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 140 毫秒
1.
Herpes simplex virus type 1 glycoprotein B (gB) is an envelope component that plays an essential role in virus infection. The biologically active form of gB is an oligomer that contributes to the process of viral envelope fusion with the cell surface membrane, resulting in viral penetration and initiation of the replication cycle. In previous studies, two discontinuous sites for oligomer formation were identified: a nonessential upstream site located between residues 93 and 282 and an essential downstream site located between residues 596 and 711. In this study, in vitro-transcribed and -translated gB test molecules were used to characterize the more active essential membrane-proximal domain. A series of gB test polypeptides mutated in this downstream oligomerization domain were assayed for their abilities to form oligomers with a mutant gB capture polypeptide containing the analogous wild-type domain. Detection of oligomers was achieved by coimmunoprecipitation of two gB mutant molecules by using a monoclonal antibody specific for a hemagglutinin epitope tag introduced into the coding sequence of the capture polypeptide. Analysis of the immune-precipitated products by sodium dodecyl sulfate-polyacrylamide gel electrophoresis demonstrated that the downstream oligomerization domain resided within residues 626 to 676. This region was further resolved into two segments, residues 626 to 653 and 653 to 675, each of which was independently sufficient to form oligomers. However, residues 626 to 653 provided for a stronger interaction between gB monomers. Moreover, this stretch of 28 amino acids was shown to form oligomers when introduced into the carboxy-terminal region of gB monomers lacking this domain at the normal site, thus indicating that this domain was functionally independent of its natural location within the gB molecule. Further analysis of the sequence within residues 596 to 653 by using mutant test polypeptides altered in individual amino acids revealed that cysteines 9 and 10 located at positions 596 and 633, respectively, were not required for oligomer formation but contributed to dimer formation and/or stabilization. The results of this study suggest that oligomerization of gB monomers is induced by interactions between contiguous residues localized within the ectodomain near the site of molecule insertion into the viral envelope membrane.  相似文献   

2.
Herpes simplex virus type 1 (HSV-1) mutants defective for envelope glycoprotein C (gC) and gB are highly impaired in the ability to attach to cell surface heparan sulfate (HS) moieties of proteoglycans, the initial virus receptor. Here we report studies aimed at defining the HS binding element of HSV-1 (strain KOS) gB and determining whether this structure is functionally independent of gB’s role in extracellular virus penetration or intercellular virus spread. A mutant form of gB deleted for a putative HS binding lysine-rich (pK) sequence (residues 68 to 76) was transiently expressed in Vero cells and shown to be processed normally, leading to exposure on the cell surface. Solubilized gBpK also had substantially lower affinity for heparin-acrylic beads than did wild-type gB, confirming that the HS binding domain had been inactivated. The gBpK gene was used to rescue a KOS gB null mutant virus to produce the replication-competent mutant KgBpK. Compared with wild-type virus, KgBpK showed reduced binding to mouse L cells (ca. 20%), while a gC null mutant virus in which the gC coding sequence was replaced by the lacZ gene (KCZ) was substantially more impaired (ca. 65%-reduced binding), indicating that the contribution of gC to HS binding was greater than that of gB. The effect of combining both mutations into a single virus (KgBpKgC) was additive (ca. 80%-reduced binding to HS) and displayed a binding activity similar to that observed for KOS virus attachment to sog9 cells, a glycosaminoglycan-deficient L-cell line. Cell-adsorbed individual and double HS mutant viruses exhibited a lower rate of virus entry following attachment, suggesting that HS binding plays a role in the process of virus penetration. Moreover, the KgBpK mutant virus produced small plaques on Vero cells in the presence of neutralizing antibody where plaque formation depended on cell-to-cell virus spread. These studies permitted the following conclusions: (i) the pK sequence is not essential for gB processing or function in virus infection, (ii) the lysine-rich sequence of gB is responsible for HS binding, and (iii) binding to HS is cooperatively linked to the process of efficient virus entry and lateral spread but is not absolutely required for virus infectivity.  相似文献   

3.
The tsB5 mutant of herpes simplex virus type 1 (HSV-1) strain HFEM was shown previously to be temperature sensitive for accumulation of the mature form of glycoprotein gB, for production or activity of a factor required in virus-induced cell fusion, and for production of virions with normal levels of infectivity. In addition, a previous study showed that virions produced by tsB5 at permissive temperature were more thermolabile than HFEM virions and contained altered gB that did not assume the dimeric conformation characteristic of HFEM. Results presented here demonstrate that, at permissive temperature, tsB5 differs from HFEM in another respect: plaques formed by tsB5 are syncytial on Vero cells (but not on HEp-2 cells), whereas plaques formed by HFEM are nonsyncytial on both cell types. In addition, our results indicate that tsB5 produces an oligomeric form of gB, but that it differs in electrophoretic mobility and stability from the gB dimers of HFEM. The major purpose of this study was to investigate the dependence of the various tsB5 mutant phenotypes on the temperature sensitivity of gB accumulation and on the alterations in oligomeric conformation of gB produced at permissive temperature. For this work the following HSV-1 strains related to tsB5 or HFEM were analyzed: (i) phenotypic revertants selected from tsB5 stocks for nonsyncytial plaque morphology on Vero cells or for ability to form plaques at restrictive temperature (38.5°C); (ii) a plaque morphology variant of HFEM selected for its syncytial phenotype on Vero cells; (iii) temperature-sensitive recombinants previously isolated from a cross between tsB5 and the non-temperature-sensitive syncytial strain HSV-1(MP); and (iv) a phenotypic revertant selected from one of the recombinant stocks for its ability to form plaques at 39°C. These strains were all compared with tsB5 and HFEM at three different temperatures in two different cell lines with respect to plaque formation, yield of infectious progeny, virus-induced cell fusion, and accumulation of gB. The results of our analyses on all the strains tested revealed the following correlations between mutant phenotypes and the accumulation and oligomeric conformation of gB. (i) There was a direct and quantitative relationship between the accumulation in infected cells of infectious progeny and of the mature form of gB, providing strong support for the hypothesis that this form of gB is necessary to the production of infectious virions. The oligomeric conformation of gB characteristic of HFEM is apparently not required for virion infectivity; nor was virion thermostability necessarily related to the presence of the HFEM-like oligomeric form of gB. (ii) The previously reported correlation between temperature sensitivity of gB accumulation and virus-induced cell fusion was confirmed for tsB5 and extended to other virus strains, and coordinate reversion of these traits was also demonstrated, providing support for the hypothesis that gB has a role in virus-induced cell fusion. At 37°C, intermediate between permissive and restrictive temperatures, some of the mutants and partial revertants induced cell fusion despite reduced accumulations of the mature form of gB, suggesting that the amount of mature gB present did not determine the extent of fusion and that other forms of gB as well as other factors should be investigated with regard to the process of cell fusion. (iii) Some of the mutants and partial revertants could form plaques at 38.5°C despite reduced ccumulations of gB and infectious progeny, indicating that the cell-to-cell transmission of viral infection may be at least in part independent of these factors.  相似文献   

4.
We studied the oligomeric properties of betaretroviral nonmyristoylated matrix protein (MA) and its R55F mutant from the Mason-Pfizer monkey virus in solution by means of chemical crosslinking and NMR spectroscopy. By analyzing crosslinked products and using concentration-dependent NMR chemical shift mapping, we have proven that the wild-type (WT) MA forms oligomers in solution. Conversely, no oligomerization was observed for the R55F mutant. Structural comparison of MAs explained their different behaviors in solution, concluding that the key residues involved in intermonomeric interaction are exposed in the WT MA but buried in the mutant, preventing the oligomerization of R55F. The final model of oligomerization of the WT MA was derived by concerted use of chemical shift mapping and diffusion-ordered spectroscopy measured on a set of protein samples with varying concentrations. We found that the Mason-Pfizer monkey virus WT MA exists in a monomer-dimer-trimer equilibrium in solution, with the corresponding dissociation constants of 2.3  and 0.24 mM, respectively. Structures of the oligomers calculated with HADDOCK software are closely related to the structures of other retroviral MA trimers.  相似文献   

5.
Herpes simplex virus (HSV) glycoproteins gB, gD, and gH/gL are necessary and sufficient for virus entry into cells. Structural features of gB are similar to those of vesicular stomatitis virus G and baculovirus gp64, and together they define the new class III group of fusion proteins. Previously, we used mutagenesis to show that three hydrophobic residues (W174, Y179, and A261) within the putative gB fusion loops are integral to gB function. Here we expanded our analysis, using site-directed mutagenesis of each residue in both gB fusion loops. Mutation of most of the nonpolar or hydrophobic amino acids (W174, F175, G176, Y179, and A261) had severe effects on gB function in cell-cell fusion and null virus complementation assays. Of the six charged amino acids, mutation of H263 or R264 also negatively affected gB function. To further analyze the mutants, we cloned the ectodomains of the W174R, Y179S, H263A, and R264A mutants into a baculovirus expression system and compared them with the wild-type (WT) form, gB730t. As shown previously, gB730t blocks virus entry into cells, suggesting that gB730t competes with virion gB for a cell receptor. All four mutant proteins retained this function, implying that fusion loop activity is separate from gB-receptor binding. However, unlike WT gB730t, the mutant proteins displayed reduced binding to cells and were either impaired or unable to bind naked, cholesterol-enriched liposomes, suggesting that it may be gB-lipid binding that is disrupted by the mutations. Furthermore, monoclonal antibodies with epitopes proximal to the fusion loops abrogated gB-liposome binding. Taken together, our data suggest that gB associates with lipid membranes via a fusion domain of key hydrophobic and hydrophilic residues and that this domain associates with lipid membranes during fusion.Herpes simplex virus (HSV) entry into cells requires four viral envelope glycoproteins (gB, gD, and the heterodimer gH/gL) as well as a cell surface gD receptor (reviewed in references 31, 42, 43, and 49). When gD binds its receptor, it undergoes conformational changes that are essential to activate the fusion machinery, gB and gH/gL. In addition to being essential for virus entry, both gH/gL and gB play important roles in primary fusion events that occur during egress of the capsid from the nuclei of infected cells (22). gB and gH/gL constitute the core fusion machinery of all members of the Herpesviridae.The mechanisms by which gB and gH/gL function individually and in concert during fusion are topics of intense investigations. Peptides based on predicted heptad repeats in gH block virus entry and have the ability to bind and disrupt model membranes (24, 26, 27). In addition, gH/gL can achieve hemifusion of adjacent cells in the absence of other herpesvirus proteins (50). These studies imply that gH/gL has fusogenic properties. Previously, we showed that both virion gB and soluble wild-type (WT) gB (gB730t), but not gD or gH/gL, bind to cells and associate with lipid rafts (10). Like gH/gL, several synthetic gB peptides induced the fusion of large unilamellar vesicles and inhibited herpesvirus infection (23, 24). Thus, it appears that both gB and gH/gL may be fusion proteins, a theory strengthened by data showing that either gB or gH/gL is sufficient for membrane fusion during nuclear egress (22). Additionally, gB730t blocks virus entry into cells deficient in heparan sulfate proteoglycans (HSPGs), suggesting that it competes with virion gB for an obligate cell surface receptor (9). A recent study suggested that paired immunoglobulin-like type 2 receptor alpha (PILRα) may serve this role for at least some cell types (47).The crystal structure of gB is now known for both HSV type 1 (HSV-1) (32) and Epstein-Barr virus (EBV) (6). Interestingly, gB is structurally related to two other viral fusion proteins, the vesicular stomatitis virus (VSV) G protein (45) and the baculovirus gp64 protein (34). VSV G, gB, and most recently, baculovirus gp64 were placed into a newly formed group of fusion proteins, the class III proteins. Class III fusion proteins have similar individual domain structures and contain a central three-stranded coiled coil reminiscent of the class I proteins. Whereas class I proteins have an N-terminal fusion peptide, class III proteins have internal bipartite fusion loops within domain I (shown in Fig. Fig.1A1A for gB) which are similar to the single fusion loop of class II fusion proteins. However, the class II fusion loop is composed entirely of hydrophobic amino acids, whereas the fusion loops of gB have both hydrophobic and charged residues (32, 34, 45). Unlike G or gp64, which are the sole fusion proteins for their respective viruses, gB requires gH/gL to function in fusion and entry.Open in a separate windowFIG. 1.HSV gB hydrophobic ridge is surrounded by charged residues on the surface of the molecule. A ribbon diagram of the HSV protomer (A) and molecular surface representation of the trimer (B) are shown. In each, one protomer is colored by secondary structure succession, using blue (domain I), green (domain II), yellow (domain III), orange (domain IV), and red (domain V). The box in panel A shows the primary amino acid sequences of the fusion loops. The box in panel B shows the base of the gB trimer, rotated 90°. For the boxes in both panels A and B, highlighted hydrophobic residues are colored in blue and charged residues are shown in red. All structural figures were generated, in part, using PyMOL Molecular Graphics System software.In our previous study, we used site-directed mutagenesis to show that three hydrophobic amino acids within the gB loops (W174, Y179, and A261) are essential for gB function (29). Similar studies of VSV G, gp64, and EBV gB support the notion that hydrophobic amino acids of both fusion loops are critical for fusion (34, 44, 51) and together constitute a fusion domain. Recently, bimolecular complementation was used to show that gB and gH/gL interact with each other concomitantly with fusion and that this interaction is triggered by binding of gD to its cellular receptor (3, 4). Thus, gB may function cooperatively with gH/gL, yet each may have some fusogenic potential on its own.The goal of the experiments reported here was twofold. First, we wanted to complete our mutagenic analysis of all of the residues in the two putative fusion loops of HSV gB. Our data show that the two fusion loops constitute a structural “subdomain” wherein key hydrophobic amino acids form a ridge that is supported on both sides by charged residues. We hypothesize that two charged residues on one side of the ridge enhance the ability of the hydrophobic residues to interact with target membranes and to function in fusion.Our second goal was to assess the effects of mutations in the fusion loops on the function of gB in cell binding, blocking of entry, and insertion into lipid membranes. Therefore, we constructed recombinant baculoviruses, with each carrying the gene for a truncated version (residues 31 to 730) of one of four mutant forms of gB (W174R, Y179S, H263A, and R264A). We found that the mutant proteins were able to efficiently block virus entry, suggesting that the fusion loops do not participate in protein-receptor binding. However, all four mutant proteins were impaired in cell binding compared to WT gB730t. Whereas WT gB730t associated with liposomes in a flotation assay, soluble truncated forms of HSV gD and gH/gL did not, consistent with our previous finding that gB730t associates with lipid rafts on cell surfaces (8). In contrast to WT gB730t, the gB mutant proteins were either impaired or unable to bind liposomes. Our data suggest that gB has an intrinsic ability to associate with a target membrane via its fusion domain.  相似文献   

6.
The human immunodeficiency virus type 1 transmembrane protein gp41 oligomer anchors the attachment protein, gp120, to the viral envelope and mediates viral envelope-cell membrane fusion following gp120-CD4 receptor-chemokine coreceptor binding. We have used mutation-directed chemical cross-linking with bis(sulfosuccinimidyl)suberate (BS3) to investigate the architecture of the gp41 oligomer. Treatment of gp41 with BS3 generates a ladder of four bands on sodium dodecyl sulfate-polyacrylamide gels, corresponding to monomers, dimers, trimers, and tetramers. By systematically replacing gp41 lysines with arginine and determining the mutant gp41 cross-linking pattern, we observed that gp41 N termini are cross-linked. Lysine 678, which is close to the transmembrane sequence, was readily cross-linked to Lys-678 on other monomers within the oligomeric structure. This arrangement appears to be facilitated by the close packing of membrane-anchoring sequences, since the efficiency of assembly of heterooligomers between wild-type and mutant Env proteins is improved more than twofold if the mutant contains the membrane-anchoring sequence. We also detected close contacts between Lys-596 and Lys-612 in the disulfide-bonded loop/glycan cluster of one monomer and lysines in the N-terminal amphipathic α-helical oligomerization domain (Lys-569 and Lys-583) and C-terminal α-helical sequence (Lys-650 and Lys-660) of adjacent monomers. Precursor-processing efficiency, gp120-gp41 association, soluble recombinant CD4-induced shedding of gp120 from cell surface gp41, and acquisition of gp41 ectodomain conformational antibody epitopes were unaffected by the substitutions. However, the syncytium-forming function was most dependent on the conserved Lys-569 in the N-terminal α-helix. These results indicate that gp160-derived gp41 expressed in mammalian cells is a tetramer and provide information about the juxtaposition of gp41 structural elements within the oligomer.  相似文献   

7.
The glycoproteins expressed by a Zaire species of Ebola virus were analyzed for cleavage, oligomerization, and other structural properties to better define their functions. The 50- to 70-kDa secreted and 150-kDa virion/structural glycoproteins (SGP and GP, respectively), which share the 295 N-terminal residues, are cleaved near the N terminus by signalase. A second cleavage event, occurring in GP at a multibasic site (RRTRR↓) that is likely mediated by furin, results in two glycoproteins (GP1 and GP2) linked by disulfide bonding. This furin cleavage site is present in the same position in the GPs of all Ebola viruses (R[R/K]X[R/K]R↓), and one is predicted for Marburg viruses (R[R/K]KR↓), although in a different location. Based on the results of cross-linking studies, we were able to determine that Ebola virion peplomers are composed of trimers of GP1-GP2 heterodimers and that aspects of their structure are similar to those of retroviruses, paramyxoviruses, and influenza viruses. We also determined that SGP is secreted from infected cells almost exclusively in the form of a homodimer that is joined by disulfide bonding.  相似文献   

8.
The C-terminal domain of human immunodeficiency virus type 1 (HIV-1) integrase (IN) is a dimer that binds to DNA in a nonspecific manner. The structure of the minimal region required for DNA binding (IN220–270) has been solved by nuclear magnetic resonance spectroscopy. The overall fold of the C-terminal domain of HIV-1 IN is similar to those of Src homology region 3 domains. Based on the structure of IN220–270, we studied the role of 15 amino acid residues potentially involved in DNA binding and oligomerization by mutational analysis. We found that two amino acid residues, arginine 262 and leucine 234, contribute to DNA binding in the context of IN220–270, as indicated by protein-DNA UV cross-link analysis. We also analyzed mutant proteins representing portions of the full-length IN protein. Amino acid substitution of residues located in the hydrophobic dimer interface, such as L241A and L242A, results in the loss of oligomerization of IN; consequently, the levels of 3′ processing, DNA strand transfer, and intramolecular disintegration are strongly reduced. These results suggest that dimerization of the C-terminal domain of IN is important for correct multimerization of IN.  相似文献   

9.
Herpes simplex virus type 1 glycoprotein B (gB) is essential for virus entry, an event involving fusion of the virus envelope with the cell surface membrane, and virus-induced cell-cell fusion, resulting in polykaryocyte, or syncytium, formation. The experiments described in this report employed a random mutagenesis strategy to develop a more complete genetic map of mutations resulting in the syn mutant phenotype. The results indicate that syn mutations occur within two essential and highly conserved hydrophilic, alpha-helical regions of the gB cytoplasmic domain. Region I is immediately proximal to the transmembrane domain and includes residues R796 to E816/817. Region II is localized centrally in the cytoplasmic domain and includes residues A855 and R858. Positively charged residues were particularly affected in both regions, suggesting that charge interactions may be required to suppress the syn mutant phenotype. No syn mutations were identified within the transmembrane domain. A virus containing a rate of entry (roe) mutation at residue A851, either within or immediately proximal to syn region II, was isolated. Since roe mutations have also been discovered in the external domain of gB, it appears likely that the external and cytoplasmic domains cooperate in virus penetration. Moreover, the observation that both roe and syn mutations occur in the cytoplasmic domain further suggests that gB functions in an analogous manner in both membrane fusion events. It might be predicted from these observations that membrane fusion involves transduction of a fusion signal along the gB molecule through the transmembrane domain. Communication between the external and cytoplasmic domain may thus be required for gB-mediated membrane fusion events.  相似文献   

10.
Synaptotagmin I (Syt I), a proposed major Ca(2+) sensor in the central nervous system, has been hypothesized as functioning in an oligomerized state during neurotransmitter release. We previously showed that Syts I, II, VII, and VIII form a stable SDS-resistant, beta-mercaptoethanol-insensitive, and Ca(2+)-independent oligomer surrounding the transmembrane domain (Fukuda, M., and Mikoshiba, K. (2000) J. Biol. Chem. 275, 28180-28185), but little is known about the molecular mechanism of the Ca(2+)-independent oligomerization by the synaptotagmin family. In this study, we analyzed the Ca(2+)-independent oligomerization properties of Syt I and found that it shows two distinct forms of self-oligomerization activity: stable SDS-resistant self-oligomerization activity and relatively unstable SDS-sensitive self-oligomerization activity. The former was found to be mediated by a post-translationally modified (i.e. fatty-acylated) cysteine (Cys) cluster (Cys-74, Cys-75, Cys-77, Cys-79, and Cys-82) at the interface between the transmembrane and spacer domains of Syt I. We also show that the number of Cys residues at the interface between the transmembrane and spacer domains determines the SDS- resistant oligomerizing capacity of each synaptotagmin isoform: Syt II, which contains seven Cys residues, showed the strongest SDS-resistant oligomerizing activity in the synaptotagmin family, whereas Syt XII, which has no Cys residues, did not form any SDS-resistant oligomers. The latter SDS-sensitive self-oligomerization of Syt I is mediated by the spacer domain, because deletion of the whole spacer domain, including the Cys cluster, abolished it, whereas a Syt I(CA) mutant carrying Cys to Ala substitutions still exhibited self-oligomerization. Based on these results, we propose that the oligomerization of the synaptotagmin family is regulated by two distinct mechanisms: the stable SDS-resistant oligomerization is mediated by the modified Cys cluster, whereas the relatively unstable (SDS-sensitive) oligomerization is mediated by the environment of the spacer domain.  相似文献   

11.
A significant proportion of the alpha2,6-sialyltransferase of protein Asn-linked glycosylation (ST6Gal I) forms disulfide-bonded dimers that exhibit decreased activity, but retain the ability to bind asialoglycoprotein substrates. Here, we have investigated the subcellular location and mechanism of ST6Gal I dimer formation, as well as the role of Cys residues in the enzyme's trafficking, localization, and catalytic activity. Pulse-chase analysis demonstrated that the ST6Gal I disulfide-bonded dimer forms in the endoplasmic reticulum. Mutagenesis experiments showed that Cys-24 in the transmembrane region is required for dimerization, while catalytic domain Cys residues are required for trafficking and catalytic activity. Replacement of Cys-181 and Cys-332 generated proteins that are largely retained in the endoplasmic reticulum and minimally active or inactive, respectively. Replacement of Cys-350 or Cys-361 inactivated the enzyme without compromising its localization or processing, suggesting that these amino acids are part of the enzyme's active site. Replacement of Cys-139 or Cys-403 generated proteins that are catalytically active and appear to be more stably localized in the Golgi, since they exhibited decreased cleavage and secretion. The Cys-139 mutant also exhibited increased dimer formation suggesting that ST6Gal I dimers may be critical in the oligomerization process involved in stable ST6Gal I Golgi localization.  相似文献   

12.
In an earlier report (S.D. Marlin, S.L. Highlander, T.C. Holland, M. Levine, and J.C. Glorioso, J. Virol. 59: 142-153), we described the production and use of complement-dependent virus-neutralizing monoclonal antibodies (MAbs) and MAb-resistant (mar) mutants to identify five antigenic sites (I to V) on herpes simplex virus type 1 glycoprotein B (gB). In the present study, the mechanism of virus neutralization was determined for a MAb specific for site III (B4), the only site recognized by MAbs which exhibited complement-independent virus-neutralizing ability. This antibody had no detectable effect on virus attachment but neutralized viruses after adsorption to cell monolayers. These findings implied that the mechanism of B4 neutralization involved blocking of virus penetration. The remaining antibodies, which recognized sites I, II, and IV, required active complement for effective neutralization. These were further studied for their ability to impede virus infectivity in the absence of complement. Antibodies to sites I (B1 and B3) and IV (B6) slowed the rate at which viruses penetrated cell surfaces, supporting the conclusion that antibody binding to gB can inhibit penetration by a virus. The data suggest that MAbs can interfere with penetration by a virus by binding to a domain within gB which is involved in this process. In another assay of virus infection, MAb B6 significantly reduced plaque development, indicating that antibody binding to gB expressed on infected-cell surfaces can also interfere with the ability of a virus to spread from cell to cell. In contrast to these results, antibodies to site II (B2 and B5) had no effect on virus infectivity; this suggests that they recognized structures which do not play a direct role in the infectious process. To localize regions of gB involved in these phenomena, antibody-binding sites were operationally mapped by radioimmunoprecipitation of a panel of truncated gB molecules produced in transient-expression assays. Residues critical to recognition by antibodies which affect penetration by a virus (sites I, III, and IV) mapped to a region of the molecule (amino acid residues 241 to 441) which is centrally located within the external domain. Antibodies which had no effect on penetration (site II) recognized sequences distal to this region (residues 596 to 737) near the transmembrane domain. The data suggest that these gB-specific MAbs recognize two major antigenic sites which reside in physically distinct components of the external domain of gB.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

13.
14.
The solute carrier transmembrane protein prestin (SLC26A5) drives an active electromechanical transduction process in cochlear outer hair cells that increases hearing sensitivity and frequency discrimination in mammals. A large intramembraneous charge movement, the nonlinear capacitance (NLC), is the electrical signature of prestin function. The transmembrane domain (TMD) helices and residues involved in the intramembrane charge displacement remain unknown. We have performed cysteine-scanning mutagenesis with serine or valine replacement to investigate the importance of cysteine residues to prestin structure and function. The distribution of oligomeric states and membrane abundance of prestin was also probed to investigate whether cysteine residues participate in prestin oligomerization and/or NLC. Our results reveal that 1) Cys-196 (TMD 4) and Cys-415 (TMD 10) do not tolerate serine replacement, and thus maintaining hydrophobicity at these locations is important for the mechanism of charge movement; 2) Cys-260 (TMD 6) and Cys-381 (TMD 9) tolerate serine replacement and are probably water-exposed; and 3) if disulfide bonds are present, they do not serve a functional role as measured via NLC. These novel findings are consistent with a recent structural model, which proposes that prestin contains an occluded aqueous pore, and we posit that the orientations of transmembrane domain helices 4 and 10 are essential for proper prestin function.  相似文献   

15.
Glycoprotein B (gB) enables the fusion of viral and cell membranes during entry of herpesviruses. However, gB alone is insufficient for membrane fusion; the gH/gL heterodimer is also required. The crystal structure of the herpes simplex virus type 1 (HSV-1) gB ectodomain, gB730, has demonstrated similarities between gB and other viral fusion proteins, leading to the hypothesis that gB is a fusogen, presumably directly involved in bringing the membranes together by refolding from its initial or prefusion form to its final or postfusion form. The only available crystal structure likely represents the postfusion form of gB; the prefusion form has not yet been determined. Previously, a panel of HSV-1 gB mutants was generated by using random 5-amino-acid-linker insertion mutagenesis. Several mutants were unable to mediate cell-cell fusion despite being expressed on the cell surface. Mapping of the insertion sites onto the crystal structure of gB730 suggested that several insertions might not be accommodated in the postfusion form. Thus, we hypothesized that some insertion mutants were nonfunctional due to being “trapped” in a prefusion form. Here, we generated five insertion mutants as soluble ectodomains and characterized them biochemically. We show that the ectodomains of all five mutants assume conformations similar to that of the wild-type gB730. Four mutants have biochemical properties and overall structures that are indistinguishable from those of the wild-type gB730. We conclude that these mutants undergo only minor local conformational changes to relieve the steric strain resulting from the presence of 5 extra amino acids. Interestingly, one mutant, while able to adopt the overall postfusion structure, displays significant conformational differences in the vicinity of fusion loops, relative to wild-type gB730. Moreover, this mutant has a diminished ability to associate with liposomes, suggesting that the fusion loops in this mutant have decreased functional activity. We propose that these insertions cause a fusion-deficient phenotype not by preventing conversion of gB to a postfusion-like conformation but rather by interfering with other gB functions.Herpes simplex virus type 1 (HSV-1) is the prototype of the diverse herpesvirus family that includes many notable human pathogens (26). In addition to the icosahedral capsid and the tegument that surround its double-stranded DNA genome, herpesviruses have an envelope—an outer lipid bilayer—bearing a number of surface glycoproteins. During infection, HSV-1 must fuse its envelope with a cellular membrane in order to deliver the capsid into a target host cell. Among its viral glycoproteins, only glycoprotein C (gC), gB, gD, gH, and gL participate in this entry process, and only the last four are required for fusion (28). Although gD is found only in alphaherpesviruses, all herpesviruses encode gB, gH, and gL, which constitute their core fusion machinery. Of these three proteins, gB is the most highly conserved.We recently determined the crystal structure of a nearly full-length ectodomain of HSV-1 gB, gB730 (18). The crystal structure of the ectodomain of gB from Epstein-Barr virus, another herpesvirus, has also been subsequently determined (4). The two structures showed similarities between gB and other viral fusion proteins, in particular, G from an unrelated vesicular stomatitis virus (VSV) (25), leading to the hypothesis that gB is a fusogen, presumably directly involved in bringing the viral and host cell membranes together to enable their fusion. However, gB alone is known to be insufficient for membrane fusion; the gH/gL heterodimer is also required. This insufficiency raises the question of exactly how gB functions during viral entry. Answering this question is critical for understanding the complex mechanism that herpesviruses use to enter their host cells.In acting as a viral fusogen, gB must undergo dramatic conformational changes, refolding through a series of conformational intermediates from its initial, or prefusion form, to its final, or postfusion form (15). These conformational changes are not only necessary to bring the two membranes into proximity; they are also thought to provide the energy for the fusion process. The prefusion form corresponds to the protein present on the viral surface prior to initiation of fusion. The postfusion form represents the protein after fusion of the viral and host cell membranes. The available gB structure likely represents its postfusion form, since it shares more in common with the postfusion rather than the prefusion structure of vesicular stomatitis virus (VSV) G (3, 17). However, the prefusion form has not yet been characterized.Recently, a panel of gB mutants was generated by using random linker-insertion mutagenesis (21). Of these mutants, 16 were particularly interesting because they were nonfunctional in cell-cell fusion assays despite being expressed on the cell surface at levels that indicate proper folding for transport. These observations suggested that each insertion somehow interfered with gB function. Insertions in 12 of these mutants are located within the available structure of the gB ectodomain, which allowed Lin and Spear to analyze their locations (21).The most prominent examples of such nonfunctional mutants are two mutants with insertions after residues I185 or E187, henceforth referred to as “cavity mutants” because both I185 and E187 point into a cavity inside the gB trimer (Fig. 1B and D). Although this cavity might accommodate a single 5-amino-acid insertion, it “is not large enough to accommodate three 5-amino-acid insertions” (21) that would be present in the trimer (one insertion per protomer).Open in a separate windowFIG. 1.Location of the insertion sites in the sequence of gB and the structure of the postfusion form of its ectodomain. (A) Linear diagram of the full-length gB with functional domains highlighted (as in reference 18). Domain I is shown in cyan, domain II in green, domain III in yellow, domain IV in orange, domain V in red, and the disordered region between domains II and III in purple. Regions absent from the crystal structure of gB730 are shown in gray. Sequences in the region of 5-amino-acid insertions (residues 181 to 190 and residues 661 to 680) are shown in black. Arrows mark the locations of 5-amino-acid insertions, shown as red text. (B) Crystal structure of gB730 (18). Residues preceding the 5-amino-acid insertions in mutants studied here are shown as spheres colored by domain, consistent with panel A. Boxes delineate the hinge region, enlarged in panel C, and the cavity region, enlarged in panel D. (C) Close-up view of the hinge region shown in molecular surface representation, with residues 663 to 675 displayed as sticks. Hydrophobic residues are colored orange. Residues preceding the 5-amino-acid insertions in mutants studied here are labeled with asterisks; remaining labels correspond to additional hydrophobic residues in the 663-675 region. (D) Enlarged view of the cavity region. Residues that line the cavity and are not solvent exposed are colored magenta. Residue E187 of each protomer is colored teal and shown as spheres. Fusion loops for two protomers are marked with asterisks; the third pair of fusion loops lies behind the crystal structure and is not visible. Panels B, C, and D were made by using Pymol (http://www.pymol.org/).Five other nonfunctional mutants have insertions after residues D663, T665, V667, I671, or L673, respectively. We refer to them as “hinge mutants.” These residues lie in the region located between domains IV and V, which has been termed the hinge region because it may play an important role during the conformational transition from the prefusion to the postfusion form (17). Lin and Spear proposed that insertions following these residues “would likely affect hinge regions” (21), with the implication that they may prevent gB from refolding into the postfusion conformation. Our analysis suggested that insertions after these residues could, perhaps, be sterically accommodated in the structure but would probably be energetically unfavorable by causing several buried hydrophobic side chains in the 665-673 region, such as F670, I671, and L673, to become exposed (Fig. 1B and C).In light of these observations, we hypothesized that the insertion mutants are “trapped” in a prefusion form. We decided to test this hypothesis by determining whether the ectodomain of gB containing one of these insertion mutations is able to assume the conformation seen in the crystal structure of the wild-type gB ectodomain, which we are referring to as the likely postfusion conformation. For this purpose, we chose one cavity mutant, containing an insertion after E187, and four hinge mutants, containing insertions after T665, V667, I671, or L673, respectively. We chose to test four hinge mutants because structure analysis suggested to us that insertions following the respective residues might not affect the structure in precisely the same way. We expressed the soluble ectodomain of each mutant by using a baculovirus expression system and characterized the purified proteins by using biochemical and biophysical methods. Surprisingly, we found that the ectodomains of all five mutants assume a conformation similar to that of the wild-type gB ectodomain. The four hinge mutants had biochemical properties and overall three-dimensional structures that were indistinguishable from those of the wild-type gB ectodomain. We conclude that these mutants undergo only minor local conformational changes to relieve the steric strain resulting from the presence of 5 extra amino acids. Interestingly, the cavity mutant, while able to adopt the overall postfusion structure, still displayed significant conformational differences relative to wild-type gB. Because these conformational differences are in the vicinity of fusion loops, we conclude that the fusion loops in this mutant have decreased functional activity.  相似文献   

16.
Mouse hepatitis virus receptor (MHVR) is a murine biliary glycoprotein (Bgp1a). Purified, soluble MHVR expressed from a recombinant vaccinia virus neutralized the infectivity of the A59 strain of mouse hepatitis virus (MHV-A59) in a concentration-dependent manner. Several anchored murine Bgps in addition to MHVR can also function as MHV-A59 receptors when expressed at high levels in nonmurine cells. To investigate the interactions of these alternative MHVR glycoproteins with MHV, we expressed and purified to apparent homogeneity the extracellular domains of several murine Bgps as soluble, six-histidine-tagged glycoproteins, using a baculovirus expression system. These include MHVR isoforms containing four or two extracellular domains and the corresponding Bgp1b glycoproteins from MHV-resistant SJL/J mice, as well as Bgp2 and truncation mutants of MHVR and Bgp1b comprised of the first two immunoglobulin-like domains. The soluble four-domain MHVR glycoprotein (sMHVR[1-4]) had fourfold more MHV-A59 neutralizing activity than the corresponding soluble Bgp1b (sBgp1b) glycoprotein and at least 1,000-fold more neutralizing activity than sBgp2. Although virus binds to the N-terminal domain (domain 1), soluble truncation mutants of MHVR and Bgp1b containing only domains 1 and 2 bound virus poorly and had 10- and 300-fold less MHV-A59 neutralizing activity than the corresponding four-domain glycoproteins. In contrast, the soluble MHVR glycoprotein containing domains 1 and 4 (sMHVR[1,4]) had as much neutralizing activity as the four-domain glycoprotein, sMHVR[1-4]. Thus, the virus neutralizing activity of MHVR domain 1 appears to be enhanced by domain 4. The sBgp1b[1-4] glycoprotein had 500-fold less neutralizing activity for MHV-JHM than for MHV-A59. Thus, MHV strains with differences in S-glycoprotein sequence, tissue tropism, and virulence can differ in the ability to utilize the various murine Bgps as receptors.  相似文献   

17.
V Huff  W Cai  J C Glorioso    M Levine 《Journal of virology》1988,62(11):4403-4406
Glycoprotein B (gB) is a virally encoded protein that is found in the envelope of herpes simplex virus type 1 and membranes of cells infected with herpes simplex virus type 1. It is essential for the production of infectious virus particles. An amber mutation was introduced into the gB gene by oligonucleotide-directed mutagenesis at the codon for amino acid 863 of the protein. Virus carrying this mutation should synthesize gB molecules lacking the last 41 amino acids of the cytoplasmic domain. Immunoprecipitation of infected cell extracts demonstrated the synthesis of appropriately truncated gB molecules. Characterization of the mutant virus indicated that the loss of the carboxy-terminal 41 amino acids has little effect on gB function.  相似文献   

18.
Disulfide bonds of herpes simplex virus type 2 glycoprotein gB.   总被引:1,自引:1,他引:0       下载免费PDF全文
Glycoprotein B (gB) is the most highly conserved envelope glycoprotein of herpesviruses. The gB protein is required for virus infectivity and cell penetration. Recombinant forms of gB being used for the development of subunit vaccines are able to induce virus-neutralizing antibodies and protective efficacy in animal models. To gain structural information about the protein, we have determined the location of the disulfide bonds of a 696-amino-acid residue truncated, recombinant form of herpes simplex virus type 2 glycoprotein gB (HSV gB2t) produced by expression in Chinese hamster ovary cells. The purified protein, which contains virtually the entire extracellular domain of herpes simplex virus type 2 gB, was digested with trypsin under nonreducing conditions, and peptides were isolated by reversed-phase high-performance liquid chromatography (HPLC). The peptides were characterized by using mass spectrometry and amino acid sequence analysis. The conditions of cleavage (4 M urea, pH 7) induced partial carbamylation of the N termini of the peptides, and each disulfide peptide was found with two or three different HPLC retention times (peptides with and without carbamylation of either one or both N termini). The 10 cysteines of the molecule were found to be involved in disulfide bridges. These bonds were located between Cys-89 (C1) and Cys-548 (C8), Cys-106 (C2) and Cys-504 (C7), Cys-180 (C3) and Cys-244 (C4), Cys-337 (C5) and Cys-385 (C6), and Cys-571 (C9) and Cys-608 (C10). These disulfide bonds are anticipated to be similar in the corresponding gBs from other herpesviruses because the 10 cysteines listed above are always conserved in the corresponding protein sequences.  相似文献   

19.
The E. coli RelA protein is a ribosome-dependent (p)ppGpp synthetase that is activated in response to amino acid starvation. RelA can be dissected both functionally and physically into two domains: The N-terminal domain (NTD) (amino acids [aa] 1 to 455) contains the catalytic domain of RelA, and the C-terminal domain (CTD) (aa 455 to 744) is involved in regulating RelA activity. We used mutational analysis to localize sites important for RelA activity and control in these two domains. We inserted two separate mutations into the NTD, which resulted in mutated RelA proteins that were impaired in their ability to synthesize (p)ppGpp. When we caused the CTD in relA+ cells to be overexpressed, (p)ppGpp accumulation during amino acid starvation was negatively affected. Mutational analysis showed that Cys-612, Asp-637, and Cys-638, found in a conserved amino acid sequence (aa 612 to 638), are essential for this negative effect of the CTD. When mutations corresponding to these residues were inserted into the full-length relA gene, the mutated RelA proteins were impaired in their regulation. In attempting to clarify the mechanism through which the CTD regulates RelA activity, we found no evidence for competition for ribosomal binding between the normal RelA and the overexpressed CTD. Results from CyaA complementation experiments of the bacterial two-hybrid system fusion plasmids (G. Karimova, J. Pidoux, A. Ullmann, and D. Ladant, Proc. Natl. Acad. Sci. USA 95:5752–5756, 1998) indicated that the CTD (aa 564 to 744) is involved in RelA-RelA interactions. Our findings support a model in which RelA activation is regulated by its oligomerization state.  相似文献   

20.
Epstein-Barr Virus (EBV) glycoprotein B (gB) is essential for viral fusion events with epithelial and B cells. This glycoprotein has been studied extensively in other herpesvirus family members, but functional domains outside of the cytoplasmic tail have not been characterized in EBV gB. In this study, a total of 28 linker insertion mutations were generated throughout the length of gB. In general, the linker insertions did not disrupt intracellular expression and variably altered cell surface expression. Oligomerization was disrupted by insertions located between residues 561 and 620, indicating the location of a potential site of oligomer contacts between EBV gB monomers. In addition, a novel N-glycosylated form of wild-type gB was identified under nonreducing Western blot conditions that likely represents a mature form of the protein. Fusion activity was abolished in all but three variants containing mutations in the N-terminal region (gB30), within the ectodomain (gB421), and in the intracellular C-terminal domain (gB832) of the protein. Fusion activity with variants gB421 and gB832 was comparable to that of the wild type with epithelial and B cells, and only these two mutants, but not gB30, were able to complement gB-null virus and subsequently function in virus entry. The mutant gB30 exhibited a low level of fusion activity with B cells and was unable to complement gB-null virus. The mutations generated here indicate important structural domains, as well as regions important for function in fusion, within EBV gB.  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