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.
Fusion of phagolysosomes (PLs) has been demonstrated to occur in vitro. Two separate cell homogenates of the ameba Acanthamoeba sp. (Neff) were prepared, each rich in PLs labeled with distinctive particulate markers. Portions of each were incubated together in vitro and fusion occurred as evidenced by the appearance of PLs containing both types of markers. Fusion was confirmed by electron microscopy, including serial sectioning. The membranes of fused vacuoles excluded the dye eosin Y. Surviving cells in the homogenates were not responsible for the observed fusion. Fusion was obtained using either synthetic markers (polystyrene and polyvinyltoluene latex) or biological markers (autoclaved yeast cells and glutaraldehyde-fixed goat red blood cells), or a combination of both. The specificity of PL fusion in vivo appeared to be maintained in vitro. As determined by light and electron microscopy, the fusion reaction was dependent on time and temperature, and on the initial presence of membrane around both marker particles. A minimum of 10% of the vacuoles fused by 10 min of incubation at 30 degrees C, and no rupture of the vacuoles was detected during this time. After 10 min of incubation, vacuole rupture began and fusion ceased. At a constant initial vacuole concentration, the extent of PL fusion in vitro was quantitatively reproducible. This appears to be a promising system for further investigation of membrane fusion in the lysosomal system.  相似文献   

2.
A quantitative real-time assay of cell-free endosomal vesicle fusion was developed and applied to study fusion mechanisms in endosomes from baby hamster kidney (BHK-21) cells. The assay is based on an irreversible approximately 10-fold increase in BODIPY-avidin fluorescence on binding of biotinylated conjugates. BODIPY-avidin and biotin-dextran were internalized for 10 min at 37 degrees C into separate populations of BHK-21 cells, and endosome fractions were prepared. Postnuclear supernatant fractions underwent ATP- and temperature-dependent fusion, as measured in a sensitive custom-built microfluorimeter by the continuous increase in BODIPY-avidin fluorescence. Fusion processes of efficiency > 2.5% could be detected with 200-ms time resolution in sample volumes of 50 microL containing endosomes derived from approximately 4 x 10(4) cells. The fusion time course consisted of a distinct lag phase (up to 10 min) in which little fusion occurred, followed by an approximately exponential rise (t 1/2 10-30 min; fusion efficiency approximately 15%). The lag phase was reduced by preincubation of separate endosome fractions with ATP at 37 degrees C and by coincubation of endosomes at 22 degrees C before the assay, suggesting a rate-limiting step involving binding of a soluble protein to the endosome membrane. Endosome fusion was strongly inhibited by GTP gamma S, N-ethylmaleimide, and AIF4-. Endosome fusion was not affected by phorbol myristate acetate but was significantly inhibited by cAMP and bovine brain calmodulin. The results establish a sensitive real-time fluorescence assay to quantify the kinetics and extent of endosome fusion in a cell-free system and demonstrate regulation of early endosome fusion by cytosolic second messengers.  相似文献   

3.
The role of the target membrane structure in fusion with Sendai virus   总被引:3,自引:0,他引:3  
Fusion between membranes of Sendai virus and liposomes or human erythrocytes ghosts was studied using an assay for lipid mixing based on the relief of self-quenching of octadecylrhodamine (R18) fluorescence. We considered only viral fusion that reflects the biological activity of the viral spike glycoproteins. The liposomes were made of phosphatidylcholine, and the effects of including cholesterol, the sialoglycolipid GD1a, and/or the sialoglycoprotein glycophorin as receptors were tested. Binding of Sendai virus to those liposomes at 37 degrees C was very weak. Fusion with the erythrocyte membranes occurred at a 30-fold faster rate than with the liposomes. Experiments with biological and liposomal targets of different size indicated that size did not account for differences in fusion efficiency.  相似文献   

4.
Fusion between membranes of 3T3 fibroblasts expressing hemagglutinin (HA) from the Japan strain of influenza virus and human red blood cells (RBC) was measured using an assay for lipid mixing based on the relief of self-quenching (dequenching) of fluorescence of the lipid probe octadecylrhodamine (R18). The probe was incorporated into the membrane of intact RBC at self-quenching concentrations, and the RBCs were bound to the 3T3 cells. Fusion, which allowed movement of R18 into 3T3 cell membranes, was monitored by spectrofluorometry as an increase in fluorescence. Upon lowering the pH below 5.4, the fluorescence increased after a delay of about 30 s at 37 degrees C, and leveled off within 2 min. In control experiments where R18 RBCs bound to 3T3 cells expressing the uncleaved precursor hemagglutinin (HA0) were incubated at 37 degrees C and low pH, no fluorescence increase was observed. This indicated that the R18 dequenching occurred as a result of HA-induced fusion of plasma membranes. Fusion showed a very steep pH dependence with a threshold at pH 5.4 and a maximum at pH 5.0, similar to HA-induced fusion seen previously using cell biological techniques. The fusion rate increased and the delay for the onset of fusion decreased as the temperature was raised above 20 degrees C. Low pH activation of the fusion process at 37 degrees C could be partially arrested by raising the pH after 2-10 s, but not after 15 s, indicating that the irreversible pH-activated conformational change of HA necessary for fusion was complete within about 15 s. Analysis of the data indicates that the pH-induced membrane fusion activity of HA is a highly cooperative event.  相似文献   

5.
Fusion of chromaffin granule ghosts was induced by synexin at pH 6, 37 degrees C, in the presence of 10(-7) M Ca2+. To study the kinetics and extent of this fusion process we employed two assays that monitored continuously mixing of aqueous contents or membrane mixing by fluorescence intensity increases. In both assays chromaffin granule ghosts were either labeled on the membrane or in the included aqueous phase. The ratios of blank to labeled chromaffin granule ghosts were varied from 1 to 10. 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 aggregation followed by a first-order fusion reaction. The model calculations gave fare simulations and predictions of the experimental results. The rate constants describing membrane mixing are more than 2-fold larger than those for volume mixing. The analysis also indicated that the initial aggregation and fusion processes, up to dimer formation, were extremely fast. The rate constant of aggregation was close to the limit in diffusion-controlled processes, whereas the fusion rate constant was about the same as found in fastest virus-liposome fusion events at pH 5. A small increase in volume was found to accompany the fusion between chromaffin granule ghosts. Using ratios of synexin to chromaffin granule ghost protein of 0.13, 0.41 and 1.15 indicated that the overall fusion rate was larger for the intermediate (0.41) case. The analysis showed that the main activity of synexin was an enhancement of the rate of aggregation. At intermediate or excessive synexin concentrations it, respectively, promoted moderately, or inhibited the actual fusion step.  相似文献   

6.
Monovalent cation-induced fusion of acidic phospholipid vesicles   总被引:1,自引:0,他引:1  
Fusion of small unilamellar vesicles (SUV) consisting of dimyristoylphosphatidylcholine (DMPC), dimyristoylphosphatidylglycerol (DMPG) and phosphatidylglycerol (PG) from egg yolk, dipalmitoylphosphatidylserine (DPPS) and phosphatidylserine (PS) from bovine brain was studied as a function of monovalent cation concentration. Fusion was detected by measuring the changes in the excimer to monomer fluorescence intensity ratio (IE/M) of pyrene-labeled phospholipid analogues upon fusion of the pyrene-labeled and unlabeled vesicles. No fusion was observed from vesicles consisting of DMPC, PS from bovine brain or PG from egg yolk upon addition of NaCl (up to 1 M). However, considerable fusion was evident for vesicles consisting of DMPG or DPPS upon addition of monovalent cations (300 mM to 1 M). Fusion kinetics were fast reaching a plateau after 5 min of addition of cations. The order of efficiency of different monovalent cations to induce the fusion of DMPG vesicles as judged by the changes of the IE/M ratio was Li+ greater than Na+ greater than K+ greater than Cs+. DSC-scan of sonicated DMPG vesicles showed, in the absence of salt, a phase transition at 19.2 degrees C with enthalpy of 1.1 kcal.mol-1. After incubation in the presence of 600 mM NaCl the DSC scan showed a narrow phase transition at 24.1 degrees C with enthalpy of 6.9 kcal.mol-1 and a pronounced pretransition, both supporting that the fusion of the vesicles had occurred in the presence of NaCl. The results indicate that sonicated vesicles consisting of acidic phospholipids with fully saturated fatty acids fuse in the presence of monovalent cations, whereas those containing unsaturated fatty acids do not.  相似文献   

7.
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   

8.
Fusion proteins of many viruses, including HIV-1 envelope protein (Env), fold into six-helix bundle structures. Fusion between individual Env-expressing cells and target cells was studied by fluorescence microscopy, and a temperature jump technique, to determine whether folding of Env into a bundle is complete by the time fusion pores have formed. Lowering temperature to 4 degrees C immediately after a pore opened halted pore growth, which quickly resumed when temperature was raised again. HIV gp41-derived peptides that inhibit bundle formation (C34 or N36) caused the cold-arrested pore to quickly and irreversibly close, demonstrating that bundle formation is not complete by the time a pore has formed. In contrast, lowering the temperature to an intermediate value also halted pore growth, but the pore was not closed by the bundle-inhibiting peptides, and it enlarged when temperature was again elevated. This latter result shows that bundle formation is definitely required for the fusion process, but surprisingly, some (if not all) bundle formation occurs after a pore has formed. It is concluded that an essential function of the bundle is to stabilize the pore against collapse and ensure its growth.  相似文献   

9.
Purified G-protein from vesicular stomatitis virus was reconstituted into egg phosphatidylcholine vesicles by detergent dialysis of octyl glucoside. A homogeneous population of reconstituted vesicles could be obtained, provided the protein to lipid ratio was high (about 0.3 mol % protein) and the detergent removal was slow. The reconstituted vesicles were assayed for fusion activity using electron microscopy and fluorescence energy transfer. The fusion activity mediated by the viral envelope protein was dependent upon pH, temperature, and target membrane lipid composition. Incubation of reconstituted vesicles at low pH with small unilamellar vesicles containing negatively charged lipids resulted in the appearance of large cochleate structures, as shown by electron microscopy using negative stain. This process did not cause leakage of a vesicle-encapsulated aqueous marker. The rate of fusion was pH-dependent with a pK of about 4 and the apparent energy of activation for the fusion was 16 +/- 1 kcal/mol. G-protein-mediated fusion showed a large preference for target membranes which contain phosphatidylserine or phosphatidic acid. Inclusion of 36% cholesterol in any of the lipid compositions had no effect on the rate of fusion. These reconstituted vesicles provide a system to study the mechanism of pH-dependent fusion induced by a viral spike protein.  相似文献   

10.
Fusion between cells expressing envelope protein (Env) of Moloney murine leukemia virus and target cells were studied by use of video fluorescence microscopy and electrical capacitance measurements. When the full-length 632-amino-acid residue Env was expressed, fusion did not occur at all for 3T3 cells as target and only somewhat for XC6 cells. Expression of Env 616*-a construct of Env with the last 16 amino acid residues (617 to 632; the R peptide) deleted from its C terminus to match the proteolytically cleaved Env produced during viral budding-resulted in high levels of fusion. Env 601*, lacking the entire cytoplasmic tail (CT) (identified by hydrophobicity), also led to fusion. Truncation of an additional six residues (Env 595*) abolished fusion. The kinetics of forming fusion pores did not depend on whether cells were first prebound at 4 degrees C and the time until fusion measured after the temperature was raised to 37 degrees C or whether cells were first brought into contact at 37 degrees C and the time until fusion immediately measured. This similarity in kinetics indicates that binding is accomplished quickly compared to subsequent steps in fusion. The fusion pores formed by Env 601* and Env 616* had the same initial size and enlarged in similar manners. Thus, once the R peptide is removed, the CT is not needed for fusion and does not affect formed pores. However, residues 595 to 601 are required for fusion. It is suggested here that the ectodomain and membrane-spanning domain of Env are directly responsible for fusion and that the R peptide affects their configurations at some point during the fusion process, thereby indirectly controlling fusion.  相似文献   

11.
A porous polymer membrane of nitrocellulose or tetrafluoroethylene (TFE) was employed for fusion of Saccharomyces cerevisiae (AH22 and D13-1A) protoplasts. Protoplasts were adsorbed on the membrane with slight suction. Some part of the protoplasts was trapped in pores of the membrane as observed by electron microscopy. The membrane retaining protoplasts was placed on a selective medium. Several colonies appeared on the medium after 5-7 days incubation at 30 degrees C. The fusion of the two strains was ascertained by DNA content and genetic markers. Fusion frequency was 1.2 X 10(-6) in the case of the TFE membrane.  相似文献   

12.
Roche S  Gaudin Y 《Journal of virology》2004,78(16):8746-8752
Fusion of rabies virus with membranes is triggered at a low pH and is mediated by a viral glycoprotein (G). Fusion of rabies virus with liposomes was monitored by using a lipid mixing assay based on fluorescence resonance energy transfer. Fusion was detected below pH 6.4, and its extent increased with H(+) concentrations to be maximal around pH 6.15. The origin of the partial fusion activity of rabies virus under suboptimal pH conditions (i.e., between pH 6.15 and 6.4) was investigated. We demonstrate unambiguously that fusion at a suboptimal pH is distinct from the phenomenon of low-pH-induced inactivation and that it is not due to heterogeneity of the virus population. We also show that viruses that do not fuse under suboptimal pH conditions are indeed bound to the target liposomes and that the fusion complexes they have formed are blocked at an early stage of the fusion pathway. Our conclusion is that along the fusion reaction, different kinds of fusion machines with different pH thresholds for fusion can be formed. Possible explanations of this difference of pH sensitivity are discussed.  相似文献   

13.
The kinetics of fusion of Sendai virus (Z strain) with the human promyelocytic leukemia cell line HL-60, and the human T lymphocytic leukemia cell line CEM was investigated. Fusion was monitored by fluorescence dequenching of octadecylrhodamine (R-18) incorporated in the viral membrane. For one virus isolate (Z/G), the overall rate of fusion (at 37 degrees C) increased as the pH was lowered, reaching a maximum at about pH 5, the lowest pH tested. For another isolate (Z/SF) the rate and extent of fusion were lower at pH 5 than at neutral pH. Lowering the pH from neutral to 5 after several minutes of incubation of either isolate with HL-60 cells resulted in an enhanced rate of fluorescence dequenching. Nevertheless, experiments utilizing NH4Cl indicated that fusion of the virus with cells was not enhanced by the mildly acidic pH of the endosome lumen. Analysis of the kinetics of fusion by means of a mass action model resulted in good simulation and predictions for the time-course of fusion. For the isolate which showed maximal fusogenic activity at pH 5, the rate constant of fusion (approx. 0.1 s-1) at neutral pH was in the range found previously for virus-liposome fusion, whereas the rate constant of adhesion was close to the upper limit for diffusion-controlled processes (1.4.10(10) M-1 s-1). However, for the other isolate (Z/SF) the rate constant of fusion at neutral pH was very small (less than 0.01 s-1), whereas the rate constant of adhesion was larger (greater than or equal to 2.10(10) M-1 s-1). Lowering the temperature decreased the fusion rate. Experiments involving competition with excess unlabeled virions indicated that not all binding sites for Sendai virus on HL-60 cells are fusion sites. The virus fusion activity towards HL-60 cells at neutral pH was not altered significantly by pre-incubation of the virus at pH 5 or 9, in contrast to earlier observations with liposomes and erythrocyte ghosts, or results based on erythrocyte hemolysis or cell-cell fusion.  相似文献   

14.
R A Parente  B R Lentz 《Biochemistry》1986,25(21):6678-6688
Poly(ethylene glycol) (PEG) of average molecular weight 8000 was used to mediate the fusion of large unilamellar vesicles composed of dipalmitoylphosphatidylcholine. Fusion was monitored by fluorescence assays of lipid mixing and aqueous contents mixing. The extent of lipid mixing, as monitored by DPHpPC fluorescence lifetime, indicated that large unilamellar vesicles underwent a single fusion cycle when incubated with PEG and subsequently diluted into buffer. The ANTS/DPX assays for contents mixing and leakage indicated that, while addition and dilution of PEG were accompanied by extensive contents leakage, this occurred on a much different time scale as compared to contents mixing. Both the lipid-mixing and contents-mixing assays gave comparable estimates for the number of rounds of fusion that occurred in a given time following PEG addition, although the contents-mixing assay always yielded an estimate 10-15% larger than the lipid-mixing assay. These assays were used to evaluate several factors purported to influence PEG-induced fusion. First, the initial rate of fusion was found to be dependent on PEG concentration in the range of 0-35 wt %, while the extent of fusion was not. In addition, a substantial rate enhancement occurred when vesicles were incubated with greater than 26% PEG. Second, the creation of an osmotic gradient upon dilution of vesicle-PEG mixtures was shown to have no effect on either the extent or the initial rate of fusion. Consistent with this observation, both contents and lipid mixing were found to occur prior to and independent of the dilution of the PEG-vesicle suspension. Third, impurities, either present in our commercially available PEG or added to vesicle-PEG mixtures, also had no effect on the rate or extent of fusion. Fourth, another dehydrating polymer, dextran (average mol wt 9000), was capable of promoting fusion, though at a much lower rate than PEG. These results suggest that even partial bilayer dehydration accompanied by vesicle collapse and close interbilayer contact may be sufficient to induce vesicle fusion.  相似文献   

15.
Membrane fusion is an obligatory step in many vital cellular processes. The well-established enrichment of bilayer-destabilizing lipids in membranes of poikilotherms subjected to growth at low temperatures leads to the prediction that such membranes will possess a greater propensity to undergo fusion. This hypothesis was explicitly tested in the present study by determining the kinetics of fusion between small unilamellar vesicles (SUVs) prepared from endoplasmic reticulum (ER) membranes of thermally-acclimated (to 5 and 20 degrees C) rainbow trout (Oncorhynchus mykiss) liver and bovine brain phosphatidylserine (BBPS). At temperatures above 10 degrees C, ER vesicles from 5 degrees C-acclimated trout, fused more rapidly and to a greater extent with BBPS vesicles (by average factors of 1.25- and 1.45-fold, respectively) than ER vesicles of 20 degrees C-acclimated trout. At temperatures >35 degrees C, apparent fusion rates declined while the extent of fusion increased in both acclimation groups. Fusion kinetics were found to be well correlated with and limited by the physical properties and phase state of the BBPS vesicles. These results indicate that dynamic attributes of biological membranes, such as the propensity to undergo fusion, are of potential regulatory significance and are partially conserved when growth or environmental temperature changes.  相似文献   

16.
We have studied the fusion activity of Sendai virus, a lipid-enveloped paramyxovirus, towards a line of adherent cells designated PC-12. Fusion was monitored by the dequenching of octadecyl-rhodamine, a fluorescent non-exchangeable probe. The results were analysed with a mass action kinetic model which could explain and predict the kinetics of virus-cell fusion. When the temperature was lowered from 37 degrees C to 25 degrees C, a sharp inhibition of the fusion process was observed, probably reflecting a constraint in the movement of viral glycoproteins at low temperatures. The rate constants of adhesion and fusion were reduced 3.5-fold and 7-fold, respectively, as the temperature was lowered from 37 degrees C to 25 degrees C. The fusion process seemed essentially pH-independent, unlike the case of liposomes and erythrocyte ghosts. Preincubation of the virus in the absence of target cell membranes at neutral and alkaline pH (37 degrees C, 30 min) did not affect the fusion process. However, a similar preincubation of the virus at pH = 5.0 resulted in marked, though slow, inhibition in fusion with the fusion rate constant being reduced 8-fold. Viral preincubation for 5 min in the same acidic conditions yielded a mild inhibition of fusogenic activity, while preincubation in the cold (4 degrees C, 30 min) did not alter viral fusion activity. These acid-induced inhibitory effects could not be fully reversed by further viral preincubation at pH = 7.4 (37 degrees C, 30 min). Changes in internal pH as well as endocytic activity of PC-12 cells had small effect on the fusion process, thus indicating that Sendai virus fuses primarily with the plasma membranes.  相似文献   

17.
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   

18.
A single molecule fluorescence assay is presented for studying the mechanism of soluble N-ethyl maleimide sensitive-factor attachment protein receptors (SNAREs)-mediated liposome fusion to supported lipid bilayers. The three neuronal SNAREs syntaxin-1A, synaptobrevin-II (VAMP), and SNAP-25A were expressed separately, and various dye-labeled combinations of the SNAREs were tested for their ability to dock liposomes and induce fusion. Syntaxin and synaptobrevin in opposing membranes were both necessary and sufficient to dock liposomes to supported bilayers and to induce thermally activated fusion. As little as one SNARE interaction was sufficient for liposome docking. Fusion of docked liposomes with the supported bilayer was monitored by the dequenching of soluble fluorophores entrapped within the liposomes. Fusion was stimulated by illumination with laser light, and the fusion probability was enhanced by raising the ambient temperature from 22 to 37 degrees C, suggesting a thermally activated process. Surprisingly, SNAP-25 had little effect on docking efficiency or the probability of thermally induced fusion. Interprotein fluorescence resonance energy transfer experiments suggest the presence of other conformational states of the syntaxin*synaptobrevin interaction in addition to those observed in the crystal structure of the SNARE complex. Furthermore, although SNARE complexes involved in liposome docking preferentially assemble into a parallel configuration, both parallel and antiparallel configurations were observed.  相似文献   

19.
We investigated cell-cell fusion induced by the envelope glycoprotein of human immunodeficiency virus type 1 strain IIIB expressed on the surface of CHO cells. These cells formed syncytia when incubated together with CD4-positive human lymphoblastoid SupT1 cells or HeLa-CD4 cells but not when incubated with CD4-negative cell lines. A new assay for binding and fusion was developed by using fluorescent phospholipid analogs that were produced in SupT1 cells by metabolic incorporation of BODIPY-labeled fatty acids. Fusion occurred as early as 10 min after mixing of labeled SupT1 cells with unlabeled CHO-gp160 cells at 37 degrees C. When both the fluorescence assay and formation of syncytia were used, fusion of SupT1 and HeLa-CD4 cells with CHO-gp160 cells was observed only at temperatures above 25 degrees C, confirming recent observations (Y.-K. Fu, T.K. Hart, Z.L. Jonak, and P.J. Bugelski, J. Virol. 67:3818-3825, 1993). This temperature dependence was not observed with influenza virus-induced cell-cell fusion, which was quantitatively similar at both 20 and 37 degrees C, indicating that cell-cell fusion in general is not temperature dependent in this range. gp120-CD4-specific cell-cell binding was found over the entire 0 to 37 degrees C range but increased markedly above 25 degrees C. The enhanced binding and fusion were reduced by cytochalasins B and D. Binding of soluble gp120 to CD4-expressing cells was equivalent at 37 and 16 degrees C. Together, these data indicate that during gp120-gp41-induced syncytium formation, initial cell-cell binding is followed by a cytoskeleton-dependent increase in the number of gp120-CD4 complexes, leading to an increase in the avidity of cell-cell binding. The increased number of gp120-CD4 complexes is required for fusion, which suggests that the formation of a fusion complex consisting of multiple CD4 and gp120-gp41 molecules is a step in the fusion mechanism.  相似文献   

20.
Fusion of cultured chick embryo myoblasts was inhibited by treatment with several lysosomotropic amines. The concentrations required for half-maximal inhibition of fusion were approximately 2 μM for chloroquine, 30 μM for tributylamine, 3.2 mM for ammonium chloride, and 3.3 mM for methylamine. All the amines inhibited fusion appreciably at concentrations lower than those that reduced cell density. Both the rate and extent of fusion were affected by the amines, which had to be present for about 20 hr before the usual onset of fusion. Inhibition of fusion was reversible by transfer of inhibited cells to fresh medium. The amines did not cause accumulation of a nondialyzable inhibitor in the culture medium. Levels of creatine kinase increased by eight-fold or more between 18 and 65 hr in cultures treated with tributylamine or chloroquine, although this increase was not as pronounced as in control cultures. The increased creatine kinase activity in amine-treated cultures was due mainly to the BB and MB isozymes, with relatively little increase in the MM isozyme.  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