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 15 毫秒
1.
To understand the relationship between conformational maturation and quality control-mediated proteolysis in the secretory pathway, we engineered the well-characterized degron from the alpha-subunit of the T-cell antigen receptor (TCRalpha) into the alpha-helical transmembrane domain of homotrimeric type I integral membrane protein, influenza hemagglutinin (HA). Although the membrane degron does not appear to interfere with acquisition of native secondary structure, as assessed by the formation of native intrachain disulfide bonds, only approximately 50% of nascent mutant HA chains (HA(++)) become membrane-integrated and acquire complex N-linked glycans indicative of transit to a post-ER compartment. The remaining approximately 50% of nascent HA(++) chains fail to integrate into the lipid bilayer and are subject to proteasome-dependent degradation. Site-specific cleavage by extracellular trypsin and reactivity with conformation-specific monoclonal antibodies indicate that membrane-integrated HA(++) molecules are able to mature to the plasma membrane with a conformation indistinguishable from that of HA(wt). These apparently native HA(++) molecules are, nevertheless, rapidly degraded by a process that is insensitive to proteasome inhibitors but blocked by lysosomotropic amines. These data suggest the existence in the secretory pathway of at least two sequential quality control checkpoints that recognize the same transmembrane degron, thereby ensuring the fidelity of protein deployment to the plasma membrane.  相似文献   

2.
J Brunner 《FEBS letters》1989,257(2):369-372
Low pH-induced binding of the bromelain-solubilized form of influenza virus hemagglutinin (BHA) to membranes occurs through the fusion peptide. From asymmetric hydrophobic photolabeling of membranes, evidence was obtained that this peptide penetrates only one leaflet of the bilayer. The asymmetrical labeling was achieved by employing a photoreactive analogue of a fatty acid whose transbilayer distribution can be manipulated by a membrane proton gradient.  相似文献   

3.
Fusion of influenza viruses with membranes is catalyzed by the viral spike protein hemagglutinin (HA). Under mildly acidic conditions (approximately pH 5) this protein undergoes a conformational change that triggers the exposure of the "fusion peptide", the hydrophobic N-terminal segment of the HA2 polypeptide chain. Insertion of this segment into the target membrane (or viral membrane?) is likely to represent a key step along the fusion pathway, but the details are far from being clear. The photoreactive phospholipid 1-palmitoyl-2-[11-[4-[3-(trifluoromethyl)diazirinyl]phenyl] [2-3H]undecanoyl]-sn-glycero-3-phosphocholine ([3H]PTPC/11), inserted into the bilayer of large unilamellar vesicles (LUVs), allowed us to investigate both the interaction of viruses with the vesicles under "prefusion" conditions (pH 5; 0 degrees C) and the fusion process itself occurring at elevated temperatures (greater than 15-20 degrees C) only. Despite the observed binding of viruses to LUVs at pH 5 and 0 degrees C, labeling of HA2 was very weak (less than 0.002% of the radioactivity originally present). In contrast, fusion could be readily monitored by the covalent labeling of that polypeptide chain. We have studied also the effect of temperature on the acid-induced (pH 5) interaction of bromelain-solubilized HA (BHA) with vesicles. Labeling of the BHA2 polypeptide chain was found to show a remarkable correlation with the temperature dependence of the fusion activity of whole viruses. A temperature-induced structural change appears to be critical for both the interaction of BHA with membranes and the expression of fusion activity of intact viruses.  相似文献   

4.
Li Y  Han X  Lai AL  Bushweller JH  Cafiso DS  Tamm LK 《Journal of virology》2005,79(18):12065-12076
Influenza virus hemagglutinin (HA)-mediated membrane fusion is initiated by a conformational change that releases a V-shaped hydrophobic fusion domain, the fusion peptide, into the lipid bilayer of the target membrane. The most N-terminal residue of this domain, a glycine, is highly conserved and is particularly critical for HA function; G1S and G1V mutant HAs cause hemifusion and abolish fusion, respectively. We have determined the atomic resolution structures of the G1S and G1V mutant fusion domains in membrane environments. G1S forms a V with a disrupted "glycine edge" on its N-terminal arm and G1V adopts a slightly tilted linear helical structure in membranes. Abolishment of the kink in G1V results in reduced hydrophobic penetration of the lipid bilayer and an increased propensity to form beta-structures at the membrane surface. These results underline the functional importance of the kink in the fusion peptide and suggest a structural role for the N-terminal glycine ridge in viral membrane fusion.  相似文献   

5.
We have simulated two conformations of the fusion domain of influenza hemagglutinin (HA) within explicit water, salt, and heterogeneous lipid bilayers composed of POPC:POPG (4:1). Each conformation has seven different starting points in which the initial peptide structure is the same for each conformation, but the location across the membrane normal and lipid arrangement around the peptide are varied, giving a combined total simulation time of 140 ns. For the HA5 conformation (primary structure from recent NMR spectroscopy at pH = 5), the peptide exhibits a stable and less kinked structure in the lipid bilayer compared to that from the NMR studies. The relative fusogenic behavior of the different conformations has been investigated by calculation of the relative free energy of insertion into the hydrophobic region of lipid bilayer as a function of the depth of immersion. For the HA7 conformations (primary structure from recent NMR spectroscopy at pH = 7.4), while the N-terminal helix preserves its initial structure, the flexible C-terminal chain produces a transient helical motif inside the lipid bilayer. This conformational change is pH-independent, and is closely related to the peptide insertion into the lipid bilayer.  相似文献   

6.
To investigate the dynamics of membrane processes that may be integral components of specific transmembrane signaling events we have synthesized several novel paramagnetic probes and their photoreactive counterparts. The structure of these probes was designed to (1) restrict “flipping” across the membrane bilayer; (2) contain paramagnetic or photoreactive moieties that could be placed at specific depths within the bilayer; (3) provide information about membrane structure as well as dynamics of protein movement; and (4) in the case of the photoreactive probes, be of high specific radioactivity. The molecules described in this paper consist of amino acid, dipeptide, or carbohydrate groups attached to arylazide- or nitroxide-bearing fatty acids. The synthesis and initial characterization of these membrane probes is described.  相似文献   

7.
Fusion of influenza viruses with target membranes is induced by acid and involves complex changes in the viral fusion protein hemagglutinin (HA) and in the contact sites between viruses and target membranes (Stegmann, T., White, J. M., and Helenius, A. (1990) EMBO J. 9, 4231-4241). At 0 degrees C, in a first, kinetically distinct step, target membranes irreversibly adhere to the viruses. Fusion itself starts only after a lag-phase of several minutes (X-31 strain viruses) or after raising the temperature (PR8/34 strain viruses). We now provide evidence that the initial conformational change resulting in virus-target membrane adhesion is restricted to a (minor) subpopulation of the HA molecules. These molecules become susceptible to bromelain digestion, and they could be labeled with the photoactivatable reagent [3H]PTPC/11, a nonexchangeable lipid present in the target lipid bilayer (Harter, C., B?chi, T., Semenza, G., and Brunner, J. (1988) Biochemistry 27, 1856-1864). Only the HA2 subunit was labeled, and analyses of 2-nitro-5-thio-cyanobenzoic acid fragments derived thereof indicate that the HA2 NH2-terminal segment (fusion peptide) inserted into the target membrane bilayer. When the temperature was raised to trigger fusion of PR8/34 viruses, labeling of HA2 increased by a factor of 130. Most (74%) of that label was incorporated into the COOH-terminal membrane anchor region, but there was also a strong increase (about 30-fold) of NH2-terminal fusion peptide labeling. This suggests that fusion is preceded., or accompanied, by further changes in HA which lead to additional extensive lipid insertions of HA2 fusion peptides.  相似文献   

8.
The effects of membrane tension on fusion between cells expressing the hemagglutinin (HA) of influenza virus and red blood cells were studied by capacitance measurements. Inflation of an HA-expressing cell was achieved by applying a positive hydrostatic pressure to its interior through a patch-clamp pipette in the whole-cell configuration. Inflating cells to the maximum extent possible without lysis created a membrane tension and completely inhibited low-pH-induced fusion at room temperature. Fully inflated cells that were subsequently deflated to normal size resumed the ability to fuse in response to low pH. At the higher temperature of 32 degrees C, fusion conditions were sufficiently optimal that full inflation did not hinder fusion, and once formed, pores enlarged more rapidly than those of never inflated cells. It is suggested that under fusogenic conditions HA causes the formation of a dimple within the membrane in which it resides, and that membrane tension hinders fusion by preventing the formation of dimples. Because dimpling bends the bilayer portion of bound membranes so that they come into intimate contact, the damping of dimpling would suppress this initial step in the fusion process.  相似文献   

9.
Influenza A virus strains adopt different host specificities mainly depending on their hemagglutinin (HA) protein. Via HA, the virus binds sialic acid receptors of the host cell and, upon endocytic uptake, HA triggers fusion between the viral envelope bilayer and the endosomal membrane by a low pH-induced conformational change leading to the release of the viral genome into the host cell cytoplasm. Both functions are crucial for viral infection enabling the genesis of new progeny virus.  相似文献   

10.
Intermediates in influenza induced membrane fusion.   总被引:34,自引:2,他引:32       下载免费PDF全文
T Stegmann  J M White    A Helenius 《The EMBO journal》1990,9(13):4231-4241
Our results show that the mechanism by which influenza virus fuses with target membranes involves sequential complex changes in the hemagglutinin (HA, the viral fusion protein) and in the contact site between virus and target membrane. To render individual steps amenable to study, we worked at 0 degree C which decreased the rate of fusion and increased the efficiency. The mechanism of fusion at 0 degree C and 37 degrees C was similar. The process began with a conformational change in HA which exposed the fusion peptides but did not lead to dissociation of the tops of the ectodomain of the trimer. The change in the protein led to immediate hydrophobic attachment of the virus to the target liposomes. Attachment was followed by a lag period (4-8 min at 0 degree C, 0.6-2 s at 37 degrees C) during which rearrangements occurred in the site of membrane contact between the virus and liposome. After a further series of changes the final bilayer merger took place. This final fusion event was not pH dependent. At 0 degree C efficient fusion occurred without dissociation of the top domains of the HA trimer, suggesting that a transient conformation of HA is responsible for fusion at physiological temperatures. The observations lead to a revised model for HA mediated fusion.  相似文献   

11.
The chronological relation between the establishment of lipid continuity and fusion pore formation has been investigated for fusion of cells expressing hemagglutinin (HA) of influenza virus to planar bilayer membranes. Self-quenching concentrations of lipid dye were placed in the planar membrane to monitor lipid mixing, and time-resolved admittance measurements were used to measure fusion pores. For rhodamine-PE, fusion pores always occurred before a detectable amount of dye moved into an HA-expressing cell. However, with DiI in the planar membrane, the relationship was reversed: the spread of dye preceded formation of small pores. In other words, by using DiI as probe, hemifusion was clearly observed to occur before pore formation. For hemifused cells, a small pore could form and subsequently fully enlarge. In contrast, for cells that express a glycosylphosphatidylinositol-anchored ectodomain of HA, hemifusion occurred, but no fully enlarged pores were observed. Therefore, the transmembrane domain of HA is required for the formation of fully enlarging pores. Thus, with the planar bilayer membranes as target, hemifusion can precede pore formation, and the occurrence of lipid dye spread does not preclude formation of pores that can enlarge fully.  相似文献   

12.
While the importance of viral fusion peptides (e.g., hemagglutinin (HA) and gp41) in virus-cell membrane fusion is established, it is unclear how these peptides enhance membrane fusion, especially at low peptide/lipid ratios for which the peptides are not lytic. We assayed wild-type HA fusion peptide and two mutants, G1E and G13L, for their effects on the bilayer structure of 1,2-dioleoyl-3-sn-phosphatidylcholine/1,2-dioleoyl-3-sn-phosphatidylethanolamine/Sphingomyelin/Cholesterol (35:30:15:20) membranes, their structures in the lipid bilayer, and their effects on membrane fusion. All peptides bound to highly curved vesicles, but fusion was triggered only in the presence of poly(ethylene glycol). At low (1:200) peptide/lipid ratios, wild-type peptide enhanced remarkably the extent of content mixing and leakage along with the rate constants for these processes, and significantly enhanced the bilayer interior packing and filled the membrane free volume. The mutants caused no change in contents mixing or interior packing. Circular dichroism, polarized-attenuated total-internal-reflection Fourier-transform infrared spectroscopy measurements, and membrane perturbation measurements all conform to the inverted-V model for the structure of wild-type HA peptide. Similar measurements suggest that the G13L mutant adopts a less helical conformation in which the N-terminus moves closer to the bilayer interface, thus disrupting the V-structure. The G1E peptide barely perturbs the bilayer and may locate slightly above the interface. Fusion measurements suggest that the wild-type peptide promotes conversion of the stalk to an expanded trans-membrane contact intermediate through its ability to occupy hydrophobic space in a trans-membrane contact structure. While wild-type peptide increases the rate of initial intermediate and final pore formation, our results do not speak to the mechanisms for these effects, but they do leave open the possibility that it stabilizes the transition states for these events.  相似文献   

13.
The hemagglutinin (HA) of influenza virus is a type I transmembrane glycoprotein which is acylated with long-chain fatty acids. In this study we have used oligonucleotide-directed mutagenesis of cloned cDNA and a simian virus 40 expression system to determine the fatty acid binding site in HA and to examine possible functions of covalently linked fatty acids. The results show that the HA is acylated through thioester linkages at three highly conserved cysteine residues located in the cytoplasmic domain and at the carboxy-terminal end of the transmembrane region, whereas a cysteine located in the middle of the membrane-spanning domain is not acylated. Mutants lacking fatty acids at individual or all three attachment sites acquire endoglycosidase H-resistant oligosaccharide side chains, are cleaved into HA1 and HA2 subunits, and are transported to the plasma membrane at rates similar to that of wild-type HA. All mutants are membrane bound and not secreted into the medium. These results exclude transport signal and membrane-anchoring functions of covalently linked fatty acids for this integral membrane glycoprotein. Furthermore, lack of acylation has no obvious influence on the biological activities of HA: cells expressing fatty acid-free HA bind to and, after brief exposure to mildly acidic pH, fuse with erythrocytes; the HA-induced polykaryon formation is not impaired, either. Other possible functions of covalently linked fatty acids in integral membrane glycoproteins which cannot be examined in conventional cDNA expression systems are discussed.  相似文献   

14.
We have studied the fusion between voltage-clamped planar lipid bilayers and influenza virus infected MDCK cells, adhered to one side of the bilayer, using measurements of electrical admittance and fluorescence. The changes in currents in-phase and 90 degrees out-of- phase with respect to the applied sinusoidal voltage were used to monitor the addition of the cell membrane capacitance to that of the lipid bilayer through a fusion pore connecting the two membranes. When ethidium bromide was included in the solution of the cell-free side of the bilayer, increases in cell fluorescence accompanied tee admittance changes, independently confirming that these changes were due to formation of a fusion pore. Fusion required acidic pH on the cell- containing side and depended on temperature. For fusion to occur, the influenza hemagglutinin (HA) had to be cleaved into HA1 and HA2 subunits. The incorporation of gangliosides into the planar bilayers greatly augmented fusion. Fusion pores developed in four distinct stages after acidification: (a) a pre-pore, electrically quiescent stage; (b) a flickering stage, with 1-2 nS pores opening and closing repetitively; (c) an irreversibly opened stage, in which pore conductances varied between 2 and 100 nS and exhibited diverse kinetics; (d) a fully opened stage, initiated by an instantaneous, time- resolution limited, increase in conductance leveling at approximately 500 nS. The expansion of pores by stages has also been shown to occur during exocytosis in mast cells and fusion of HA-expressing cells and erythrocytes. We conclude that essential features of fusion pores are produced with proteins in just one of the two fusing membranes.  相似文献   

15.
Membrane fusion intermediates induced by the glycosylphosphatidylinositol-linked ectodomain of influenza hemagglutinin (GPI-HA) were investigated by rapid freeze, freeze-substitution, thin section electron microscopy, and with simultaneous recordings of whole-cell admittance and fluorescence. Upon triggering, the previously separated membranes developed numerous hourglass shaped points of membrane contact (∼10–130 nm waist) when viewed by electron microscopy. Stereo pairs showed close membrane contact at peaks of complementary protrusions, arising from each membrane. With HA, there were fewer contacts, but wide fusion pores. Physiological measurements showed fast lipid dye mixing between cells after acidification, and either fusion pore formation or the lack thereof (true hemifusion). For the earliest pores, a similar conductance distribution and frequency of flickering pores were detected for both HA and GPI-HA. For GPI-HA, lipid mixing was detected prior to, during, or after pore opening, whereas for HA, lipid mixing is seen only after pore opening. Our findings are consistent with a pathway wherein conformational changes in the ectodomain of HA pull membranes towards each other to form a contact site, then hemifusion and pore formation initiate in a small percentage of these contact sites. Finally, the transmembrane domain of HA is needed to complete membrane fusion for macromolecular content mixing.  相似文献   

16.
The N-terminal domain of the influenza hemagglutinin (HA) is the only portion of the molecule that inserts deeply into membranes of infected cells to mediate the viral and the host cell membrane fusion. This domain constitutes an autonomous folding unit in the membrane, causes hemolysis of red blood cells and catalyzes lipid exchange between juxtaposed membranes in a pH-dependent manner. Combining NMR structures determined at pHs 7.4 and 5 with EPR distance constraints, we have deduced the structures of the N-terminal domain of HA in the lipid bilayer. At both pHs, the domain is a kinked, predominantly helical amphipathic structure. At the fusogenic pH 5, however, the domain has a sharper bend, an additional 3(10)-helix and a twist, resulting in the repositioning of Glu 15 and Asp 19 relative to that at the nonfusogenic pH 7.4. Rotation of these charged residues out of the membrane plane creates a hydrophobic pocket that allows a deeper insertion of the fusion domain into the core of the lipid bilayer. Such an insertion mode could perturb lipid packing and facilitate lipid mixing between juxtaposed membranes.  相似文献   

17.
The mechanism of bilayer unification in biological fusion is unclear. We reversibly arrested hemagglutinin (HA)-mediated cell–cell fusion right before fusion pore opening. A low-pH conformation of HA was required to form this intermediate and to ensure fusion beyond it. We present evidence indicating that outer monolayers of the fusing membranes were merged and continuous in this intermediate, but HA restricted lipid mixing. Depending on the surface density of HA and the membrane lipid composition, this restricted hemifusion intermediate either transformed into a fusion pore or expanded into an unrestricted hemifusion, without pores but with unrestricted lipid mixing. Our results suggest that restriction of lipid flux by a ring of activated HA is necessary for successful fusion, during which a lipidic fusion pore develops in a local and transient hemifusion diaphragm.  相似文献   

18.
Comparative analysis of carbohydrate chains variations in influenza virus A/Leningrad/385/80 (H3N2) hemagglutinin (HA) and its heavy (HA1) and light (HA2) chains has been carried out. The carbohydrate chains of these three glycoproteins were eliminated by reductive cleavage of N-glucosaminidic linkages under LiBH4 - tert-BuOH treatment. Fractionation of the oligosaccharides thus obtained by means of gel chromatography and HPLC resulted in isolation of 21 individual oligosaccharides from each glycoprotein. Their monosaccharide composition revealed almost identical pattern of high-mannose as well as complex chains in HA1 and HA2 in spite of different number (6-7 in HA1 and only 1 in HA2) of glycosilated sites. The possibility of a great number of both high-mannose and complex chains attached at the same site of glycoprotein is shown.  相似文献   

19.
Viral fusion peptides are short N-terminal regions of type-1 viral fusion proteins that are critical for virus entry. Although the importance of viral fusion peptides in virus-cell membrane fusion is established, little is known about how they function. We report the effects of wild-type (WT) hemagglutinin (HA) fusion peptide and its G1S, G1V, and W14A mutants on the kinetics of poly(ethylene glycol)(PEG)-mediated fusion of small unilamellar vesicles composed of dioleoylphosphatidylcholine, dioleoylphosphatidylethanolamine, sphingomyelin, and cholesterol (molar ratio of 35:30:15:20). Time courses of lipid mixing, content mixing, and content leakage were obtained using fluorescence assays at multiple temperatures and analyzed globally using either a two-step or three-step sequential ensemble model of the fusion process to obtain the rate constant and activation thermodynamics of each step. We also monitored the influence of peptides on bilayer interfacial order, acyl chain order, bilayer free volume, and water penetration. All these data were considered in terms of a recently published mechanistic model for the thermodynamic transition states for each step of the fusion process. We propose that WT peptide catalyzes Step 1 by occupying bilayer regions vacated by acyl chains that protrude into interbilayer space to form the Step 1 transition state. It also uniquely contributes a positive intrinsic curvature to hemi-fused leaflets to eliminate Step 2 and catalyzes Step 3 by destabilizing the highly stressed edges of the hemi-fused microstructures that dominate the ensemble of the intermediate state directly preceding fusion pore formation. Similar arguments explain the catalytic and inhibitory properties of the mutant peptides and support the hypothesis that the membrane-contacting fusion peptide of HA fusion protein is key to its catalytic activity.  相似文献   

20.
Viral fusion peptides are short N-terminal regions of type-1 viral fusion proteins that are critical for virus entry. Although the importance of viral fusion peptides in virus-cell membrane fusion is established, little is known about how they function. We report the effects of wild-type (WT) hemagglutinin (HA) fusion peptide and its G1S, G1V, and W14A mutants on the kinetics of poly(ethylene glycol)(PEG)-mediated fusion of small unilamellar vesicles composed of dioleoylphosphatidylcholine, dioleoylphosphatidylethanolamine, sphingomyelin, and cholesterol (molar ratio of 35:30:15:20). Time courses of lipid mixing, content mixing, and content leakage were obtained using fluorescence assays at multiple temperatures and analyzed globally using either a two-step or three-step sequential ensemble model of the fusion process to obtain the rate constant and activation thermodynamics of each step. We also monitored the influence of peptides on bilayer interfacial order, acyl chain order, bilayer free volume, and water penetration. All these data were considered in terms of a recently published mechanistic model for the thermodynamic transition states for each step of the fusion process. We propose that WT peptide catalyzes Step 1 by occupying bilayer regions vacated by acyl chains that protrude into interbilayer space to form the Step 1 transition state. It also uniquely contributes a positive intrinsic curvature to hemi-fused leaflets to eliminate Step 2 and catalyzes Step 3 by destabilizing the highly stressed edges of the hemi-fused microstructures that dominate the ensemble of the intermediate state directly preceding fusion pore formation. Similar arguments explain the catalytic and inhibitory properties of the mutant peptides and support the hypothesis that the membrane-contacting fusion peptide of HA fusion protein is key to its catalytic activity.  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