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 218 毫秒
1.
The fusion protein of avian sarcoma and leukosis virus is likely to fold into a six-helix bundle as part of its final configuration. A peptide, R99, inhibits fusion, probably by binding into the grooves of the triple-stranded coiled coil that becomes the central core of the six-helix bundle. The stages at which the envelope protein (Env) of avian sarcoma and leukosis virus subgroup A folds into a bundle during low pH-induced fusion were determined. Effector cells expressing Env were bound to target cells expressing the cognate receptor Tva, and intermediates of fusion were created. R99 was added and the extent of fusion inhibition was used to distinguish between a prebundle state with exposed grooves and a state in which the grooves were no longer exposed. The native conformation of Env was not sensitive to R99. But adding a soluble form of Tva to effector cells conferred sensitivity. Acidic pH applied at low temperature created an intermediate state of local hemifusion. Surprisingly, R99 caused these locally hemifused membranes to separate. This indicates that the grooves of Env were still exposed, that prebundle configurations of Env stabilized hemifused states, and that binding of R99 altered the conformation of Env. In the presence of an inhibitory lipid that blocks fusion before hemifusion, applying low pH at 37 degrees C created an intermediate in which R99 was without effect. This suggests that the six-helix bundle can form before hemifusion and that subsequent conformational changes, such as formation of the trimeric hairpin, are responsible for pore formation and/or growth.  相似文献   

2.
Nucleocytoplasmic large DNA viruses are a steadily growing group of viruses that infect a wide range of hosts and are characterized by large particle dimensions and genome sizes. Understanding how they enter into the host cell and deliver their genome in the cytoplasm is therefore particularly intriguing. Here, we review the current knowledge on the entry of two of the best-characterized nucleocytoplasmic large DNA viruses: the poxvirus Vaccinia virus (VACV) and the giant virus Mimivirus. While previous studies on VACV had proposed both direct fusion at the plasma membrane and endocytosis as entry routes, more recent biochemical and morphological data argue for macropinocytosis as well. Notably, direct imaging by electron microscopy (EM) also supported the existence of parallel ways of entry for VACV. Instead, all the giant viruses studied so far only enter cells by phagocytosis as observed by EM, and we discuss the mechanisms for opening of the particle, fusion of the viral and phagosomal membranes and genome delivery via a unique portal, specific for each giant virus. VACV core uncoating, in contrast, remains a morphologically ill-defined process. We argue that correlated light and electron microscopy methods are required to study VACV entry and uncoating in a direct and systematic manner. Such EM studies should also address whether entry of single particles and viral aggregates is different and thus provide an explanation for the different modes of entry described in the literature.  相似文献   

3.
Using two-photon-induced fluorescence lifetime imaging microscopy, we corroborate an interaction (previously demonstrated by yeast two-hybrid domain analysis) of full-length vaccinia virus (VACV; an orthopoxvirus) A36 protein with the cellular microtubule motor protein kinesin. Quenching of enhanced green fluorescent protein (EGFP), fused to the C terminus of VACV A36, by monomeric red fluorescent protein (mDsRed), fused to the tetratricopeptide repeat (TPR) domain of kinesin, was observed in live chicken embryo fibroblasts infected with either modified vaccinia virus Ankara (MVA) or wild-type fowlpox virus (FWPV; an avipoxvirus), and the excited-state fluorescence lifetime of EGFP was reduced from 2.5 ± 0.1 ns to 2.1 ± 0.1 ns due to resonance energy transfer to mDsRed. FWPV does not encode an equivalent of intracellular enveloped virion surface protein A36, yet it is likely that this virus too must interact with kinesin to facilitate intracellular virion transport. To investigate possible interactions between innate FWPV proteins and kinesin, recombinant FWPVs expressing EGFP fused to the N termini of FWPV structural proteins Fpv140, Fpv168, Fpv191, and Fpv198 (equivalent to VACV H3, A4, p4c, and A34, respectively) were generated. EGFP fusions of intracellular mature virion (IMV) surface protein Fpv140 and type II membrane protein Fpv198 were quenched by mDsRed-TPR in recombinant FWPV-infected cells, indicating that these virion proteins are found within 10 nm of mDsRed-TPR. In contrast, and as expected, EGFP fusions of the IMV core protein Fpv168 did not show any quenching. Interestingly, the p4c-like protein Fpv191, which demonstrates late association with preassembled IMV, also did not show any quenching.  相似文献   

4.
An in vitro fusion assay uses fluorescence microscopy of labeled lipids to monitor single v-SNARE vesicle docking and fusion events on a planar lipid bilayer containing t-SNAREs. For vesicles and bilayer comprising phosphatidylcholine (POPC, 84-85% by mol) and phosphatidylserine (DOPS, 15% by mol), previous work demonstrated prompt, full fusion (τfus = 25 ms). Substitution of 20-60% phosphatidylethanolamine (DOPE) for phosphatidylcholine in the v-SNARE vesicle with either 0 or 20% DOPE included in the t-SNARE bilayer gives rise to hemifusion events. Labeled lipids diffuse into the planar bilayer as two temporally distinct waves, presumably hemifusion of the outer leaflet followed by inner leaflet (core) fusion. The fusion kinetics with DOPE is markedly heterogeneous. Some vesicle/docking site pairs exhibit prompt, full fusion while others exhibit hemifusion. Hemifusion events are roughly half productive (leading to subsequent core fusion within 20 s) and half dead-end. In qualitative accord with expectations from studies of protein-free vesicle-vesicle fusion, the hemifusion rate khemi is 15-20 times faster than the core fusion rate kcore, and the fraction of hemifusion events increases with increasing percentage of DOPE. This suggests similar underlying molecular pathways for protein-free and neuronal SNARE-driven fusion. Removal of phosphatidylserine from the v-SNARE vesicle has no effect on docking or fusion.  相似文献   

5.
We have performed full-field cryo X-ray microscopy in the water window photon energy range on vaccinia virus (VACV) infected cells to produce tomographic reconstructions. PtK2 cells were infected with a GFP-expressing VACV strain and frozen by plunge fast freezing. The infected cells were selected by light fluorescence microscopy of the GFP marker and subsequently imaged in the X-ray microscope under cryogenic conditions. Tomographic tilt series of X-ray images were used to yield three-dimensional reconstructions showing different cell organelles (nuclei, mitochondria, filaments), together with other structures derived from the virus infection. Among them, it was possible to detect viral factories and two types of viral particles related to different maturation steps of VACV (immature and mature particles), which were compared to images obtained by standard electron microscopy of the same type of cells. In addition, the effect of radiation damage during X-ray tomographic acquisition was analyzed. Thin sections studied by electron microscopy revealed that the morphological features of the cells do not present noticeable changes after irradiation. Our findings show that cryo X-ray nano-tomography is a powerful tool for collecting three-dimensional structural information from frozen, unfixed, unstained whole cells with sufficient resolution to detect different virus particles exhibiting distinct maturation levels.  相似文献   

6.
The transmembrane subunits of viral envelope proteins are thought to perform all of the functions required for membrane fusion during entry of enveloped viruses. However, changes in a conserved SPHQ motif near the N terminus of the receptor binding subunit of a murine leukemia virus (MLV) envelope protein block infection and induction of cell-cell fusion but not receptor binding. Here we report evidence that a histidine-to-arginine change at position 8 (H8R) in the SPHQ motif of Moloney MLV blocks infection by arresting virus-cell fusion at the hemifusion state. In cell-cell fusion assays, H8R envelope protein induced mixing of membrane outer leaflet lipids but did not lead to content mixing, a finding indicative of fusion pore formation. Kinetic studies of virus-cell fusion showed that lipid mixing of H8R virus membranes begins much later than for wild-type virus. The length of the delay in lipid mixing decreased upon addition of two second-site changes that increase H8R virus infection to 100-fold less than the wild-type virus. Finally, chlorpromazine, dibucaine, and trifluoperazine, agents that induce pores in an arrested hemifusion state, rescued infection by H8R virus to within 2.5-fold of the level of wild-type virus infection and cell-cell fusion to half that mediated by wild-type envelope protein. We interpret these results to indicate that fusion progressed to the hemifusion intermediate but fusion pore formation was inhibited. These results establish that membrane fusion of Moloney MLV occurs via a hemifusion intermediate. We also interpret these findings as evidence that histidine 8 is a key switch-point residue between the receptor-induced conformation changes that expose fusion peptide and those that lead to six-helix bundle formation.  相似文献   

7.
Influenza virus hemagglutinin (HA) has three highly conserved acylation sites close to the carboxyl terminus of the HA2 subunit, one in the transmembrane domain and two in the cytoplasmic domain. Each site is modified by palmitic acid through a thioester linkage to cysteine. To elucidate the biological significance of HA acylation, the acylation sites of HA of influenza virus strain A/USSR/77 (H1N1) were changed by site-directed mutagenesis, and the membrane fusion activity of mutant HAs lacking the acylation site(s) was examined quantitatively using transfer assays of lipid (R18) and aqueous (calcein) dyes. Lipid mixing, so-called hemifusion, activity was not affected by deacylation, whereas transfer of aqueous dye, so-called fusion pore formation, was dramatically restricted. When the fusion reaction was induced by a lower pH than the optimal one, calcein transfer with the mutant HAs was improved, but simultaneously a considerable calcein leakage into the medium was observed. From these results, we conclude that the palmitic acids on the H1 subtype HA facilitate the transition from hemifusion to fusion pore formation.  相似文献   

8.
Synaptic transmission requires the controlled release of neurotransmitter from synaptic vesicles by membrane fusion with the presynaptic plasma membrane. SNAREs are the core constituents of the protein machinery responsible for synaptic membrane fusion. The mechanism by which SNAREs drive membrane fusion is thought to involve a hemifusion intermediate, a condition in which the outer leaflets of two bilayers are combined and the inner leaflets remain intact; however, hemifusion has been observed only as an end point rather than as an intermediate. Here, we examined the kinetics of membrane fusion of liposomes mediated by recombinant neuronal SNAREs using fluorescence assays that monitor both total lipid mixing and inner leaflet mixing. Our results demonstrate that hemifusion is dominant at the early stage of the fusion reaction. Over time, hemifusion transitioned to complete fusion, showing that hemifusion is a true intermediate. We also show that hemifusion intermediates can be trapped, likely as unproductive outcomes, by modulating the surface concentration of the SNARE proteins.  相似文献   

9.
A hemagglutinin (HA) of influenza virus having a single semiconserved Gly residue within the transmembrane domain mutated to Leu (G520L) was expressed on cells; these cells were bound to red blood cells. By decreasing pH at 23 degrees C rather than 37 degrees C, an intermediate with properties expected of hemifusion just as the membranes are about to transit to full fusion was captured. As evidence: 1) increasing temperature to 37 degrees C at neutral pH allowed fusion to proceed; 2) after achieving the intermediate, the two membranes did not separate from each other after proteolytic cleavage of G520L because cells treated with proteinase K could not fuse upon temperature increase but could fuse upon the addition of chlorpromazine; and 3) at the point of the intermediate, adding exogenous lipids known to promote or inhibit the creation of hemifusion did not significantly alter the lipid dye spread that occurred upon increasing temperature, implying that at the intermediate, contacting membrane leaflets had already merged. A stable intermediate of hemifusion that could transit to fusion was also generated for wild-type HA, but pH had to be reduced at the significantly lower temperature of 4 degrees C. The fusion pores generated by G520L did not enlarge, whereas those induced by wild-type HA did. The finding that a state of transitional hemifusion can be readily obtained via a point mutation without the need for unusually low temperature supports the hypothesis that hemifusion occurs before pore formation.  相似文献   

10.
S Nir  T Stegmann  J Wilschut 《Biochemistry》1986,25(1):257-266
The kinetics and extent of low pH induced fusion between influenza virus and large unilamellar cardiolipin liposomes were investigated with an assay for lipid mixing based on fluorescence resonance energy transfer. The results were analyzed in terms of a mass action kinetic model, which views the overall fusion reaction as a sequence of a second-order process of virus-liposome adhesion or aggregation followed by the first-order fusion reaction itself. The fluorescence development during the course of the fusion process was calculated by numerical integration, employing separate rate constants for the initial aggregation step and for the subsequent fusion reaction. Analytical solutions were found for several limiting cases. Deaggregation of virus--liposome aggregates was explicitly taken into account but was found to be a minor effect under the conditions studied. The calculations gave good simulations and predictions for the kinetics and extent of fusion at different virus/liposome concentrations and ratios. At pH 5.0 and 37 degrees C, very high rate constants for aggregation and fusion were obtained, and essentially all of the virus particles were involved in the fusion process. Experiments at different virus/liposome ratios showed that fusion products may consist of a single virus particle and several liposomes but not of a single liposome and several virus particles. At pH 6.0, the rate constant for aggregation was the same as at pH 5.0, but the rate constant of fusion was about 5-fold lower, and only 25-40% of the virus particles were capable of fusing with the liposomes. The analytical procedure presented enables elucidation of the crucial role of the composition of target membrane vesicles in the initial adhesion and subsequent fusion of the virus at various pH values.  相似文献   

11.
Vacuolar carboxypeptidase Y (CPY) from Aspergillus nidulans was used to construct a CPY-EGFP fusion protein and expressed in A. oryzae to study vacuolar morphology and functions in A. oryzae. While the fluorescence of EGFP was barely detectable in A. oryzae expressing CPY-EGFP grown under normal conditions at pH 5-6, the increase in pH of the growth medium towards alkalinity restored the fluorescence. In accordance with such an observation, the fluorescence of CPY-EGFP fusion protein in cell extract decreased in acidic pH condition, concomitant with lowered content of EGFP detected in A. oryzae grown under acidic pH conditions. The pH sensitivity of EGFP fluorescence and enhanced degradation of proteins in vacuoles under acidic pH conditions are thus proposed to result in the reduction of fluorescence in A. oryzae. Further, visualization of vacuoles revealed the presence of peculiar ring- or tube-like structures as distinct from normal spherical-shaped vacuoles.  相似文献   

12.
Epstein-Barr virus (EBV) infects two cell types, B lymphocytes and epithelial cells. Electron microscopic studies have shown that the virus fuses with the lymphoblastoid cell line Raji but is endocytosed into thin-walled non-clathrin-coated vesicles in normal B cells before fusion takes place. To compare early interactions of EBV with epithelial cells and B cells, a fluorescence dequenching assay of fusion was employed, using virus labeled either with the pH-insensitive probe octadecyl rhodamine B chloride (R18) or with 5(N-octadecanoyl) aminofluorescein (AF), which loses emission intensity at a pH below 7.4. Fusion of virus labeled with R18 could be monitored with B cells, Raji cells, and epithelial cells. Lowering the extracellular pH or pretreatment of cells with ammonium chloride or methylamine had no effect on these measurements. In contrast, fusion of virus labeled with AF could be measured with Raji cells and epithelial cells, but not with normal B cells unless cells were previously treated with ammonium chloride. Fusion of virus with normal B cells was inhibited with chlorpromazine, chloroquine, and sodium azide, but none of these reagents had any effect on fusion with Raji or epithelial cells. These results suggest that entry of EBV into nonpolarized suspensions of epithelial cells occurs by fusion at the cell surface, that EBV may be incapable of fusing with normal B cells unless it has first been endocytosed, and that pH appears to be irrelevant to either event. A combination of the two probes, R18 and AF, may have general use for determining the sites of entry of enveloped viruses that fuse in a pH-independent manner.  相似文献   

13.
The decrease of the intrinsic tryptophan fluorescence intensity of purified influenza (X31 strain) hemagglutinin (HA) was used to monitor the low pH-induced conformational change of this protein. The kinetics of the fluorescence decrease depended strongly on the pH. At pH optimal for fusion, the change in tryptophan fluorescence was fast and could be fitted to a monoexponential function. We measured a rate constant of 5.78 s-1 (t1/2 = 120 ms) at pH 4.9 using rapid stopped-flow mixing. Under suboptimal conditions (higher pH), the rate constant was decreased by an order of magnitude. In addition, a slow component appeared and the fluorescence decrease followed a sum of two exponentials. The kinetics of conformational changes were compared with those of the fusion of influenza virus with red blood cell membranes as assessed by the R18-dequenching assay. At optimal pH the HA conformational change was not rate-limiting for the fusion process. However, at sub-optimal pH, the slow transition to the fusogenic conformational of HA resulted in slower kinetics and decreased extent of fusion.  相似文献   

14.
Membrane fusion mediated by influenza virus hemagglutinin (HA) yields different phenotypes depending on the surface density of activated HAs. A key question is whether different phenotypes arise from different fusion machines or whether different numbers of identical fusion machines yield different probabilistic outcomes. If fusion were simply a less probable event than hemifusion, requiring a larger number of identical fusion machines to occur first, then two predictions can be made. First, fusion should have a shorter average delay time than hemifusion, since there are more machines. Second, fusion should have a longer execution time of lipid mixing after it begins than hemifusion, since the full event cannot be faster than the partial event. Using a new automated video microscopy technique, we simultaneously monitored many HA-expressing cells fusing with erythrocytes and identified individual cell pairs with either full or only partial redistribution of fluorescent lipids. The full lipid mixing phenotype also showed contents mixing, i.e., fusion. Kinetic screening of the digitized fluorescence data showed that the execution of lipid mixing after the onset is faster for fusion than hemifusion. We found no correlation between the delay times before the onset of lipid mixing and the final fusion phenotype. We also found that the execution time for fusion was faster than that for hemifusion. Thus, we provide the first experimental evidence for fusion and hemifusion arising from different machines.  相似文献   

15.
It has been reported that low-pH-induced fusion of influenza virus with liposomes results in rapid and extensive release of both low- and high-molecular-weight substances from the liposomes [Günther-Ausborn et al., J. Biol. Chem. 270 (1995) 29279-29285; Shangguan et al., Biochemistry 35 (1996) 4956-4965]. Here, we demonstrate retention of encapsulated water-soluble compounds during fusion of Semliki Forest virus (SFV) or Sindbis virus with liposomes at low pH. Under conditions allowing complete fusion of the liposomes, a limited fluorescence dequenching of liposome-encapsulated calcein was observed, particularly for SFV. Also, radioactively labeled inulin or sucrose were largely retained. Freezing and thawing of the viruses in the absence of sucrose resulted in an enhanced leakiness of fusion. These results support the notion that the alphavirus fusion event per se is non-leaky and may well involve a discrete hemifusion intermediate.  相似文献   

16.
Fusion between influenza virus and target membranes is mediated by the viral glycoprotein hemagglutinin (HA). Replacement of the transmembrane domain of HA with a glycosylphosphatidylinositol (GPI) membrane anchor allows lipid mixing but not the establishment of cytoplasmic continuity. This observation led to the proposal that the fusion mechanism passes through an intermediate stage corresponding to hemifusion between outer monolayers. We have used confocal fluorescence microscopy to study the movement of probes for specific bilayer leaflets of erythrocytes fusing with HA-expressing cells. N-Rh-PE and NBD-PC were used for specific labeling of the outer and inner membrane leaflet, respectively. In the case of GPI-HA-induced fusion, different behaviors of lipid transfer were observed, which include 1) exclusive movement of N-Rh-PE (hemifusion), 2) preferential movement of N-Rh-PE relative to NBD-PC, and 3) equal movement of both lipid analogs. The relative population of these intermediate states was dependent on the time after application of a low pH trigger for fusion. At early time points, hemifusion was more common and full redistribution of both bilayers was rare, whereas later full redistribution of both probes was frequently observed. In contrast to wild-type HA, the latter was not accompanied by mixing of the cytoplasmic marker Lucifer Yellow. We conclude that 1) the GPI-HA-mediated hemifusion intermediate is meta-stable and 2) expansion of an aqueous fusion pore requires the transmembrane and/or cytoplasmic domain of HA.  相似文献   

17.
The characteristics of fusion of respiratory syncytial virus (RSV) with HEp-2 cells were studied by the R18 fluorescence dequenching assay of membrane fusion. A gradual increase in fluorescence intensity indicative of virion-cell fusion was observed when R18-labeled RSV was incubated with HEp-2 cells. Approximately 35% dequenching of the probe fluorescence was observed in 1 h at 37 degrees C. Fusion showed a temperature dependence, with significant dequenching occurring above 18 degrees C. The dequenching was also dependent on the relative concentration of target membrane. Thus, increasing the concentration of target membrane resulted in increased levels of dequenching. In addition, viral glycoproteins were shown to be involved in this interaction, since dequenching was significantly reduced by pretreatment of labeled virus at 70 degrees C for 5 min or by trypsinization of R18-labeled virions prior to incubation with HEp-2 cells at 37 degrees C. The fusion of RSV with HEp-2 cells was unaffected over a pH range of 5.5 to 8.5, with some increase seen at lower pH values. Treatment of HEp-2 cells with ammonium chloride (20 and 10 mM), a lysosomotropic agent, during early stages of infection did not inhibit syncytium formation or progeny virion production by RSV. At the same concentrations of ammonium chloride, the production of vesicular stomatitis virus was reduced approximately 4 log10 units. These results suggest that fusion of the virus with the cell surface plasma membrane is the principal route of entry.  相似文献   

18.
We have studied fusion between membranes of vesicular stomatitis virus (VSV) and Vero cells using an assay for lipid mixing based on the relief of self-quenching of octadecylrhodamine (R18) fluorescence. We could identify the two pathways of fusion by the kinetics of R18 dequenching, effects of inhibitors, temperature dependence, and dependence on osmotic pressure. Fusion at the plasma membrane began immediately after lowering the pH below 6 and showed an approximately exponential time course, whereas fusion via the endocytic pathway (pH 7.4) became apparent after a time delay of about 2 min. Fusion via the endocytic pathway was attenuated by treating cells with metabolic inhibitors and agents that raise the pH of the endocytic vesicle. A 10-fold excess of unlabeled virus arrested R18VSV entry via the endocytic pathway, whereas R18 dequenching below pH 6 (fusion at the plasma membrane) was not affected by the presence of unlabeled virus. The temperature dependence for fusion at pH 7.4 (in the endosome) was much steeper than that for fusion at pH 5.9 (with the plasma membrane). Fusion via the endocytic pathway was attenuated at hypo-osmotic pressures, whereas fusion at the plasma membrane was not affected by this treatment. The pH profile of Vero-VSV fusion at the plasma membrane, as measured by the dequenching method, paralleled that observed for VSV-induced cell-cell fusion. Fusion was blocked by adding neutralizing antibody to the Vero-VSV complexes. Activation of the fusion process by lowering the pH was reversible, in that the rate of fusion was arrested by raising the pH back to 7.4. The observation that pH-dependent fusion occurred at similar rates with fragments and with intact cells indicates that pH, voltage, or osmotic gradients are not required for viral fusion.  相似文献   

19.
S Nir  K Klappe  D Hoekstra 《Biochemistry》1986,25(8):2155-2161
The kinetics and extent of fusion between Sendai virus and erythrocyte ghosts were investigated with an assay for lipid mixing based on the relief of self-quenching of fluorescence. The results were analyzed in terms of a mass action kinetic model, which views the overall fusion reaction as a sequence of a second-order process of virus-cell adhesion followed by the first-order fusion reaction itself. The fluorescence development during the course of the fusion process was calculated by numerical integration, employing separate rate constants for the adhesion step and for the subsequent fusion reaction. Dissociation of virus particles from the cells was found to be of minor importance when fusion was initiated by mixing the particles at 37 degrees C. However, besides the initiation of fusion, extensive dissociation does occur after a preincubation of a concentrated suspension of particles at 4 degrees C followed by a transfer of the sample to 37 degrees C. The conclusion drawn from the levels of fluorescence increase obtained after 20 h of incubation is that in principle most virus particles can fuse with the ghosts at 37 degrees C and pH 7.4. However, the number of Sendai virus particles that actually fuse with a single ghost is limited to 100-200, despite the fact more than 1000 particles can bind to one cell. This finding may imply that 100-200 specific fusion sites for Sendai virus exist on the erythrocyte membrane. A simple equation can yield predictions for the final levels of fluorescence for a wide range of ratios of virus particles to ghosts.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
S Nir  K Klappe  D Hoekstra 《Biochemistry》1986,25(25):8261-8266
The kinetics and extent of fusion between Sendai virus particles and liposomes were investigated with an assay for lipid mixing based on the relief of self-quenching of fluorescence. The measurements, which were carried out at pH 7.4 and 5.0, included liposomes of three compositions, cardiolipin (CL), CL/dioleoylphosphatidylcholine (CL/DOPC 1:1), and phosphatidylserine (PS). Liposomal lipid concentrations varied from 2.5 to 50 microM. In addition, the effect of low concentrations of the dehydrating agent poly(ethylene glycol) (PEG) on fusion between the virus and the liposomes at pH 7.4 was studied. The results were analyzed in terms of a mass action kinetic model which views the overall fusion reaction as a sequence of a second-order process of virus-liposome adhesion or aggregation, followed by the first-order fusion reaction itself. The fusion products were shown to consist of a single virus particle and several liposomes. Analytical solutions were found for the final extent of fusion and increase in fluorescence intensity following the fusion of fluorescently labeled virus particles with liposomes. The final extents of fluorescence intensity were explained by assuming an essentially irreversible binding of liposomes to inactive virus particles. The percents of active virus particles and the rate constants of fusion and aggregation were larger at pH 5 than at pH 7.4, increased when PEG was included in the medium, and varied with liposomal lipid composition according to the sequence CL greater than CL/DOPC greater than PS.(ABSTRACT TRUNCATED AT 250 WORDS)  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