition, a third, aromatic region adjacent to the transmembrane domain (the membrane-proximal domain) has been shown to be important in SARS-CoV fusion (15, 20, 25, 30). This membrane-proximal domain likely acts in concert with a fusion peptide in the S2 ectodomain to mediate final bilayer fusion once conformational changes have exposed the fusion peptide in the ectodomain. To date, there is little or no information on the fusion peptides of CoVs other than SARS-CoV, except for the identification of the N-terminal part of the mouse hepatitis virus (MHV) S HR1 domain as a putative fusion peptide based on sequence analysis (6). In none of these cases (for SARS-CoV or MHV) is the role of these sequences as bone fide fusion peptides established.The majority of class I fusion proteins prime fusion activation by proteolytic processing, with the cleavage event occurring immediately N-terminal to the fusion peptide (21). In the case of SARS-CoV, early reports analyzing heterologously expressed SARS-CoV spike protein indicated that most of the protein was not cleaved (31, 39) but that there was some possibility of limited cleavage at the S1-S2 boundary (39). However, it is generally considered that S1-S2 cleavage is not directly linked to fusion peptide exposure in the case of SARS-CoV or any other CoV (4). Recently, however, it has been shown that SARS-CoV S can be proteolytically cleaved at a downstream position in S2, at residue 797 (2, 36). Here, we investigated whether cleavage at this internal position in S2 might expose a domain with properties of a viral fusion peptide. We carried out a mutagenesis study of SARS-CoV S residues 798 to 815 using cell-cell fusion and pseudovirus assays, as well as lipid mixing and structural studies of an isolated peptide, and we show the importance of this region as a novel fusion peptide for SARS-CoV.  相似文献   

14.
Fusogenic reoviruses utilize the FAST proteins, a novel family of nonstructural viral membrane fusion proteins, to induce cell-cell fusion and syncytium formation. Unlike the paradigmatic enveloped virus fusion proteins, the FAST proteins position the majority of their mass within and internal to the membrane in which they reside, resulting in extended C-terminal cytoplasmic tails (CTs). Using tail truncations, we demonstrate that the last 8 residues of the 36-residue CT of the avian reovirus p10 FAST protein and the last 20 residues of the 68-residue CT of the reptilian reovirus p14 FAST protein enhance, but are not required for, pore expansion and syncytium formation. Further truncations indicate that the membrane-distal 12 residues of the p10 and 47 residues of the p14 CTs are essential for pore formation and that a residual tail of 21 to 24 residues that includes a conserved, membrane-proximal polybasic region present in all FAST proteins is insufficient to maintain FAST protein fusion activity. Unexpectedly, a reextension of the tail-truncated, nonfusogenic p10 and p14 constructs with scrambled versions of the deleted sequences restored pore formation and syncytiogenesis, while reextensions with heterologous sequences partially restored pore formation but failed to rescue syncytiogenesis. The membrane-distal regions of the FAST protein CTs therefore exert multiple effects on the membrane fusion reaction, serving in both sequence-dependent and sequence-independent manners as positive effectors of pore formation, pore expansion, and syncytiogenesis.The only examples of nonenveloped viruses that induce cell-cell fusion and syncytium formation occur within the family Orthoreoviridae, an extremely diverse group of viruses containing segmented double-stranded RNA genomes (9). In recent years, the viral proteins responsible for the syncytiogenic phenotype of the fusogenic orthoreoviruses and aquareoviruses have been identified and characterized (14, 18, 41, 46). These fusion-associated small transmembrane (FAST) proteins define a new family of viral fusogens with several unique biological and biophysical properties. Unlike the well-characterized enveloped virus fusion proteins, reovirus FAST proteins are nonstructural viral proteins and are therefore not involved in mediating virus-cell fusion and virus entry (18, 21, 46). The FAST proteins are instead dedicated to inducing cell-cell fusion and syncytium formation following their expression and trafficking to the plasma membrane of virus-infected or transfected cells (14, 17, 46). Data from previously reported studies also suggest that the FAST proteins serve as virulence factors for the fusogenic reoviruses, promoting virus dissemination and increased tissue destruction (6, 43). How this atypical family of viral fusogens functions to mediate cell-cell membrane fusion remains unclear.The unusual biological role of the FAST proteins as nonstructural, virus-encoded, “cellular” fusogens is embodied in structural features that clearly distinguish the FAST proteins from the membrane fusion proteins of enveloped viruses. There are currently four distinct members of the FAST protein family, named according to their molecular masses: the homologous p10 proteins of avian reovirus (ARV) and Nelson Bay reovirus and the unrelated p14, p15, and p22 proteins of reptilian reovirus (RRV), baboon reovirus, and Atlantic salmon aquareovirus, respectively (14, 18, 41, 46). These proteins are the smallest known fusogens, ranging from 95 to 198 amino acids in size, and assume an asymmetric topology in the plasma membrane, with a single transmembrane domain that separates small N-terminal ectodomains of ∼20 to 41 residues from equal-sized or considerably larger C-terminal endodomains of ∼36 to 141 residues (Fig. (Fig.1A).1A). A number of structural motifs in both the ecto- and endodomains of the FAST proteins have been identified, including sites of acylation, hydrophobic patches, a membrane-proximal polybasic region, and regions rich in proline, cysteine, or arginine, proline, and histidine. Each of the FAST proteins has its own signature repertoire and arrangement of these motifs. Determining how these various motifs contribute to the fusogenic activity of the FAST proteins remains an area of active investigation.Open in a separate windowFIG. 1.ARV p10 and RRV p14 FAST protein topologies and tail truncations. (A) Diagrammatic representation of the p10 and p14 FAST proteins showing their topology in the plasma membrane. Both are single-pass transmembrane proteins with N-terminal ectodomains on the surface of cells and C-terminal endodomains in the cytoplasm. Structural motifs include hydrophobic patches (HP), polybasic motifs (PB), fatty acid modifications (indicated by squiggly lines) that are either the N-terminal myristoylation or palmitoylation of a dicysteine motif (CC), and a polyproline motif (PP). The total number of residues in each protein is indicated by the numbers. (B) The amino acid sequences of the p10 and p14 endodomains are shown, along with the motifs described above. Progressive truncations of the CTs were constructed (arrows), with the numbers indicating the last amino acid present in the full-length proteins or each truncation.Numerous studies of diverse fusion processes define five general steps of the pathway for membrane fusion and syncytium formation: membrane binding, close membrane apposition, hemifusion (i.e., the mixing of the outer leaflets of the two bilayers), stable pore formation, and pore expansion (12, 13, 44). The well-characterized enveloped virus fusion proteins utilize extensive structural rearrangement of their complex ectodomains to provide mechanical energy to draw membranes into close proximity and promote membrane merger (21, 53). The limited size of the FAST protein ectodomains precludes such a mechanical model for membrane fusion, necessitating the development of alternate models to explain how the diminutive FAST proteins breach the thermodynamic barriers that prevent the spontaneous merger of biological membranes. The FAST proteins are both necessary and sufficient to mediate membrane fusion (51). However, data from recent studies indicate that for maximal cell fusion activity, the FAST proteins rely on surrogate adhesins to mediate close membrane apposition (42). Data from recent studies also indicate that a small percentage of the p14 FAST protein expressed in virus-infected or transfected cells is proteolytically processed to generate a bioactive, soluble endodomain that recruits cellular pathways to drive the expansion of stable fusion pores into the extended fusion apertures needed for syncytium formation (50). The FAST proteins therefore utilize accessory proteins to mediate the prefusion (membrane binding and apposition) and postfusion (pore expansion) stages of syncytiogenesis, retaining within their rudimentary structures all that is required to mediate the actual process of membrane merger. This subdivision of the multistep process of syncytium formation is reflected in, and is perfectly suited to, the evolution of the FAST proteins as virus-encoded cellular fusogens.The small size of the FAST protein ectodomains and their donor membrane-focused topology contrast markedly with enveloped virus fusion proteins that position the majority of their mass external to the membrane. While the complex ectodomains of the enveloped virus fusion proteins clearly play an essential role in the fusion reaction, the involvement of their cytoplasmic tails (CTs) is far less certain, and no consistent picture of the role of these C-terminal tails has emerged. The CTs of many enveloped viral fusion proteins, including baculovirus (31), severe acute respiratory syndrome coronavirus (5), vesicular stomatitis virus (36), parainfluenza virus type 2 (56), and influenza A virus subtype H3 (10), play no role in the membrane fusion reaction. Of the fusion protein tails that do modulate the fusion reaction, the majority serve inhibitory roles, including the F proteins of measles virus and parainfluenza virus type 5 SER (7, 45, 52), glycoprotein B from several herpesviruses (22, 24, 28), and the fusion proteins of numerous retroviruses (1, 8, 30, 32, 34, 47, 48). These inhibitory cytoplasmic domains alter the conformation of the fusion protein ectodomains, thereby coupling virion maturation to fusion competence (1, 2, 35, 52, 54). In the few cases where extensive tail truncations adversely affect fusion, these truncations generally decrease but do not eliminate syncytiogenesis, and it is the membrane-proximal portion of the tail that promotes pore formation or pore expansion (20, 25, 26, 32).Since the FAST proteins are nonstructural viral proteins, their CTs (also referred to as endodomains) are not required to suppress fusion activity until after virus particle assembly. At the same time, the disproportionate size of their endodomains strongly suggests that these CTs play an important role in membrane fusion activity. Although one such role of the p14 CT is the generation of a soluble endodomain that recruits cellular factors involved in pore expansion, the majority of p14 is not proteolytically processed, suggesting that FAST protein CTs may serve additional roles as components of the intact protein (50). We now show that C-terminal truncations of the p10 and p14 FAST proteins reduced and eventually eliminated cell-cell fusion. Fluorescence-based pore formation assays coupled with tail reextension studies further revealed that FAST protein CTs drive fusion pore formation and expansion in both sequence-dependent and sequence-independent manners. The membrane-distal regions of FAST protein CTs therefore exert multiple effects on the mechanism of membrane fusion.  相似文献   

15.
Wild-type measles virus (MV) isolated in B95a cells could be adapted to Vero cells after several blind passages. In this study, we have determined the complete nucleotide sequences of the genomes of the wild type (T11wild) and its Vero cell-adapted (T11Ve-23) MV strain and identified amino acid substitutions R516G, E271K, D439E and G464W (D439E/G464W), N481Y/H495R, and Y187H/L204F in the nucleocapsid, V, fusion (F), hemagglutinin (H), and large proteins, respectively. Expression of mutated H and F proteins from cDNA revealed that the H495R substitution, in addition to N481Y, in the H protein was necessary for the wild-type H protein to use CD46 efficiently as a receptor and that the G464W substitution in the F protein was important for enhanced cell-cell fusion. Recombinant wild-type MV strains harboring the F protein with the mutations D439E/G464W [F(D439E/G464W)] and/or H(N481Y/H495R) protein revealed that both mutated F and H proteins were required for efficient syncytium formation and virus growth in Vero cells. Interestingly, a recombinant wild-type MV strain harboring the H(N481Y/H495R) protein penetrated slowly into Vero cells, while a recombinant wild-type MV strain harboring both the F(D439E/G464W) and H(N481Y/H495R) proteins penetrated efficiently into Vero cells, indicating that the F(D439E/G464W) protein compensates for the inefficient penetration of a wild-type MV strain harboring the H(N481Y/H495R) protein. Thus, the F and H proteins synergistically function to ensure efficient wild-type MV growth in Vero cells.Measles virus (MV), which belongs to the genus Morbillivirus in the family Paramyxoviridae, is an enveloped virus with a nonsegmented negative-strand RNA genome. The MV genome encodes six structural proteins: the nucleocapsid (N), phosphoprotein (P), matrix (M), fusion (F), hemagglutinin (H), and large (L) proteins. The P gene also encodes two other accessory proteins, the C and V proteins. The C protein is translated from an alternative translational initiation site leading a different reading frame, and the V protein is synthesized from an edited mRNA. MV has two envelope glycoproteins, the F and H proteins. The former is responsible for envelope fusion, and the latter is responsible for receptor binding (12).Wild-type MV strains isolated in B95a cells and laboratory-adapted MV strains have distinct phenotypes (18). Wild-type MV strains can grow in B95a cells but not in Vero cells, while laboratory-adapted MV strains can grow in both B95a and Vero cells. Wild-type MV strains do not cause hemadsorption (HAd) in African green monkey red blood cells (AGM-RBC), while most of laboratory-adapted MV strains cause HAd. Importantly, wild-type MV strains are pathogenic and induce clinical signs that resemble human measles in experimentally infected monkeys while laboratory-adapted MV strains do not.One approach to identify amino acid substitutions responsible for these phenotypic differences is the comparison of a wild-type MV strain with a standard laboratory-adapted MV strain such as the Edmonston strain. With regard to the H protein, amino acid substitutions important for HAd activity and cell-cell fusion in tissue culture cells were identified by expressing the H proteins in mammalian cells (15, 21). Recently, Tahara et al. revealed that the M, H, and L proteins are responsible for efficient growth in Vero cells by constructing a series of recombinant viruses in which part of the genome of the wild-type MV was replaced with the corresponding sequences of the Edmonston strain (45, 46, 47).Another approach is the comparison of wild-type MV strains with their Vero cell-adapted MV strains. It was reported that Vero cell-adapted MV strains could be obtained by successive blind passages of wild-type MV strains in Vero cells (18, 24, 30, 43). Interestingly, in vivo and in vitro phenotypes of Vero cell-adapted MV strains were similar to those of laboratory-adapted standard MV strains (18, 19, 24, 30, 43). Comparison of the complete nucleotide sequences of the genomes of wild-type MV strains with those of Vero cell-adapted wild-type MV strains revealed amino acid substitutions in the P, C, V, M, H, and L proteins (27, 42, 48, 53).At present, these phenotypic differences are explained mainly by the receptor usage of MV. Wild-type MV strains can use signaling lymphocyte activation molecule (SLAM; also called CD150) but not CD46 as a cellular receptor, whereas laboratory-adapted MV strains can use both SLAM and CD46 as cellular receptors (7, 10, 16, 29, 56, 60).However, receptor usage per se cannot explain all of the phenotypic differences (20, 25, 48, 53). For example, recombinant Edmonston strains expressing wild-type H proteins can grow in Vero cells to some extent (17, 54). Several reports suggested the presence of the third MV receptor on Vero cells (14, 44, 54, 60). Other reports indicated the contribution of the M protein on cell-cell fusion and growth of MV in Vero cells (4, 27, 47). Recently, the unidentified epithelial cell receptor for MV was predicted in primary culture of human cells (1, 55) and several epithelial cell lines (23, 51). However, the identity of the third receptor on Vero cells and the unidentified epithelial cell receptor is not clear yet. Thus, the mechanism of Vero cell adaptation of wild-type MV is not completely understood.In order to understand the molecular mechanism of these phenotypic changes of wild-type MV strains during adaptation in Vero cells, we determined the complete nucleotide sequences of the genomes of the wild-type (T11wild) and its Vero cell-adapted (T11Ve-23) MV strains (43) and examined the effect of individual amino acid substitutions using a mammalian cell expression system and reverse genetics. We show here that previously unrecognized new amino acid substitutions in the H and F proteins are important for MV adaptation and HAd activity.  相似文献   

16.
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   

17.
Epstein-Barr virus (EBV) membrane glycoprotein 42 (gp42) is required for viral entry into B lymphocytes through binding to human leukocyte antigen (HLA) class II on the B-cell surface. EBV gp42 plays multiple roles during infection, including acting as a coreceptor for viral entry into B cells, binding to EBV glycoprotein H (gH) and gL during the process of membrane fusion, and blocking T-cell recognition of HLA class II-peptide complexes through steric hindrance. EBV gp42 occurs in two forms in infected cells, a full-length membrane-bound form and a soluble form generated by proteolytic cleavage that is secreted from infected cells due to loss of the N-terminal transmembrane domain. Both the full-length and the secreted gp42 forms bind to gH/gL and HLA class II, and the functional significance of gp42 cleavage is currently unclear. We found that in a virus-free cell-cell fusion assay, enhanced secretion of gp42 promoted fusion with B lymphocytes, and mutation of the site of gp42 cleavage inhibited membrane fusion activity. The site of gp42 cleavage was found to be physically distinct from the residues of gp42 necessary for binding to gH/gL. These results suggest that cleavage and secretion of gp42 are necessary for the process of membrane fusion with B lymphocytes, providing the first indicated functional difference between full-length and cleaved, secreted gp42.Epstein-Barr virus (EBV) is a large DNA virus belonging to the human gammaherpesvirus subfamily. EBV is orally transmitted through saliva and persists for the lifetime of its human host, establishing a latency reservoir in B lymphocytes with intermittent viral reactivation (1, 27). More than 90% of the world''s adult population is infected with EBV, although in healthy individuals, viral reactivation from latency is quickly controlled by the immune system. During primary infection and viral reactivation from latency, EBV infects epithelial cells as well as B lymphocytes (27). Primary infection with EBV can lead to development of infectious mononucleosis, and EBV has also been strongly associated with a number of human malignancies of epithelial and B-cell origin, including Burkitt''s lymphoma and nasopharyngeal carcinoma (4, 9, 10, 33, 36).EBV encodes a number of membrane glycoproteins important in a variety of viral processes, including entry of the virus into target host cells and virus-induced cell-cell fusion. The membrane glycoproteins necessary for fusion with both epithelial and B cells are glycoprotein B (gB), gH, and gL, and together, they form the core virus fusion machinery (7, 20, 24, 29). In addition to these glycoproteins, glycoprotein 42 (gp42) has been shown to play an essential role in membrane fusion with B cells (7, 18, 20). Attachment of EBV virions to B cells occurs through binding of the main envelope protein gp350/220 to CD21 (also known as complement receptor type 2) (5, 23, 34). This interaction enhances the efficiency of EBV infection of B cells but is not required for viral entry (12, 30). Antibodies to gp350/220 inhibit EBV infection of B cells but enhance infection of epithelial cells, possibly by facilitating the access of other viral glycoproteins to the epithelial cell membrane (35). Virus-cell membrane fusion is subsequently triggered by binding of gp42 to human leukocyte antigen (HLA) class II on the B-cell surface (6, 8, 11, 17, 31). Interestingly, gp42 appears to function as a switch of cellular tropism between epithelial and B cells. The presence of gp42 in the viral envelope is necessary for infection of B lymphocytes, and virions that are low in gp42 are better able to infect HLA class II-negative epithelial cells (3). Aside from its role in membrane fusion, gp42 plays a significant role in evasion of the host immune system. Gp42 binds to HLA class II-peptide complexes in infected cells, sterically hindering T-cell recognition of the complex by the T-cell receptor (25). This inhibition may allow EBV to delay detection by the host immune system.Two different mature forms of gp42 are produced by EBV-positive B lymphocytes in the lytic cycle (26). The first form is a full-length type II membrane protein, and the second is a truncated soluble form (s-gp42) (26). s-gp42 is generated by posttranslational cleavage (most likely mediated by a cellular protease resident in the endoplasmic reticulum) and is secreted (26). Both forms of gp42 associate with HLA class II intracellularly, and both inhibit HLA class II-restricted antigen presentation to T cells (26). Both forms of gp42 produced by EBV-positive B cells in the lytic cycle were found to be present in gH-gL-gp42 complexes, indicating that s-gp42 retains the ability to bind gH/gL (26). The physiological significance of s-gp42 is currently unclear, but this form has been suggested to function in infection and immune evasion, blocking EBV entry receptors on lytically infected B cells to prevent reinfection and neutralizing gp42-specific antibodies following its secretion from infected cells (26).Both forms of gp42 have been examined for their functions in mediating evasion from T-cell immunity through binding to HLA class II complexes (26), but the functions of the two forms of the protein in membrane fusion are unknown. To examine how each form of gp42 functions during membrane fusion, we have assayed the effect of gp42 cleavage site mutation on this process. Also, to distinguish residues important for gp42 cleavage from those necessary for association with gH/gL, we have constructed a number of fully secreted gp42 truncation mutants and examined their interaction with gH/gL and their ability to mediate fusion. Mutation or deletion of the gp42 cleavage site inhibited or eliminated cleavage of the protein, which had a direct effect on gp42 function in membrane fusion. An assay of N-terminal truncations of gp42 indicated that the region of gp42 necessary for cleavage was physically distinct from the region of gp42 necessary for association with gH/gL. We show that membrane association of gp42 has an inhibitory effect on gp42 function in membrane fusion and that increased secretion of gp42 stimulates membrane fusion in vitro. Cleavage of gp42 may be necessary for EBV gp42 to assume a functional position, interaction, or conformation for participation in membrane fusion.  相似文献   

18.
Vaccinia virus A26 protein is an envelope protein of the intracellular mature virus (IMV) of vaccinia virus. A mutant A26 protein with a truncation of the 74 C-terminal amino acids was expressed in infected cells but failed to be incorporated into IMV (W. L. Chiu, C. L. Lin, M. H. Yang, D. L. Tzou, and W. Chang, J. Virol 81:2149-2157, 2007). Here, we demonstrate that A27 protein formed a protein complex with the full-length form but not with the truncated form of A26 protein in infected cells as well as in IMV. The formation of the A26-A27 protein complex occurred prior to virion assembly and did not require another A27-binding protein, A17 protein, in the infected cells. A26 protein contains six cysteine residues, and in vitro mutagenesis showed that Cys441 and Cys442 mediated intermolecular disulfide bonds with Cys71 and Cys72 of viral A27 protein, whereas Cys43 and Cys342 mediated intramolecular disulfide bonds. A26 and A27 proteins formed disulfide-linked complexes in transfected 293T cells, showing that the intermolecular disulfide bond formation did not depend on viral redox pathways. Finally, using cell fusion from within and fusion from without, we demonstrate that cell surface glycosaminoglycan is important for virus-cell fusion and that A26 protein, by forming complexes with A27 protein, partially suppresses fusion.Vaccinia virus, the prototype of the Orthopoxvirus genus of the family Poxviridae, infects many cell lines and animals (13) and produces several forms of infectious particles, among which the intracellular mature virus (IMV) is the most abundant form inside cells. The IMV can be wrapped with additional Golgi membrane, transported through microtubules, and released from cells as extracellular enveloped viruses (10). The IMV has evolved to enter host cells through plasma membrane fusion (1, 3, 12, 29, 47) or endocytosis (11, 48). Recently, Mercer et al. reported that IMV entered HeLa cells through apoptotic mimicry and macropinocytosis (32), and Huang et al. reported that IMV enters into HeLa cells through a dynamin-dependent fluid-phase endocytosis that required the cellular protein VPEF (22).The IMV contains more than 75 viral proteins. Of these, more than 10 viral envelope proteins are known to be involved in vaccinia virus entry into cells (6, 34, 55). Vaccinia virus contains at least five attachment proteins, with H3, A27, and D8 binding to cell surface glycosaminoglycans (GAGs) (7, 21, 28), A26 protein binding to the extracellular matrix protein laminin (5), and L1 protein binding to unidentified cell surface molecules (14). A27 protein also binds to the viral A17 protein through its C-terminal region (35, 50), and it was recently shown that the coexpression of A17 and A27 proteins resulted in cell fusion in transiently transfected 293T cells (27). In this study, we demonstrate the formation through disulfide bonds of complexes between two viral attachment proteins, A26 and A27, and we determine the cysteine residues that are critical for these disulfide bonds. We also address the biological role of the A26-A27 protein complex formation in cell fusion regulation.  相似文献   

19.
Herpes simplex virus (HSV) entry into cells requires four membrane glycoproteins: gD is the receptor binding protein, and gB and gH/gL constitute the core fusion machinery. Crystal structures of gD and its receptors have provided a basis for understanding the initial triggering steps, but how the core fusion proteins function remains unknown. The gB crystal structure shows that it is a class III fusion protein, yet unlike other class members, gB itself does not cause fusion. Bimolecular complementation (BiMC) studies have shown that gD-receptor binding triggers an interaction between gB and gH/gL and concurrently triggers fusion. Left unanswered was whether BiMC led to fusion or was a by-product of it. We used gB monoclonal antibodies (MAbs) to block different aspects of these events. Non-virus-neutralizing MAbs to gB failed to block BiMC or fusion. In contrast, gB MAbs that neutralize virus blocked fusion. These MAbs map to three functional regions (FR) of gB. MAbs to FR1, which contains the fusion loops, and FR2 blocked both BiMC and fusion. In contrast, MAbs to FR3, a region involved in receptor binding, blocked fusion but not BiMC. Thus, FR3 MAbs separate the BiMC interaction from fusion, suggesting that BiMC occurs prior to fusion. When substituted for wild-type (wt) gB, fusion loop mutants blocked fusion and BiMC, suggesting that loop insertion precedes BiMC. Thus, we postulate that each of the gB FRs are involved in different aspects of the path leading to fusion. Upon triggering by gD, gB fusion loops are inserted into target lipid membranes. gB then interacts with gH/gL, and this interaction is eventually followed by fusion.Entry of herpes simplex virus (HSV) into cells requires four viral glycoproteins, gB, gD, gH, and gL, plus one of several cell receptors, either herpesvirus entry mediator (HVEM), nectin-1, or 3-OST (45). Crystal structures and other studies have documented that receptor binding triggers conformational changes to gD that trigger the downstream events leading to fusion (10, 11, 18, 26, 28, 52). Moreover, when HSV receptor-bearing cells are transfected with expression plasmids for glycoproteins gB, gD, gH, and gL, the cells fuse to form multinucleated giant cells or syncytia (39, 48). However, the precise series of events that take place after receptor binding have not yet been fully elucidated. What we do know is that both gB and a heterodimer of gH/gL constitute the core fusion machinery that is conserved and required for the fusion step of entry of all herpesviruses (18, 26, 30, 46, 49).Thus far, we know the crystal structure of one form of the gB ectodomain of HSV type 1 (HSV-1) (19). This protein has the characteristics of a fusion protein and is a charter member of the class III group of viral fusion proteins (4). Others in this class include Epstein-Barr virus gB, vesicular stomatitis virus (VSV) G, and baculovirus gp64 (5, 22, 41). Like VSV G and gp64, gB has two putative fusion loops at the base of each protomer of the crystallized trimer. Single-amino-acid mutations in many of the hydrophobic residues of the putative fusion loops of gB ablate its ability to function in cell-cell fusion assays (16, 17). Moreover, these mutants are unable to complement the entry of a gB-null virus (16). Finally, the ectodomains of these mutants, unlike wild-type protein, failed to coassociate with liposomes, indicating that the putative fusion loops do insert into membranes (16, 17). Recently, it was shown that several of these mutants are also defective for fusion events involved in virus egress (51). Together, these studies provide compelling evidence that HSV gB functions as a fusion protein and that the fusion loops are critical for this function. However, unlike VSV G and baculovirus gp64, gB does not function on its own in entry but, rather, requires the participation of gH/gL. In the absence of crystallographic data for gH/gL, it is not yet clear what role it plays in herpesvirus fusion. In a previous study, we used bimolecular complementation (BiMC) to examine protein-protein interactions that occur among the viral glycoproteins during fusion (1). A similar study was carried out by Avitabile et al. (2). The BiMC assay is based on the observation that N- and C-terminal fragments of green fluorescent protein (GFP) (and derivatives such as enhanced yellow fluorescent protein [EYFP]) do not spontaneously reconstitute a functional fluorophore (20, 29, 40). However, the codons for each half can be appended to the genes for two interacting proteins (23, 24). When these are cotransfected, an interaction between the two proteins of interest brings the two halves of the fluorophore in close enough contact to restore fluorescence.When HSV receptor-bearing cells, such as B78H1 cells that are engineered to express nectin-1, are transfected with plasmids that express gB, gD, gH, and gL, they undergo cell-cell fusion (13, 15, 27, 31, 48). When gD is omitted, no fusion occurs. We found that fusion of these transfected cells could be triggered by addition of a soluble form of gD (the gD ectodomain). We then used this approach to examine interactions between gB and gH/gL during cell fusion (1). Therefore, we tagged gB with the C-terminal half of EYFP and gH with the N-terminal half. When plasmids bearing these forms were cotransfected into C10 cells along with a plasmid for untagged gL, no fusion occurred, but importantly, no BiMC occurred. However, when we added gD306, cells began to fuse within 10 min, and all of the syncytia that formed exhibited bright EYFP fluorescence indicative of BiMC. We concluded that gD triggers both fusion and a physical interaction between gB and gH/gL. However, these experiments did not separate these two events, so we were unable to determine if the interaction preceded fusion or merely was a by-product of it.The purpose of this study was to determine if the gB-gH/gL interaction is essential for fusion and if it occurs prior to fusion. We focused on gB because its structure is known and we have a panel of well-characterized monoclonal antibodies (MAbs) to gB. Our approach was to determine which of these MAbs, if any, could block fusion and also block the interaction with gH/gL. We also examined the effect of mutations to the fusion loops of gB on its interaction with gH/gL. We previously mapped these MAbs to four functional regions (FR) of gB, three of which were resolved in the crystal structure (6, 19). Of these, FR1 contains the fusion loops, FR2 is in the center of the gB structure with no known function, and FR3 is at in the crown of the protein and may be involved in binding to cells (7). Our rationale was that if the interaction between gB and gH/gL is important for fusion, then it should not be blocked by nonneutralizing anti-gB MAbs. At the same time, we thought that some neutralizing MAbs might not only block fusion but also block BiMC. We found that neutralizing MAbs to FR1 and FR2 inhibited both BiMC and fusion. In contrast, we found that neutralizing MAbs that map to FR3 blocked fusion but failed to block the interaction between gB and gH/gL, thereby dissociating the two events. Finally, we found that gB mutants with changes in the fusion loops that were fusion negative were also unable to bind to gH/gL. The latter results suggest that insertion of gB into the target membrane precedes its interaction with gH/gL.  相似文献   

20.
Clade B of the New World arenaviruses contains both pathogenic and nonpathogenic members, whose surface glycoproteins (GPs) are characterized by different abilities to use the human transferrin receptor type 1 (hTfR1) protein as a receptor. Using closely related pairs of pathogenic and nonpathogenic viruses, we investigated the determinants of the GP1 subunit that confer these different characteristics. We identified a central region (residues 85 to 221) in the Guanarito virus GP1 that was sufficient to interact with hTfR1, with residues 159 to 221 being essential. The recently solved structure of part of the Machupo virus GP1 suggests an explanation for these requirements.Arenaviruses are bisegmented, single-stranded RNA viruses that use an ambisense coding strategy to express four proteins: NP (nucleoprotein), Z (matrix protein), L (polymerase), and GP (glycoprotein). The viral GP is sufficient to direct entry into host cells, and retroviral vectors pseudotyped with GP recapitulate the entry pathway of these viruses (5, 13, 24, 31). GP is a class I fusion protein comprising two subunits, GP1 and GP2, cleaved from the precursor protein GPC (4, 14, 16, 18, 21). GP1 contains the receptor binding domain (19, 28), while GP2 contains structural elements characteristic of viral membrane fusion proteins (8, 18, 20, 38). The N-terminal stable signal peptide (SSP) remains associated with the mature glycoprotein after cleavage (2, 39) and plays a role in transport, maturation, and pH-dependent fusion (17, 35, 36, 37).The New World arenaviruses are divided into clades A, B, and C based on phylogenetic relatedness (7, 9, 11). Clade B contains the human pathogenic viruses Junin (JUNV), Machupo (MACV), Guanarito (GTOV), Sabia, and Chapare, which cause severe hemorrhagic fevers in South America (1, 10, 15, 26, 34). Clade B also contains the nonpathogenic viruses Amapari (AMAV), Cupixi, and Tacaribe (TCRV), although mild disease has been reported for a laboratory worker infected with TCRV (29).Studies with both viruses and GP-pseudotyped retroviral vectors have shown that the pathogenic clade B arenaviruses use the human transferrin receptor type 1 (hTfR1) to gain entry into human cells (19, 30). In contrast, GPs from nonpathogenic viruses, although capable of using TfR1 orthologs from other species (1), cannot use hTfR1 (1, 19) and instead enter human cells through as-yet-uncharacterized hTfR1-independent pathways (19). In addition, human T-cell lines serve as useful tools to distinguish these GPs, since JUNV, GTOV, and MACV pseudotyped vectors readily transduce CEM cells, while TCRV and AMAV GP vectors do not (27; also unpublished data). These properties of the GPs do not necessarily reflect a tropism of the pathogenic viruses for human T cells, since viral tropism is influenced by many factors and T cells are not a target for JUNV replication in vivo (3, 22, 25).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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