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 62 毫秒
1.
Hendra virus F protein-promoted membrane fusion requires the presence of the viral attachment protein, G. However, events leading to the association of these glycoproteins remain unclear. Results presented here demonstrate that Hendra virus G undergoes slower secretory pathway trafficking than is observed for Hendra virus F. This slowed trafficking is not dependent on the G protein cytoplasmic tail, the presence of the G receptor ephrin B2, or interaction with other viral proteins. Instead, Hendra virus G was found to undergo intrinsically slow oligomerization within the endoplasmic reticulum. These results suggest that the critical F-G interactions occur only after the initial steps of synthesis and cellular transport.The Henipavirus genus of the paramyxovirus family comprises two recently emerged, zoonotic pathogens. Hendra virus, first identified in Australia in 1994, caused respiratory illness in and the subsequent death of over one dozen horses and two of the three humans infected (12, 21, 25). Nipah virus led to an outbreak of respiratory and encephalitic illnesses in Malaysia in 1999, affecting both swine and humans and leading to fatality in 105 of the 265 human cases (11). Additional periodic outbreaks of infections with these viruses have occurred (11), and evidence indicates human-to-human transmission of Nipah virus in at least one outbreak (16). Henipavirus, like many paramyxoviruses, requires the presence of two surface glycoproteins for virus-cell and cell-cell fusion (8, 9, 36): the fusion protein, F, which mediates the membrane fusion event, and the attachment protein, G, which binds cellular receptors ephrin B2 (6, 22) and ephrin B3 (23) and which is required for F-mediated membrane fusion. Interactions between the fusion and attachment proteins of a number of paramyxoviruses have been observed previously (30, 33, 35), and interactions between the henipavirus F and G proteins have been demonstrated by coimmunoprecipitation (1-5, 18). However, important questions remain concerning the timing of these interactions and the mechanism by which the attachment protein regulates F-mediated fusion. Results from studies of measles virus (30), Newcastle disease virus (33), and human parainfluenza virus (35) have suggested that the initial interaction between the two glycoproteins occurs within the endoplasmic reticulum (ER) at the time of synthesis, potentially allowing the attachment protein to hold the F protein in its prefusion conformation. In contrast, the retention of the parainfluenza virus type 5 (PIV5) F protein in the ER does not lead to the retention of the PIV5 attachment protein (29), suggesting that interaction between these glycoproteins does not occur soon after synthesis. Recently, henipavirus F proteins have been shown to undergo processing through a complex intracellular trafficking pathway, with expression on the cell surface in a nonfusogenic precursor form (F0), subsequent endocytosis, cleavage by cathepsin L into the fusogenic (F1+F2) form, and retrafficking to the plasma membrane (10, 19, 26-28). While the half-life (t1/2) of F protein uncleaved by cathepsin L is approximately 2 h (28), results from initial studies of Hendra virus G trafficking indicate much slower trafficking of this protein through the secretory pathway (37), a result inconsistent with the formation of an F-G complex in the ER.To more closely examine the trafficking of the henipavirus glycoproteins, endoglycosidase H (endo H) analysis was used as a marker for trafficking time to the medial-Golgi compartment. We previously reported that wild-type (wt) Hendra virus G protein becomes endo H resistant, with a t1/2 of between 2 and 3 h (37), suggesting slow trafficking through the secretory pathway. Porotto et al. (31) described a mutant G protein lacking the first 32 residues of the cytoplasmic tail (G-Δ32), which showed enhanced fusion promotion. This Hendra virus G-Δ32 mutant also exhibited higher overall expression than the wt Hendra virus G (S. D. Whitman and R. E. Dutch, unpublished results). To determine if this mutant exhibited altered trafficking kinetics, Vero cells transfected with pCAGGS-Hendra G-Δ32 were examined by pulse-chase analysis followed by endo H treatment as described previously (37). Trafficking kinetics similar to those of wt G were observed, as G-Δ32 became endo H resistant, with a t1/2 of approximately 2 to 3 h (Fig. (Fig.1A).1A). In contrast, endo H analysis of Hendra virus F revealed a resistant population of F0 within 30 min (Fig. (Fig.1B),1B), with the majority of F converted to either an F0 endo H-resistant form or to the F1 cleaved form by 2 h. The cleavage of F0 was observed concomitantly with the appearance of two partially endo H-resistant F1 bands, suggestive of differential complex sugar additions within the Golgi compartment. These data confirm that Hendra virus F and G traffic through the secretory pathway at different rates, with Hendra virus G trafficking unaffected by the 32-amino-acid deletion. Endo H analysis of the Nipah virus G dimeric form (Fig. (Fig.1C),1C), for which the mobility shift after endo H treatment was most apparent, mirrors that of wt Hendra virus G, suggesting that slow trafficking through the secretory pathway is a property of both henipavirus G proteins.Open in a separate windowFIG. 1.Endo H digestion indicates differential rates of trafficking for the henipavirus F and G proteins. (A) Vero cells were transfected with pCAGGS-Hendra G-Δ32, and 24 h posttransfection, the cells were labeled for 30 min and chased for various times and Hendra virus G was immunoprecipitated using an antibody directed to a soluble form of G (7). Endo H digestion was performed as described previously (37). Proteins were separated via sodium dodecyl sulfate-polyacrylamide gel electrophoresis and visualized using the Typhoon imaging system. (B) Vero cells were transfected with pCAGGS-Hendra F. Twenty-four hours later, the cells were labeled for 30 min. Hendra virus F was immunoprecipitated with a Hendra virus F-specific antibody (28) and subjected to endo H treatment. (C) Pulse-chase analysis of Vero cells transfected with pCAGGS-Nipah G was performed. Immunoprecipitation with antibody to a soluble form of G (7) and endo H analysis were performed as described previously. R and S denote the endo H-resistant and -sensitive species, respectively. +, present; −, absent.Slow trafficking of Hendra virus G through the secretory pathway may be caused by interactions with other viral proteins or with its receptor ephrin B2. Chinese hamster ovary (CHO) cells do not express ephrin B2 (39) and thus were utilized to examine Hendra virus G trafficking in the absence of its cognate receptor. When Hendra virus G-Δ32 was expressed in CHO cells, an endo H-resistant population appeared at 2 h (Fig. (Fig.2A)2A) and was found to have increased at subsequent time points, consistent with results obtained using Vero cells. These data suggest that the low trafficking rate of Hendra virus G is not due to association with ephrin B2 and is not cell type specific. Coexpression of either the Hendra virus matrix (M) protein or the F protein with G-Δ32 did not alter trafficking kinetics, indicating that an interaction with either M or F is not responsible for the low rate of G-Δ32 trafficking (Whitman and Dutch, unpublished).Open in a separate windowFIG. 2.Endo H analysis indicates that slow trafficking through the secretory pathway is not dependent on ephrin B2 or on sequences present in the Hendra virus G cytoplasmic tail. (A) CHO cells expressing Hendra virus G-Δ32 were metabolically labeled with 35S for 30 min and chased for the times indicated. Following immunoprecipitation and endo H analysis, proteins were separated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and visualized via the Typhoon imaging system. (B) Pulse-chase analysis of Vero cells expressing the Hendra virus G-Δ41 tail was performed, followed by immunoprecipitation and endo H analysis. R and S denote the endo H-resistant and -sensitive species, respectively. +, present; −, absent.Lysine-rich motifs (KKXX or KXKXX) have been shown previously to be involved in the retention of type I glycoproteins in the ER (17), while motifs containing multiple arginines have been implicated in the retention of type II integral membrane proteins (20, 32). The type II Hendra and Nipah virus G proteins contain no and one arginine residue in their cytoplasmic domains, respectively, but do have several lysine-rich motifs. To determine if these motifs facilitated ER retention of Hendra virus G, thus resulting in a lowered trafficking rate, an additional mutant (Hendra virus G-Δ41) with the removal of the first 41 amino acids of the cytoplasmic tail, deleting both KKXX and KXKXX motifs and the majority of the cytoplasmic tail, was constructed. Consistent with the results from endo H analyses of wt Hendra virus G and Hendra virus G-Δ32, an endo H-resistant form of Hendra virus G-Δ41 appeared after only 2 h (Fig. (Fig.2B),2B), suggesting that Hendra virus G is not retained in the ER by the putative ER retention motifs KKXX and KXKXX or by any other sequence in the cytoplasmic tail.Endo H data strongly indicated differential rates of trafficking through the secretory pathway for Hendra virus G and Hendra virus F, suggesting that interactions between these two proteins likely occur subsequent to transport to the cell surface. To verify the different trafficking kinetics of these two proteins, the rates at which Hendra virus F and G appear on the cell surface were determined by using surface biotinylation, with the more highly expressed mutant Hendra virus G-Δ32 utilized in these experiments to facilitate visualization of the surface population. Hendra virus F was present on the cell surface as F0 at the completion of the 30-min labeling (Fig. (Fig.3A)3A) (time zero), with F1+F2 appearing on the surface at approximately 2 h (Fig. (Fig.3A),3A), consistent with the complex trafficking pathway observed for F. In contrast, the majority of Hendra virus G-Δ32 arrived on the cell surface after the 2-h time point (Fig. (Fig.3B).3B). Similar cell surface profiles were observed when the proteins were coexpressed (Fig. (Fig.3C).3C). These data confirm the differential trafficking rates and suggest that Hendra virus F and G do not interact within the secretory pathway but traffic independently to the cell surface.Open in a separate windowFIG. 3.Analysis of cell surface populations confirms the differential trafficking rates of the Hendra virus F and G proteins. (A) Vero cells were transfected with pCAGGS-Hendra F. Twenty-four hours later, the cells were metabolically labeled with 35S for 30 min and chased for the indicated times. Analysis of biotinylated surface proteins was performed as described previously (37). (B) Surface biotinylation of Vero cells expressing Hendra virus G-Δ32 was performed, and samples were analyzed as described previously (37). (C) Biotinylation analyses of Vero cells expressing both Hendra virus F and Hendra virus G-Δ32 were performed as described previously.Exit out of the ER is a carefully controlled process, allowing only properly folded proteins to be transported to the Golgi compartment (15). Thus, slow folding kinetics for Hendra virus G may explain the delayed appearance of an endo H-resistant population. Unlike Hendra virus F, which folds as a trimer (13, 14), Hendra virus G is a tetramer composed of two disulfide-linked dimers (7). To examine folding kinetics, cross-linking analysis (13) was performed at various time points post-metabolic labeling. By the end of the 30-min labeling, the majority of Hendra virus F was folded into a trimeric state which could be stabilized by the addition of a cross-linker (Fig. (Fig.4A),4A), and little of the F protein was present in the monomeric form after the addition of the cross-linker. No further increases in trimers were subsequently observed, suggesting very rapid oligomerization of Hendra virus F. Rapid oligomerization of paramyxovirus fusion proteins is consistent with the prefusion structure (38), in which monomers are tightly folded together to form the trimeric unit. The presence of a dimeric species with the addition of a cross-linker is due likely to incomplete cross-linking and does not represent an intermediate conformational population. The folding of Hendra virus G into a tetramer, however, occurs at a lower rate than that of F. In the absence of a cross-linker, monomeric G is present until 2 h postlabeling, suggesting that the formation of the disulfide-linked dimer is intrinsically slow (Fig. (Fig.4B).4B). The formation of the disulfide-linked dimer occurs at a rate similar to that of tetramer formation, as the majority of G can be cross-linked to a tetramer by the 2-h time point. Tetramerization of other paramyxovirus attachment proteins, such as PIV5 hemagglutinin-neuraminidase, occurs more rapidly, with t1/2s for these proteins of 25 to 30 min (24), contrasting greatly with the t1/2 of 1 to 2 h observed for Hendra virus G. These results suggest that, unlike other paramyxovirus attachment proteins, Hendra virus G undergoes very slow tetramerization and that the slow trafficking of this protein through the secretory pathway is likely a direct reflection of the low intrinsic rate of folding and oligomerization.Open in a separate windowFIG. 4.Cross-linking analysis of Hendra virus glycoproteins indicates slow tetramerization of the Hendra virus G protein. (A) Vero cells expressing Hendra virus F were metabolically labeled for 30 min and chased for the indicated times. Cross-linking with DTSSP [3,3′-dithiobis(sulfosuccinimidyl propionate)] was performed as described previously (14). (B) Vero cells expressing Hendra virus G-Δ32 were labeled for 30 min and chased for the indicated times, and cross-linking analysis with DTSSP was performed as described previously (14). +, present; −, absent.While the majority of paramyxovirus F proteins require their homotypic attachment protein for fusogenic activity, the role of the attachment protein in controlling F protein function remains unclear. The attachment protein has been proposed to hold the F protein in its prefusion conformation until receptor binding occurs (34), and research from several systems has suggested that this F protein-attachment protein interaction occurs in the ER during initial protein folding (30, 33, 35). Data from experiments presented here demonstrating differential rates of oligomerization and secretory pathway transport for Hendra virus F and G strongly indicate that the association of the newly synthesized proteins does not occur in the ER but that they instead traffic independently through the secretory pathway. Thus, at least for the henipaviruses, F-G interactions are unlikely to play a role in preventing premature triggering of the newly synthesized F protein. An alternative model for paramyxovirus fusion has suggested that attachment protein-fusion protein interactions occur only after the attachment protein binds the receptor. However, analysis of henipavirus G and F mutants suggests that F-G avidity inversely correlates with fusion (2, 3, 5), and Hendra virus G protein mutants deficient in receptor binding also lose the ability to coimmunoprecipitate Hendra virus F (4). These data support a model in which F-G interactions occur prior to receptor binding, with the subsequent G-ephrin B2 interaction leading to the release of the F-G interaction and the triggering of fusion. Taken together, these results support a model in which the henipavirus F and G associate only after trafficking to the cell surface. The mechanisms by which this interaction is promoted and/or regulated represent an exciting area of future research.  相似文献   

2.
The entry of enveloped viruses into host cells is preceded by membrane fusion, which in paramyxoviruses is triggered by the fusion (F) protein. Refolding of the F protein from a metastable conformation to a highly stable postfusion form is critical for the promotion of fusion, although the mechanism is still not well understood. Here we examined the effects of mutations of individual residues of the F protein of Newcastle disease virus, located at critical regions of the protein, such as the C terminus of the N-terminal heptad repeat (HRA) and the N terminus of the C-terminal heptad repeat (HRB). Seven of the mutants were expressed at the cell surface, showing differences in antibody reactivity in comparison with the F wild type. The N211A, L461A, I463A, and I463F mutants showed a hyperfusogenic phenotype both in syncytium and in dye transfer assays. The four mutants promoted fusion more efficiently at lower temperatures than the wild type did, meaning they probably had lower energy requirements for activation. Moreover, the N211A, I463A, and I463F mutants exhibited hemagglutinin-neuraminidase (HN)-independent activity when influenza virus hemagglutinin (HA) was coexpressed as an attachment protein. The data are discussed in terms of alterations of the refolding pathway and/or the stability of the prefusion and fusion conformations.Newcastle disease virus (NDV) is an avian enveloped virus belonging to the family Paramyxoviridae. Two viral membrane-associated proteins are responsible for the entry of the virus into the host cell: they are hemagglutinin-neuraminidase (HN), a receptor-binding protein that interacts with sialoglycoconjugates at the cell surface, and F, a trimeric class I fusion protein that, upon activation, triggers the fusion of the viral and target membranes. F protein is activated after the attachment of its homotypic HN protein to the proper receptor; however, how HN activates F is not well understood. F protein is synthesized as an inactive precursor, F0, that is activated by proteolytic cleavage to the disulfide-linked F1-F2 fusion-competent form (Fig. (Fig.1)1) (10). The crystal structures of several paramyxoviral fusion proteins, in both the prefusion and postfusion conformations (3, 26, 27), have revealed that these proteins undergo major conformational changes, from a metastable conformation to a highly stable, postfusion form. Several regions in the ectodomain of class I viral fusion proteins are involved in these conformational conversions, including a hydrophobic fusion peptide at the N terminus of the F1 protein and two hydrophobic heptad repeat motifs, HRA and HRB, located at its N and C termini, respectively (Fig. (Fig.1).1). In the prefusion form, HRB shows a triple-stranded coiled-coil conformation forming the stalk of the mushroom-like protein (3, 19, 27). Its globular head contains three domains, DI to DIII (Fig. (Fig.1),1), with the base of the head being formed by the DI and DII domains, with residues predominantly located between HRA and HRB. The top of the head is formed by DIII, consisting mainly of HRA and the fusion peptide, located on the side of the head sequestered between adjacent subunits. In this prefusion state, HRA is folded as two antiparallel β-strands and four (h1 to h4) helices (27) (see Fig. Fig.6).6). The DIII domain undergoes major structural changes from the prefusion to the final postfusion conformation. HRA refolds as an α-helix, propelling the fusion peptide into the target membrane and generating a prehairpin intermediate (see Fig. Fig.6).6). The final, stable conformation consists of a six-helical bundle (6HB), comprising a dimer of trimers in which the trimeric HRA coiled coil forms the core, packed along the outside by three antiparallel HRB α-helices (1, 3, 19, 27).Open in a separate windowFIG. 1.Schematic representation of the structure of the NDV fusion protein. (A) Domain structure of F protein (27). (B) Locations of the fusion peptide, HR regions, and sequences studied. Mutated residues are indicated in bold.Open in a separate windowFIG. 6.Scheme of conformational changes in HRA from prefusion to postfusion state. (A) Ribbon model of PIV5 F protein in its metastable prefusion conformation (PDB accession number 2b9b) (27), showing some residues (named in white) from the A subunit and the corresponding residues in the NDV F protein (named in yellow). Subunits B and C are depicted in gray for clarity. (B) In the metastable, prefusion conformation, HRA is folded as a spring-loaded mixture of α-helices, turns, and β-strands, comprising 11 segments in the DIII head domain of the trimer (27). (C) After fusion, HRA is presented as a single long helix that allows the fusion peptide to be buried in the target membrane. The approximate positions of HRC and the core β-sheet are shown as dashed lines for both conformations.The refolding mechanism that triggers F protein activation is still not well understood. Mutational analysis of the HRA and HRB domains of paramyxovirus F proteins (3, 13, 18, 19, 22, 23), as well as the use of HRA- and HRB-derived peptides (6, 17), has led to the proposal of a series of discrete refolding intermediates of the F protein, from the metastable native conformation, through the prehairpin intermediate, and to the final postfusion hairpin structure (6HB) (17, 19, 27). To gain further insight into the individual residues critical for this mechanism, in this work we mutated several residues of the head and stalk of the NDV F protein (Fig. (Fig.1).1). The mutations disrupted F protein antibody reactivity, fusogenicity, and HN dependence in different ways. Interestingly, a mutant of the C-terminal h4 α-helix of HRA (N211A mutant) and two mutants of a residue located at the most N-terminal position of HRB (I463A and I463F mutants) exhibited a hyperfusogenic phenotype and HN-independent activity when influenza virus hemagglutinin (HA) was coexpressed as an attachment protein. The data are discussed in terms of alterations of the refolding pathway and/or the stability of the prefusion and fusion conformations.  相似文献   

3.
The deadly paramyxovirus Nipah virus (NiV) contains a fusion glycoprotein (F) with canonical structural and functional features common to its class. Receptor binding to the NiV attachment glycoprotein (G) triggers F to undergo a two-phase conformational cascade: the first phase progresses from a metastable prefusion state to a prehairpin intermediate (PHI), while the second phase is marked by transition from the PHI to the six-helix-bundle hairpin. The PHI can be captured with peptides that mimic F''s heptad repeat regions, and here we utilized a NiV heptad repeat peptide to quantify PHI formation and the half-lives (t1/2) of the first and second fusion cascade phases. We found that ephrinB2 receptor binding to G triggered ∼2-fold more F than that triggered by ephrinB3, consistent with the increased rate and extent of fusion observed with ephrinB2- versus ephrinB3-expressing cells. In addition, for a series of hyper- and hypofusogenic F mutants, we quantified F-triggering capacities and measured the kinetics of their fusion cascade phases. Hyper- and hypofusogenicity can each be manifested through distinct stages of the fusion cascade, giving rise to vastly different half-lives for the first (t1/2, 1.9 to 7.5 min) or second (t1/2, 1.5 to 15.6 min) phase. While three mutants had a shorter first phase and a longer second phase than the wild-type protein, one mutant had the opposite phenotype. Thus, our results reveal multiple critical parameters that govern the paramyxovirus fusion cascade, and our assays should help efforts to elucidate other class I membrane fusion processes.Nipah (NiV) and Hendra (HeV) viruses are emerging members of the new Paramyxoviridae genus Henipavirus (12, 19). The Paramyxoviridae family comprises important viral pathogens, such as measles, mumps, human parainfluenza, respiratory syncytial, and Newcastle disease viruses and the henipaviruses (HNV), and NiV is its deadliest known member (4, 5). NiV has a broad host range and causes respiratory and neurological symptoms that often lead to encephalitis and a mortality rate of up to 75% in humans (21, 47). It can also spread efficiently and cause morbidity in economically important livestock (21). NiV is a biosafety level 4 (BSL4) pathogen and is considered a select agent with bio- and agro-terrorism potential. Both animal-to-human and human-to-human transmissions have been documented (4, 5), underscoring the need for research and treatment development. Since microvascular endothelial cell-cell fusion (syncytium formation) is a pathognomonic hallmark of NiV infection (50), understanding virus-cell and cell-cell membrane fusion should assist in the development of therapeutics to target this aspect of NiV pathobiology.Paramyxovirus membrane fusion requires the coordinated action of the attachment (G, HN, or H) and fusion (F) glycoproteins, and numerous canonical structural and functional features of G/HN/H and F proteins are conserved among paramyxoviruses (20, 23, 46, 48). G/HN/H proteins have a receptor-binding globular domain formed by a six-bladed beta-propeller connected to its transmembrane anchor via a flexible stalk domain (10, 51). For NiV and HeV, both ephrinB2 (B2) and ephrinB3 (B3) can be used as cell receptors (8, 33, 34), although B2 appears to be the higher-affinity receptor (34). B2 or B3 receptors bind to and activate G, which in turn triggers a conformation cascade in F that leads to membrane fusion (1). HNV F proteins are trimeric class I fusion proteins with structural/functional features common to their class (23, 52). HNV F proteins are synthesized as precursors that are cleaved and hence activated into a metastable conformation, poised for enabling membrane fusion. Cleavage generates a new N terminus that contains a hydrophobic fusion peptide (48). For NiV and HeV, the precursor (F0) reaches the plasma membrane uncleaved, but endocytosis exposes F0 to cathepsin L in the endosomes, cleaving F0 to generate mature disulfide-linked F1 and F2 subunits that are trafficked back to the cell surface (14, 31). The structures of the retroviral Moloney murine leukemia virus p15E, lentiviral human immunodeficiency virus type 1 (HIV-1) gp41, Ebola virus GP2, influenza virus HA, and paramyxovirus SV5 and NiV-F fusion proteins all share similar trimeric coiled-coil core structures (6, 11, 17, 27, 53) and, in general, similar membrane fusion mechanisms (22, 23, 48).Receptor binding to paramyxoviral G/HN/H triggers a conformational cascade in F, leading to membrane fusion (Fig. (Fig.1).1). Although the determinants for F triggering on G/HN/H have not been defined clearly, evidence suggests that the stalk domain (7, 13, 24, 28, 29) and, at least for NiV, a region at the base of the globular domain of G (1) are involved in F triggering. Additionally, recent evidence indicates an interaction between the stalk region of the measles virus H protein and the globular domain of the cognate F protein (35). Once triggered, F progresses through a prehairpin intermediate (PHI) (Fig. 1A and B). In the PHI conformation, the fusion peptide is harpooned into the host cell membrane, and the N- and C-terminal heptad repeat domains (HR1 and HR2, respectively) are exposed. The HR domains then coalesce into the postfusion six-helix-bundle (6HB) hairpin conformation. In the 6HB, the transmembrane and fusion peptide domains are juxtaposed, bringing viral and target cell membranes together and driving membrane fusion (Fig. (Fig.1C)1C) (30, 48). Much evidence suggests that 6HB formation is coincident with membrane merger and that synthetic HR1 and HR2 peptides only bind to and inhibit fusion intermediates (e.g., PHI) prior to 6HB formation (9, 30, 37, 43, 48). Additionally, HR1 peptides can inhibit an earlier fusion intermediate than that inhibited by HR2 peptides (43), and HR2 peptides are invariably more potent inhibitors of fusion than HR1 peptides. HR2 peptides trap the PHI by binding to the radial interstices formed by the trimeric HR1 core, inhibiting 6HB formation and membrane fusion (22, 23, 48). Altogether, there is much evidence to support the fusion cascade shown in Fig. Fig.11 and the use of HR2 peptides to physically capture fusion intermediates (9, 30, 43, 48).Open in a separate windowFIG. 1.Nipah virus fusion cascade. The schematic shows the NiV fusion cascade broken down into three major stages. (A) EphrinB2 or ephrinB3 binding to NiV-G triggers the metastable NiV-F protein through allosteric mechanisms that are still being elucidated. (B) After F is triggered, it forms the PHI, in which a fusion peptide is harpooned into the host cell membrane. The PHI can be captured by peptides that mimic the NiV-F HR1 (orange-striped cylinder) or HR2 (green-striped cylinder) region and bind the F HR2 or HR1 region, respectively. (C) The HR1 and HR2 regions in the PHI coalesce to form the 6HB conformation, bringing the viral and cell membranes together and facilitating virus-host membrane fusion and viral entry. The viral membrane can be replaced by a cell membrane expressing the F and G glycoproteins in cell-cell fusion, resulting in syncytium formation. We term the transitions from A to B and from B to C phases I and II, respectively, of the fusion cascade. (D) Schematic representation of the F-triggering assay, showing its four main steps: (1) receptor binding at 4°C, (2) biotinylated HR2 peptide addition and induction of F triggering at 37°C, (3) fixation at 4°C with paraformaldehyde, and (4) signal amplification at 4°C. In the “time-of-addition” and “time-of-stopping” experiments, step 2 was modified as indicated in the text. The HR2 peptide (green hatched column) is shown with its N-terminal biotin modification (red star). Blue stars, streptavidin-APC; black, three-pronged symbols, activator; blue symbols with red octagons, enhancer.We previously developed a fluorescence-activated cell sorting (FACS)-based NiV-F-triggering assay by measuring the amount of HR2 peptide binding to F/G-expressing cells triggered by cell surface ephrinB2 (1). In this study, we further optimized our assay for robust quantification of HR2 peptide binding and used this assay to monitor the differential degree of F triggering induced by B2 or B3. In addition, through “time-of-addition” and “time-of-stopping” experiments (described below), we show that this HR2 binding assay can measure the half-lives of various fusion intermediates, i.e., the transition times from the prefusion (PF) state to PHI and from PHI to 6HB. Using a panel of hyper- and hypofusogenic mutants, we show that hyper- and hypofusogenicity can each be manifested through distinct effects on the half-lives of these fusion intermediates and/or the absolute amounts of F triggering. Thus, we elucidated the impacts of different mutations on individual steps of the fusion cascade. Since HR2 peptides can generally capture the PHI of class I fusion proteins, our assays should help efforts to understand fusion processes mediated by other class I fusion proteins.  相似文献   

4.
The envelope protein E of flaviviruses mediates both receptor-binding and membrane fusion. At the virion surface, 180 copies of E are tightly packed and organized in a herringbone-like icosahedral structure, whereas in noninfectious subviral particles, 60 copies are arranged in a T=1 icosahedral symmetry. In both cases, the basic building block is an E dimer which exposes the binding sites for neutralizing antibodies at its surface. It was the objective of our study to assess the dependence of the antigenic structure of E on its quaternary arrangement, i.e., as part of virions, recombinant subviral particles, or soluble dimers. For this purpose, we used a panel of 11 E protein-specific neutralizing monoclonal antibodies, mapped to distinct epitopes in each of the three E protein domains, and studied their reactivity with the different soluble and particulate forms of tick-borne encephalitis virus E protein under nondenaturing immunoassay conditions. Significant differences in the reactivities with these forms were observed that could be related to (i) limited access of certain epitopes at the virion surface; (ii) limited occupancy of epitopes in virions due to steric hindrance between antibodies; (iii) differences in the avidity to soluble forms compared to the virion, presumably related to the flexibility of E at its domain junctions; and (iv) modulations of the external E protein surface through interactions with its stem-anchor structure. We have thus identified several important factors that influence the antigenicity of the flavivirus E protein and have an impact on the interaction with neutralizing antibodies.Flaviviruses form a genus in the family Flaviviridae (52) and comprise a number of important human pathogens such as yellow fever, dengue, Japanese encephalitis, West Nile, and tick-borne encephalitis (TBE) viruses (30). They are small, enveloped viruses with only three structural proteins, designated C (capsid), M (membrane), and E (envelope). The E protein is oriented parallel to the viral membrane and forms a head-to-tail homodimeric complex (Fig. 1A and B). The structure of the E ectodomain (soluble E [sE])—consisting of about 400 amino acids and lacking the 100 C-terminal amino acids (including the so-called stem and two transmembrane helices)—has been determined by X-ray crystallography for several flaviviruses (Fig. (Fig.1A)1A) (25, 34, 36, 38, 44, 55). Both of the essential entry functions—receptor-binding and membrane fusion after uptake by receptor-mediated endocytosis—are mediated by E, which is therefore the primary target for virus-neutralizing antibodies (11, 42, 43, 45).Open in a separate windowFIG. 1.Structures and schematic representations of the TBE virus E protein, virions, and RSPs. In all panels, DI, DII, and DIII of the E protein are shown in red, yellow, and blue, respectively, and the fusion peptide (FP) is in orange. (A) Ribbon diagram of the sE dimer (top view). (B) Schematic of the full-length E dimer in a top view (upper panel) and side view (lower panel). The position of the two transmembrane helices of the membrane anchor and the two helices of the stem are based on Zhang et al. (54) and are shown in green and purple, respectively. (C) Pseudo-atomic structure of the virion based on cryo-EM reconstructions of dengue and West Nile viruses (27, 37, 54). One of the 30 rafts, each consisting of three parallel dimers, is highlighted. DIIIs of three monomers belonging to one icosahedral asymmetric unit are labeled by white stars. (D) Pseudo-atomic structure of RSP based on cryo-EM reconstructions (12).As revealed by cryo-electron microscopy (cryo-EM), mature infectious virions have smooth surfaces, comparable to a golf ball (27, 37). Their envelopes are icosahedrally symmetric and consist of a closed shell of 180 E monomers that are arranged in a herringbone-like pattern of 30 rafts of three dimers each (Fig. (Fig.1C)1C) (27). On the other hand, capsid-lacking subviral particles, which can be produced in recombinant form by the coexpression of prM and E, have a different symmetry, with 30 E dimers in a T=1 icosahedral structure (Fig. (Fig.1D)1D) (12, 49).The peculiar organization of E in virions is reminiscent of the tight packing of capsid proteins in nonenveloped viruses, for which it was shown that the native antigenic structure is strongly dependent on the intact capsid structure and not completely represented by isolated forms of capsid proteins (1, 41, 53). Such modulations of antigenic structure may be due to conformational changes in the course of packaging the capsid proteins into virions and/or to the fact that antibody binding sites at the virion surface are composed of residues that come together only through the juxtaposition of capsid proteins or neighboring protein subunits. Even in the case of spiky viral envelope proteins, the dependence of certain epitopes on the quaternary organization of the envelope glycoproteins has been described (8, 47).For flaviviruses, structural studies provide evidence for the considerable flexibility of E, especially at the junctions between the individual domains I, II, and III (DI, DII, and DIII) (7, 35, 55), suggesting that soluble forms may display differences in antigenic structure compared to those fixed in the closed envelope shell of whole virions. Furthermore, because of the tight packing of E at the virion surface, certain epitopes may be cryptic in the context of whole virus particles but accessible in soluble forms of E (40, 51).Studies on the antigenic structure of flaviviruses have used different antigen preparations including virions, recombinant subviral particles (RSPs), and soluble forms and subunits of E (10, 15-17, 32, 39, 40, 46, 49, 51), but so far no systematic comparative analysis of E in different physical forms and quaternary arrangements has been conducted. It was therefore the objective of our study, using TBE virus as a model, to investigate possible structural and/or antigenic differences between (i) soluble dimeric forms of E, including C-terminally truncated sE and detergent-solubilized full-length E (Fig. 1A and B); (ii) E in the context of whole virions (Fig. (Fig.1C);1C); and (iii) E in the context of RSPs (Fig. (Fig.1D).1D). For this purpose we used, and further characterized, a set of monoclonal antibodies (MAbs) directed to each of the three domains of E. All of these MAbs have neutralizing activity (17, 24) and therefore, by definition, react with infectious virions.Through these analyses, we demonstrate that the reactivity of several MAbs is significantly dependent on the quaternary arrangement of E and differs between virions, RSPs, and/or sE dimers. We thus provide evidence for previously unrecognized structural factors that have an impact on the antigenicity of the flavivirus E protein.  相似文献   

5.
T20 (generic name, enfuvirtide; brand name, Fuzeon) is a first-generation human immunodeficiency virus (HIV) fusion inhibitor approved for salvage therapy of HIV-infected patients refractory to current antiretroviral drugs. However, its clinical use is limited because of rapid emergence of T20-resistant viruses in T20-treated patients. Therefore, T1249 and T1144 are being developed as the second- and third-generation HIV fusion inhibitors, respectively, with improved efficacy and drug resistance profiles. Here, we found that combinations of T20 with T1249 and/or T1144 resulted in exceptionally potent synergism (combination index, <0.01) against HIV-1-mediated membrane fusion by 2 to 3 orders of magnitude in dose reduction. Highly potent synergistic antiviral efficacy was also achieved against infection by laboratory-adapted and primary HIV-1 strains, including T20-resistant variants. The mechanism underlying the synergistic effect could be attributed to the fact that T20, T1249, and T1144 all contain different functional domains and have different primary binding sites in gp41. As such, they may work cooperatively to inhibit gp41 six-helix bundle core formation, thereby suppressing virus-cell fusion. Therefore, these findings strongly imply that, rather than replacing T20, combining it with HIV fusion inhibitors of different generations might produce synergistic activity against both T20-sensitive and -resistant HIV-1 strains, suggesting a new therapeutic strategy for the treatment of HIV-1 infection/AIDS.In the early 1990s, a number of highly potent anti-human immunodeficiency virus type 1 (HIV-1) peptides derived from the C-heptad repeat (CHR) domain of the HIV-1 envelope glycoprotein (Env) transmembrane subunit gp41 were discovered (21, 22, 35, 59, 61). Biophysical and biochemical analyses suggest that the CHR peptides inhibit HIV-1 Env-mediated membrane fusion by interacting with the viral gp41 N-heptad repeat (NHR) domain to form heterologous trimer-of-heterodimer complexes, thus blocking gp41 six-helix bundle (6-HB) core formation, a critical step in virus-cell fusion (4, 5, 31, 52, 57).T20 (generic name, enfuvirtide; brand name, Fuzeon), a 36-mer CHR peptide (amino acids [aa] 638 to 673) containing a heptad repeat (HR) sequence-binding domain (HBD) and a tryptophan-rich domain (TRD) (Fig. (Fig.1)1) (30, 61), was licensed by the U.S. FDA as a first-generation HIV fusion inhibitor. T20 is very effective in inhibiting infection by HIV-1, especially the strains resistant to current antiretroviral therapies (24). However, many patients are now failing to respond to T20 because the viruses have developed T20 resistance (34, 51, 56, 62).Open in a separate windowFIG. 1.Functional domains of HIV fusion inhibitors and the interaction model. (A) Schematic view of the HIV-1HXB2 gp41 molecule and sequences of the first-, second-, and third-generation HIV fusion inhibitors. FP, fusion peptide; TM, transmembrane domain; CP, cytoplasmic domain. (B) Interaction between the NHR and CHR peptides. The dashed lines between the NHR and CHR domains indicate the interaction between the residues located at the e and g and a and d positions in the NHR and CHR, respectively. The PBD, HBD, and TRD in the CHR peptides are shown in blue, light blue, and orange, respectively. The HR sequence, the region of aa 36 to 45 (determinant for T20 resistance and the primary binding site for T20), and the pocket-forming sequence in the NHR are shown in red, purple, and green, respectively. The interaction between the PBD and pocket-forming sequence is critical for stabilization of the 6-HB (3).T1249, a second-generation HIV fusion inhibitor, is a 39-mer peptide consisting of a pocket-binding domain (PBD), an HBD, and a TRD (Fig. (Fig.1).1). T1249 was shown to have a longer half-life than T20 in primates (7) and greater anti-HIV-1 potency than T20 in clinical studies and to be active against some T20-resistant HIV-1 variants (7, 14, 27, 38). However, the clinical development of T1249 was discontinued due to formulation difficulties (37).T1144, a third-generation HIV fusion inhibitor, is a 38-mer peptide containing a PBD and an HBD (Fig. (Fig.1).1). T1144 was designed by modifying the amino acid sequence of T651 (peptide C38; aa 626 to 673) to increase α-helicity and 6-HB stability and to improve pharmacokinetic properties (10). T1144 and its analog peptides are effective against viruses that are resistant to T20 (11).Sifuvirtide, a new generation of HIV fusion inhibitor, is a 34-mer peptide analogue of C34 containing a PBD and an HBD. Our previous studies have shown that sifuvirtide is more effective than T20 against both primary and laboratory-adapted HIV-1 strains. Pharmacokinetic studies of sifuvirtide demonstrated longer decay half-lives than T20 (19). Sifuvirtide is under phase II clinical trial (www.fusogen.com). Most recently, we found that the combination of sifuvirtide with T20 resulted in potent synergistic effect against T20-sensitive and -resistant HIV-1 strains (43). These findings encouraged us to test whether combining T20 with T1249 and/or T1144 would also have synergistic anti-HIV-1 activity since next-generation HIV fusion inhibitors, like C34 and sifuvirtide, also contain a PBD that can interact with pocket-forming sequence in the gp41 NHR. In this study, we were also motivated to address the mechanism(s) underlying a synergic effect. Once this effect is confirmed, a novel combination therapy could be designed for the treatment of HIV/AIDS patients who have failed to respond to T20 or other antiretroviral drugs.  相似文献   

6.
7.
In the nitrate-responsive, homodimeric NarX sensor, two cytoplasmic membrane α-helices delimit the periplasmic ligand-binding domain. The HAMP domain, a four-helix parallel coiled-coil built from two α-helices (HD1 and HD2), immediately follows the second transmembrane helix. Previous computational studies identified a likely coiled-coil-forming α-helix, the signaling helix (S helix), in a range of signaling proteins, including eucaryal receptor guanylyl cyclases, but its function remains obscure. In NarX, the HAMP HD2 and S-helix regions overlap and apparently form a continuous coiled-coil marked by a heptad repeat stutter discontinuity at the distal boundary of HD2. Similar composite HD2-S-helix elements are present in other sensors, such as Sln1p from Saccharomyces cerevisiae. We constructed deletions and missense substitutions in the NarX S helix. Most caused constitutive signaling phenotypes. However, strongly impaired induction phenotypes were conferred by heptad deletions within the S-helix conserved core and also by deletions that remove the heptad stutter. The latter observation illuminates a key element of the dynamic bundle hypothesis for signaling across the heptad stutter adjacent to the HAMP domain in methyl-accepting chemotaxis proteins (Q. Zhou, P. Ames, and J. S. Parkinson, Mol. Microbiol. 73:801-814, 2009). Sequence comparisons identified other examples of heptad stutters between a HAMP domain and a contiguous coiled-coil-like heptad repeat sequence in conventional sensors, such as CpxA, EnvZ, PhoQ, and QseC; other S-helix-containing sensors, such as BarA and TorS; and the Neurospora crassa Nik-1 (Os-1) sensor that contains a tandem array of alternating HAMP and HAMP-like elements. Therefore, stutter elements may be broadly important for HAMP function.Transmembrane signaling in homodimeric bacterial sensors initiates upon signal ligand binding to the extracytoplasmic domain. In methyl-accepting chemotaxis proteins (MCPs), the resulting conformational change causes a displacement of one transmembrane α-helix (TM α-helix) relative to the other. This motion is conducted by the HAMP domain to control output domain activity (reviewed in references 33 and 39).Certain sensors of two-component regulatory systems share topological organization with MCPs. For example, the paralogous nitrate sensors NarX and NarQ contain an amino-terminal transmembrane signaling module similar to those in MCPs, in which a pair of TM α-helices delimit the periplasmic ligand-binding domain (Fig. (Fig.1)1) (24) (reviewed in references 32 and 62). The second TM α-helix connects to the HAMP domain. Hybrid proteins in which the NarX transmembrane signaling module regulates the kinase control modules of the MCPs Tar, DifA, and FrzCD demonstrate that NarX and MCPs share a mechanism for transmembrane signaling (73, 74, 81, 82).Open in a separate windowFIG. 1.NarX modular structure. Linear representation of the NarX protein sequence, from the amino (N) to carboxyl (C) termini, drawn to scale. The four modules are indicated at the top of the figure and shown in bold typeface, whereas domains within each module are labeled with standard (lightface) typeface. The nomenclature for modules follows that devised by Swain and Falke (67) for MCPs. Overlap between the HAMP domain HD2 and S-helix elements is indicated in gray. The three conserved Cys residues within the central module (62) are indicated. TM1 and TM2 denote the two transmembrane helices. Helices H1 to H4 of the periplasmic domain (24), and the transmitter domain H, N, D, G (79), and X (41) boxes, are labeled. The HPK 7 family of transmitter sequences, including NarX, have no F box and an unconventional G box (79). The scale bar at the bottom of the figure shows the number of aminoacyl residues.The HAMP domain functions as a signal conversion module in a variety of homodimeric proteins, including histidine protein kinases, adenylyl cyclases, MCPs, and certain phosphatases (12, 20, 77). This roughly 50-residue domain consists of a pair of amphiphilic α-helices, termed HD1 and HD2 (formerly AS1 and AS2) (67), joined by a connector (Fig. (Fig.2A).2A). Results from nuclear magnetic resonance (NMR) and electron paramagnetic resonance (EPR) spectroscopy, Cys and disulfide scanning, and mutational analysis converge on a model in which the HD1 and HD2 α-helices form a four-helix parallel coiled-coil (7, 20, 30, 42, 67, 75, 84). The mechanisms through which HAMP domains mediate signal conduction remain to be established (30, 42, 67, 84) (for commentary, see references 43, 49, and 50).Open in a separate windowFIG. 2.HAMP domain extensions. (A) Sequences from representative MCPs (E. coli Tsr and Salmonella enterica serovar Typhimurium Tar) and S-helix-containing sensors (E. coli NarX, NarQ, and BarA, and S. cerevisiae Sln1p). The HAMP domain, S-helix element, and the initial sequence of the MCP adaptation region are indicated. Flanking numbers denote positions of the terminal residues within the overall sequence. Sequential heptad repeats are indicated in alternating bold and standard (lightface) typeface. Numbering for heptad repeats in the methylation region and S-helix sequences has been described previously (4, 8). Numbers within the HD1 and HD2 helices indicate interactions within the HAMP domain (42). Residues at heptad positions a and d are enclosed within boxes, residues at the stutter position a/d are enclosed within a thickly outlined box, and residues in the S-helix ERT signature are in bold typeface. (B) NarX mutational alterations. Deletions are depicted as boxes, and missense substitutions are shown above the sequence. Many of these deletions were reported previously (10) and are presented here for comparison. The phenotypes conferred by the alterations are indicated as follows: impaired induction, black box; constitutive and elevated basal, light gray box; reversed response, dark gray box; wild-type, white box; null, striped box.Coiled-coils result from packing of two or more α-helices (27). The primary sequence of coiled-coils exhibits a characteristic heptad repeat pattern, denoted as a-b-c-d-e-f-g (52, 61), in which positions a and d are usually occupied by nonpolar residues (reviewed in references 1, 47, and 80). For example, the coiled-coil nature of the HAMP domain can be seen in the heptad repeat patterns within the HD1 and HD2 sequences (Fig. (Fig.2A2A).Coiled-coil elements adjacent to the HAMP domain have been identified in several sensors, including Saccharomyces cerevisiae Sln1p (69) and Escherichia coli NarX (60). Recently, this element was defined as a specific type of dimeric parallel coiled-coil, termed the signaling helix (S helix), present in a wide range of signaling proteins (8). Sequence comparisons delimit a roughly 40-residue element with a conserved heptad repeat pattern (Fig. (Fig.2A).2A). Based on mutational analyses of Sln1p and other proteins, the S helix is suggested to function as a switch that prevents constitutive activation of adjacent output domains (8).The term “signaling helix” previously was used to define the α4-TM2 extended helix in MCPs (23, 33). Here, we use the term S helix to denote the element described by Anantharaman et al. (8).The NarX and NarQ sensors encompass four distinct modules (Fig. (Fig.1):1): the amino-terminal transmembrane signaling module, the signal conversion module (including the HAMP domain and S-helix element), the central module of unknown function, and the carboxyl-terminal transmitter module (62). The S-helix element presumably functions together with the HAMP domain in conducting ligand-responsive motions from the transmembrane signaling module to the central module, ultimately regulating transmitter module activity.Regulatory output by two-component sensors reflects opposing transmitter activities (reviewed in reference 55). Positive function results from transmitter autokinase activity; the resulting phosphosensor serves as a substrate for response regulator autophosphorylation. Negative function results from transmitter phosphatase activity, which accelerates phosphoresponse regulator autodephosphorylation (reviewed in references 64 and 65). We envision a homogeneous two-state model for NarX (17), in which the equilibrium between these mutually exclusive conformations is modulated by ligand-responsive signaling.Previous work from our laboratory concerned the NarX and other HAMP domains (9, 10, 26, 77) and separately identified a conserved sequence in NarX and NarQ sensors, the Y box, that roughly corresponds to the S helix (62). Therefore, we were interested to explore the NarX S helix and to test some of the predictions made for its function. Results show that the S helix is critical for signal conduction and suggest that it functions as an extension of the HAMP HD2 α-helix in a subset of sensors exemplified by Sln1p and NarX. Moreover, a stutter discontinuity in the heptad repeat pattern was found to be essential for the NarX response to signal and to be conserved in several distinct classes of HAMP-containing sensors.  相似文献   

8.
9.
Complex N-glycans flank the receptor binding sites of the outer domain of HIV-1 gp120, ostensibly forming a protective “fence” against antibodies. Here, we investigated the effects of rebuilding this fence with smaller glycoforms by expressing HIV-1 pseudovirions from a primary isolate in a human cell line lacking N-acetylglucosamine transferase I (GnTI), the enzyme that initiates the conversion of oligomannose N-glycans into complex N-glycans. Thus, complex glycans, including those that surround the receptor binding sites, are replaced by fully trimmed oligomannose stumps. Conversely, the untrimmed oligomannoses of the silent domain of gp120 are likely to remain unchanged. For comparison, we produced a mutant virus lacking a complex N-glycan of the V3 loop (N301Q). Both variants exhibited increased sensitivities to V3 loop-specific monoclonal antibodies (MAbs) and soluble CD4. The N301Q virus was also sensitive to “nonneutralizing” MAbs targeting the primary and secondary receptor binding sites. Endoglycosidase H treatment resulted in the removal of outer domain glycans from the GnTI- but not the parent Env trimers, and this was associated with a rapid and complete loss in infectivity. Nevertheless, the glycan-depleted trimers could still bind to soluble receptor and coreceptor analogs, suggesting a block in post-receptor binding conformational changes necessary for fusion. Collectively, our data show that the antennae of complex N-glycans serve to protect the V3 loop and CD4 binding site, while N-glycan stems regulate native trimer conformation, such that their removal can lead to global changes in neutralization sensitivity and, in extreme cases, an inability to complete the conformational rearrangements necessary for infection.The intriguing results of a recent clinical trial suggest that an effective HIV-1 vaccine may be possible (97). Optimal efficacy may require a component that induces broadly neutralizing antibodies (BNAbs) that can block virus infection by their exclusive ability to recognize the trimeric envelope glycoprotein (Env) spikes on particle surfaces (43, 50, 87, 90). Env is therefore at the center of vaccine design programs aiming to elicit effective humoral immune responses.The amino acid sequence variability of Env presents a significant challenge for researchers seeking to elicit broadly effective NAbs. Early sequence comparisons revealed, however, that the surface gp120 subunit can be divided into discrete variable and conserved domains (Fig. (Fig.1A)1A) (110), the latter providing some hope for broadly effective NAb-based vaccines. Indeed, the constraints on variability in the conserved domains of gp120 responsible for binding the host cell receptor CD4, and coreceptor, generally CCR5, provide potential sites of vulnerability. However, viral defense strategies, such as the conformational masking of conserved epitopes (57), have made the task of eliciting bNAbs extremely difficult.Open in a separate windowFIG. 1.Glycan biosynthesis and distribution on gp120 and gp41. (A) Putative carbohydrate modifications are shown on gp120 and gp41 secondary structures, based on various published works (26, 42, 63, 74, 119, 128). The gp120 outer domain is indicated, as are residues that form the SOS gp120-gp41 disulfide bridge. The outer domain is divided into neutralizing and silent faces. Symbols distinguish complex, oligomannose, and unknown glycans. Generally, the complex glycans of the outer domain line the receptor binding sites of the neutralizing face, while the oligomannose glycans of the outer domain protect the silent domain (105). Asterisks denote sequons that are unlikely to be utilized, including position 139 (42), position 189 (26, 42), position 406 (42, 74), and position 637 (42). Glycans shown in gray indicate when sequon clustering may lead to some remaining unused, e.g., positions 156 and 160 (42, 119), positions 386, 392, and 397 (42), and positions 611 and 616 (42). There is also uncertainty regarding some glycan identities: glycans at positions 188, 355, 397, and 448 are not classified as predominantly complex or oligomannose (26, 42, 63, 128). The number of mannose moieties on oligomannose glycans can vary, as can the number of antennae and sialic acids on complex glycans (77). The glycan at position 301 appears to be predominantly a tetra-antennary complex glycan, as is the glycan at position 88, while most other complex glycans are biantennary (26, 128). (B) Schematic of essential steps of glycan biosynthesis from the Man9GlcNAc2 precursor to a mature multiantennary complex glycan. Mannosidase I progressively removes mannose moieties from the precursor, in a process that can be inhibited by the drug kifunensine. GnTI then transfers a GlcNAc moiety to the D1 arm of the resulting Man5GlcNAc2 intermediate, creating a hybrid glycan. Mannose trimming of the D2 and D3 arms then allows additional GlcNAc moieties to be added by a series of GnT family enzymes to form multiantennary complexes. This process can be inhibited by swainsonine. The antennae are ultimately capped and decorated by galactose and sialic acid. Hybrid and complex glycans are usually fucosylated at the basal GlcNAc, rendering them resistant to endo H digestion. However, NgF is able to remove all types of glycan.Carbohydrates provide a layer of protection against NAb attack (Fig. (Fig.1A).1A). As glycans are considered self, antibody responses against them are thought to be regulated by tolerance mechanisms. Thus, a glycan network forms a nonimmunogenic “cloak,” protecting the underlying protein from antibodies (3, 13, 20, 29, 39, 54, 65, 67, 74, 85, 96, 98, 117, 119, 120). The extent of this protection can be illustrated by considering the ways in which glycans differ from typical amino acid side chains. First, N-linked glycans are much larger, with an average mass more than 20 times that of a typical amino acid R-group. They are also usually more flexible and may therefore affect a greater volume of surrounding space. In the more densely populated parts of gp120, the carbohydrate field may even be stabilized by sugar-sugar hydrogen bonds, providing even greater coverage (18, 75, 125).The process of N-linked glycosylation can result in diverse structures that may be divided into three categories: oligomannose, hybrid, and complex (56). Each category shares a common Man3GlcNAc2 pentasaccharide stem (where Man is mannose and GlcNAc is N-acetylglucosamine), to which up to six mannose residues are attached in oligomannose N-glycans, while complex N-glycans are usually larger and may bear various sizes and numbers of antennae (Fig. (Fig.1B).1B). Glycan synthesis begins in the endoplasmic reticulum, where N-linked oligomannose precursors (Glc3Man9GlcNAc2; Glc is glucose) are transferred cotranslationally to the free amide of the asparagine in a sequon Asn-X-Thr/Ser, where X is not Pro (40). Terminal glucose and mannose moieties are then trimmed to yield Man5GlcNAc2 (Fig. (Fig.1B).1B). Conversion to a hybrid glycan is then initiated by N-acetylglucosamine transferase I (GnTI), which transfers a GlcNAc moiety to the D1 arm of the Man5GlcNAc2 substrate (19) (Fig. (Fig.1B).1B). This hybrid glycoform is then a substrate for modification into complex glycans, in which the D2 and D3 arm mannose residues are replaced by complex antennae (19, 40, 56). Further enzymatic action catalyzes the addition of α-1-6-linked fucose moiety to the lower GlcNAc of complex glycan stems, but usually not to oligomannose glycan stems (Fig. (Fig.1B)1B) (21, 113).Most glycoproteins exhibit only fully mature complex glycans. However, the steric limitations imposed by the high density of glycans on some parts of gp120 lead to incomplete trimming, leaving “immature” oligomannose glycans (22, 26, 128). Spatial competition between neighboring sequons can sometimes lead to one or the other remaining unutilized, further distancing the final Env product from what might be expected based on its primary sequence (42, 48, 74, 119). An attempt to assign JR-FL gp120 and gp41 sequon use and types, based on various studies, is shown in Fig. Fig.1A1A (6, 26, 34, 35, 42, 63, 71, 74, 119, 128). At some positions, the glycan type is conserved. For example, the glycan at residue N301 has consistently been found to be complex (26, 63, 128). At other positions, considerable heterogeneity exists in the glycan populations, in some cases to the point where it is difficult to unequivocally assign them as predominantly complex or oligomannose. The reasons for these uncertainties might include incomplete trimming (42), interstrain sequence variability, the form of Env (e.g., gp120 or gp140), and the producer cell. The glycans of native Env trimers and monomeric gp120 may differ due to the constraints imposed by oligomerization (32, 41, 77). Thus, although all the potential sequons of HXB2 gp120 were found to be occupied in one study (63), some are unutilized or variably utilized on functional trimers, presumably due to steric limitations (42, 48, 75, 96, 119).The distribution of complex and oligomannose glycans on gp120 largely conforms with an antigenic map derived from structural models (59, 60, 102, 120), in which the outer domain is divided into a neutralizing face and an immunologically silent face. Oligomannose glycans cluster tightly on the silent face of gp120 (18, 128), while complex glycans flank the gp120 receptor binding sites of the neutralizing face, ostensibly forming a protective “fence” against NAbs (105). The relatively sparse clustering of complex glycans that form this fence may reflect a trade-off between protecting the underlying functional domains from NAbs by virtue of large antennae while at the same time permitting sufficient flexibility for the refolding events associated with receptor binding and fusion (29, 39, 67, 75, 98, 117). Conversely, the dense clustering of oligomannose glycans on the silent domain may be important for ensuring immune protection and/or in creating binding sites for lectins such as DC-SIGN (9, 44).The few available broadly neutralizing monoclonal antibodies (MAbs) define sites of vulnerability on Env trimers (reviewed in reference 52). They appear to fall into two general categories: those that access conserved sites by overcoming Env''s various evasion strategies and, intriguingly, those that exploit these very defensive mechanisms. Regarding the first category, MAb b12 recognizes an epitope that overlaps the CD4 binding site of gp120 (14), and MAbs 2F5 and 4E10 (84, 129) recognize adjacent epitopes of the membrane-proximal external region (MPER) at the C-terminal ectodomain of gp41. The variable neutralizing potencies of these MAbs against primary isolates that contain their core epitopes illustrate how conformational masking can dramatically regulate their exposure (11, 118). Conformational masking also limits the activities of MAbs directed to the V3 loop and MAbs whose epitopes overlap the coreceptor binding site (11, 62, 121).A second category of MAbs includes MAb 2G12, which recognizes a tight cluster of glycans in the silent domain of gp120 (16, 101, 103, 112). This epitope has recently sparked considerable interest in exploiting glycan clusters as possible carbohydrate-based vaccines (2, 15, 31, 70, 102, 116). Two recently described MAbs, PG9 and PG16 (L. M. Walker and D. R. Burton, unpublished data), also target epitopes regulated by the presence of glycans that involve conserved elements of the second and third variable loops and depend largely on the quaternary trimer structure and its in situ presentation on membranes. Their impressive breadth and potency may come from the fact that they target the very mechanisms (variable loops and glycans) that are generally thought to protect the virus from neutralization. Like 2G12, these epitopes are likely to be constitutively exposed and thus may not be subject to conformational masking (11, 118).The above findings reveal the importance of N-glycans both as a means of protection against neutralization as well as in directly contributing to unique neutralizing epitopes. Clearly, further studies on the nature and function of glycans in native Env trimers are warranted. Possible approaches may be divided into four categories, namely, (i) targeted mutation, (ii) enzymatic removal, (iii) expression in the presence of glycosylation inhibitors, and (iv) expression in mutant cell lines with engineered blocks in the glycosylation pathway. Much of the available information on the functional roles of glycans in HIV-1 and simian immunodeficiency virus (SIV) infection has come from the study of mutants that eliminate glycans either singly or in combination (20, 54, 66, 71, 74, 91, 95, 96). Most mutants of this type remain at least partially functional (74, 95, 96). In some cases these mutants have little effect on neutralization sensitivity, while in others they can lead to increased sensitivity to MAbs specific for the V3 loop and CD4 binding site (CD4bs) (54, 71, 72, 74, 106). In exceptional cases, increased sensitivity to MAbs targeting the coreceptor binding site and/or the gp41 MPER has been observed (54, 66, 72, 74).Of the remaining approaches for studying the roles of glycans, enzymatic removal is constrained by the extreme resistance of native Env trimers to many common glycosidases, contrasting with the relative sensitivity of soluble gp120 (67, 76, 101). Alternatively, drugs can be used to inhibit various stages of mammalian glycan biosynthesis. Notable examples are imino sugars, such as N-butyldeoxynojirimycin (NB-DNJ), that inhibit the early trimming of the glucose moieties from Glc3Man9GlcNAc2 precursors in the endoplasmic reticulum (28, 38, 51). Viruses produced in the presence of these drugs may fail to undergo proper gp160 processing or fusion (37, 51). Other classes of inhibitor include kifunensine and swainsonine, which, respectively, inhibit the trimming of the Man9GlcNAc2 precursor into Man5GlcNAc2 or inhibit the removal of remaining D2 and D3 arm mannoses from the hybrid glycans, thus preventing the construction of complex glycan antennae (Fig. (Fig.1B)1B) (17, 33, 76, 104, 119). Unlike NB-DNJ, viruses produced in the presence of these drugs remain infectious (36, 76, 79, 100).Yet another approach is to express virus in insect cells that can only modify proteins with paucimannose N-glycans (58). However, the inefficient gp120/gp41 processing by furin-like proteases in these cells prevents their utility in functional studies (123). Another option is provided by ricin-selected GnTI-deficient cell lines that cannot transfer GlcNAc onto the mannosidase-trimmed Man5GlcNAc2 substrate, preventing the formation of hybrid and complex carbohydrates (Fig. (Fig.1B)1B) (17, 32, 36, 94). This arrests glycan processing at a well-defined point, leading to the substitution of complex glycans with Man5GlcNAc2 rather than with the larger Man9GlcNAc2 precursors typically obtained with kifunensine treatment (17, 32, 33, 104). With this in mind, here we produced HIV-1 pseudoviruses in GnTI-deficient cells to investigate the role of complex glycan antennae in viral resistance neutralization. By replacing complex glycans with smaller Man5GlcNAc2 we can determine the effect of “lowering the glycan fence” that surrounds the receptor binding sites, compared to the above-mentioned studies of individual glycan deletion mutants, whose effects are analogous to removing a fence post. Furthermore, since oligomannose glycans are sensitive to certain enzymes, such as endoglycosidase H (endo H), we investigated the effect of dismantling the glycan fence on Env function and stability. Our results suggest that the antennae of complex glycans protect against certain specificities but that glycan stems regulate trimer conformation with often more dramatic consequences for neutralization sensitivity and in extreme cases, infectious function.  相似文献   

10.
Alphavirus particles are covered by 80 glycoprotein spikes that are essential for viral entry. Spikes consist of the E2 receptor binding protein and the E1 fusion protein. Spike assembly occurs in the endoplasmic reticulum, where E1 associates with pE2, a precursor containing E3 and E2 proteins. E3 is a small, cysteine-rich, extracellular glycoprotein that mediates proper folding of pE2 and its subsequent association with E1. In addition, cleavage of E3 from the assembled spike is required to make the virus particles efficiently fusion competent. We have found that the E3 protein in Sindbis virus contains one disulfide bond between residues Cys19 and Cys25. Replacing either of these two critical cysteines resulted in mutants with attenuated titers. Replacing both cysteines with either alanine or serine resulted in double mutants that were lethal. Insertion of additional cysteines based on E3 proteins from other alphaviruses resulted in either sequential or nested disulfide bond patterns. E3 sequences that formed sequential disulfides yielded virus with near-wild-type titers, while those that contained nested disulfide bonds had attenuated activity. Our data indicate that the role of the cysteine residues in E3 is not primarily structural. We hypothesize that E3 has an enzymatic or functional role in virus assembly, and these possibilities are further discussed.Alphaviruses are members of the Togaviradae family and are single-stranded, positive-sense RNA, enveloped viruses (17). The lipid membranes of the viruses have 80 glycoprotein spikes which are required for viral entry. Each spike is comprised of three copies of a heterodimer which consists of the E2 and E1 proteins (22, 54). E2 and E1 are glycoproteins with a single transmembrane helix that traverses the host-derived lipid bilayer. E2 interacts with the nucleocapsid core at the C terminus (12, 16, 27, 43) and contains the receptor binding site at the N terminus (5, 21, 45). E1 is the viral fusion protein responsible for mediating fusion between the virus membrane and the host cell membrane during an infection (13, 39, 47). Specific interactions in both the ectodomain and transmembrane regions are critical for heterodimer formation (30, 35, 46, 54). The assembly of each heterodimer, its subsequent assembly into a spike, and the interaction of the cytoplasmic tail of the spike with the nucleocapsid core are all essential for the efficient production of infectious particles.Glycoprotein spike assembly requires four structural proteins, E3, E2, 6K, and E1, which are expressed as a single polyprotein. E3 is a small, 64-amino-acid protein (Sindbis virus [SINV] numbering) and contains a signal sequence that translocates the protein into the endoplasmic reticulum (ER) (3, 4, 15). Early in translation, glycosylation of N14 (SINV numbering) occurs and this promotes E3''s release from the ER membrane into the lumen. As a result, the signal sequence is not cleaved from the E3 protein (14). Cellular enzymes cleave the polyprotein to yield pE2 (an uncleaved protein consisting of E3 and E2), 6K, and E1 (23, 55) proteins. In the ER, E1 is found in several conformations, only one of which will form a functional heterodimer with pE2, allowing its transport to the Golgi apparatus (1, 2, 6, 7, 36). After pE2-E1 heterodimerization, self-association between three heterodimers occurs and each individual spike is formed (25, 26, 36). As observed with Semliki Forest virus, disulfide bonds reshuffle within pE2 during protein folding (34), possibly forming intermolecular disulfide bonds between E3 and E2 residues. However, no intermolecular disulfide bonds between pE2 and E1 have been identified (34). Once the viral spikes have been assembled, they are transported to the plasma membrane (11) and are thus exposed to subcellular changes of pH, from pH 7.2 in the ER to pH 5.7 in the vesicles constitutively transporting the spikes to the plasma membrane. In the trans-Golgi network, the E3 protein is cleaved from pE2 by the cellular protein furin (18, 44, 55). E3 remains noncovalently attached to the released virus particle, while in other species E3 is found in the medium of virus-infected cells (32, 49).E3 is required for efficient particle assembly, both in mediating spike folding and in spike activation for viral entry. When an ER signal sequence was substituted for the E3 protein, heterodimerization of pE2 and E1 was abolished (26). Furthermore, when E2 and E1 were expressed individually, low levels of E2 were transported to the cell surface while E1 remained in the ER, suggesting that heterodimerization with pE2 is necessary for E1 to be transported to the cell surface (24, 26, 46). These results are consistent with E3 playing a critical role in mediating the folding of pE2 and the association of pE2 and E1 proteins during spike assembly (7, 38). In viruses where the furin cleavage site was mutated, the virus particles were correctly assembled but severely reduced in infectivity, presumably because the fusion protein was unable to dissociate from pE2 and initiate fusion (44, 55).A comparison of an amino acid sequence alignment of E3 proteins from different alphaviruses (Fig. (Fig.1)1) shows that the E3 protein is a small protein with four conserved cysteine (Cys) residues. A subset of E3 proteins contains an additional two Cys residues in a narrow cysteine/proline-rich region, PPCXPCC (Fig. (Fig.1).1). We have purified recombinant E3 protein from SINV and have determined that a disulfide bond is present and, furthermore, that these Cys residues are important in virus assembly. Within the alphavirus E3 proteins, we have identified a region that is important for mediating spike transport to the plasma membrane and thus is critical for spike assembly.Open in a separate windowFIG. 1.E3 amino acid sequence alignment from a representative group of alphaviruses. The cysteines marked with asterisks are conserved in all alphavirus species. The ⋄ indicates the conserved but nonessential glycosylation site. The PPCXPCC motif present in ∼50% of alphaviruses is underlined. SFV, Semliki Forest virus; RRV, Ross River virus; BFV, Barmah Forest virus; EEE, eastern equine encephalitis virus; ONN, O''nyong nyong virus; IGB, Igbo Ora virus; OCK, Ockelbo virus; WEE, western equine encephalitis virus; AUR, Aura virus; VEE, Venezuelan equine encephalitis virus.  相似文献   

11.
l-2-Amino-4-methoxy-trans-3-butenoic acid (AMB) is a potent antibiotic and toxin produced by Pseudomonas aeruginosa. Using a novel biochemical assay combined with site-directed mutagenesis in strain PAO1, we have identified a five-gene cluster specifying AMB biosynthesis, probably involving a thiotemplate mechanism. Overexpression of this cluster in strain PA7, a natural AMB-negative isolate, led to AMB overproduction.The Gram-negative bacterium Pseudomonas aeruginosa is an opportunistic pathogen that causes a wide range of human infections and is considered the main pathogen responsible for chronic pneumonia in cystic fibrosis patients (7, 23). P. aeruginosa also infects other organisms, such as insects (4), nematodes (6), plants (18), and amoebae (20). Its ability to thrive as a pathogen and to compete in aquatic and soil environments can be partly attributed to the production and interplay of secreted virulence factors and secondary metabolites. While the importance of many of these exoproducts has been studied, the antimetabolite l-2-amino-4-methoxy-trans-3-butenoic acid (AMB; methoxyvinylglycine) (Fig. (Fig.1)1) has received only limited attention. Identified during a search for new antibiotics, AMB was found to reversibly inhibit the growth of Bacillus spp. (26) and Escherichia coli (25) and was later shown to inhibit the growth and metabolism of cultured Walker carcinosarcoma cells (28). AMB is a γ-substituted vinylglycine, a naturally occurring amino acid with a β,γ-C=C double bond. Other members of this family are aminoethoxyvinylglycine from Streptomyces spp. (19) and rhizobitoxine, made by Bradyrhizobium japonicum (16) and Pseudomonas andropogonis (15) (Fig. (Fig.1).1). As inhibitors of pyridoxal phosphate-dependent enzymes (13, 17, 21, 22), γ-substituted vinylglycines have multiple targets in bacteria, animals, and plants (3, 5, 10, 21, 22, 29). However, the importance of AMB as a toxin in biological interactions with P. aeruginosa has not been addressed, as AMB biosynthesis and the genes involved have not been elucidated.Open in a separate windowFIG. 1.Chemical structures of the γ-substituted vinylglycines AMB, aminoethoxyvinylglycine, and rhizobitoxine.  相似文献   

12.
13.
The biofilm matrix contributes to the chemistry, structure, and function of biofilms. Biofilm-derived membrane vesicles (MVs) and DNA, both matrix components, demonstrated concentration-, pH-, and cation-dependent interactions. Furthermore, MV-DNA association influenced MV surface properties. This bears consequences for the reactivity and availability for interaction of matrix polymers and other constituents.The biofilm matrix contributes to the chemistry, structure, and function of biofilms and is crucial for the development of fundamental biofilm properties (46, 47). Early studies defined polysaccharides as the matrix component, but proteins, lipids, and nucleic acids are all now acknowledged as important contributors (7, 15). Indeed, DNA has emerged as a vital participant, fulfilling structural and functional roles (1, 5, 6, 19, 31, 34, 36, 41, 43, 44). The phosphodiester bond of DNA renders this polyanionic at a physiological pH, undoubtedly contributing to interactions with cations, humic substances, fine-dispersed minerals, and matrix entities (25, 41, 49).In addition to particulates such as flagella and pili, membrane vesicles (MVs) are also found within the matrices of gram-negative and mixed biofilms (3, 16, 40). MVs are multifunctional bilayered structures that bleb from the outer membranes of gram-negative bacteria (reviewed in references 4, 24, 27, 28, and 30) and are chemically heterogeneous, combining the known chemistries of the biofilm matrix. Examination of biofilm samples by transmission electron microscopy (TEM) has suggested that matrix material interacts with MVs (Fig. (Fig.1).1). Since MVs produced in planktonic culture have associated DNA (11, 12, 13, 20, 21, 30, 39, 48), could biofilm-derived MVs incorporate DNA (1, 39, 40, 44)?Open in a separate windowFIG. 1.Possible interactions between matrix polymers and particulate structures. Shown is an electron micrograph of a thin section through a P. aeruginosa PAO1 biofilm. During processing, some dehydration occurred, resulting in collapse of matrix material into fibrillate arrangements (black filled arrows). There is a suggestion of interactions occurring with particulate structures such as MVs (hollow white arrow) and flagella (filled white arrows) (identified by the appearance and cross-dimension of these highly ordered structures when viewed at high magnification), which was consistently observed with other embedded samples and also with whole-mount preparations of gently disrupted biofilms (data not shown). The scale bar represents 200 nm.  相似文献   

14.
Flaviviruses are a group of single-stranded, positive-sense RNA viruses causing ∼100 million infections per year. We have recently shown that flaviviruses produce a unique, small, noncoding RNA (∼0.5 kb) derived from the 3′ untranslated region (UTR) of the genomic RNA (gRNA), which is required for flavivirus-induced cytopathicity and pathogenicity (G. P. Pijlman et al., Cell Host Microbe, 4: 579-591, 2008). This RNA (subgenomic flavivirus RNA [sfRNA]) is a product of incomplete degradation of gRNA presumably by the cellular 5′-3′ exoribonuclease XRN1, which stalls on the rigid secondary structure stem-loop II (SL-II) located at the beginning of the 3′ UTR. Mutations or deletions of various secondary structures in the 3′ UTR resulted in the loss of full-length sfRNA (sfRNA1) and production of smaller and less abundant sfRNAs (sfRNA2 and sfRNA3). Here, we investigated in detail the importance of West Nile virus Kunjin (WNVKUN) 3′ UTR secondary structures as well as tertiary interactions for sfRNA formation. We show that secondary structures SL-IV and dumbbell 1 (DB1) downstream of SL-II are able to prevent further degradation of gRNA when the SL-II structure is deleted, leading to production of sfRNA2 and sfRNA3, respectively. We also show that a number of pseudoknot (PK) interactions, in particular PK1 stabilizing SL-II and PK3 stabilizing DB1, are required for protection of gRNA from nuclease degradation and production of sfRNA. Our results show that PK interactions play a vital role in the production of nuclease-resistant sfRNA, which is essential for viral cytopathicity in cells and pathogenicity in mice.Arthropod-borne flaviviruses such as West Nile virus (WNV), dengue virus (DENV), and Japanese encephalitis virus (JEV) cause major outbreaks of potentially fatal disease and affect over 50 million people every year. The highly pathogenic North American strain of WNV (WNVNY99) has already claimed more than 1,000 lives with over 27,000 cases reported since its emergence in New York in 1999 and has raised global public health concerns (9). In contrast, the closely related Australian strain of WNV, WNVKUN, is highly attenuated and does not cause overt disease in humans and animals (11). WNVKUN has been used extensively as a model virus to study flavivirus replication and flavivirus-host interactions (13, 14, 16-19, 26, 38, 39).The ∼11-kb positive-stranded, capped WNV genomic RNA (gRNA) lacks a poly(A) tail and consists of 5′ and 3′ untranslated regions (UTRs) flanking one open reading frame, which encodes the viral proteins required for the viral life cycle (6, 15, 38, 39). Flavivirus UTRs are involved in translation and initiation of RNA replication and likely determine genome packaging (13, 14, 16, 21, 30, 39-41). Both the 5′ UTR (∼100 nucleotides [nt] in size) and the 3′ UTR (from ∼400 to 700 nucleotides) can form secondary and tertiary structures which are highly conserved among mosquito-borne flaviviruses (1, 8, 10, 14, 29, 32, 34). More specifically, the WNVKUN 3′ UTR consists of several conserved regions and secondary structures (Fig. (Fig.1A)1A) which were previously predicted or shown to exist in various flaviviruses by computational and chemical analyses, respectively (4, 10, 25, 26, 29-32). The 5′ end of the 3′ UTR starts with an AU-rich region which can form stem-loop structure I (SL-I) followed by SL-II, which we previously showed to be vitally important for subgenomic flavivirus RNA (sfRNA) production (26; see also below). SL-II is followed by a short, repeated conserved hairpin (RCS3) and SL-III (26). Further downstream of SL-III are the SL-IV and CS3 structures, which are remarkably similar to the preceding SL-II-RCS3 structure (26, 29). Further downstream of the SL-IV-CS3 structure are dumbbells 1 and 2 (DB1 and DB2, respectively) followed by a short SL and the 3′ SL (25, 26).Open in a separate windowFIG. 1.(A) Model of the WNVKUN 3′ UTR RNA structure. Highlighted in bold are the secondary structures investigated here. Dashed lines indicate putative PKs. The two sites of the putative PK interactions are shown in open boxes. sfRNA1, -2, -3, and -4 start sites are indicated by arrows. (R)CS, (repeated) conserved sequence; DB, dumbbell structure; PK, pseudoknot; SL, stem-loop. (B) Structural model of PK1 in SL-II with disruptive mutations. Nucleotide numbering is from the end of the 3′ UTR. The sfRNA1 start is indicated by an arrow. Nucleotides forming PK1 are on a gray background, and mutated nucleotides are white on a black background. (C) Sequences mutated in the different constructs. Nucleotides in the wt PK sequences used for mutations are bold and underlined. Introduced mutations are shown under the corresponding nucleotides in the wt sequence.The described structures have been investigated in some detail for their requirement in RNA replication and translation. Generally, a progressive negative effect on viral growth was shown with progressive deletions into the 3′-proximal region of the JEV 3′ UTR (41). However, only a relatively short region of the JEV 3′ UTR, consisting of the 3′-terminal 193 nt, was shown to be absolutely essential for gRNA replication (41). The minimal region for DENV replication was reported to be even shorter (23). Extensive analysis has shown that the most 3′-terminal, essential regions of the 3′ UTR include the cyclization sequence and 3′ SL, which are required for efficient RNA replication (2, 14, 16, 23, 35). As we showed, deletion of SL-II or SL-I did not overtly affect WNVKUN replication (26). However, deletion of CS2, RCS2, CS3, or RCS3 in WNV replicon RNA significantly reduced RNA replication but not translation (20), indicating that these elements facilitate but are not essential for RNA replication. In addition, it was shown that deletion of DB1 or DB2 resulted in a viable mutant virus that was reduced in growth efficiency, while deletion of both DB structures resulted in a nonviable mutant (23).In addition to the above-mentioned secondary stem-loop structures, computational and chemical analysis of the flavivirus 3′ UTR suggested the presence of 5 pseudoknot (PK) interactions (Fig. (Fig.1A)1A) (25, 26, 32). A PK is a structure formed upon base pairing of a single-stranded region of RNA in the loop of a hairpin to a stretch of complementary nucleotides elsewhere in the RNA chain (Fig. (Fig.1B).1B). These structures are referred to as hairpin type (H-type) PKs (3), and they usually stabilize secondary RNA structures. Typically, the final tertiary structure does not significantly alter the preformed secondary structure (5). In general, PK interactions have been shown to be important in biological processes such as initiation and/or elongation of translation, initiation of gRNA replication, and ribosomal frameshifting for a number of different viruses, including flaviviruses (reviewed in references 3 and 22). The first PK in the WNV 3′ UTR was predicted to form in SL-II, followed by a similar PK in SL-IV (26) (PK1 and PK2 in Fig. Fig.1A).1A). For the DENV, yellow fever virus (YFV), and JEV subgroup of flaviviruses, two PKs further downstream were predicted to form between DB1 and DB2 and corresponding single-stranded RNA regions located further downstream (25) (PK3 and PK4 in Fig. Fig.1A).1A). The formation of these structures is supported by covariations in the WNV RNAs. In addition, a PK was proposed to form between a short SL and the 3′ SL at the 3′ terminus of the viral genome (32) (PK5 in Fig. Fig.1A1A).Importantly, in addition to its role in viral replication and translation, we have shown that the WNVKUN 3′ UTR is important for the production of a small noncoding RNA fragment designated sfRNA (26). This short RNA fragment of ∼0.5 kb is derived from the 3′ UTR of the gRNA and exclusively produced by the members of the Flavivirus genus of the Flaviviridae family, where it is required for efficient viral replication, cytopathicity, and pathogenicity (26). Our studies suggested that sfRNA is a product of incomplete degradation of the gRNA presumably by the cellular 5′-3′ exoribonuclease XRN1, resulting from XRN1 stalling on the rigid secondary/tertiary structures located at the beginning of the 3′ UTR (26). XRN1 is an exoribonuclease which usually degrades mRNA from the 5′ to the 3′ end as part of cellular mRNA decay and turnover (33), and it was shown previously that XRN1 can be stalled by SL structures (28). Mutations or deletions of WNV 3′ UTR secondary structures resulted in the loss of full-length sfRNA (sfRNA1) and production of smaller and less abundant sfRNAs (sfRNA2 and sfRNA3) (26). In particular, SL-II (Fig. (Fig.1A)1A) was shown to be important for sfRNA1 production; deletion of this structure either alone or in conjunction with other structures located downstream of SL-II abolished sfRNA1 production, leading to the production of the smaller RNA fragments sfRNA2 and sfRNA3.Here, we extended our investigation and studied the importance of several predicted 3′ UTR secondary structures and PK interactions for the production of sfRNA. To further understand the generation mechanism of sfRNA and its requirements, we deleted or mutated a number of RNA structures in the WNVKUN 3′ UTR and investigated the size and amount of sfRNA generated from these mutant RNAs. The results show that not only SLs but also PK interactions play a vital role in stabilizing the 3′ UTR RNA and preventing complete degradation of viral gRNA to produce nuclease-resistant sfRNA, which is required for efficient virus replication and cytopathicity in cells and virulence in mice.  相似文献   

15.
16.
17.
18.
19.
20.
Cyanophycin (multi-l-arginyl-poly-l-aspartic acid; also known as cyanophycin grana peptide [CGP]) is a putative precursor for numerous biodegradable technically used chemicals. Therefore, the biosynthesis and production of the polymer in recombinant organisms is of special interest. The synthesis of cyanophycin derivatives consisting of a wider range of constituents would broaden the applications of this polymer. We applied recombinant Saccharomyces cerevisiae strains defective in arginine metabolism and expressing the cyanophycin synthetase of Synechocystis sp. strain PCC 6308 in order to synthesize CGP with citrulline and ornithine as constituents. Strains defective in arginine degradation (Car1 and Car2) accumulated up to 4% (wt/wt) CGP, whereas strains defective in arginine synthesis (Arg1, Arg3, and Arg4) accumulated up to 15.3% (wt/wt) of CGP, which is more than twofold higher than the previously content reported in yeast and the highest content ever reported in eukaryotes. Characterization of the isolated polymers by different analytical methods indicated that CGP synthesized by strain Arg1 (with argininosuccinate synthetase deleted) consisted of up to 20 mol% of citrulline, whereas CGP from strain Arg3 (with ornithine carbamoyltransferase deleted) consisted of up to 8 mol% of ornithine, and CGP isolated from strain Arg4 (with argininosuccinate lyase deleted) consisted of up to 16 mol% lysine. Cultivation experiments indicated that the incorporation of citrulline or ornithine is enhanced by the addition of low amounts of arginine (2 mM) and also by the addition of ornithine or citrulline (10 to 40 mM), respectively, to the medium.Cyanophycin (multi-l-arginyl-poly-[l-aspartic acid]), also referred to as cyanophycin grana peptide (CGP), represents a polydisperse nonribosomally synthesized polypeptide consisting of poly(aspartic acid) as backbone and arginine residues bound to each aspartate (49) (Fig. (Fig.1).1). One enzyme only, referred to as cyanophycin synthetase (CphA), catalyzes the synthesis of the polymer from amino acids (55). Several CphAs originating from different bacteria exhibit specific features (2, 7, 5, 32, 50, 51). CphAs from the cyanobacteria Synechocystis sp. strain PCC 6308 and Anabaena variabilis ATCC 29413, respectively, exhibit a wide substrate range in vitro (2, 7), whereas CphA from Acinetobacter baylyi or Nostoc ellipsosporum incorporates only aspartate and arginine (23, 24, 32). CphA from Thermosynechococcus elongatus catalyzes the synthesis of CGP primer independently (5); CphA from Synechococcus sp. strain MA19 exhibits high thermostability (22). Furthermore, two types of CGP were observed concerning its solubility behavior: (i) a water-insoluble type that becomes soluble at high or low pH (34, 48) and (ii) a water-soluble type that was only recently observed in recombinant organisms (19, 26, 42, 50, 56). In the past, bacteria were mainly applied for the synthesis of CGP (3, 14, 18, 53), whereas recently there has been greater interest in synthesis in eukaryotes (26, 42, 50). CGP was accumulated to almost 7% (wt/wt) of dry matter in recombinant Nicotiana tabacum and Saccharomyces cerevisiae (26, 50).Open in a separate windowFIG. 1.Chemical structures of dipeptide building blocks of CGP variants detected in vivo. Structure: 1, aspartate-arginine; 2, aspartate-lysine; 3, aspartate-citrulline; 4, aspartate-ornithine. Aspartic acid is presented in black; the second amino acid of the dipeptide building blocks is shown in gray. The nomenclature of the carbon atoms is given.In S. cerevisiae the arginine metabolism is well understood and has been investigated (30) (see Fig. Fig.2).2). Arginine is synthesized from glutamate via ornithine and citrulline in eight successive steps. The enzymes acetylglutamate synthase, acetylglutamate kinase, N-acetyl-γ-glutamylphosphate reductase, and acetylornithine aminotransferase are involved in the formation of N-α-acetylornithine. The latter is converted to ornithine by acetylornithine acetyltransferase. In the next step, ornithine carbamoyltransferase (ARG3) condenses ornithine with carbamoylphosphate, yielding citrulline. Citrulline is then converted to l-argininosuccinate by argininosuccinate synthetase. The latter is in the final step cleaved into fumarate and arginine by argininosuccinate lyase (ARG4). The first five steps occur in the mitochondria, whereas the last three reactions occur in the cytosol (28, 54). Arginine degradation is initiated by arginase (CAR1) and ornithine aminotransferase (CAR2) (10, 11, 38, 39).Open in a separate windowFIG. 2.Schematic overview of the arginine metabolism in S. cerevisiae. Reactions shown in the shaded area occur in the mitochondria, while the other reactions are catalyzed in the cytosol. Abbreviations: ARG2, acetylglutamate synthase; ARG6, acetylglutamate kinase; ARG5, N-acetyl-γ-glutamyl-phosphate reductase; ARG8, acetylornithine aminotransferase; ECM40, acetylornithine acetyltransferase; ARG1, argininosuccinate synthetase; ARG3, ornithine carbamoyltransferase; ARG4, argininosuccinate lyase; CAR1, arginase; CAR2, ornithine aminotransferase.A multitude of putative applications for CGP derivatives are available (29, 41, 45, 47), thus indicating a need for efficient biotechnological production and for further investigations concerning the synthesis of CGP with alternative properties and different constituents. It is not only the putative application of the polymer as a precursor for poly(aspartic acid), which is used as biodegradable alternative for poly(acrylic acid) or for bulk chemicals, that makes CGP interesting (29, 45-47). In addition, a recently developed process for the production of dipeptides from CGP as a precursor makes the synthesis of CGP variants worthwhile (43). Dipeptides play an important role in medicine and pharmacy, e.g., as additives for malnourished patients, as treatments against liver diseases, or as aids for muscle proliferation (43). Because dipeptides are synthesized chemically (40) or enzymatically (6), novel biotechnological production processes are welcome.  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