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.
Differential polarized phase fluorometry has been used to investigate the depolarizing rotations of 1,6-diphenyl-1,3,5-hexatriene (DPH) in isotropic solvents and in lipid bilayers. For DPH dissolved in isotropic solvents, there is a precise agreement between the observed and predicted values for maximum differential tangents, indicating that in these media DPH is a free isotropic rotator. In lipid bilayers the tangent defects (i.e., the differences between the calculated and the observed maximum differential tangents) are too large to be explained by anisotropy in the depolarizing rotations but are accounted for by hindered isotropic torsional motions for the fluorophore [Weber, G (1978) Acta Phys. Pol A 54, 173]. This theory describes the depolarizing rotations of the fluorophore by its rotational rate R (in radians/second) and the limiting fluorescence anisotropy (r) at times long compared with the fluorescence lifetime. Through the combined use of both steady-state anisotropy measurements and differential phase measurements, we have demonstrated that one may obtain unique solutions for both R and r. For DPH embedded in vesicles prepared from dimyristoyl-, dipalmitoyl-, and distearoylphosphatidylcholines, the depolarizing motions are highly hindered at temperatures below the transition temperature (Tc) but are unhindered above Tc. The apparent rotational rates of the probe do not change significantly at Tc. These data suggest that the changes observed in the steady-state anisotropy near Tc derive primarily from changes in the degree to which the probe's rotations are hindered, and only to a small extent from changes in rotational rate. For DPH embedded in bilayers that contained 25 mol % cholesterol, no clear transition occurred and the rotations appeared to be hindered at all temperatures. The rotational motions of DPH embedded in dioleolyphosphatidylcholine were found to be far less hindered, but the rotational rates were similar to those obtained in the saturated phosphatidylcholines. Finally, the data show that in an anisotropic environment, such as that of a lipid bilayer, steady-state fluorescence anisotropy measurements alone cannot yield quantitatively meaningful rotational rates. Extrapolation of steady-state aniosotropy data to the quantitation of membrane viscosity is therefore difficult, if not invalid; however, qualitative comparisons can be useful.  相似文献   

2.
The lipid-phase structures of brush border membrane vesicles (BBMV) and basolateral membrane vesicles (BLMV) isolated from rabbit renal cortex were compared by steady-state and phase-modulation measurements of diphenylhexatriene (DPH) and trans- and cis-parinaric acid (tPnA and cPnA) fluorescence. A temperature-scanning system was used which gave reproducible temperature profiles of steady-state and dynamic fluorescence parameters with a resolution of 0.1 degrees C. Steady-state anisotropy of DPH showed a triphasic dependence on temperature with slope discontinuities at 22 +/- 4 and 47 +/- 3 degrees C (BBMV) and at 23 +/- 2 and 48 +/- 1 degrees C (BLMV). At all temperatures, DPH anisotropy in BBMV was greater than that in BLMV. Ground-state heterogeneity analysis of tPnA and cPnA fluorescence lifetime data demonstrated the presence of long (greater than 12 ns) and short (less than 5 ns) lifetime components, interpreted in terms of solid-phase and fluid-phase lipid domains. The fraction of solid-phase phospholipid decreased from 0.9 to 0.1 for BBMV and from 0.7 to 0.3 in BLMV with increasing temperature (10-50 degrees C). In both membranes, tryptophan-PnA fluorescence energy-transfer measurements showed that membrane proteins were surrounded by a fluidlike phospholipid phase. These results demonstrate the inadequacy of steady-state DPH anisotropy data in defining the structural characteristics of complex biological membranes. Results obtained with the phase-sensitive parinaric acid probes demonstrate major differences in the phase structure of the two opposing cell membranes in both the bulk lipid and the lipid microenvironment around membrane proteins.  相似文献   

3.
R A Parente  B R Lentz 《Biochemistry》1985,24(22):6178-6185
We have investigated the behavior of 1-palmitoyl-2-[[2-[4- (6-phenyl-trans-1,3,5-hexatrienyl)phenyl]ethyl]carbonyl]-3-sn -phosphatidylcholine (DPHpPC) in synthetic, multilamellar phosphatidylcholine vesicles. This fluorescent phospholipid has photophysical properties similar to its parent fluorophore, diphenylhexatriene (DPH). DPHpPC preferentially partitioned into fluid phase lipid (Kf/s = 3.3) and reported a lower phase transition temperature as detected by fluorescence anisotropy than that observed by differential scanning calorimetry. Calorimetric measurements of the bilayer phase transition in samples having different phospholipid to probe ratios demonstrated very slight changes in membrane phase transition temperature (0.1-0.2 degree C) and showed no measurable change in transition width. Nonetheless, measurements of probe fluorescence properties suggested that DPHpPC disrupts its local environment in the membrane and may even induce perturbed probe-rich local domains below the phospholipid phase transition. Temperature profiles of steady-state fluorescence anisotropy, limiting anisotropy, differential tangent, and rotational rate were similar to those of DPH below the main lipid phase transition but indicated more restricted rotational motion above the lipid phase transition temperature. As for DPH, the fluorescence decay of DPHpPC could be described by either a single or double exponential both above and below the DPPC phase transition. The choice seemed dependent on the treatment of the sample. The intensity-weighted average lifetime of DPHpPC was roughly 1.5 ns shorter than that of DPH. In summary, the measured properties of DPHpPC and its lipid-like structure make it a powerful probe of membrane structure and dynamics.  相似文献   

4.
Steady-state fluorescence polarization measurements obtained with a flow cytometer were compared with those obtained with an SLM subnanosecond fluorometer. Measurements were made over time after exposure of HeLa cells to the membrane probe 1,6-diphenyl-1,3,5-hexatriene (DPH), 1-[4-(trimethylamino)phenyl]-6-phenyl-1,3,5-hexatriene (TMA-DPH), or [12-(9:anthroyloxy) stearate (12-AS). After 1 min, anisotropy values of 0.28 (DPH), 0.28 (TMA-DPH), and 0.21 (12-AS) were obtained. Thereafter, the anisotropy of DPH- and 12-AS-labelled cells rapidly decreased (0.18 and 0.12 after 5 min), while that of TMA-DPH-labelled cells changed only slightly (0.27 after 30 min), suggesting that DPH and 12-AS, unlike TMA-DPH, do not remain anchored in the HeLa plasma membrane, but translocate to more fluid environments inside the cell. These suggestions were confirmed by visual observation with fluorescence microscopy. There was no significant difference between the results obtained with the flow cytometer and those obtained with the fluorometer.  相似文献   

5.
In this study, we report the effect of cholesterol content on the dynamic and structural properties of a dimyristoyl-phosphatidylcholine and distearoyl-phosphatidylcholine mixture in large unilamellar vesicles. The range of cholesterol concentrations studied varied around approximately 33.3 mol%, where it has been postulated that an abrupt change in bilayer organization occurs. Steady-state fluorescence measurements demonstrated a typical behavior; at low temperatures in the main phase transition, the cholesterol concentration did not affect the gel phase, but at 37.5 °C (phase coexistence) and in the liquid crystalline phase, the presence of cholesterol produced an increase in the fluorescence anisotropy of DPH and the generalized polarization of Laurdan. The greater effect was observed in the liquid crystalline phase, in which the bilayer became a mixture of fluid-like and liquid-ordered phases. The results obtained at approximately 33.3 mol% of Cholesterol demonstrated that the Generalized Polarization of Laurdan, the DPH lifetime, the limiting anisotropy and the rotational correlation time, as well as the fluorescence quenching of DPH by TEMPO, are at maxima, while the fluorescence intensity of dehydroergosterol and the lipid solubility in TritonX-100 are at minima. These results correlate well with the hypothesis of domain segregation in the DMPC/DSPC/Cholesterol LUV system. In this context, we postulate that at 33.3 mol% of Cho, the proportion of ordered domains reaches a maximum.  相似文献   

6.
The chemical composition and the physical properties of lipoproteins (VLDL, LDL and HDL) were studied in two groups of patients: 14 healthy normolipidemic subjects and 15 type IIa familial hypercholesterolemic patients. The steady-state fluorescence anisotropy rs was estimated in lipoproteins by the fluorescence depolarization of two fluorescent probes: the DPH (1,6-diphenyl-1,3,5-hexatriene) and the TMA-DPH (1,4-trimethylammonium phenyl-6-1,3,5-hexatriene). A structured order parameter S was calculated from the DPH fluorescence anisotropy. The flow activation energies were calculated for LDL and HDL from both groups from the Arrhenius plots (log r DPH versus 1/T). By using TNBS (trinitrobenzene sulfonic acid) as a distance control quencher, the two probes were located in the outer shell of LDL. In HDL, TMA-DPH remained at the surface of the particles, while DPH was more deeply embedded in the lipid core. There was no difference in the physico-chemical properties of VLDL between the two groups studied. DPH fluorescence anisotropies were significantly increased in LDL and HDL from the hypercholesterolemic group compared to the control particles (P less than 0.05 and P less than 0.01, respectively). In LDL this modification of the fluorescence anisotropy can be related to a change in the lipid composition of particles. LDL from hypercholesterolemic patients contained significantly less triacylglycerol (P less than 0.01) and more cholesteryl ester (N.S.). Their cholesteryl ester to triacylglycerol ratio was significantly higher. In HDL, there was no difference in chemical composition between the two groups. The increase in DPH fluorescence anisotropy can be related to the presence of smaller particles in HDL from HC group. No difference was noted in the TMA-DPH fluorescence anisotropy at 37 degrees C in the LDL from the two groups. In contrast, TMA-DPH fluorescence anisotropy in HDL from hypercholesterolemic group was significantly higher than in control HDL. The flow activation energy of DPH was also significantly higher in both LDL and HDL from the hypercholesterolemic group than in control group particles. In both LDL and HDL from the control group, DPH fluorescence anisotropy was negatively correlated with TG/protein and TG/PL ratios and positively correlated with the CE/TG ratio. No correlation was observed between lipid composition and DPH fluorescence anisotropy values in hypercholesterolemic particles. The modification in fluidity parameters, especially the increase in the flow activation energies in LDL and HDL from hypercholesterolemic patients, could lead to a restriction of cholesterol movements in these particles. From a physiological point of view, this could represent a loss of functional capacity.  相似文献   

7.
Charged anesthetics selectively alter plasma membrane order   总被引:5,自引:0,他引:5  
W D Sweet  W G Wood  F Schroeder 《Biochemistry》1987,26(10):2828-2835
Although indirect evidence supporting differential lipid fluidity in the two monolayers of plasma membranes has accumulated, unambiguous demonstration of this difference has been difficult to obtain. In the present study, the fluorescent probe 1,6-diphenyl-1,3,5-hexatriene (DPH), selective quenching of fluorescence by trinitrophenyl groups, and differential polarized phase fluorescence techniques were used to directly examine the static (order) and dynamic (rotational rate) components of lipid motion in the exofacial and cytofacial leaflets of LM fibroblast plasma membranes. The limiting anisotropy (0.137), the order parameter (0.590), and the rotational relaxation time (1.20 ns) of DPH in the plasma membranes (inner plus outer leaflet) indicated rapid but restricted probe motion in the lipid environment. However, the statics and dynamics of DPH motion in the individual monolayers were significantly (p less than 0.025) different. The limiting anisotropy, order parameter, and rotational relaxation time of DPH in the cytofacial monolayer were 0.036, 0.08, and 0.16 ns, respectively, greater than calculated for the exofacial monolayer of the LM plasma membrane. At appropriate concentrations, phenobarbital and, to a lesser degree, pentobarbital preferentially reduced the limiting anisotropy of DPH calculated for the exofacial leaflet while prilocaine reduced the limiting anisotropy of DPH in the cytofacial leaflet of LM fibroblast plasma membranes. In contrast, the putative cytofacial anesthetic procaine failed to show any preference for either leaflet. Arrhenius plots of DPH fluorescence in LM plasma membranes showed a prominent characteristic break point near 30-32 degrees C. Phenobarbital, pentobarbital, and procaine did not affect this break point while prilocaine selectively abolished it. The break point was therefore assigned to the inner monolayer of the LM plasma membrane.  相似文献   

8.
The fluorescence anisotropy (r) of diphenylhexatriene (DPH) was measured in different preparations (bovine spinal cord phosphatidylserine liposomes, rat brain microsomes, liposomes made with rat brain microsomal lipid having different phospholipid:cholesterol ratios) at temperatures ranging from 10° to 55°C. Phosphatidylserine liposomes exhibited an exponential relationship of rversus temperature, whereas the relationship shown by microsomes and liposomes prepared with microsomal lipid extracts was a linear one. The removal of protein and high phospholipid:cholesterol ratios decreased the slope of the lines (fluidity increased), although the intercept was unaffected. This means that differences were better appreciated at high temperatures and were well evident at 37°C. Acetyl-l-carnitine decreased r in rat brain microsomes and in liposomes made with microsomal lipids with different phospholipid:cholesterol ratios. The fluidifying effect of acetyl-l-carnitine was mild but statistically significant and could explain, at least in part, the data reported in the literature of acetyl-l-carnitine acting on some parameters affected by ageing. Besides, acetyl-l-carnitine seemed to oppose the changes of viscosity due to lipid peroxidation, which has been reported to increase in ageing and dementia.l-carnitine shares the properties of its acetyl ester, but only in part.Abbreviations DPH diphenylhexatriene - HEPES 4-(2-hydroxyethyl-l-piperazineethansulfonic) acid - r fluorescence anisotropy - SHB sucrose-HEPES-buffer (0.32 M sucrose, 2 mM HEPES, pH 7.0)  相似文献   

9.
The apparent steady-state fluorescence anisotropy of DPH- or TMA-DPH-labeled washed rat platelets is strongly affected by factors that also influence the turbidity by these platelet suspensions. Sonicated preparations from platelet lipids have a low turbidity and give anisotropy values which are hardly affected by the experimental conditions. We studied the effect of four high-fat diets on membrane fluidity, lipid composition and activation tendency of washed platelets. The diets contained 50 energy% of oils with different levels of saturated and (poly)unsaturated fatty acids. Only small diet-induced differences in DPH fluorescence anisotropy were found, which were comparable for intact platelets and platelet lipids. These differences were unrelated to the degree of saturation of the dietary fatty acids. Platelets from rats fed mainly saturated fatty acids differed significantly from other diet groups in a higher unsaturation degree of phospholipids and a lower cholesterol/phospholipid ratio, but this was not detected by DPH in terms of decreased anisotropy. These platelets aggregated less than other platelets in response to thrombin or collagen. The lower response to collagen persisted in indomethacin-treated platelets activated with the thromboxane A2 mimetic U46619, indicating a different sensitivity of these platelets for thromboxane A2. We conclude that in rat platelets: (a) the overall membrane fluidity and phospholipid unsaturation degree are subject to strong homeostatic control; (b) steady-state anisotropy with DPH or TMA-DPH label is inadequate to reveal subtile changes in lipid profile; (c) changes in platelet responsiveness to thrombin and thromboxane A2, rather than (plasma) membrane fluidity, determine the effect of dietary fatty acids on platelet aggregation.  相似文献   

10.
Changes in the molecular organization of membranes in pericarp cells of ripening tomato fruit were examined by fluorescence depolarization after labeling with fluorescent lipid-soluble probes. The fluorescent labels were partitioned into isolated protoplasts and purified plastids from fruit at various stages of senescence. Values for steady-state anisotropy (rss) of 1,6-diphenyl-1,3,5-hexatriene (DPH)-labeled protoplasts rose progressively during the early stages of ripening over a time frame that overlapped the climacteric rise in ethylene production. This can be interpreted as reflecting a decrease in the lipid fluidity of primarily plasma membrane. By contrast, there was no significant change during ripening in rss for plastid membranes labeled with DPH, 1-[4-trimethylamino)phenyl]-6-phenyl-1,3,5-hexatriene (TMA-DPH), and cis- or trans-parinaric acid. Nor was there any change during ripening in the limiting fluorescence anisotropy (roo) and order parameter (S) for plastids labeled with DPH or TMA-DPH, parameters that are corrected for any differences in lifetime. Some degree of lifetime heterogeneity, possibly reflecting structurally distinct domains, was discerned in both young and senescent plastids that had been labeled with DPH or TMA-DPH, but this also did not change as ripening progressed. Thus membranes of the pericarp cells sustain different fates as the tomato fruit ripens, implying that there are distinguishable mechanisms of membrane deterioration in senescing tissues.  相似文献   

11.
Nanosecond decays of the fluorescence anisotropy, r, were studied for the emission of 1,6-diphenyl-1,3,5-hexatriene (DPH) embedded in a series of mixed multilamellar liposomes containing egg yolk phosphatidylcholine, phosphatidylethanolamine and cholesterol in varying molar ratios, as well as in membranes of intact cells and in virus envelopes. The relative contributions of the fast and the infinitely slow decaying component to the steady-state value r, of the fluorescence anisotropy were very similar for artifical and biological membranes. Angles, theta, of the cone, by which the motion of the fluorescent molecule is limited, were calculated from the intensity of the infinitely slow decaying anisotropy component and compared with steady-state fluorescence anisotropies and with 'microviscosities', (eta). An increase in (eta) from 1.5 to 5.2 P in our systems was accompanied by a decrease in theta from 49 degrees to 30 degrees while the decrease in the mean motional relaxation times, phi f, of the label molecule was not more than 1 ns and due mainly to changes in the potential, by which the diffusion of DPH in the membrane is restricted. From these observations we conclude that differences in the steady-state fluorescence anisotropy and in 'microviscosities' of cholesterol-containing membranes (r greater than 0.15) represent changes in the degree of static orientational constraint rather than changes in diffusion rates of the label.  相似文献   

12.
Fluorescence steady-state anisotropy and phase-modulation lifetime techniques have been utilized to study the interactions of pyrethroid compounds with fluid-phase phosphatidylcholine membranes containing the polypeptide gramicidin. This polypeptide is considered to be a model of hydrophobic regions of cellular integral membrane proteins. The pyrethroids disorder lipid packing in cellular membranes and gel-phase liposomes but do not disorder lipid packing in fluid-phase lipid (Stelzer, K.J. and Gordon, M.A. (1984) J. Immunopharmacol. 6, 381-410; (1985) Biochim. Biophys. Acta 812, 361-368) Irrespective of liposomal size, gramicidin incorporation resulted in a substantial increase in anisotropy of the fluorescent probe, 1,6-diphenyl-1,3,5-hexatriene (DPH), in fluid phase lipid. In the absence of gramicidin, permethrin and three other pyrethroids, allethrin, cypermethrin and fenpropathrin, increased DPH anisotropy. In these fluid phase systems, as the protein:lipid ratio was increased, the extent of the pyrethroid-mediated increase in fluorescence anisotropy diminished. Also, the pyrethroids shortened DPH fluorescence lifetimes. At high gramicidin:lipid ratios, permethrin substantially lowered anisotropy in the fluid phase lipid, relative to controls. The data suggest that pyrethroids disturb fluid-phase lipids which have been promoted to a relative state of order by proximity to an integral membrane protein. This type of order is one which is represented by DPH fluorescence anisotropy. A model based on these results is proposed to explain the effects of pyrethroids on lipid packing order in cellular membranes, as determined by DPH fluorescence anisotropy.  相似文献   

13.
Summary Static polarization and differential polarized phase fluorimetry studies on rat renal cortical brush border (BBM) and basolateral membranes (BLM) were undertaken to determine the membrane components responsible for differences in BBM and BLM fluidity, whether these differences were due to the order or dynamic components of membrane fluidity and if a fluidity gradient existed within the bilayer. Surface membrane proteins rigidified both BBM and BLM fluidity. Neutral lipid extraction, on the other hand, caused a larger decrease in BBM than BLM fluorescence polarization (0.104vs. 0.60,P<0.01) using diphenyl hexatriene (DPH). Cholesterol addition to phospholipid fractions restored membrane fluidity to total lipid values in both BBM and BLM phospholipids. The response to cholesterol in the BBM was biphasic, while the BLM response was linear. Lateral mobility, quantitated using dipyrenylpropane, was similar in both BBM and BLM fractions at 35°C. BBM and BLM differed primarily in the order component of membrane fluidity as DPH-limiting anisotropy (r ) (0.212vs. 0.154,P<0.01) differed markedly between the two membrane fractions. The two membrane components also differed with respect to 2 and 12-anthroyloxy stearate (2-AS, 12-AS) probes, indicating a difference in the dynamic component of membrane fluidity may also be present. DPH and 12-As probes were also used to quantitate inner core membrane fluidity and showed the BBM was less fluid than the BLM for intact membranes, total lipid extracts and phospholipids. Results obtained using the surface membrane probes trimethylammonium-DPH (TMA-DPH) and 2-AS suggested a fluidity gradient existed in both BBM and BLM bilayers with the inner core being more fluid in both membranes. These data indicate cholesterol is in large part responsible for fluidity differences between BBM and BLM and that these membranes, while clearly differing in the order component of membrane fluidity, may also difer in the dynamic component as well.  相似文献   

14.
In the present study, the in vitro effect of polyphenol rich plant extract, flavonoid--Pycnogenol (Pyc), on erythrocyte membrane fluidity was studied. Membrane fluidity was determined using 1-[4-trimethyl-aminophenyl]-6-phenyl-1,3,5-hexatriene (TMA-DPH), 1,6-diphenyl-1,3,5-hexatriene (DPH) and 12-(9-anthroyloxy) stearic acid (12-AS) fluorescence anisotropy. After Pyc action (50 microg/ml to 300 microg/ml), we observed decreases in the anisotropy values of TMA-DPH and DPH in a dose-dependent manner compared with the untreated erythrocyte membranes. Pyc significantly increased the membrane fluidity predominantly at the membrane surface. Further, we observed the protective effect of Pyc against lipid peroxidation, TBARP generation and oxidative hemolysis induced by H2O2. Pyc can reduce the lipid peroxidation and oxidative hemolysis either by quenching free radicals or by chelating metal ions, or by both. The exact mechanism(s) of the positive effect of Pyc is not known. We assume that Pyc efficacy to modify effectively some membrane dependent processes is related not only to the chemical action of Pyc but also to its ability to interact directly with cell membranes and/or penetrate the membrane thus inducing modification of the lipid bilayer and lipid-protein interactions.  相似文献   

15.
The emission and polarization spectra of 1-phenyl-3-(2-naphthyl)-2-pyrazoline (PNP) in various environments were studied. Compared to the widely used orientational membrane probe 1,6-diphenylhexatriene (DPH), PNP is five times less photolabile and since its fluorescence emission maximum is at longer wavelengths max ≈ 445 nm), it is more suitable for use with intact erythrocytes. The limiting fluorescence anisotropy of PNP is 0.385. In erythrocyte ghosts, the steady-state emission anisotropy of PNP is a decreasing function of wavelength and its temperature dependence parallels that of DPH, dropping from 0.298 at 2°C to 0.185 at 38°C when averaged between 420 and 470 nm.  相似文献   

16.
S Kawato  K Kinosita  A Ikegami 《Biochemistry》1977,16(11):2319-2324
Molecular motions in liposomes of dipalmitoyl-phosphatidylcholine (DPPC) were studied by nanosecond fluorescence techniques. As a fluorescent probe for the hydrocarbon region, 1,6-diphenyl-1,3,5-hexatriene (DPH) was used. Time courses of fluorescence intensity IT(t) and emission anisotropy r(t) of DPH embedded in DPPC liposomes were measured at various temperatures. The value of the fluorescence lifetime tau obtained froma single exponential decay of IT(t) was somewhat higher than that in liquid paraffin below the transition temperature Tt and decreased above Tt. Higher values of tau below Tt indicate the almost complete hydrophobic environment. The decay curves of r(t) were separated into two phases: an initial fast decreasing phase of the order of one nanosecond and a second almost constant phase. This indicates that the orientational motion of DPH in the hydrocarbon region is described by a wobbling diffusion restricted by a certain anisotropic potential. The results were analyzed on the model that the wobbling diffusion is confined in a cone with a uniform diffusion constant. Though temperature dependence of the cone angle was sigmoidal, that of the wobbling diffusion constant was like the exponential function. The change in the cone angle at Tt was sharper than that in the wobbling diffusion constant at Tt. Estimated values of the viscosity in the cone were an order of magnitude smaller than the values of "microviscosity" which were estimated from the steady-state emission anisotropy without considering the restrictions on the rotational motion.  相似文献   

17.
The influence of mature lysozyme fibrils on the structural and physical properties of model membranes composed of phosphatidylcholine (PC) and its mixtures with cardiolipin (CL) (10 mol%) and cholesterol (Chol) (30 mol%) was studied using fluorescent probes DPH, pyrene, Laurdan and MBA. Analysis of pyrene fluorescence spectra along with the measurements of DPH fluorescence anisotropy revealed that the structure of hydrocarbon chains region of lipid bilayer is not affected by the fibrillar aggregates of lysozyme. In contrast, probing the membrane effects by Laurdan and MBA showed the rise of both the generalized polarization of Laurdan and the MBA fluorescence anisotropy, suggesting that amyloid protein induces reduction of bilayer hydration and increase of lipid packing in the interfacial region of model membranes.  相似文献   

18.
The steady state fluorescence anisotropy (rs) of 1-acyl-2-cis parinaroyl phosphatidylcholine (PnPC) was compared with that of diphenylhexatriene (DPH) in a variety of model- and biological membrane systems. The fluorescence anisotropy of both probes responded similarly to temperature changes and variations in the acyl chain composition in phosphatidylcholine (PC) liposomes. The presence of proteins and cholesterol increased rs for both DPH and PnPC in the biological membranes as compared to the isolated polar membrane lipids. Comparison of DPH and PnPC in dipalmitoyl-PC-liposomes with and without 50 mol% cholesterol, showed at temperatures above the phase transition of pure dipalmitoyl-PC the presence of cholesterol increased the rs-value for DPH strongly, whereas the rs-value for PnPC was much less affected. In the cholesterol-rich erythrocyte membrane as well as in microsomes from Morris hepatoma 7787, which have an increased cholesterol content as compared to normal rat liver microsomes, the rs of DPH was higher than that of PnPC. No large differences between the rs-values of both probes were evident in the normal cholesterol-poor rat liver microsomes. These effects are discussed in terms of structural differences between the probes and variation of cholesterol content. Alterations in the fatty acid composition of PC present in human erythrocyte membranes were introduced with the aid of a PC-specific transfer protein. Fluorescence anisotropy values of both probes hardly changed upon enrichment of the red cell membrane with either dipalmitoyl PC or 1-palmitoyl-2-arachidonyl PC.  相似文献   

19.
Phospholipid-sterol interactions were investigated using parinaric acid fluorescence spectroscopy. Cholesterol and cholesterol analogues which were modified in the sterol nucleus or side chain were added at 50 mol % to multilamellar vesicles of model phospholipids selected to be representative of major components in an LM cell plasma membrane. These included sphingomyelins and saturated and monounsaturated phosphatidylcholines and phosphatidylethanolamines. Based on the changes in cis-parinaric acid steady-state fluorescence polarization observed with addition of sterol, 50 mol % cholesterol abolished the phase transition of all the model phospholipids. Dihydrocholesterol and trans-22-dehydrocholesterol behaved like cholesterol in the two systems studied. 24-Methylcholesterols interacted well with all phospholipids except phosphatidylethanolamine which contained an unsaturated fatty acid. 24-Alkyl,trans-22-dehydrocholesterols abolished the phase transition in only two systems: sphingomyelins and phosphatidylcholines possessing relatively short saturated acyl chains. Since steady-state anisotropy is a function of fluorescence lifetime, rotational diffusion rates, and limiting anisotropy, we determined these parameters for two of the phospholipid systems. The results show that steady-state anisotropy values for phospholipid-sterol interactions correlate closely with limiting anisotropy and to a lesser extent with rotational relaxation time. The behavior of the sterols in the model phospholipids are used to interpret 1) fluorescence polarization measurements made with phospholipids extracted from LM cell plasma membranes, and 2) changes in membrane lipid composition which accompany growth of LM cells on various sterols.  相似文献   

20.
Alterations of membrane lipid biophysical properties of sensitive A549 and resistant A549/DDP cells to the Cis-dichlorodiammine platinum (Cisplatin) were performed by measurements of fluorescence and flow cytometry approaches using fluorescence dyes of DPH, N-AS and Mero-cyanine 540 (MC 540) respectively. Fatty acids of membrane lipid of the two cell lines were analyzed by gas chromatography. The results indicated clearly that fluorescence polarization (P) of the DPH probe is 0.169 for the sensitive A549 cell and 0.194 for the resistant A549/DDP cells. Statistical analysis showed significant difference between the two cell lines. The polarizations of 2-AS and 7-AS which reflect the fluidity of surface and middle of lipid bilayer are 0.134 and 0.144 for the sensitive A549 cells as well as 0.171 and 0.178 for the resistant A549/DDP cells respectively, but there is no significant difference of the polarization of 12-AS between the two cell lines. This shows that alterations of the membrane fluidity of both  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