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.
The kinetics of redistribution of lipid-like molecules between the membranes of two fused spherical vesicles is studied by solving the time-dependent diffusion equation of the system. The effects on the probe redistribution rate of pore size at the fusion junction and the relative sizes of the vesicles are examined. It is found that the redistribution rate constant decreases significantly, but not drastically, as the relative size of the pore to that of the vesicles decreases (the bottleneck effect). In general, the time scale of the probe redistribution rate is determined by the size of the vesicles that is loaded with the probe before the activation of the fusion. For a pore size 50 A in diameter and a typical diffusion coefficient of 10(-8) cm2/s for lipids, the mixing half times for typical virus-cell and cell-cell fusion systems are less than 30 ms and above 200 s, respectively. Thus, although the redistribution of lipid-like probes by diffusion is not rate limiting in virus-cell fusion, redistribution by diffusion is close to rate limiting in spike-protein mediated cell-cell fusion.  相似文献   

2.
In glycoprotein-mediated pH-induced fusion of virus to animal cells, the mixing of materials between membranes or between cytoplasmic spaces occurs after the virus-cell complex has gone through a number of activation reactions. The monitoring of the fluorescence changes measured in a fusing system using self-quenching probes could reflect not only the kinetics of activation, but also the redistribution reaction of probes. For instance, time delay seen in the onset of fluorescence changes after triggering the fusion reaction (S.J. Morris, D.P. Sarkar, J.M. White and R. Blumenthal, J. Biol. Chem. (1989) 3972), could be due to rate-limiting probe redistribution kinetics. In this paper we examined in detail the effect of probe redistribution rates on fusion kinetics. Simulations were performed using a very simple model with two fusion-activation steps and an exponential probe redistribution kinetics. We conclude that if the rates of probe redistribution are faster than or equal to those of viral glycoprotein activation, the kinetics of the fusion reaction are not significantly affected.  相似文献   

3.
The theoretical basis of a new technique for measuring equilibrium adsorption/desorption kinetics and surface diffusion of fluorescent-labeled solute molecules at solid surfaces has been developed. The technique combines total internal reflection fluorescence (TIR) with either fluorescence photobleaching recovery (FPR) or fluorescence correlation spectroscopy (FCS). A laser beam totally internally reflects at a solid/liquid interface; the shallow evanescent field in the liquid excites the fluorescence of surface adsorbed molecules. In TIR/FPR, adsorbed molecules are bleaching by a flash of the focused laser beam; subsequent fluorescence recovery is monitored as bleached molecules exchange with unbleached ones from the solution or surrounding nonilluminated regions of the surface. In TIR/FCS, spontaneous fluorescence fluctuations due to individual molecules entering and leaving a well-defined portion of the evanescent field are autocorrelated. Under appropriate experimental conditions, the rate constants and surface diffusion coefficient can be readily obtained from the TIR/FPR and TIR/FCS curves. In general, the shape of the theoretical TIR/FPR and TIR/FCS curves depends in a complex manner upon the bulk and surface diffusion coefficients, the size of the iluminated or observed region, and the adsorption/desorption/kinetic rate constants. The theory can be applied both to specific binding between immobilized receptors and soluble ligands, and to nonspecific adsorption processes. A discussion of experimental considerations and the application of this technique to the adsorption of serum proteins on quartz may be found in the accompanying paper (Burghardt and Axelrod. 1981. Biophys. J. 33:455).  相似文献   

4.
Total internal reflection with fluorescence correlation spectroscopy (TIR-FCS) is a method for measuring the surface association/dissociation rates and absolute densities of fluorescent molecules at the interface of solution and a planar substrate. This method can also report the apparent diffusion coefficient and absolute concentration of fluorescent molecules very close to the surface. An expression for the fluorescence fluctuation autocorrelation function in the absence of contributions from diffusion through the evanescent wave, in solution, has been published previously (N. L. Thompson, T. P. Burghardt, and D. Axelrod. 1981, Biophys. J. 33:435-454). This work describes the nature of the TIR-FCS autocorrelation function when both surface association/dissociation kinetics and diffusion through the evanescent wave contribute to the fluorescence fluctuations. The fluorescence fluctuation autocorrelation function depends in general on the kinetic association and dissociation rate constants, the surface site density, the concentration of fluorescent molecules in solution, the solution diffusion coefficient, and the depth of the evanescent field. Both general and approximate expressions are presented.  相似文献   

5.
S K Huang  M Cheng    S W Hui 《Biophysical journal》1990,58(5):1119-1126
Monolayers of human erythrocytes, immobilized on a cover slip, were induced to fuse by polyethylene glycol (mol wt 8,000). The mobility of fluorescent probes, 1-oleoyl-2-[12-[(7-nitro-2,1,3-benzoxadizol-4-yl)amino]dodecanoyl] phosphatidyl-choline (C12-NBD-PC), from labeled cells to unlabeled cells was monitored by video-enhanced fluorescence microscopy. A dequenching curve was obtained from the measurement of fluorescence intensities of pairs of fused cells over time. The dequenching curve and the curve obtained from macroscopic measurements of a cell monolayer (described in the preceding article) were compared and discussed. The slow probe transfer rate between a pair of fused cells was explained by a diffusion model based on membrane area conservation and the geometry of the fusion lumen. An equivalent lumen between two fused cells, thought to be the main rate limitation of probe mobility after fusion, was calculated to be approximately 130 nm in diameter. Lumens of 75 nm in diameter were observed by electron microscopy. Thus, the rate of macroscopic fluorescence dequenching depends not only upon the fusion efficiency, but also upon the number of simultaneous fusion partners, the geometry of their contact points, and the lateral mobility of the fluorescent probes through these points. The relative fusion efficiency can be derived only from the saturation dequenching values.  相似文献   

6.
Fluorescence photobleaching recovery with total internal reflection illumination (TIR-FPR) has been used to measure the dissociation kinetics of a fluorescein-labeled anti-dinitrophenyl monoclonal Fab specifically bound to supported monolayers composed of a mixture of dipalmitoylphosphatidylcholine and dinitrophenyl-conjugated dipalmitoylphosphatidylethanolamine. The fluorescence recovery curves were not monoexponential; when analyzed as a sum of two exponentials, the rates and fractional recoveries were approximately 1 s-1 (approximately 50%) and approximately 0.1 s-1 (approximately 30%). The data did not change as a function of the Fab solution concentration, indicating that the fluorescence recovery curves were not influenced by the rate of diffusion in bulk solution. Also, the recovery curves were independent of the size of the illuminated area, indicating that surface diffusion did not significantly contribute to the rate and shape of the fluorescence recovery. The measured off rates and apparent association constant (1.6 x 10(5) M-1) were analyzed with the theoretical formalism for a proposed mechanism that accounts for the nonmonoexponential kinetics.  相似文献   

7.
Mass action kinetics of virus-cell aggregation and fusion.   总被引:1,自引:2,他引:1       下载免费PDF全文
J Bentz  S Nir    D G Covell 《Biophysical journal》1988,54(3):449-462
A simple approximate solution for the mass action kinetics of small particles (viruses or vesicles) binding to large particles (cells) and their subsequent fusion has been derived. The solution is evaluated in terms of the measurable fluorescence changes expected when the virus or vesicles are labeled with fluorescent probes, which are diluted into the cellular membrane by fusion. Comparison with numerical integrations shows that the approximate solution is extremely accurate. Analytic simplifications for a variety of special cases of this general problem are also shown.  相似文献   

8.
From the study of the kinetics of the fluorescence self-quenching of pyrene in multilamellar vesicles of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) in the fluid phase we obtain the molecular diffusion parameters, diffusion coefficients, and activation energies for lateral diffusion for three probes: 1-palmitoyl-2-(1-pyrenedecanoyl)-sn-glycero-3-phosphocholine (py(10)-PC), pyrene, and 1-pyrenebutanoic acid (py-but). The experiments are done in a range of temperatures for which the reversibility of excimer formation is negligible and the probe/phospholipid ratios used are low, avoiding non-ideal mixing of the probe. The time-resolved fluorescence decays are, in all cases, accurately and consistently described by the two-dimensional (2D) formalism for bimolecular diffusion-controlled reactions. From the parameters obtained in this way we conclude that the primary step of the diffusion of the pyrene-labeled phospholipid involves the simultaneous jump of phospholipid and fluorophore moieties, and also that although in the case of py(10)-PC the pyrene from molecules pertaining to different layers may interact during the lifetime of the exited state, this is not the case for free pyrene.  相似文献   

9.
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   

10.
Total internal reflection with fluorescence correlation spectroscopy is a method for measuring the surface association/dissociation rate constants and absolute densities of fluorescent molecules at the interface of a planar substrate and solution. This method can also report the apparent diffusion coefficient and absolute concentration of fluorescent molecules very close to the surface. Theoretical expressions for the fluorescence fluctuation autocorrelation function when both surface association/dissociation kinetics and diffusion through the evanescent wave, in solution, contribute to the fluorescence fluctuations have been published previously. In the work described here, the nature of the autocorrelation function when both surface association/dissociation kinetics and diffusion through the evanescent wave contribute to the fluorescence fluctuations, and when fluorescent and nonfluorescent molecules compete for surface binding sites, is described. The autocorrelation function depends in general on the kinetic association and dissociation rate constants of the fluorescent and nonfluorescent molecules, the surface site density, the concentrations of fluorescent and nonfluorescent molecules in solution, the solution diffusion coefficients of the two chemical species, the depth of the evanescent field, and the size of the observed area on the surface. Both general and approximate expressions are presented.  相似文献   

11.
Fluorescence photobleaching recovery has been employed to study the lateral mobility of the Sendai virus envelope glycoproteins (HN, neuraminidase/hemagglutinin protein (HN) fusion protein (F) on the surface of human erythrocytes. Our results indicate that the two viral glycoproteins are laterally immobile on the cell surface prior to fusion, and become mobile during the fusion process. The two fused glycoproteins are mobilized to the same extent (diffusion coefficients of 3.1-3.3 X 10(-10) cm2/sec with mobile fractions of 0.53-0.57 for both HN and F). Their mobilization is blocked under conditions that allow virus adsorption and hemagglutination, but not virus-cell or cell-cell fusion. These findings suggest a possible role for the lateral diffusion of the viral glycoproteins in the mechanism of cell-cell fusion, enabling them to perturb the membranes of adjacent cells and lead to cell-cell fusion.  相似文献   

12.
The mobility of cell surface MHC class I molecules on HeLa cells was measured by fluorescence recovery after photobleaching (FRAP). The probe used for these studies was the phycobiliprotein R-phycoerythrin coupled to Fab fragments of a monoclonal antibody specific for human monomorphic MHC class I molecules. It was found that the recovery curves could be equally well fitted by either a random diffusion model with an immobile component or by an anomalous diffusion model. In the latter case, the anomalous diffusion exponent was consistent with that previously determined by single-particle tracking (SPT) experiments using the same probe (P. R. Smith, I. E. G. Morrison, K. M. Wilson, N. Fernandez, and R. J. Cherry. 1999. Biophys. J. 76:3331-3344). The FRAP experiments, however, yielded a considerably higher value of D(0), the diffusion coefficient for a time interval of 1 s. To determine whether the results were probe dependent, FRAP measurements were also performed with the same monoclonal antibody labeled with Oregon Green. These experiments gave similar results to those obtained with the phycoerythrin probe. FRAP experiments with the lipid probe 5-N-(octadecanoyl) aminofluoroscein (ODAF) bound to HeLa cells gave typical results for lipid diffusion. Overall, our observations and analysis are consistent with anomalous diffusion of MHC class I diffusion on HeLa cells, but quantitative differences between FRAP and SPT data remain to be explained.  相似文献   

13.
The refolding kinetics of ribonuclease S have been measured by tyrosine absorbance, by tyrosine fluorescence emission, and by rapid binding of the specific inhibitor 2′CMP 2 to folded RNAase S. The S-protein is first unfolded at pH 1.7 and then either mixed with S-peptide as refolding is initiated by a stopped-flow pH jump to pH 6.8, or the same results are obtained if S-protein and S-peptide are present together before refolding is initiated. The refolding kinetics of RNAase S have been measured as a function of temperature (10 to 40 °C) and of protein concentration (10 to 120 μm). The results are compared to the folding kinetics of S-protein alone and to earlier studies of RNAase A. A thermal folding transition of S-protein has been found below 30 °C at pH 1.7; its effects on the refolding kinetics are described in the following paper (Labhardt &; Baldwin, 1979).In this paper we characterize the refolding kinetics of unfolded S-protein, as it is found above 30 °C at pH 1.7, together with the kinetics of combination between S-peptide and S-protein during folding at pH 6.8. Two classes of unfolded S-protein molecules are found, fast-folding and slow-folding molecules, in a 20: 80 ratio. This is the same result as that found earlier for RNAase A; it is expected if the slow-folding molecules are produced by the slow cis-trans isomerization of proline residues after unfolding, since S-protein contains all four proline residues of RNAase A.The refolding kinetics of the fast-folding molecules show clearly that combination between S-peptide and S-protein occurs before folding of S-protein is complete. If combination occurred only after complete folding, then the kinetics of formation of RNAase S should be rather slow (5 s and 100 s at 30 °C) and nearly independent of protein concentration, as shown by separate measurements of the folding kinetics of S-protein, and of the combination between S-peptide and folded S-protein. The observed folding kinetics are faster than predicted by this model and also the folding rate increases strongly with protein concentration (apparent 1.6 order kinetics). The fact that RNAase S is formed more rapidly than S-protein alone is sufficient by itself to show that combination with S-peptide precedes complete folding of S-protein. Computer simulation of a simple, parallel-pathway scheme is able to reproduce the folding kinetics of the fast-folding molecules. All three probes give the same folding kinetics.These results exclude the model for protein folding in which the rate-limiting step is an initial diffusion of the polypeptide chain into a restricted range of three-dimensional configurations (“nueleation”) followed by rapid folding (“propagation”). If this model were valid, one would expect comparable rates of folding for RNAase A and for S-protein and one would also expect to find no populated folding intermediates, so that combination between S-peptide and S-protein should occur after folding is complete. Instead, RNAase A folds 60 times more rapidly than S-protein and also combination with S-peptide occurs before folding of S-protein is complete. The results demonstrate that the folding rate of S-protein increases after the formation, or stabilization, of an intermediate which results from combination with S-peptide. They support a sequential model for protein folding in which the rates of successive steps in folding depend on the stabilities of preceding intermediates.The refolding kinetics of the slow-folding molecules are complex. Two results demonstrate the presence of folding intermediates: (1) the three probes show different kinetic progress curves, and (2) the folding kinetics are concentration-dependent, in contrast to the results expected if complete folding of S-protein precedes combination with S-peptide. A faster phase of the slow-refolding reaction is detected both by tyrosine absorbance and fluorescence emission but not by 2′CMP binding, indicating that native RNAase S is not formed in this phase. Comparison of the kinetic progress curves measured by different probes is made with the use of the kinetic ratio test, which is defined here.  相似文献   

14.
Total internal reflection fluorescence (TIRF) microscopy reveals highly mobile structures containing enhanced green fluorescent protein-tagged glucose transporter 4 (GLUT4) within a zone about 100 nm beneath the plasma membrane of 3T3-L1 adipocytes. We developed a computer program (Fusion Assistant) that enables direct analysis of the docking/fusion kinetics of hundreds of exocytic fusion events. Insulin stimulation increases the fusion frequency of exocytic GLUT4 vesicles by approximately 4-fold, increasing GLUT4 content in the plasma membrane. Remarkably, insulin signaling modulates the kinetics of the fusion process, decreasing the vesicle tethering/docking duration prior to membrane fusion. In contrast, the kinetics of GLUT4 molecules spreading out in the plasma membrane from exocytic fusion sites is unchanged by insulin. As GLUT4 accumulates in the plasma membrane, it is also immobilized in punctate structures on the cell surface. A previous report suggested these structures are exocytic fusion sites (Lizunov et al., J. Cell Biol. 169:481-489, 2005). However, two-color TIRF microscopy using fluorescent proteins fused to clathrin light chain or GLUT4 reveals these structures are clathrin-coated patches. Taken together, these data show that insulin signaling accelerates the transition from docking of GLUT4-containing vesicles to their fusion with the plasma membrane and promotes GLUT4 accumulation in clathrin-based endocytic structures on the plasma membrane.  相似文献   

15.
Orientational exchange approach to fluorescence anisotropy decay.   总被引:1,自引:1,他引:0       下载免费PDF全文
Fluorescence depolarization is a powerful technique in resolving dynamics of molecular systems. Data obtained in fluorescence depolarization experiments are highly complex. Mathematical models for analyzing data from depolarization due to rotational motion have been largely based on the rotational diffusion equation. These results have been verified by Monte Carlo simulations. It has been implicitly stated that a 90 degrees jump model between predefined orientations such as presented by G. Weber (1971. J. Chem. Phys. 55:2399-2411) should, for the specific case of fluorescence depolarization, give the same answer as the diffusion equation. Since the highly symmetric cases considered by G. Weber gave the same result as the diffusion equation, it has been desirable to use this method in cases where depolarization arises from both discrete processes and rotational diffusion. We have derived, in a compartmental formalism, the general result for excitation and emission dipoles not necessarily coincident with any of the principal rotational axes of the fluorophore from this exchange model, and have found it to be different from that of the diffusion equation approach. We have also verified this difference with a Monte Carlo simulation of our exchange model. This derivation allows us to define the limits of validity of the 90 degrees exchanges to model rotational diffusion. Also, for systems where movements may be jumps between a few preferred orientations, the actual physical mechanism of depolarization may not be accurately represented by continuous diffusion. The compartmental formalism developed here can be used to easily combine rotational motions with discrete position jumps or other level kinetics.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The kinetics of poly(ethylene glycol) (PEG)-induced fusion between intact human erythrocytes was continuously monitored by a fluorescence lipid mixing method, utilizing the dequenching of the fluorescence probe, 1-oleoyl-2-[12-[(7-nitro-2,1,3-benzoxadiazol-4-yl)amino]dodecanoyl ] phosphatidylcholine (C12-NBD-PC). The steady-state fluorescence intensity was detected from the surface of cells in a monolayer on an alcian blue-coated glass coverslip. The relief of fluorescence self-quenching after fusion between C12-NBD-PC labeled and unlabeled intact erythrocytes was measured. The extent of fluorescence dequenching was normalized based on the measured concentration of probes in membranes, the projected partial dequenching due both to dilution by intercellular fusion, and the dilution between the inner and outer leaflets of membranes (flip-flop). There was no significant increase in fluorescence intensity during PEG treatment of 5 min, at 4 degrees C. Intensity increased immediately after the dilution of PEG, and reached saturation in 30 min. The efficiency of fusion increased with the increasing of PEG concentrations. Only 4% enhancement of saturated relative fluorescence intensity was detected in 25 wt% PEG-induced cell fusion; 23% enhancement in 30 wt%; and 66% enhancement in 35 wt%. The transfer of fluorescent probes between membrane bilayer leaflets (flip-flop) was also monitored during the fusion process. Flip-flop was monitored in confluent monolayers as well as in isolated cells. There was no significant spontaneous flip-flop within 30 min of dilution. The relative fluorescence intensity enhancement contributed by the dilution of probes between fused labeled and unlabeled cells (at a 1:1 ratio) was found to account for only 39% of the observed final dequenching, whereas the contribution by flip-flop associated with cell fusion was found to account for 9%, and flip-flop without fusion contributed approximately 18%. A portion of the flip-flop is a consequence of hemolysis. Therefore, fluorescence dequenching measurements of fusion of whole cells must be interpreted with caution.  相似文献   

17.
Fluorescence assays for viral membrane fusion employ lipidic probes whose kinetics of fluorescence dequenching should mimic the actual kinetics of membrane merging. We examined the fusion of influenza virus with CEM cells, erythrocyte ghosts or liposomes by monitoring the fluorescence dequenching of each one of the three probes, octadecylrhodamine B chloride (R18), N-(lissamine rhodamine B sulfonyl)phosphatidylethanolamine (Rh-PE), or rac-2,3-dioleoylglycerol ester of rhodamine B (DORh-B), inserted into the virus membrane. Experimental conditions were designed to allow a clear distinction between membrane mixing and non-specific probe transfer. Fluorescence dequenching observed with Rh-PE was much slower than with R18, unless a particular experimental procedure was used. Using liposomes as a target membrane, the kinetics and extent of the decrease in resonance energy transfer between N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine (NBD-PE) and Rh-PE, initially embedded in the liposome membrane, were matched by that of the dequenching of viral R18, but not of viral Rh-PE. DORh-B was found not to be appropriate to follow membrane merging. Our results indicate that on a time scale of several minutes R18 more accurately reflects the kinetics of membrane fusion. Nevertheless, control experiments should be performed to evaluate non-specific probe transfer of R18 molecules, whose contribution to fluorescence dequenching can become significant after long incubation times.  相似文献   

18.
Y I Henis  O Gutman 《Biochemistry》1987,26(3):812-819
We have recently employed fluorescence photobleaching recovery (FPR) to demonstrate that the envelope glycoproteins of Sendai virions become laterally mobile on the surface of human erythrocytes following fusion [Henis, Y. I., Gutman, O., & Loyter, A. (1985) Exp. Cell Res. 160, 514-526]. In order to investigate whether this lateral mobilization is involved in the mechanism of virally mediated cell-cell fusion, or is merely a result of viral envelope-cell fusion, we have now performed FPR studies on erythrocytes fused with reconstituted Sendai virus envelopes (RSVE). These RSVE, which were prepared by solubilization of Sendai virions with Triton X-100 followed by removal of the detergent through adsorption to SM-2 Bio-beads, fused with human erythrocytes as efficiently as native virions but induced cell-cell fusion to a much lower degree. The fraction of the viral envelope glycoproteins that became laterally mobile in the erythrocyte membrane following fusion was markedly lower in the case of RSVE than in the case of native virions. The lower cell-cell fusion activity of the RSVE does not appear to be due to inactivation of the viral fusion protein, since the envelope-cell fusion and hemolytic activities of the RSVE were similar to those of native virions. Moreover, fusion with RSVE or with native virions resulted in the incorporation of rather similar amounts of viral glycoproteins into the cell membrane. Since the reduced fraction of laterally mobile viral glycoproteins correlates with the lower cell-cell fusion activity of the RSVE.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Proton diffusion along the surface of a planar bilayer lipid membrane was measured by means of acid/base injection with a micropipette and recording of the kinetics of fluorescence changes of fluorescein-labelled lipid on the surface. The dimensionality of the process was assayed by fitting the kinetic curves with two-dimensional (2D) or three-dimensional (3D) diffusion equations. In agreement with Serowy et al. (Biophys J 84:1031-1037, 2003), lateral proton diffusion proceeded via bulk phase by means of buffer molecules as proton carriers (D = 600 microm2/s) under the conditions of 1 mM buffer in the solution. Introduction of proton binding sites on the membrane surface led to the appearance of a considerable contribution of two-dimensional proton diffusion on the membrane surface with D = 1,100 mum(2)/s. The system described can be used to study the dependence of the proton diffusion rate on the phospholipid and protein composition of the membrane.  相似文献   

20.
Interactions of cell surface components with one another and with structures inside and outside the cell may have important physiological functions in the transmission of signals and the assembly of specialized structures. These interactions may be detected and analyzed through their effects on the lateral mobility of cell surface molecules. Measurements by a fluorescence photobleaching method have shown that in general lipid-like molecules diffuse rapidly and freely through the plasma membrane, whereas proteins move much more slowly or appear to be immobile. This dichotomy has been supposed to result from forces beyond the viscosity of the lipid bilayer, which specifically retard the diffusion of membrane proteins. This general picture should be qualified, however, by noting that the lateral mobility of lipid-like molecules can be influenced in detail by changes in the state of the plasma membrane such as result from mitosis or fertilization. The interactions of cell surface proteins that limit their lateral mobility are unknown. The effects of binding concanavalin A to localized regions of cell surface show that these interactions can vary in subtle and complex ways. It may soon be useful to interpret mobility experiments in terms of simple reaction models that attempt to describe surface interactions in physicochemical terms. More experimental data are needed to carry out this program and to relate interactions that affect mobility to the structural connections between cell surface components and the cytoskeleton, which have been detected by biochemical methods and electron and immunofluorescence microscopy.  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