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 31 毫秒
1.
Fourier transform-infrared (IR) spectroscopic and electron microscopic studies are reported for 1,2-dimyristoylphosphatidylserine (DMPS) and for DMPS/1,2-dimyristoylphosphatidylcholine mixtures in the presence and absence of Ca2+ ion. The frequency of the methyl symmetric deformation mode near 1,378 cm-1, previously assumed insensitive to changes in lipid morphology, has been found to respond to cochleate phase formation by undergoing an approximately 8 cm-1 increase. The new IR spectroscopic marker at 1,386 cm-1 has been used to identify and verify structures suggested from the phase diagram of J. R. Silvius and J. Gagné (1984. Biochemistry. 23:3241-3247) for this system. In addition, the ability of Mg2+ ion to induce cochleate formation has been demonstrated. Higher Mg2+ than Ca2+ levels are required for this process. Finally, IR spectroscopy has been used to monitor dehydration of the lipid surface through changes in the asymmetric PO2- stretching mode. Dehydration precedes cochleate phase formation (i.e., occurs at a lower Ca2+/phosphatidylserine level).  相似文献   

2.
Temperature dependence of Na+/Ca2+ exchange activity was studied in beef cardiac sarcolemmal vesicles in the absence and presence of the inhibitor amiloride and in proteoliposomes reconstituted with different lipid mixtures. Arrhenius plots for Na+/Ca2+ exchange activity in both control and amiloride-treated vesicles revealed an apparent energy of activation of 9665 +/- 585 (SE, n = 4) cal/mol, corresponding to a temperature coefficient (Q10) value of 1.70 +/- 0.05 (SE, n = 4) over the range 25-37 degrees C. When Na+/Ca2+ exchange was reconstituted into phosphatidylcholine (PC):phosphatidylserine (PS) (52:48, mol/mol), PC:PS:cholesterol (25:39:36, mol/mol), and PC:PS:distearoylphosphatidylcholine (DSPC) (31:48:21, mol/mol) proteoliposomes, the highest activity was found in PC:PS:cholesterol proteoliposomes. Arrhenius plots of Na+/Ca2+ exchange activity exhibited breakpoints at 23 degrees C (PC:PS), 33 degrees C (PC:PS:cholesterol), and 23 degrees C (PC:PS:DSPC). The increase in the thermotropic transition temperature with cholesterol could result from the condensing effect of this sterol, whereas the breaks observed with PC:PS and PC:PS:DSPC could be caused by a non-lipid-mediated membrane protein conformational change. These results indicate that the lipid microenvironment around the Na+/Ca2+ exchanger and the nature of the specific lipid-protein interactions influence the activity of this antiporter. Further evidence supporting the hypothesis that cholesterol behaves as a specific positive effector for the exchanger is also given.  相似文献   

3.
Na+-Ca2+ exchange activity in cardiac sarcolemmal vesicles is known to be sensitive to charged, membrane lipid components. To examine the interactions between membrane components and the exchanger in more detail, we have solubilized and reconstituted the Na+-Ca2+ exchanger into membranes of defined lipid composition. Our results indicate that optimal Na+-Ca2+ exchange activity requires the presence of certain anionic phospholipids. In particular, phosphatidylserine (PS), cardiolipin, or phosphatidic acid at 50% by weight results in high Na+-Ca2+ exchange activity, whereas phosphatidylinositol and phosphatidylglycerol provide a poor environment for exchange. In addition, incorporation of cholesterol at 20% by weight greatly facilitates Na+-Ca2+ exchange activity. Thus, for example, an optimal lipid environment for Na+-Ca2+ exchange is phosphatidylcholine (PC, 30%)/PS (50%)/cholesterol (20%). Na+-Ca2+ exchange activity is also high when cardiac sarcolemma is solubilized and then reconstituted into asolectin liposomes. We fractionated the lipids of asolectin into subclasses for further reconstitution studies. When sarcolemma is reconstituted into vesicles formed from the phospholipid component of asolectin, Na+-Ca2+ exchange activity is low. When the neutral lipid fraction of asolectin (including sterols) is also included in the reconstitution medium, Na+-Ca2+ exchange activity is greatly stimulated. This result is consistent with the requirement for cholesterol described above. Proteinase treatment, high pH, intravesicular Ca2+ and dodecyl sulfate all stimulate Na+-Ca2+ exchange in native sarcolemmal vesicles. We examined the effects of these interventions on exchange activity in reconstituted vesicles of varying lipid composition. In general, Na+-Ca2+ exchange could be stimulated only when reconstituted into vesicles of a suboptimal lipid composition. That is, when reconstituted into asolectin or PC/PS/cholesterol (30:50:20), the exchanger is already in an activated state and can no longer be stimulated. The one exception was that the Na+-Ca2+ exchanger responded to altered pH in an identical manner, independent of vesicle lipid composition. The mechanism of action of altered pH on the exchanger thus appears to be different from other interventions.  相似文献   

4.
We have examined the influence of different sterols and phospholipids on the activities of the cardiac sarcolemmal Na+-Ca2+ exchanger and Na+,K+-ATPase and the sarcoplasmic reticular Ca2+-ATPase in reconstituted proteoliposomes. When either the solubilized Na+-Ca2+ exchanger or the Na+,K+-ATPase is reconstituted into phosphatidylcholine (PC):phosphatidylserine (30:50 by weight) vesicles, high cholesterol levels (20% by weight) are required for activity to be expressed. This sterol requirement is highly specific for cholesterol. Several cholesterol analogues with minor structural changes are unable to support Na+-Ca2+ exchange or Na+,K+-ATPase activities. When solubilized sarcolemma is reconstituted into PC:cardiolipin vesicles, however, the requirement for cholesterol is lost. Substantial activity can be obtained in the complete absence of cholesterol or in the presence of several cholesterol analogues. Thus, sterol/protein interactions can be highly dependent on the phospholipid environment. In contrast, the skeletal muscle sarcoplasmic reticular Ca2+-ATPase functions equally well in the presence or absence of cholesterol after reconstitution into either PC:phosphatidylserine or PC:cardiolipin proteoliposomes. Phospholipid requirements of the transporters were also examined. The sarcolemmal Na+-Ca2+ exchanger, Na+,K+-ATPase, and the sarcoplasmic reticular Ca2+-ATPase all function optimally in the presence of phosphatidylserine or cardiolipin after reconstitution. Thus, the sarcolemmal cation transporters have similar sterol and phospholipid requirements and may have structural similarities in their hydrophobic regions. The sarcoplasmic reticular Ca2+ pump evolved in a low cholesterol membrane and has different lipid interactions. These findings may have general applicability to other plasma membrane and endoplasmic reticular enzymes.  相似文献   

5.
R A Parente  B R Lentz 《Biochemistry》1986,25(5):1021-1026
The sensitivity of the fluorescence lifetime of 1-palmitoyl-2-[[2-[4- (6-phenyl-trans-1,3,5-hexatrienyl)phenyl]ethyl]carbonyl]- 3-sn-phosphatidylcholine (DPHpPC) to its local concentration in lipid bilayers was used to monitor both lipid mixing and phase separation occurring during membrane vesicle fusion. Vesicles containing 2 mol % DPHpPC were mixed with a 10-fold excess of vesicles devoid of probe. Upon addition of a fusogen, mixing of bilayer lipids associated with fusion was followed as an increase in the fluorescence lifetime of DPHpPC. Ca2+-induced fusion of phosphatidylserine vesicles served to test the method and was shown to have an exponential half-time of 7 s. Phase separation (between the phosphatidylserine head groups of bulk lipid and the phosphatidylcholine head groups of the probe) was monitored by DPHpPC under the same conditions used to follow lipid mixing due to fusion. Phase separation was not significant until 10 min after Ca2+ addition and was completely reversible by disodium ethylenediaminetetraacetate addition. Vesicle aggregation induced by Ca2+ addition to mixed phosphatidylserine/phosphatidylcholine vesicles did not alter the DPHpPC lifetime, indicating that close association of vesicles did not promote intervesicular exchange of the probe. In addition, we have investigated the effects of CA2+ on the fluorescence properties of this probe and of the head-group-labeled fluorescent probes N-(4-nitro-2,1,3-benzoxadiazolyl)phosphatidylethanolamine and N-(lissamine Rhodamine B sulfonyl)dioleoyl-phosphatidylethanolamine, which are used in the fluorescence energy transfer assay of Struck et al.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The behavior of fluorescent and spin-label probes is examined in several fluid and gel phospholipid phases, with particular focus on the Ca2+-induced gel phase in phosphatidylserine (PS). These probes have behavior characteristic of the type of probe and of the type of lipid environment. Anthroyloxy- and doxyl-labeled PS [12-AS-PS and (7,6)PS, respectively] exhibit greatly restricted and/or slow probe motion in Ca(PS)2, even compared to thermotropic gel-phase lipid at the same temperature. In contrast, anthroyloxy- and doxyl-labeled phosphatidylcholine (PC), as well as fluorescent-labeled and spin-labeled fatty acid derivatives, show no apparent change in probe motion in Ca(PS)2 compared to fluid lamellar lipid. Doxyl-labeled phosphatidic acid, phosphatidylethanolamine, and phosphatidylglycerol show restricted motion in Ca(PS)2 relative to fluid-phase lipid, but the electron paramagnetic resonance (EPR) spectra could not be interpreted in terms of simple models for probe ordering. The fluorescent probes diphenylhexatriene (DPH) and trans-parinaric acid methyl ester (tPNA-Me) show motional behavior in Ca(PS)2 that is intermediate between that observed in fluid and in thermotropic gel-phase lipid. When Ca(PS)2 and fluid PS/PC phases coexist, probe molecules distribute between the two phases. Experiments using fluorescence quenching by spin-labeled PC in PS/PC in excess Ca2+ yield the distribution of several fluorophore probes between fluid liquid-crystal and Ca(PS)2 gel phases, expressed as a concentration ratio, RLC/G. The value of RLC/G = 100 in favor of the fluid phase is obtained for 12-AS-PC, 18 for 12-AS-Me, 12 for DPH, 3 for tPnA-Me, and 1 for 12-AS-PS.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
G W Feigenson 《Biochemistry》1989,28(3):1270-1278
Ca2+ binding between lamellae of phosphatidylserine (PS) and phosphatidylcholine (PC) gives rise to a rigid phase of Ca(PS)2. When aqueous Ca2+, hydrated PS/PC, and Ca(PS)2 coexist at equilibrium, the aqueous Ca2+ concentration is invariant and is characteristic of the PS/PC ratio. This characteristic Ca2+ concentration is 0.040 microM for palmitoyloleoylphosphatidylserine without PC and increases as the inverse square of the PS mole fraction at high PS concentration (Raoult's law) and as the inverse square of the PS mole fraction multiplied by a constant at low PS concentration (Henry's law). For example, for palmitoyloleoylphosphatidylserine/palmitoyloleoylphosphatidylcholi ne = 0.6/0.4 or 0.2/0.8, this characteristic Ca2+ concentration is about 0.1 or about 6 microM, respectively. These observations at constant temperature are summarized in a quaternary phase diagram for the four-component system CaCl2/PS/PC/water.  相似文献   

8.
Modulation of vitronectin receptor binding by membrane lipid composition.   总被引:5,自引:0,他引:5  
The vitronectin (Vn) receptor belongs to the integrin family of proteins and although its biochemical structure is fully characterized little is known about its binding affinity and specificity. We report here that Vn receptor binding to different matrix proteins is influenced by the surrounding lipid composition of the membrane. Human placenta affinity purified Vn receptor was inserted into liposomes of different composition: (i) phosphatidylcholine (PC); (ii) PC+phosphatidylethanolamine (PE); (iii) PC+PE+phosphatidylserine (PS) + phosphatidylinositol (PI) + cholesterol (chol). The amount of purified material that could be incorporated into the three lipid vesicle preparations was proportional to the efficiency of the vesicle formation that increased from PC (38%) to PC+PE and PC+PE+PS+PI+chol (about 50%) vesicles. Electron microscopy analysis showed that the homogeneity and size of the three liposome preparations were comparable (20-nm diameter) but their binding capacity to a series of substrates differed widely. Vn receptor inserted in PC liposomes bound only Vn, but when it was inserted in PC+PE and PC+PE+PS+PI+chol liposomes it also attached to von Willebrand factor (vWF) and fibronectin (Fn). Vn receptor had higher binding capacity for substrates when it was inserted in PC+PE+PS+PI+chol than PC+PE liposomes. Antibodies to Vn receptor blocked Vn receptor liposome binding to Vn, vWF, and Fn. The intrinsic emission fluorescence spectrum of the Vn receptor reconstituted in PC+PE+PS+PI+chol liposomes was blue-shifted in relation to PC liposomes, suggesting a conformational change of the receptor in the membranes. These data provide direct evidence that the Vn receptor is "promiscuous" and can associate with Vn, vWF and Fn. The nature of the membrane lipid composition surrounding the receptor could thus influence its binding affinity, possibly by changing its conformation or exposure or both.  相似文献   

9.
A microscopic study has allowed the analysis of modifications of various shapes acquired by phospholipid vesicles during a hydrostatic pressure treatment of up to 300 MPa. Giant vesicles of dimyristoylphosphatidylcholine / phosphatidylserine (DMPC/PS) prepared at 40°C mainly presented a shape change resembling budding during pressure release. This comportment was reinforced by the incorporation of 1,2-dioleyl-sn-glycero-3-phosphatidylethanolamine (DOPE) or by higher temperature (60°C) processing. The thermotropic main phase transition (Lα to Pβ′) of the different vesicles prepared was determined under pressure through a spectrofluorimetric study of 6-dodecanoyl-2-dimethylamino-naphtalene (Laurdan) incorporated into the vesicles’ bilayer. This analysis was performed by microfluorescence observation of single vesicles. The phase transition was found to begin at about 80 MPa and 120 MPa for DMPC/PS vesicles at, respectively, 40°C and 60°C. At 60°C the liquid-to-gel transition phase was not complete within 250 MPa. Addition of DMPE at 40°C does not significantly shift the onset boundary of the phase transition but extends the transition region. At 40°C, the gel phase was obtained at, respectively, 110 MPa and 160 MPa for DMPC/PS and DMPC/PS/DOPE vesicles. In comparing volume data obtained from image analysis and Laurdan signal, we assume the shape change is a consequence of the difference between lateral compressibility of the membrane and bulk water. The phase transition contributes to the membrane compression but seems not necessary to induce shape change of vesicles. The high compressibility of the Lα phase at 60°C allows induction on DMPC/PS vesicles of a morphological transition without phase change.  相似文献   

10.
To investigate the importance of the physical state of phospholipids for activation of protein kinase C, we have used short chain phospholipids, which, depending on their concentration, can exist as either monomers or micelles. We previously reported that short chain phosphatidylcholines (PC) can activate protein kinase C at concentrations that correlate with the critical micelle concentration of the activating lipid (Walker, J. M., and Sando, J. J. (1988) J. Biol. Chem. 263, 4537-4540). We have now expanded this work to short chain phosphatidylserine (PS) systems in order to examine the role of Ca2(+)-phospholipid interactions in the activation process. Short chain PS were synthesized from corresponding PC and purified by reverse-phase high pressure liquid chromatography. Use of the short chain system has revealed significant differences in the activation of type II and type III protein kinase C isozymes. The type II isozyme required Ca2+ in the presence of long chain PS vesicles; in the presence of the short chain phospholipid micelles (PC or PS), most of the activity was Ca2+ independent. Addition of diacylglycerol caused a small increase in type II activity in all phospholipid systems. In contrast, type III protein kinase C was Ca(+)-dependent in all of the lipid systems. The concentration of Ca2+ required to activate type III protein kinase C was independent of the phospholipid type despite large differences in the ability of these lipids to bind Ca2+. This isozyme required diacylglycerol only in the PC micelle system or with vesicles composed of long chain saturated PS. The presence of short chain PS micelles or long chain PS with unsaturated fatty acyl chains rendered this Ca2(+)-dependent protein kinase C virtually diacylglycerol independent. These results are consistent with a model in which type II protein kinase C requires Ca2+ primarily for membrane association, a requirement which is bypassed with the micelle system, whereas type III protein kinase C has an additional Ca2+ requirement for activity that does not involve Ca2(+)-phospholipid interactions.  相似文献   

11.
Gd3+ was evaluated as a probe for Ca2+ sites on protein kinase C (PKC) by studying its ability to replace Ca2+ in activation of PKC isozymes II (beta) and III (alpha) in the lipid systems phosphatidylserine/1,2-dioleoyl-sn-glycerol (PS/DO) and diheptanoylphosphatidylcholine (PC7)/DO. PKC beta was stimulated by Ca2+ or Gd3+ in PS/DO whereas activity in PC7/DO was independent of these metals. Thus, it is suggested that Gd3+ replaces Ca2+ at a site involving metal-lipid interactions. High concentrations of Ca2+ or Gd3+ inhibited activity in both lipid systems. Analysis of the Gd3+ inhibition in the PC7/DO system suggests that it is due to formation of GdATP, which competes at the MgATP site. Activity of PKC alpha was dependent on low concentrations of Ca2+ in both lipid systems. The ability of Gd3+ to substitute for Ca2+ could not be evaluated in the PS system due to the inability to completely remove contaminating Ca2+ without chelating buffers. Successful reduction of contaminating Ca2+ was achieved in the PC7 system but Gd3+ failed to substitute for Ca2+ in activating PKC alpha and only caused inhibition. This is consistent with binding of Gd3+ to a Ca2+ site at or near the active site of the enzyme rather than to a site on the lipid. These results indicate that interactions between PKC and Gd3+ are complex, involving occupation of more than one class of sites. Conditions for separately evaluating the individual sites can be manipulated by selection of isozyme and lipid system.  相似文献   

12.
Sonicated vesicles of phosphatidylserine and phosphatidylserine/phosphatidylcholine mixtures were recombined with spectrin-actin from human erythrocyte ghosts. Morphological properties and physicochemical characteristics of the recombinates were studied with freeze etch electron microscopy, 31P NMR and differential scanning calorimetry. Sonicated dimyristoyl phosphatidylserine vesicles show a decrease in enthalpy change of the lipid phase transition upon addition of spectrin-actin. These vesicles collapse and fuse, into multilamellar structures in the presence of spectrin-actin, as demonstrated by freeze fracturing and NMR. Spectrin-actin cannot prevent the salt formation between phosphatidylserine and Ca2+, all phosphatidylserine is withdrawn from the lipid phase transition. In contrast a protection against the action of Mg2+ could be observed. Mixed bilayers of dimyristoyl phosphatidylserine/dimyristoyl phosphatidylcholine show phase separations at molar ratios above 1/1 (van Dijck, P.W.M., de Kruijff, B., Verkleij, A.J., van Deenen, L.L.M. and de Gier, J. (1978) Biochim. Biophys. Acta 512, 84--96). These phase spearations can be prevented by spectrin-actin. Ca2+-induced lateral phase separations in cocrystallizing phosphatidylserine/phosphatidylcholine mixtures, can be reduced by spectrin-actin. Formation of the Ca2+-phosphatidylserine salt, occurring in addition to lateral phase separation when mixtures contain more than 30 mol % phosphatidylserine, cannot be prevented by spectrin-actin.  相似文献   

13.
Electrophoretic light scattering (ELS) and depolarization of fluorescence have been used to determine the effect of membrane fluidity on the binding of Ca2+ to liposomes. ELS was used to measure the electrophoretic mobilities of the liposomes. Fluorescence depolarization was used to determine membrane fluidity. Zero to 30 mol% phosphatidylserine (PS) was incorporated into liposomes containing, as bulk phospholipids, one of the following: dimyristoyl-phosphatidylcholine (DMPC), dipalmitoylphosphatidylcholine (DPPC), egg phosphatidylcholine (PC), or hydrogenated egg phosphatidylcholine (H egg PC). The binding of Ca2+ to the liposomes appears to be influenced by membrane fluidity. Liposomes containing bulk phospholipids whose phase transition temperature is higher than the experimental temperature exhibit enhanced binding of CA2+.  相似文献   

14.
A M Haywood  B P Boyer 《Biochemistry》1984,23(18):4161-4166
How the lipid composition of liposomes determines their ability to fuse with Sendai virus membranes was tested. Liposomes were made of compositions designed to test postulated mechanisms of membrane fusion that require specific lipids. Fusion does not require the presence of lipids that can form micelles such as gangliosides or lipids that can undergo lamellar to hexagonal phase transitions such as phosphatidylethanolamine (PE), nor is a phosphatidylinositol (PI) to phosphatidic acid (PA) conversion required, since fusion occurs with liposomes containing phosphatidylcholine (PC) and any one of many different negatively charged lipids such as gangliosides, phosphatidylserine (PS), phosphatidylglycerol, dicetyl phosphate, PI, or PA. A negatively charged lipid is required since fusion does not occur with neutral liposomes containing PC and a neutral lipid such as globoside, sphingomyelin, or PE. Fusion of Sendai virus membranes with liposomes that contain PC and PS does not require Ca2+, so an anhydrous complex with Ca2+ or a Ca2+-induced lateral phase separation is not required although the possibility remains that viral binding causes a lateral phase separation. Sendai virus membranes can fuse with liposomes containing only PS, so a packing defect between domains of two different lipids is not required. The concentration of PS required for fusion to occur is approximately 10-fold higher than that required for ganglioside GD1a, which has been shown to act as a Sendai virus receptor. When cholesterol is added as a third lipid to liposomes containing PC and GD1a, the amount of fusion decreases if the GD1a concentration is low.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Effects of phorbol ester and teleocidin on Ca2+-induced fusion of liposomes   总被引:1,自引:0,他引:1  
The effects of different types of lipid membrane defects on Ca2+-induced fusion of liposomes containing phosphatidylserine (PS) were investigated using fluorescent probes. Teleocidin enhanced the fusion of phospholipid vesicles in an assay system using terbium/dipicolinic acid during mixing of internal aqueous phases of vesicles upon fusion. 12-O-Tetradecanoylphorbol-13-acetate (TPA) suppressed the fusion. This latter phenomenon was also observed by measuring the excitation energy transfer. The promotion of membrane fusion by teleocidin was ascribed to dehydration of the membrane surface, the suppressive effect of TPA to desorption of Ca2+ from the membrane surface. Thus, Ca2+-induced fusion of PS vesicles was shown to be sensitive to defects of the membrane surface, but insensitive to defects of the hydrophobic core of the lipid membrane.  相似文献   

16.
The concentration of free Ca(2+) and the composition of nonsubstrate phospholipids profoundly affect the activity of phospholipase C delta1 (PLCdelta1). The rate of PLCdelta1 hydrolysis of phosphatidylinositol 4,5-bisphosphate was stimulated 20-fold by phosphatidylserine (PS), 4-fold by phosphatidic acid (PA), and not at all by phosphatidylethanolamine or phosphatidylcholine (PC). PS reduced the Ca(2+) concentration required for half-maximal activation of PLCdelta1 from 5.4 to 0.5 microM. In the presence of Ca(2+), PLCdelta1 specifically bound to PS/PC but not to PA/PC vesicles in a dose-dependent and saturable manner. Ca(2+) also bound to PLCdelta1 and required the presence of PS/PC vesicles but not PA/PC vesicles. The free Ca(2+) concentration required for half-maximal Ca(2+) binding was estimated to be 8 microM. Surface dilution kinetic analysis revealed that the K(m) was reduced 20-fold by the presence of 25 mol % PS, whereas V(max) and K(d) were unaffected. Deletion of amino acid residues 646-654 from the C2 domain of PLCdelta1 impaired Ca(2+) binding and reduced its stimulation and binding by PS. Taken together, the results suggest that the formation of an enzyme-Ca(2+)-PS ternary complex through the C2 domain increases the affinity for substrate and consequently leads to enzyme activation.  相似文献   

17.
Phosphatidylcholine (PC) alone or with phosphatidylethanolamine (PE) are sufficient for the reconstitution of Na+ channels in planar lipid bilayers. However, when Na+ channels were first reconstituted into liposomes using the freeze-thaw-sonication method, addition of acidic phospholipids, such as phosphatidylserine (PS), to the neutral phospholipids was necessary to obtain a significant toxin-modulated 22Na uptake. To further investigate the acidic phospholipid effect on reconstitution into liposomes, Na+ channels purified from Electrophorus electricus electrocytes were reconstituted into liposomes of different composition by freeze-thaw sonication and the effect of batrachotoxin and tetrodotoxin on the 22Na flux was measured. The results revealed that, under our experimental conditions, the presence of an acidic phospholipid was also necessary to obtain a significant neurotoxin-modulated 22Na influx. Though neurotoxin-modulated 22Na fluxes have been reported in proteoliposomes made with purified Na+ channels and PC alone, the 22Na fluxes were smaller than those found using lipid mixtures containing acidic phospholipids. Electron microscopy of negatively stained proteoliposomes prepared with PC, PC/PS (1:1 molar ratio), and PS revealed that the acidic phospholipid increases the size of the reconstituted proteoliposomes. The increment in size caused by the acidic phospholipid, due to the associated increase in internal volume for 22Na uptake and in area for Na+ channel incorporation, appears to be responsible for the large neurotoxin-modulated 22Na fluxes observed.  相似文献   

18.
We have investigated the fusion behavior of a mixed vesicle system consisting of vesicles prepared from the simple synthetic surfactants di(n-dodecyl)phosphate (DDP) or di(n-tetradecyl)phosphate (DTP) and vesicles prepared from the phospholipids phosphatidylserine (PS) or dioleoylphosphatidylcholine (DOPC). Fusion between the vesicles, induced by Ca2+, was determined by a resonance energy transfer assay for lipid mixing, sucrose density gradient analysis, and electron microscopy. We demonstrate that synthetic surfactant vesicles can specifically engage in asymmetric fusion events, provided that the incubation temperature is kept below the gel-liquid crystalline phase-transition temperature (Tc) of the synthetic amphiphile (29 and 48 degrees C for DDP and DTP, respectively) and that the physical state of the target membrane is fluid. Asymmetric fusion of DDP or DTP vesicles was most efficient with PS vesicles, but it also occurred with zwitterionic PC vesicles. In the latter case, fusion proceeded spontaneously, but the process was markedly accelerated upon addition of Ca2+. Furthermore, in contrast to a massive transformation of bilayer into nonbilayer hexagonal HII tubular structures, as occurs upon symmetric Ca(2+)-induced fusion of DDP vesicles, asymmetric fusion with phospholipid bilayers predominantly leads to the formation of larger vesicles. This indicates that both PS and DOPC stabilize the DDP bilayer structure in the fusion product.  相似文献   

19.
The interaction of Ca2+ and Mg2+ with phosphatidylserine (PS) vesicles in 0.1 M NaCl aqueous solution was studied by equilibrium dialysis binding, X-ray diffraction, batch microcalorimetry, kinetics of cation-induced vesicle aggregation, release of vesicle contents, and fusion. Addition of either cation causes aggregation of PS vesicles and produces complexes with similar stoichiometry (1:2 cation/PS) at saturating concentrations, although the details of the interactions and the resulting complexes are quite different. Addition of Ca2+ to PS vesicles at T greater than or equal to 25 degrees C induces the formation of an "anhydrous" complex of closely apposed membranes with highly ordered crystalline acyl chains and a very high transition temperature (Tc greater than 100 degrees C). The formation of this complex is accompanied by a release of heat (5.5 kcal/mol), rapid release of vesicle contents, and fusion of the vesicles into larger membranous structures. By contrast, addition of Mg2+ produces a complex with PS which is much more hydrated, has no crystallization of the acyl chains at T greater than or equal to 20 degrees C, and has comparatively little fusion. Studies with both Ca2+ and Mg2+ added simultaneously indicate that there is a synergistic effect between the two cations, which results in an enhancement of the ability of Ca2+ to form its specific complex with PS at lower concentrations. The presence of the erythrocyte protein "spectrin" inhibits this synergism and interferes with the formation of the specific PS/Ca complex. It also inhibits the fusion of PS vesicles. It is proposed that the unique PS/Ca complex, which involves close apposition of vesicle membranes, is an intermembrane "trans" complex. We further propose that such a complex is a key step for the resultant phase transition and fusion of PS vesicles. By contrast, the PS/Mg complex is proposed to be a "cis" complex with respect to each membrane. The results are discussed in terms of the mechanism of membrane fusion.  相似文献   

20.
The kinetics of lipid mixing during membrane aggregation and fusion was monitored by two assays employing resonance energy transfer between N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine (NBD-PE) and N-(lissamine Rhodamine B sulfonyl)phosphatidylethanolamine (Rh-PE). For the "probe mixing" assay, NBD-PE and Rh-PE were incorporated into separate populations of phospholipid vesicles. For the "probe dilution" assay, both probes were incorporated into one population of vesicles, and the assay monitored the dilution of the molecules into the membrane of unlabeled vesicles. The former assay was found to be very sensitive to aggregation, even when the internal aqueous contents of the vesicles did not intermix. Examples of this case were large unilamellar vesicles (LUV) composed of phosphatidylserine (PS) in the presence of Mg2+ and small unilamellar vesicles (SUV) composed of phosphatidylserine in the presence of high concentrations of Na+. No lipid mixing was detected in these cases by the probe dilution assay. Under conditions where membrane fusion (defined as the intermixing of aqueous contents with concomitant membrane mixing) was observed, such as LUV (PS) in the presence of Ca2+, the rate of probe mixing was faster than that of probe dilution, which in turn was faster than the rate of contents mixing. Two assays monitoring the intermixing of aqueous contents were also compared. The Tb/dipicolinic acid assay reported slower fusion rates than the 1-aminonaphthalene-3,6,8-trisulfonic acid/N,N'-p-xylylene-bis(pyridinium bromide) assay for PS LUV undergoing fusion in the presence of Ca2+. These observations point to the importance of utilizing contents mixing assays in conjunction with lipid mixing assays to obtain the rates of membrane destabilization and fusion.(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