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.
Heme transfer between phospholipid membranes and uptake by apohemoglobin   总被引:4,自引:0,他引:4  
The incorporation of CO-heme into single bilayer, egg lecithin vesicles was examined by following the spectral changes that occur when the porphyrin becomes embedded in the membranes. The rate of CO-heme uptake by liposomes is extremely fast (t1/2 less than or equal to 20 ms at 10 degrees C), and the maximum extent is roughly 1 heme/5 phospholipid molecules. This limiting stoichiometry is due to unfavorable electrostatic interactions between the propionate groups of the bound CO-heme. This effect was treated theoretically by attenuating the intrinsic heme partitioning equilibrium constant with an exponential term reflecting the surface potential of the membranes. The surface potential was assumed to be proportional to the concentration of CO-heme in the membranes, and the final expression is Kp = Kop exp[-AHb/VpCp], where Kp is the observed partition constant; Kop, the intrinsic constant; Hb, the concentration of bound heme in the suspension; Vp, the partial molar volume of egg lecithin; Cp, the concentration of lipid phosphate; and A, an empirical constant representing the capacitance of the membrane for heme. For the analysis of kinetic data, the electrostatic term is assumed to apply only to the membrane dissociation rate constant, k-1, and not the association rate constant, k1. The dissociation rate was measured independently either by following the transfer of CO-heme from one vesicle fraction to another or by monitoring heme efflux from the membranes and incorporation into apohemoglobin at high protein concentrations. The data for all three sets of experiments, heme uptake, transfer, and incorporation into globin at 10 degrees C, were fitted quantitatively to the partitioning mechanism using A = 15 M-1, Kop = 5 X 10(5), k1 = 2 X 10(6) s-1, and k0(-1) = 4 s-1. Thus, heme can spontaneously migrate across lipid-water interfaces and hence diffuse rapidly from the mitochondrial inner membrane where it is synthesized to the rough endoplasmic reticulum where it is incorporated into hemoglobin.  相似文献   

2.
Three structurally related crown compounds and cryptands have been synthesized that differ by the number and linkage of coronand units and anthracene moieties. The interaction of the fluorescent dyes with sonicated dimyristoylphosphatidylcholine (DMPC) vesicles is characterized by the relative quantum yields, uptake kinetics, binding curves, lifetimes, fluorescence titrations with water- and lipid-soluble quenching agents, fluorescence anisotropy, and equilibrium cooling curves. The most lipophilic compound II, which displays a similar quantum yield as the parent fluorophore 9,10-dimethylanthracene, shows a nearly equal distribution between solid and fluid lipid and is located at the bilayer surface. The least lipophilic compound IV is excluded from the hydrocarbon phase. The anthracenophane cryptand III preferentially partitions into solid-phase lecithins with the highest affinity for the phases L epsilon and L beta. The binding constant to DMPC amounts to (5.4 +/- 1.3) X 10(2) M-1 at 0 degrees C. From fluorescence quenching titrations it is concluded that the average position of the anthracenoyl dye III discontinuously shifts during the gel to liquid crystalline transition from the glycerol backbone to the choline head group. Electron microscopy and NMR experiments revealed that the anthracenophane induces in the liquid crystalline phase the fusion of small unilamellar DMPC vesicles to unilamellar medium-sized vesicles and macrovesicles, which subsequently fuse at the transition temperature to large multilamellar coacervates. Due to its large change of fluorescence intensity, the anthracenophane cryptand is a very sensitive probe for the detection of the pretransition of symmetrically substituted and of the subtransition of asymmetrically substituted phosphatidylcholines.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Cholesterol, when sequestered in saturated liposomes of dimyristoylphosphatidylcholine (DMPC) or dipalmitoylphosphatidylcholine (DPPC), undergoes peroxidation thermally initiated either by a lipid-soluble or a water-soluble azo initiator and in both cases the reaction is inhibited effectively by the water-soluble antioxidant, 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylate (Trolox). Quantitative kinetic methods of autoxidation show that the oxidizability, kp/(2kt)1/2 (where kp and 2kt are the rate constants of radical chain propagation and termination, respectively) of cholesterol in DMPC or DPPC multilamellar liposomes, where kp/(2kt)1/2 is 3.0.10(-3) to 4.3.10(-3) M-1/2 s-1/2 at 37-45 degrees C, is similar to that measured in homogeneous solution in chlorobenzene, where kp/(2kt)1/2 is 3.32.10(-3). However, its oxidizability in smaller unilamellar vesicles of DMPC or DPPC increases by at least 3-times that measured in multilamellar systems. Autoxidation/antioxidant methods show that cholesterol partitions directly from the solid state into DMPC or DPPC liposomes by shaking and this is confirmed by 31P and 2H quadrupole NMR spectra of deuterated cholesterol when membrane bound. Analytical studies indicate that up to 21 mol% cholesterol will partition into the membranes by shaking.  相似文献   

4.
Interaction of the local anesthetic dibucaine with small unilamellar vesicles of dimyristoylphosphatidylcholine (DMPC) and dioleoyl phosphatidylcholine (DOPC) containing different mol percents of cholesterol has been studied by fluorescence spectroscopy. Fluorescence measurements on dibucaine in presence of phospholipid vesicles containing various amounts of cholesterol yielded a pattern of variation of wavelength at emission maximum and steady-state anisotropy which indicated that the microenvironment of dibucaine is more polar and flexible in membranes that contain cholesterol than in membranes without cholesterol. Experiments on quenching of fluorescence from membrane-associated dibucaine by potassium iodide showed a marked increase in quenching efficiency as the cholesterol content of the vesicles was increased, demonstrating increased accessibility of the iodide quenchers to dibucaine in the presence of cholesterol, when compared to that in its absence. Total emission intensity decay profiles of dibucaine yielded two lifetime components of approximately 1 ns and approximately 2.8--3.1 ns with mean relative contributions of approximately 25 and approximately 75%, respectively. The mean lifetime in vesicles was 20--30% smaller than in the aqueous medium and showed a moderate variation with cholesterol content. Fluorescence measurements at two different temperatures in DMPC SUVs, one at 33 degrees C, above the phase transition temperature and another at 25 degrees C, around the main phase transition, indicated two different mode of dibucaine localization. At 25 degrees C dibucaine partitioned differentially in presence and absence of cholesterol. However, at 33 degrees C the apparent partition coefficients remained unaltered indicating differences in the microenvironment of dibucaine in presence and absence of cholesterol in the phospholipid membranes.  相似文献   

5.
The submillisecond kinetics for phloretin binding to unilamellar phosphatidylcholine (PC) vesicles was investigated using the temperature-jump technique. Spectrophotometric studies of the equilibrium binding performed at 328 nm demonstrated that phloretin binds to a single set of independent, equivalent sites on the vesicle with a dissociation constant of 8.0 microM and a lipid/site ratio of 4.0. The temperature of the phloretin-vesicle solution was jumped by 4 degrees C within 4 microseconds producing a monoexponential, concentration-dependent relaxation process with time constants in the 30--200-microseconds time range. An analysis of the concentration dependence of relaxation time constants at pH 7.30 and 24 degrees C yielded a binding rate constant of 2.7 X 10(8) M-1 s-1 and an unbinding constant of 2,900 s-1; approximately 66 percent of total binding sites are exposed at the outer vesicle surface. The value of the binding rate constant and three additional observations suggest that the binding kinetics are diffusion limited. The phloretin analogue, naringenin, which has a diffusion coefficient similar to phloretin yet a dissociation constant equal to 24 microM, bound to PC vesicle with the same rate constant as phloretin did. In addition, the phloretin-PC system was studied in buffers made one to six times more viscous than water by addition of sucrose or glycerol to the differ. The equilibrium affinity for phloretin binding to PC vesicles is independent of viscosity, yet the binding rate constant decreases with the expected dependence (kappa binding alpha 1/viscosity) for diffusion-limited processes. Thus, the binding rate constant is not altered by differences in binding affinity, yet depends upon the diffusion coefficient in buffer. Finally, studies of the pH dependence of the binding rate constant showed a dependence (kappa binding alpha [1 + 10pH-pK]) consistent with the diffusion-limited binding of a weak acid.  相似文献   

6.
Ubiquinone (UQn with n = 2, 3, or 10 isoprenoid groups) was incorporated into small, sonicated vesicles made of dipalmitoylphosphatidylcholine (DPPC) or dimyristoylphosphatidylcholine (DMPC). (1) The accessibility of oxidized UQ in DPPC or DMPC vesicles to the reductant sodium borohydride (NaBH4), measured by UV spectroscopy, was UQ2 greater than UQ3 greater than UQ10 (DPPC) and UQ2 greater than UQ3 approximately UQ10 (DMPC). (2) Catalysis of the reduction of entrapped ferricyanide by exogenous NaBH4 was more effective with UQ2 than UQ10 but was slower with all quinones than reduction by added dithionite. (3) The methoxy protons of UQ2 and UQ3 in DPPC and DMPC vesicles exhibited a single NMR resonance centered at approximately 3.95 ppm, whereas the methoxy groups of UQ10 gave rise to two separate proton resonances, at 3.93 ppm and a more narrow resonance at 3.78 ppm. The UQ10 population characterized by the 3.78 ppm resonance was present at a higher concentration in DPPC than in DMPC vesicles and was relatively insensitive to reduction by NaBH4. (4) UQ10 perturbed the melting temperature (Tm) of DPPC vesicles to a smaller extent (delta Tm = -1 degrees C) than did UQ2 and UQ3 (delta Tm = -3 to -4 degrees C). The combined UV and NMR data imply the following: The UQ10 pool characterized by the 3.78 ppm peak corresponds to a more mobile UQ10 fraction that is not reduced by NaBH4 in 2-3 min and is thought to be localized close to the center of the DPPC bilayer since it has little effect on the DPPC Tm.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Kinetic and equilibrium studies on the interaction of steroids with human corticosteroid-binding globulin (CBG, transcortin) were performed with pH, temperature, and steroid structure as variables. Dissociation rate constants were determined fluorometrically; the values for cortisol, corticosterone, deoxycorticosterone, and progesterone are 0.031, 0.047, 0.10, and 0.16 s-1, respectively, at 20 degrees C, pH 7.4. The pH dependence of the dissociation rate constant for the corticosterone complex below pH 10.5 at 20 degrees C is given by koff = 0.043 (1 + [H+]/10(-6.50)) s-1; above pH 11, koff = 0.030 (1 + 10(-12.15/[H+] s-1. A temperature-dependence study of koff for the cortisol and progesterone complexes gave values of 0.0028 s-1 and 0.012 s-1 at 4 degrees C, respectively, and 0.88 s-1 and 4.5 s-1 at 37 degrees C, with progesterone dissociating about four to five times faster over the entire temperature range. The affinity constants, determined by equilibrium dialysis, for the binding of cortisol, corticosterone, and progesterone at 4 degrees C were 7.9, 7.2, and 7.0 X 10(8) M-1; values of 0.40 and 0.26 X 10(8) M-1 were determined at 37 degrees C for cortisol and progesterone. The close similarity of the affinity constants of the three steroids combined with differing dissociation rates implies that the association rate changes with steroid structure, in contrast to our earlier findings with progesterone-binding globulin.  相似文献   

8.
Membrane fluidity as affected by the organochlorine insecticide DDT   总被引:4,自引:0,他引:4  
Fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene (DPH) was used to study the interaction of DDT with model and native membranes. DDT decreases the phase transition midpoint temperature (Tm) of liposomes reconstituted with dimyristoyl-, dipalmitoyl- and distearoylphosphatidylcholines (DMPC, DPPC and DSPC), and broadens the thermotropic profile of the transition. The effects of DDT are concentration dependent and are more pronounced in bilayers of short-chain lipids, e.g., DMPC. The insecticide fails to alter DPH polarization in the fluid phase of the above lipids. Similar effects were observed in binary mixtures of DMPC plus DPPC. Furthermore, DDT alters the single broad transition of the equimolar mixture of DMPC plus DSPC into a biphasic transition. The lower temperature component has a midpoint at 25 degrees C, i.e., a value close to the Tm of DMPC. DDT inhibits to some extent the cholesterol-induced ordering in DMPC bilayers and high cholesterol concentrations (greater than or equal to 30 mol%) do not prevent insecticide interaction, conversely to the effect observed for lindane (Antunes-Madeira, M.C. and Madeira, V.M.C. (1989) Biochim. Biophys. Acta 982, 161-166). Apparently, the bilayer order is not disturbed by DDT in fluid native membranes of mitochondria and sarcoplasmic reticulum, but moderate disordering effects are noticed in membranes enriched in cholesterol, namely, brain microsomes and erythrocytes.  相似文献   

9.
Rate constants for the hydration of bilirubin bound to unilamellar bilayers of dioleoylphosphatidylcholine and albumin were measured by stopped-flow methods. Rate constants for association of bilirubin with these vesicles and albumin were calculated from measured rate constants for dissociation and the equilibrium binding constants of bilirubin and lipids or albumin. Rate constants for hydration (dissociation) for bilirubin bound to dioleoylphosphatidylcholine and albumin were 71 s-1 and 1.8 s-1 respectively. Rate constants for association were 4.0 10(7) s-1 and 1.1 10(9) M-1 s-1, respectively. Both rates for interactions of bilirubin with bilayers were essentially independent of temperature in the range 0-40 degrees C, indicating that barriers to entry and exit of bilirubin from bilayers were entropic. Rates of transbilayer movement of bilirubin in dioleoylphosphatidylcholine were too fast to resolve by measuring rates of hydration of bilirubin. Rate constants for hydration of bilirubin bound to bilayers with less avidity for bilirubin as compared with dioleoylphosphatidylcholine also were too fast to measure with stopped-flow methods. In addition to providing details of the energetic basis for interactions between bilirubin and membranes, the data allow for calculating the maximal rates at which bilirubin could transfer spontaneously from sites on albumin in blood to the interior of cells. The data show, in this regard, that this rate is 10-50 fold faster than measured rates of uptake of bilirubin by intact liver.  相似文献   

10.
We have measured the lateral diffusion coefficient (D), of active dansyl-labeled gramicidin C (DGC), using the technique of fluorescence photobleaching recovery, under conditions in which the cylindrical dimer channel of DGC predominates. In pure, hydrated, dimyristoylphosphatidylcholine (DMPC) multibilayers (MBL), D decreases from 6 X 10(-8) cm2/s at 40 degrees C to 3 X 10(-8) cm2/s at 25 degrees C, and drops 100-fold at 23 degrees C, the phase transition temperature (Tm) of DMPC. Above Tm, addition of cholesterol decreases D; a threefold stepwise drop occurs between 10 and 20 mol %. Below Tm, increasing cholesterol increases D; a 10-fold increase occurs between 10 and 20 mol % at 21 degrees C, between 20 and 25 mol % at 15 degrees C, and between 25 and 30 mol % at 5 degrees C. In egg phosphatidylcholine (EPC) MBL, D decreases linearly from 5 X 10(-8) cm2/s at 35 degrees C to 2 X 10(-8) cm2/s at 5 degrees C; addition of equimolar cholesterol reduces D by a factor of 2. Thus this transmembrane polypeptide at low membrane concentrations diffuses quite like a lipid molecule. Its diffusivity in lipid mixtures appears to reflect predicted changes of lateral composition. Increasing gramicidin C (GC) in DMPC/GC MBL broadened the phase transition, and the diffusion coefficient of the lipid probe N-4-nitrobenzo-2-diazole phosphatidylethanolamine (NBD-PE) at 30 degrees C decreases from 8 X 10(-8) cm2/s below 5 mol % GC to 2 X 10(-8) cm2/s at 14 mol % GC; D for DGC similarly decreases from 4 X 10(-8) cm2/s at 2 mol % GC to 1.4 X 10(-8) cm2/s at 14 mol % GC. Hence, above Tm, high concentrations of this polypeptide restrict the lateral mobility of membrane components.  相似文献   

11.
Giant liposomes obtained by electroformation and observed by phase-contrast video microscopy show spontaneous deformations originating from Brownian motion that are characterized, in the case of quasispherical vesicles, by two parameters only, the membrane tension sigma and the bending elasticity k(c). For liposomes containing dimyristoyl phosphatidylcholine (DMPC) or a 10 mol% cholesterol/DMPC mixture, the mechanical property of the membrane, k(c), is shown to be temperature dependent on approaching the main (thermotropic) phase transition temperature T(m). In the case of DMPC/cholesterol bilayers, we also obtained evidence for a relation between the bending elasticity and the corresponding temperature/cholesterol molecular ratio phase diagram. Comparison of DMPC/cholesterol with DMPC/cholesterol sulfate bilayers at 30 degrees C containing 30% sterol ratio shows that k(c) is independent of the surface charge density of the bilayer. Finally, bending elasticities of red blood cell (RBC) total lipid extracts lead to a very low k(c) at 37 degrees C if we refer to DMPC/cholesterol bilayers. At 25 degrees C, the very low bending elasticity of a cholesterol-free RBC lipid extract seems to be related to a phase coexistence, as it can be observed by solid-state (31)P-NMR. At the same temperature, the cholesterol-containing RBC lipid extract membrane shows an increase in the bending constant comparable to the one observed for a high cholesterol ratio in DMPC membranes.  相似文献   

12.
W Knoll  G Schmidt  K Ibel  E Sackmann 《Biochemistry》1985,24(19):5240-5246
The small-angle neutron scattering (SANS) technique developed previously is used to study the lateral phase separation in dimyristoylphosphatidylcholine (DMPC)-cholesterol mixed vesicles in the L alpha (35 degrees C) and L beta' (7 degrees C) phase of DMPC. To increase the sensitivity of the previous method, we apply the so-called inverse contrast variation technique where contrast matching is performed at a constant H2O/D2O ratio by varying the ratio of DMPC with deuterated and protonated hydrocarbon chains. Phase boundaries can be determined to an accuracy of +/- 0.5 mol %. In parallel experiments phase separation in the L beta' phase was also studied by freeze-fracture electron microscopy. For DMPC in the L alpha phase complete miscibility is clearly established up to cholesterol molar fractions of xc = 0.14. Strong evidence is provided that this is also the case up to xc approximately equal to 0.45. Cholesterol is no longer soluble above this limit and precipitates as small crystallites. For the L beta' phase (7 degrees C) phase boundaries are clearly established at xc1 = 0.08 and xc2 = 0.24, and very strong evidence is provided for two additional boundaries at xc3 = 0.435 and xc4 approximately equal to 1.0. At 0 less than or equal to xc less than or equal to xc1 the mixture forms a tilted solid solution in both the L beta' and P beta' phase while at xc1 less than or equal to xc less than or equal to xc2 this phase coexists with a nontilted mixture containing 24 mol % cholesterol.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The S-adenosyl-l-homocysteine (AdoHcy) hydrolases catalyze the reversible conversion of AdoHcy to adenosine and homocysteine, making use of a catalytic cycle in which a tightly bound NAD+ oxidizes the 3-hydroxyl group of the substrate at the beginning of the cycle, activating the 4-CH bond for elimination of homocysteine, followed by Michael addition of water to the resulting intermediate and a final reduction by the tightly bound NADH to give adenosine. The equilibrium and kinetic properties of the association and dissociation of the cofactor NAD+ from the enzymes of Homo sapiens (Hs-SAHH) and Trypanosoma cruzi (Tc-SAHH) are qualitatively similar but quantitatively distinct. Both enzymes bind NAD+ in a complex scheme. The four active sites of the homotetrameric apoenzyme appear to divide into two numerically equal classes of active sites. One class of sites binds cofactor weakly and generates full activity very rapidly (in less than 1 min). The other class binds cofactor more strongly but generates activity only slowly (>30 min). In the case of Tc-SAHH, the final affinity for NAD+ is roughly micromolar and this affinity persists as the equilibrium affinity. In the case of Hs-SAHH, the slow-binding phase terminates in micromolar affinity also, but over a period of hours, the dissociation rate constant decreases until the final equilibrium affinity is in the nanomolar range. The slow binding of NAD+ by both enzymes exhibits saturation kinetics with respect to the cofactor concentration; however, binding to Hs-SAHH has a maximum rate constant around 0.06 s-1, while the rate constant for binding to Tc-SAHH levels out at 0.006 s-1. In contrast to the complex kinetics of association, both enzymes undergo dissociation of NAD+ from all four sites in a single first-order reaction. The equilibrium affinities of both Hs-SAHH and Tc-SAHH for NADH are in the nanomolar range. The dissociation rate constants and the slow-binding association rate constants for NAD+ show a complex temperature dependence with both enzymes; however, the cofactor always dissociates more rapidly from Tc-SAHH than from Hs-SAHH, the ratio being around 80-fold at 37 degrees C, and the cofactor binds more rapidly to Hs-SAHH than to Tc-SAHH above approximately 16 degrees C. These features present an opening for selective inhibition of Tc-SAHH over Hs-SAHH, demonstrated with the thioamide analogues of NAD+ and NADH. Both analogues bind to Hs-SAHH with approximately 40 nM affinities but much more weakly to Tc-SAHH (0.6-15 microM). Nevertheless, both analogues inactivated Tc-SAHH 60% (NAD+ analogue) or 100% (NADH analogue) within 30 min, while the degree of inhibition of Hs-SAHH approached 30% only after 12 h. The rate of loss of activity is equal to the rate of dissociation of the cofactor and thus 80-fold faster at 37 degrees C for Tc-SAHH.  相似文献   

14.
We examined the consequences of membrane heterogeneity for the association of a simple amphiphilic molecule with phospholipid vesicles with solid-liquid and liquid-liquid phase coexistence. To address this problem we studied the association of a single-chain, fluorescent amphiphile with dimyristoylphosphatidylcholine (DMPC) vesicles containing varying amounts of cholesterol. DMPC bilayers containing 15 mol% cholesterol show a region of solid-liquid-ordered (s-l(o)) coexistence below the T(m) of pure DMPC (23.9 degrees C) and a region of liquid-disordered-liquid-ordered coexistence (l(d)-l(o)) above the T(m). We first examined equilibrium binding and kinetics of amphiphile insertion into single-phase vesicles (s, l(d), and l(o) phase). The data obtained were then used to predict the behavior of the equivalent process in a two-phase system, taking into account the fractions of phases present. Next, the predicted kinetics were compared to experimental kinetics obtained from a two-phase system. We found that association of the amphiphile with lipid vesicles is not influenced by the existence of l(d)-l(o) phase boundaries but occurs much more slowly in the s-l(o) phase coexistence region than expected on the basis of phase composition.  相似文献   

15.
A W Scotto  D Zakim 《Biochemistry》1986,25(7):1555-1561
The presence of cholesterol in small unilamellar vesicles (ULV) of dimyristoylphosphatidylcholine (DMPC) catalyzes fusion of the vesicles at temperatures below the upper limit for the gel to liquid-crystalline phase transition of the DMPC. The extent to which ULV grow depends on the concentration of cholesterol in the vesicles and on temperature. Maximum growth occurs at 21 degrees C. It decreases as the temperature is lowered below 21 degrees C. Growth does not occur at temperatures above the phase transition. In addition, the presence of cholesterol in ULV of DMPC catalyzes the insertion of integral membrane proteins into the vesicles. Thus, bacteriorhodopsin from Halobacterium halobrium, UDPglucuronosyltransferase (EC 2.4.1.17) from pig liver microsomes, and cytochrome oxidase from beef heart mitochondria formed stable lipid-protein complexes spontaneously when added to ULV containing cholesterol at temperatures under which these vesicles would fuse. Incorporation of these proteins into the ULV of DMPC did not occur in the absence of cholesterol or in the presence of cholesterol when the temperature of the system was above that for the phase transition. It appears that cholesterol lowers the energy barrier for fusion of ULV of DMPC and for insertion of integral membrane proteins into these bilayers. Studies with bacteriorhodopsin suggest that the energy barrier for insertion of proteins into ULV containing cholesterol is smaller than the energy barrier for fusion of the ULV with each other.  相似文献   

16.
Effect of lipid physical state on the rate of peroxidation of liposomes.   总被引:2,自引:0,他引:2  
The effect of cholesterol on the rate of peroxidation of arachidonic acid and 1-palmitoyl-2-arachidonoyl phosphatidylcholine (PAPC) in dimyristoylphosphatidylcholine (DMPC) liposomes was examined above and below the phase transition temperature (Tm) of the lipid. The rate of peroxidation of arachidonic acid was more rapid below the phase transition temperature of the host lipid. At a temperature below the Tm (4 degrees C), increasing concentrations of cholesterol reduced the rate of peroxidation of arachidonic acid as judged by the production of thiobarbituric acid reactive substances. Above Tm (37 degrees C), cholesterol increased the rate of peroxidation of the fatty acid. Similarly, PAPC was peroxidized more rapidly at 4 degrees C than at 37 degrees C. However, cholesterol had little effect on the rate of peroxidation of PAPC at 4 degrees C. The rate of peroxidation of arachidonic acid was related to the lipid bilayer fluidity as judged by fluorescence anisotropy measurements of diphenylhexatriene. The rate of peroxidation increased slowly with increasing rigidity of the probe environment when the bilayer was relatively fluid and more rapidly as the environment became more rigid. The increase in the rate of peroxidation of arachidonic acid in the less fluid host lipid was unrelated to differences in iron binding or to transfer of arachidonic acid to the aqueous phase. Decreasing the concentration of arachidonic acid in DMPC to less than 2 mol% dramatically decreased the rate of peroxidation at 4 degrees C, suggesting that formation of clusters of fatty acids at 4 degrees C is required for rapid peroxidation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Erythrocyte ghosts were incubated with sonicated vesicles and the uptake of cholesterol by vesicles allowed to proceed to equilibrium. The experiments were carried out for a series of phospholipids at different temperatures. The equilibrium partition of cholesterol between ghosts and single shelled vesicles provided a measure of the relative affinities of cholesterol for the different phospholipids studied. It was found that the affinity of cholesterol for dipalmitoyl phosphatidylcholine was the same as that for N-palmitoyl sphingomyelin both at temperatures above and below the gel to liquid crystalline transition temperature of these phospholipids.  相似文献   

18.
The changes in steady-state fluorescence lifetimes and anisotropy decay parameters, as well as enzyme activities, of dansyl-labeled cytochrome b5 (DNS-cytochrome b5), on interaction with NADH-cytochrome-b5 reductase in DMPC vesicles, have been measured as a function of temperature. Steady-state fluorescence of DNS-cytochrome b5 in DMPC vesicles with and without cholesterol was increased on interaction with reductase at temperatures both above and below the DMPC phase transition. In all systems three fluorescence decay components of the dansyl label in DNS-cytochrome b5 were observed. In the reductase-containing system, the long (major) decay time component of DNS-cytochrome b5 and the fraction of the total fluorescence associated with this component increased over the temperature range 15-30 degrees C. In time-resolved anisotropy measurements, the order parameters of DNS-cytochrome b5 in DMPC vesicles increased on interaction with reductase at temperatures above the DMPC phase transition, and this increase was even more pronounced in cholesterol-containing vesicles, at temperatures from 15-30 degrees C. The enzyme activity of the DNS-cytochrome-b5 reductase system in DMPC vesicles was also greatly increased in the presence of cholesterol. These results show that interaction of vesicle-bound DNS-cytochrome b5 and NADH-cytochrome-b5 reductase leads to an increased degree of order of the dansyl-labeled cytochrome with little change in its rotational flexibility, and suggests that the increased order can be correlated with increased enzyme activity.  相似文献   

19.
The passive leakage of glucose across bilayers of dimyristoylphosphatidylcholine (DMPC), cholesterol (variable), and dicetyl phosphate (constant 5.9 mol%) has been measured as efflux over 30 min from multilamellar vesicles. Bilayer cholesterol was varied from 20 mol% to 40 mol%. Glucose permeation rates were measured from 10 degrees C to 36 degrees C, and showed a maximum in permeability at 24 degrees C, the DMPC phase transition temperature. Increasing the bilayer cholesterol content above 20 mol% reduced that permeability peak. These results are quite consistent with a large number of similar bilayer permeability studies over the past 25 years. However, they are not consistent with a previous study of these same systems, which reported increased glucose permeability with temperature, without any maximum at or near the lipid chain melting temperature (K. Inoue, Biochim. Biophys. Acta 339 (1974) 390-402).  相似文献   

20.
M Masserini  E Freire 《Biochemistry》1987,26(1):237-242
The transfer of ganglioside GM1 from micelles to membranes and between different membrane populations has been examined by using a pyrene fatty acid derivative of the ganglioside. The transfer of gangliosides from micelles to membranes depends on the physical state as well as the molecular composition of the acceptor vesicles. At 30 degrees C, the transfer of micellar gangliosides to dipalmitoylphosphatidylcholine (DPPC) large unilameller vesicles (Tm = 41.3 degrees C) is characterized by a rate constant of 0.01 min-1; at 48 degrees C, however, the rate constant is 0.11 min-1. Below the phase transition temperature, the activation energy is 25 kcal/mol whereas above the phase transition it is 17 kcal/mol. Similar experiments performed with synaptic plasma membranes yielded a rate constant of 0.05 min-1 at 37 degrees C. The rate of transfer of ganglioside molecules, asymmetrically located on the outer layer of donor vesicles, to acceptor vesicles lacking ganglioside depends on the physical state of both the donor and acceptor vesicles. For the transfer of ganglioside from DPPC (donor) vesicles to dimyristoylphosphatidylcholine (DMPC) (acceptor) vesicles, the rates were essentially zero at 15 degrees C in which both vesicle populations were in the gel phase, 0.008 min-1 at 30 degrees C in which DPPC is in the gel phase and DMPC is in the fluid phase, and 0.031 min-1 at 48 degrees C in which both vesicle populations are in the fluid phase. The transfer of ganglioside from DPPC vesicles to synaptic plasma membranes was also dependent on the physical state of the donor vesicles and showed an inflection point at the phase transition temperature of DPPC.(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