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.
Addition of the local anesthetic tetracaine (TTC) to multilamellar dispersions of natural phosphatidylserine (PS) causes changes in the thermotropic properties of the membrane, which can be detected by differential scanning calorimetry, and in the structure of the membrane as detected by X-ray diffraction. At molar ratio [PS]/[TTC]8.5, the melting temperature of the phospholipid shifts downwards by approximately 2.5 °C. The melting endotherm is broadened; however, there is little change in the enthalpy of melting. In ternary mixtures (PS–TTC–cholesterol), the thermotropic changes are enhanced. At [PS]/[TTC]13, the onset of phase separation of cholesterol crystals from PS in the liquid crystalline state occurs at molar fraction cholesterol (Xchol)0.28, marginally smaller than that found in the absence of the anesthetic.  相似文献   

2.
Phosphatidylserine and cholesterol are two major components of the cytoplasmic leaflet of the plasma membrane. The arrangement of cholesterol is markedly affected by the presence of phosphatidylserine in model membranes. At relatively low mol fractions of cholesterol in phosphatidylserine, compared with other phospholipids, cholesterol crystallites are formed that exhibit both thermotropic phase transitions as well as diffraction of x-rays. In the present study we have observed and characterized a novel thermotropic transition occurring in mixtures of phosphatidylserine and cholesterol. This new transition is observed at 96 degrees C by differential scanning calorimetry (DSC), using a heating scan rate of 2 degrees C/min. Observation of the transition requires that the hydrated lipid mixture be incubated for several days, depending on the temperature of incubation. The rate of formation of the material exhibiting a transition at 96 degrees C is more rapid at higher incubation temperatures. At 37 degrees C the half-time of conversion is approximately 7 days. Concomitant with the appearance of the 96 degrees C peak the previously known transitions of cholesterol, occurring at approximately 38 degrees C and 75 degrees C on heating scans of freshly prepared suspensions, disappear. These two transitions correspond to the polymorphic transition of anhydrous cholesterol and to the dehydration of cholesterol monohydrate, respectively. The loss of the 75 degrees C peak takes a longer time than that of the 38 degrees C peak, indicating that anhydrous cholesterol first gets hydrated to the monohydrate form exhibiting a transition at 75 degrees C and subsequently is converted by additional time of incubation to an altered form of the monohydrate, showing a phase transition at 96 degrees C. After several weeks of incubation at 37 degrees C, only the form with a phase transition at 96 degrees C remains. If such a sample undergoes several successive heating and cooling cycles, the 96 degrees C peak disappears and the 38 degrees C transition reappears on heating. For samples of 1-palmitoyl-2-oleoyl phosphatidylserine or of 1-stearoyl-2-oleoyl phosphatidylserine having mol fractions of cholesterol between 0.4 and 0.7, the 38 degrees C transition that reappears after the melting of the 96 degrees C component generally has the same enthalpy as do freshly prepared samples. This demonstrates that, at least for these samples, the amount of anhydrous cholesterol crystallites formed is indeed a property of the lipid mixture. We have also examined variations in the method of preparation of the sample and find similar behavior in all cases, although there are quantitative differences. The 96 degrees C transition is partially reversible on cooling and reheating. This transition is also scan rate dependent, indicating that it is, at least in part, kinetically determined. The enthalpy of the 96 degrees C transition, after incubation of the sample for 3 weeks at 37 degrees C is dependent on the ratio of cholesterol to 1-palmitoyl-2-oleoyl phosphatidylserine or to 1-stearoyl-2-oleoyl phosphatidylserine, with the enthalpy per mole cholesterol increasing between cholesterol mol fractions of 0.2 and 0.5. Dimyristoyl phosphatidylserine at a 1:1 molar ratio with cholesterol, after incubation at 37 degrees C, exhibits a transition at 95 degrees C that reverses on cooling at 44 degrees C, instead of 60 degrees C, as observed with either 1-palmitoyl-2-oleoyl phosphatidylserine or 1-stearoyl-2-oleoyl phosphatidylserine. These findings along with the essential absence of the 96 degrees C transition in pure cholesterol or in cholesterol/phosphatidylcholine mixtures, indicates that the phospholipid affects the characteristics of the transition, and therefore the cholesterol crystallites must be in direct contact with the phospholipid and are not simply in the form of pure crystals of cholesterol. These observations are particularly important in view of recent observations of the presence of cholesterol crystals in biological systems.  相似文献   

3.
We have examined the effects of cholesterol on the thermotropic phase behavior and organization of aqueous dispersions of a homologous series of linear disaturated phosphatidylserines by high-sensitivity differential scanning calorimetry and Fourier transform infrared spectroscopy. We find that the incorporation of increasing quantities of cholesterol progressively reduces the temperature, enthalpy, and cooperativity of the gel-to-liquid-crystalline phase transition of the host phosphatidylserine bilayer, such that a cooperative chain-melting phase transition is completely or almost completely abolished at 50 mol % cholesterol, in contrast to the results of previous studies. We are also unable to detect the presence of a separate anhydrous cholesterol or cholesterol monohydrate phase in our binary mixtures, again in contrast to previous reports. We further show that the magnitude of the reduction in the phase transition temperature induced by cholesterol addition is independent of the hydrocarbon chain length of the phosphatidylserine studied. This result contrasts with our previous results with phosphatidylcholine bilayers, where we found that cholesterol increases or decreases the phase transition temperature in a chain length-dependent manner (1993. Biochemistry, 32:516-522), but is in agreement with our previous results for phosphatidylethanolamine bilayers, where no hydrocarbon chain length-dependent effects were observed (1999. Biochim. Biophys. Acta, 1416:119-234). However, the reduction in the phase transition temperature by cholesterol is of greater magnitude in phosphatidylethanolamine as compared to phosphatidylserine bilayers. We also show that the addition of cholesterol facilitates the formation of the lamellar crystalline phase in phosphatidylserine bilayers, as it does in phosphatidylethanolamine bilayers, whereas the formation of such phases in phosphatidylcholine bilayers is inhibited by the presence of cholesterol. We ascribe the limited miscibility of cholesterol in phosphatidylserine bilayers reported previously to a fractional crystallization of the cholesterol and phospholipid phases during the removal of organic solvent from the binary mixture before the hydration of the sample. In general, the results of our studies to date indicate that the magnitude of the effect of cholesterol on the thermotropic phase behavior of the host phospholipid bilayer, and its miscibility in phospholipid dispersions generally, depend on the strength of the attractive interactions between the polar headgroups and the hydrocarbon chains of the phospholipid molecule, and not on the charge of the polar headgroups per se.  相似文献   

4.
The effect of cholesterol on vibrational spectra in the non polar and in the polar region of dimyristoyl phosphatidylserine (DMPS) and of phosphatidylserine from bovine spinal cord (PS) has been investigated. The small shifts in the methylene CH stretching frequencies after taking into account the contribution of the cholesterol spectrum were interpreted as a combined effect of cholesterol on the conformation of the chains and of the lesser contributions of the cholesterol methyl groups. Cholesterol also influences the ratio of the trans (1465 cm–1) to the lower wavelength (1457 cm–1) CH2 bending bands. No significant direct effect of cholesterol on the vibration of the polar residues was discerned. The small shift of the carboxylate band observed below the phase transition is probably due to the change in the intermolecular zwitterions when the average distance between the neighboring polar groups increases due to incorporation of cholesterol molecules.Abbreviations PS phosphatidylserine natural - DMPS dimyristoyl phosphatidylserine - DPPC dipalmitoyl phosphatidylcholine - FTIR Fourier transform infrared spectroscopy - DSC differential scanning calorimetry - PE phosphatidylethanolamine Offprint requests to: D. Bach  相似文献   

5.
Summary High sensitivity, differential scanning calorimetry studies of vovine retinal rod outer segment (ROS) disk membranes and aqueous dispersions of the extracted ROS phospholipids have been performed. ROS disk membranes were found to exhibit a broad peak of excess heat capacity with a maximum at less than about 3°C, ascribable to a gel-to-liquid crystalline phase transition of traction of the phospholipids. A similar thermotropic transition was observed for aqueous dispersions of the total extracted and purified ROS phospholipids. Comparison of the results obtained for the dispersion of total ROS phospholipids to those of the purified head group fractions. suggests that the thermotropic behavior reffects a gel-to-liquid crystalline transition, leading to lateral phase separation, involving those phosphatidylcholine (PC) molecules containing saturated fatty acylchains, possibley together with the highest melting ROS phosphatidylethanolamine (PE) and phosphatidylserine (PS) components. The interpretation of the thermal behavior of the ROS disk membranes depends on whether the transition is assumed to derive from the ROS PC and/or PE/PS fractions, and whether the transbilayer arrangement of the ROS phospholipids is assumed to be symmetric or asymmetric. The calorimetric data can be simply explained in terms of an asymmetric distribution of the major ROS disk membrane phospholipids (G.P. Miljanich et al.,J. Membrane Biol. 60:249–255, 1981). In this case, the transition would arise from the PE/PS fractions in the outer ROS disk membrane monolyer, and the anticipated transition from the PC in the inner monolayer would be broadened due to interaction with cholesterol. For the ROS membranes at higher temperatures, two additional, irreversible transitions are observed at 57 and 72°C, corresponding to the thermal denauturation of opsin and rhodopsin, respectively.  相似文献   

6.
Mixtures of cholesterol with dipalmitoylphosphatidylserine or phosphatidic acid were investigated by differential scanning calorimetry. As in mixtures of natural phosphatidylserine with cholesterol (Bach, D. (1984) Chem. Phys. Lipids 35, 385-392), also here phase separation of cholesterol at molar ratios of 2:1 (phospholipid:cholesterol) and below is observed. The limited solubility of cholesterol in negatively charged phospholipids is found to be independent of the nature of the acyl chain residues, and independent of whether the negative charge resides on both COO- and PO- groups (as in phosphatidylserine) or on PO- only (as in phosphatidic acid). The separate cholesterol phase is also seen by DSC in mixtures of natural phosphatidylserine or phosphatidic acid with cholesterol in the presence of Ca2+; and in phosphatidylserine/cholesterol mixtures in the presence of Li+, by DSC and X-ray diffraction.  相似文献   

7.
The influence of protons or calcium ions on the miscibility of cholesterol in phosphatidylserine has been examined using differential scanning calorimetry and X-ray diffraction. At pH 2.6, where the carboxyl group of the serine moiety is protonated, two endothermic transitions are observed in cholesterol-phosphatidylserine mixtures. The midpoint of the first is at 35 degrees C in the absence of cholesterol and decreases to approx. 15 degrees C for molar fraction of cholesterol 0.5. The second transition is centered at approx. 44 degrees C, almost independent of cholesterol content. The two lower temperature phases are lamellar and the high temperature phase has hexagonal symmetry. Cholesterol is more miscible in protonated phosphatidylserine than in the sodium form: cholesterol crystals are detected at a molar ratio of phosphatidylserine to cholesterol of about 1.7:1 as compared to about 2.3:1 at neutral pH. In the presence of calcium ions (1.3 Ca2+ per phosphatidylserine), a lamellar phase is observed with layer spacing 53 A which is independent of temperature (25 degrees C-65 degrees C) and of cholesterol content. Calcium ions cause reduced cholesterol solubility: crystallites are detected already at a molar ratio of 4:1.  相似文献   

8.
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   

9.
We have used a computer-controlled differential scanning calorimeter to determine the phases present in mixtures of the brain galactocerebrosides with other representative brain lipids. There are two types of brain galactocerebroside, those which possess an alpha-hydroxy substituent on the acyl chain (HFA) and those that do not (NFA). In the liquid crystalline state both cerebrosides were miscible with all the lipids studied, but in the gel state they were immiscible with cholesterol and the brain phosphatidylcholines. However, cholesterol mixtures in which the cholesterol mole fraction exceeded one third formed homogeneous metastable gel states on cooling from above the melting point of the cerebroside. Relaxation to the stable two phase state took place slowly over several hours. The solubilities of the galactocerebrosides in the other main brain sphingolipid, sphingomyelin, were much higher. Only in the case of the NFA galactocerebroside and at low mole fractions of sphingomyelin was immiscibility detected. Ternary mixtures of the two cerebrosides with sphingomyelin/cholesterol and phosphatidylcholine/cholesterol (PC/Chol) showed different miscibility characteristics. On cooling from 80 degrees C all mixtures formed homogeneous gel states. However, on standing the cerebrosides separated into discrete gel phases in all mixtures but one, that in which HFA galactocerebrosides were mixed with sphingomyelin and cholesterol. The cerebroside in the mixture with the composition closest to that of myelin, HFA/PC/Chol, melted at 38 degrees C. On scanning guinea pig CNS myelin which had been equilibrated at 5 degrees C a transition was detected with Tmax 33 degrees C. On the basis of comparison with the HFA/PC/Chol mixture we propose that the transition in myelin at this temperature is due to the melting of a galactocerebroside gel phase.  相似文献   

10.
Apoptosis is generally accompanied by a late phase of ceramide (Cer) production, the significance of which is unknown. This study describes a previously unrecognized link between Cer accumulation and phosphatidylserine (PS) exposure at the cell surface, a characteristic of the execution phase of apoptosis resulting from a loss of plasma membrane phospholipid asymmetry. Using a fluorescent sphingomyelin (SM) analogue, N-(N-[6-[(7-nitrobenz-2-oxa-1, 3-diazol-4-yl)amino]caproyl]-sphingosylphosphorylcholine (C(6)-NBD-SM), we show that Cer is derived from SM, initially located in the outer leaflet of the plasma membrane, which gains access to a cytosolic SMase by flipping to the inner leaflet in a process of lipid scrambling paralleling PS externalization. Lipid scrambling is both necessary and sufficient for SM conversion: Ca(2+) ionophore induces both PS exposure and SM hydrolysis, whereas scrambling-deficient Raji cells do not show PS exposure or Cer formation. Cer is not required for mitochondrial or nuclear apoptotic features since these are still observed in Raji cells. SM hydrolysis facilitates cholesterol efflux to methyl-beta-cyclodextrin, which is indicative of a loss of tight SM-cholesterol interaction in the plasma membrane. We provide evidence that these biophysical alterations in the lipid bilayer are essential for apoptotic membrane blebbing/vesiculation at the cell surface: Raji cells show aberrant apoptotic morphology, whereas replenishment of hydrolyzed SM by C(6)- NBD-SM inhibits blebbing in Jurkat cells. Thus, SM hydrolysis, during the execution phase of apoptosis, results from a loss of phospholipid asymmetry and contributes to structural changes at the plasma membrane.  相似文献   

11.
Synaptotagmin 1 (syt1) functions as the Ca(2+) sensor in neuronal exocytosis, and it has been proposed to act by modulating lipid bilayer curvature. Here we examine the effect of the two C2 domains (C2A and C2B) of syt1 on membrane lipid order and lateral organization. In mixtures of phosphatidylcholine and phosphatidylserine (PS), attenuated total internal reflection Fourier transform infrared spectroscopy indicates that a fragment containing both domains (C2AB) or C2B alone disorders the lipid acyl chains, whereas the C2A domain has little effect upon chain order. Two observations suggest that these changes reflect a demixing of PS. First, the changes in acyl chain order are reversed at higher protein concentration; second, selective lipid deuteration demonstrates that the changes in lipid order are associated only with the PS component of the bilayer. Independent evidence for lipid demixing is obtained from fluorescence self-quenching of labeled lipid and from natural abundance (13)C NMR, where heteronuclear single quantum correlation spectra reveal Ca(2+)-dependent chemical shift changes for PS, but not for phosphatidylcholine, in the presence of the syt1 C2 domains. The ability of syt1 to demix PS is observed in a range of lipid mixtures that includes cholesterol, phosphatidylethanolamine, and varied PS content. These data suggest that syt1 might facilitate SNARE (soluble N-ethylmaleimide-sensitive factor attachment protein receptors)-mediated membrane fusion by phase separating PS, a process that is expected to locally buckle bilayers and disorder lipids due to the curvature tendencies of PS.  相似文献   

12.
The effect of cholesterol on the Ca2+-induced aggregation and fusion of large unilamellar phosphatidylserine (PS) vesicles has been investigated. Mixing of aqueous vesicle contents was followed continuously with the Tb/dipicolinate assay, while the dissociation of pre-encapsulated Tb/dipicolinate complex was taken as a measure of the release of vesicle contents. Vesicles consisting of pure PS or PS/cholesterol mixtures at molar ratios of 4:1, 2:1 and 1:1 were employed at three different lipid concentrations, each at four different Ca2+ concentrations. The results could be well simulated in terms of a mass-action kinetic model, providing separately the rate constants of vesicle aggregation, c11, and of the fusion reaction itself, f11. In the analyses the possibility of deaggregation of aggregated vesicles was considered explicitly. Values of both c11 and f11 increase steeply with the Ca2+ concentration increasing from 2 to 5 mM. With increasing cholesterol content of the vesicles the value of c11 decreases, while the rate of the actual fusion reaction, f11, increases. Remarkably, the effect of cholesterol on both aggregation and fusion is quite moderate. The presence of cholesterol in the vesicle bilayer does not affect the leakage of vesicle contents during fusion.  相似文献   

13.
We have examined the effects of cholesterol (Chol) on the thermotropic phase behavior and organization of aqueous dispersions of a homologous series of linear disaturated phosphatidylglycerols (PGs) by high-sensitivity differential scanning calorimetry and Fourier transform infrared and 31P NMR spectroscopy. We find that the incorporation of increasing quantities of Chol alters the temperature and progressively reduces the enthalpy and cooperativity of the gel-to-liquid-crystalline phase transition of the host PG bilayer. With dimyristoyl-PG:Chol mixtures, cooperative chain-melting phase transitions are completely or almost completely abolished at Chol concentrations near 50 mol%, whereas with the dipalmitoyl- and distearoyl-PG:Chol mixtures, cooperative hydrocarbon chain-melting phase transitions are still discernable at Chol concentrations near 50 mol%. We are also unable to detect the presence of significant populations of separate domains of the anhydrous or monohydrate forms of Chol in our binary mixtures, in contrast to previous reports. We ascribe the previously reported large scale formation of Chol crystallites to the fractional crystallization of the Chol and phospholipid phases during the removal of organic solvent from the binary mixture before the hydration of the sample. We further show that the direction and magnitude of the change in the phase transition temperature induced by Chol addition is dependent on the hydrocarbon chain length of the PG studied. This finding agrees with our previous results with phosphatidylcholine bilayers, where we found that Chol increases or decreases the phase transition temperature in a hydrophobic mismatch-dependent manner (Biochemistry 1993, 32:516-522), but is in contrast to our previous results for phosphatidylethanolamine (Biochim. Biophys. Acta 1999, 1416:119-234) and phosphatidylserine (Biophys. J. 2000, 79:2056-2065) bilayers, where no such hydrophobic mismatch-dependent effects were observed. We also show that the addition of Chol facilitates the formation of the lamellar crystalline phase in PG bilayers, as it does in phosphatidylethanolamine and phosphatidylserine bilayers, whereas the formation of such phases in phosphatidylcholine bilayers is inhibited by the presence of Chol. Moreover, the formation of the lamellar crystalline phase in PG bilayers at lower temperatures excludes Chol, resulting in an apparent Chol immiscibility in gel-state PG bilayers. We suggest that the magnitude of the effect of Chol on the thermotropic phase behavior of the host phospholipid bilayer, and its miscibility in phospholipids dispersions generally, depend on the strength of the attractive interactions between the polar headgroups and the hydrocarbon chains of the phospholipid molecule, and not on the charge of the polar headgroups per se.  相似文献   

14.
S Das  R P Rand 《Biochemistry》1986,25(10):2882-2889
The effects of incorporating diacylglycerol (DG) derived from egg phosphatidylcholine (PC) into PC, egg phosphatidylethanolamine (PE), and bovine phosphatidylserine (PS) have been measured. In excess solution DG induces a multilamellar-to-hexagonal (L-H) structural transition in PE and PC that is temperature dependent. At 37 degrees C it begins at about 3 and 30 mol%, respectively. In PC at lower DG concentrations a modified lamellar phase is formed; at about 70 mol% DG a single primitive cubic phase forms. An L-H transition induced by 20-30 mol% DG in PS is dependent on ionic strength and degree of lipid hydration, with the appearance of crystalline acyl chains at the higher DG levels. Calcium precipitates of DG/PS (1/1) mixtures have melted chains. Structural parameters were derived for the lamellar phases at subtransition levels of DG in PE and PC. The area per polar group is increased, but by contrast with cholesterol, the polar group spreading is not accompanied by an increase in bilayer thickness. DG does not affect the equilibrium separation of PC or PE bilayers. Measured interbilayer forces as they vary with bilayer separation show that DG at 20 mol% does not effect closer apposition of PC bilayers at any separation. Spreading the polar groups may effect the binding of protein kinase C or the activation of phospholipases; the nonlamellar phases may be linked to the biochemical production of DG in cellular processes involving membrane fusion.  相似文献   

15.
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   

16.
The fluorescence of chlorophyll a incorporated into liposomes of mixtures of phosphatidylcholines and phosphatidylethanolamines is reported. Plots of fluorescence intensities against temperature show breaks at characteristic temperatures which can be attributed to the onset and completion of solid phase lipid formation. These temperatures can be plotted to give diagrams analogous to the phase diagrams obtained for macroscopic systems. Complications due to "small-system effects" are discussed, and the experimental diagrams are compared with theoretical phase diagrams calculated for ideal mixing. Introduction of cholesterol leads to a reduction in fluorescence intensity, most readily explained by a 1:1 lipid:cholesterol interaction with exclusion of monomeric, fluorescent, chlorophyll a. Interaction of divalent ions with mixtures of dipalmitoyl phosphatidylcholine and dipalmitoyl phosphatidylserine leads to exclusion of chlorophyll a from the phosphatidylserine.  相似文献   

17.
M Z Lai  N Düzgüne?  F C Szoka 《Biochemistry》1985,24(7):1646-1653
The role of the hydroxyl groups of cholesterol and tocopherol in mediating their interaction with phospholipid bilayers has been a subject of considerable interest. We have examined this question by using derivatives of cholesterol and tocopherol in which the hydroxyl group is esterified to succinate. The hemisuccinate esters of cholesterol and alpha-tocopherol can be readily incorporated into phospholipid membranes and in fact can by themselves form closed membrane vesicles as demonstrated by the encapsulation of [3H]sucrose. The thermotropic behavior of mixtures containing each succinate ester and phospholipid was studied by differential scanning calorimetry. The effect of cholesteryl hemisuccinate on the thermotropic properties of dipalmitoylphosphatidylcholine and dimyristoylphosphatidylethanolamine is very similar to that of cholesterol. This indicates that the 3 beta-OH is not required for the formation of a cholesterol-phospholipid complex. In mixtures of tocopherol acid succinate and phospholipids the peak transition temperature is progressively shifted to lower temperatures as the mole fraction of alpha-tocopherol succinate is increased, while the enthalpy of the transition is only slightly affected. At a tocopherol succinate/phospholipid molar ratio of 9/1 a phase transition is still detectable. A comparison between tocopherol succinate and tocopherol indicates that the substitution of the hydroxyl group reduces the interaction of tocopherol with phospholipids to a small but measurable extent. Thus, the hydroxyl group of tocopherol is more important than the hydroxyl group of cholesterol in influencing their interactions with phospholipids.  相似文献   

18.
The effects of low concentrations of cholesterol in mixtures of a negatively charged phospholipid (phosphatidylserine or phosphatidylglycerol) and another phospholipid (phosphatidylcholine, sphingomyelin or phosphatidylethanolamine) have been studied by differential scanning calorimetry. Only mixtures which showed a gel phase miscibility gap have been employed. It was demonstrated that in mixtures with phosphatidylethanolamine, cholesterol was preferentially associated with the negatively charged phospholipid, regardless whether this species represented the component with the high or with the low transition temperature in the mixture. In mixtures of a negatively charged phospholipid and phosphatidylcholine, cholesterol associated with the negatively charged phospholipid; when the phosphatidylcholine was the species with the low transition temperature, cholesterol had an affinity for the phosphatidylcholine and for the negatively charged phospholipid as well. Cholesterol, in a mixture of sphingomyelin with a high and phosphatidylserine with a low transition temperature, was preferentially associated with sphingomyelin.From these experiments it is concluded that phospholipids show a decrease in affinity for cholesterol in the following order: sphingomyelin ? phosphatidylserine, phosphatidylglycerol > phosphatidylcholine ? phosphatidylethanolamine.  相似文献   

19.
Cholesterol is essential for cell physiology. Transport of the “accessible” pool of cholesterol from the plasma membrane (PM) to the endoplasmic reticulum (ER) by ER‐localized GRAMD1 proteins (GRAMD1a/1b/1c) contributes to cholesterol homeostasis. However, how cells detect accessible cholesterol within the PM remains unclear. We show that the GRAM domain of GRAMD1b, a coincidence detector for anionic lipids, including phosphatidylserine (PS), and cholesterol, possesses distinct but synergistic sites for sensing accessible cholesterol and anionic lipids. We find that a mutation within the GRAM domain of GRAMD1b that is associated with intellectual disability in humans specifically impairs cholesterol sensing. In addition, we identified another point mutation within this domain that enhances cholesterol sensitivity without altering its PS sensitivity. Cell‐free reconstitution and cell‐based assays revealed that the ability of the GRAM domain to sense accessible cholesterol regulates membrane tethering and determines the rate of cholesterol transport by GRAMD1b. Thus, cells detect the codistribution of accessible cholesterol and anionic lipids in the PM and fine‐tune the non‐vesicular transport of PM cholesterol to the ER via GRAMD1s.  相似文献   

20.
Translocation of phosphatidylserine (PS) to the outer leaflet of the cellular membrane seems to be a key step in apoptosis and cell activation. In this paper, the production and characterization of a monoclonal antibody designated as Mab 1H6 is described which does not show cross reactivity with others anionic phospholipids. It is demonstrated that Mab1H6 can recognize externalized PS at early stages after the induction of apoptosis shown by both flow cytometry and immunofluorescence. Our results show that translocation of PS can be detected as early as 5 min by immunofluorescence and 10 min by flow cytometry after the treatment of cells and a specific dynamics is observed concerning the location and distribution of the staining. These data prove that antibody Mab 1H6 can be used as a specific probe for detection of PS translocation.M.M. and D.K. contributed equally to this work  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