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.
L Powers  B M Kincaid 《Biochemistry》1989,28(10):4461-4468
Differences in the methods of analysis of X-ray absorption data used by Powers et al. [Powers, L., Blumberg, W. E., Chance, B., Barlow, C., Leigh, J., Jr., Smith, J., Yonetani, T., Vik, S., & Peisach, J. (1979) Biochim. Biophys. Acta 547, 520-538; Powers, L., Chance, B., Ching, Y., & Angiolillo, P. (1981) Biophys. J. 34, 465-498] and Scott et al. [Scott, R., Schwartz, J., & Cramer S. (1986) Biochemistry 25, 5546-5555] are clarified. In addition, we compare the X-ray absorption data and results for resting cytochrome c oxidase reported by both groups using the same analysis method and conclude apart from any assumptions that the data are not identical.  相似文献   

2.
The model of Cruzeiro-Hansson et al. (Biochim. Biophys. Acta (1989) 979, 166-1176) for lipid-cholesterol bilayers at low cholesterol concentrations is used to predict the thermodynamic properties and the passive ion permeability of lipid bilayers as a function of acyl-chain length and cholesterol concentration. Numerical simulations based on the Monte Carlo method are used to determine the equilibrium state of the system near the main gel-fluid phase transition. The permeability is calculated using an ansatz which relates the passive permeability to the amount of interfaces formed in the bilayer when cholesterol is present. The model predicts at low cholesterol contents an increase in the membrane permeability in the transition region both for increasing cholesterol concentration and for decreasing chain length at a given value of the reduced temperature. This is in contrast to the case of lipid bilayers containing high cholesterol concentrations where the cholesterol strongly suppresses the permeability. Experimental results for the Na+ permeability of C15PC and DPPC (C16PC) bilayers containing cholesterol are presented which confirm the theoretical predictions at low cholesterol concentrations.  相似文献   

3.
The dynamics of lipid hydrocarbon chains in phosphatidylcholine (dimyristoyl- or dipalmitoyl-) and cholesterol/dimyristoylphosphatidylcholine membranes were investigated by nanosecond time-resolved fluorescence depolarization measurements on a lipophilic fluorescent probe 1,6-diphenyl-1,3,5-hexatriene embedded in the membranes. In the pure lipid membranes, both the range (amplitude) and the rate of the wobbling motion of the probe increased sigmoidally with temperature reflecting the thermotropic phase transition of the lipid. The rise in the rate slightly preceded the increase in the range, suggesting that the fluctuation of lipid chains is activated to a high level before the ordered array of chains melt into the liquid-crystalline phase. Above the transition temperature, incorporation of cholesterol resulted in a dramatic decrease in the range of wobbling motion while the rate remained high. Below the transition, on the other hand, cholesterol had little effect on the range, whereas the rate was greatly increased. These effects of cholesterol are remarkably similar to the effects of cytochrome oxidase on lipid chain dynamics (Kinosita, K., Jr., Kawato, S., Ikegami, A., Yoshida, S. and Orii, Y. (1981) Biochim. Biophys. Acta 647, 7–17).  相似文献   

4.
The binding of cytochrome b5 to single-walled liposomes of egg phosphatidylcholine was inhibited by the presence of cholesterol in the lipid bilayer under conditions where a limited amount of liposomes was incubated with the cytochrome. Since similar conditions seem to apply for the binding of cytochrome b5 to erythrocyte ghosts, this observation supports the conclusion of Enomoto and Sato (Enomoto, K. and Sato, R. (1977) Biochim. Biophys. Acta 466, 136–147) that the localization of cholesterol on the outer surface of the ghost membrane prevents the binding of cytochrome b5 to this surface. The finding reported by Roseman et al. (Roseman, M.A., Holloway, P.W. and Calabro, M.A. (1978) Biochim. Biophys. Acta 507, 552–556) that cholesterol did not prevent the cytochrome binding to phosphatidylcholine liposomes in the presence of a large excess of liposomes could be confirmed in the present study, but this does not contradict the abovementioned conclusion.  相似文献   

5.
We recently reported the equilibrium maximum solubility of cholesterol in a lipid bilayer, chi*chol, to be 0.66 in four different phosphatidylcholines, and 0.51 in a phosphatidylethanolamine (Huang, J.,J.T. Buboltz, and G. W. Feigenson. 1999. Biochim. Biophys. Acta. in press). Here we present a model of cholesterol-phospholipid mixing that explains these observed values of chi*chol. Monte Carlo simulations show that pairwise-additivity of nearest-neighbor interactions is inadequate to describe all the chi*chol values. Instead, if cholesterol multibody interactions are assigned highly unfavorable energy, then jumps occur in cholesterol chemical potential that lead to its precipitation from the bilayer. Cholesterol precipitation is most likely to occur near three discrete values of cholesterol mole fraction, 0.50, 0.57, and 0.67, which correspond to cholesterol/phospholipid mole ratios of 1/1, 4/3, and 2/1, respectively. At these solubility limits, where cholesterol chemical potential jumps, the cholesterol-phospholipid bilayer mixture forms highly regular lipid distributions in order to minimize cholesterol-cholesterol contacts. This treatment shows that dramatic structural and thermodynamic changes can occur at particular cholesterol mole fractions without any stoichiometric complex formation. The physical origin of the unfavorable cholesterol multibody interaction is explained by an "umbrella model": in a bilayer, nonpolar cholesterol relies on polar phospholipid headgroup coverage to avoid the unfavorable free energy of cholesterol contact with water. Thus, at high cholesterol mole fraction, this unfavorable free energy, not any favorable cholesterol-phospholipid interaction, dominates the mixing behavior. This physical origin also explains the "cholesterol condensing effect" and the increase in acyl chain order parameter in cholesterol-phospholipid mixtures.  相似文献   

6.
According to the liquid hydrocarbon model, the lipid bilayer is viewed simply as a thin slice of bulk hydrocarbon liquid. This allows the water permeability of the bilayer to be calculated from bulk properties. In this paper the prediction of the liquid hydrocarbon model is compared with the known water permeability coefficient of the glycerol monoolein/n-hexadecane bilayer (Fettiplace, R. (1978) Biochim. Biophys. Acta 513, 1–10). As the alkyl chain of glycerol monoolein is equivalent to 8-heptadecene, the water permeability coefficient of 8-heptadecene/n-hexadecane mixtures was measured for temperatures between 20 and 35°C. The mole fraction of n-hexadecane in the bulk liquid was chosen at each temperature to match the known mole fraction of n-hexadecane in the bilayer (White, S. (1976) Nature 262, 421–422). The predicted water permeability coefficient agrees with the measured value at 32°C but is 40% above the measured value at 20°C. The apparent activation energy predicted by the liquid hydrocarbon model is 9.0 ± 0.3 kcal/mol, while the measured value is 14.2 ± 1.0 kcal/mol. The failure of the liquid hydrocarbon model probably results from a different molecular organization of the hydrocarbon chains in the bilayer and in the bulk liquid.  相似文献   

7.
A L MacDonald  D A Pink 《Biochemistry》1987,26(7):1909-1917
We have developed a model of glycophorin in a phospholipid bilayer membrane in order to study the thermodynamics of this system and to understand the detailed behavior of recent calorimetric data. We assume that the larger glycophorin polar group can be considered as either adopting a pancakelike conformation at the bilayer interface (D state) or be directed generally away from the interface (U state) [Ruppel, D., Kapitza, H.G., Galla, H.J., Sixl, F., & Sackmann, E. (1982) Biochim. Biophys. Acta 692, 1-17]. Lipid hydrocarbon chains are described either as excited (e state) with high energy and relatively many gauche conformers or as generally extended (g state) with low energy. We performed a Monte-Carlo simulation using the Glauber and Kawasaki procedures on a triangular lattice which represents the plane of half of the bilayer. Lattice sites can be occupied either by lipid hydrocarbon chains or by model glycophorin alpha-helical hydrophobic cores. The states D and U are represented by hexagons of different sizes in the plane of the lattice, and the hard core repulsion between two such polar groups is accounted for by forbidding hexagon-hexagon overlap. We have studied the effect of having the glycophorin polar group interact in various ways with the lipid bilayer. We find that the protein polar group in its D state interacts, either directly or indirectly, with the lipid bilayer so as to reduce the effective lateral pressure acting on the lipid hydrocarbon chains by about 3 dyn/cm. Polar groups in their U states do not reduce this lateral pressure.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The water permeability of the lipid bilayer can be used as a probe of membrane structure. A simple model of the bilayer, the liquid hydrocarbon model, views the membrane as a thin slice of bulk hydrocarbon liquid. A previous study (Petersen, D. (1980) Biochim. Biophys. Acta 600, 666–677) showed that this model does not accurately predict the water permeability of the monoolein/n-hexadecane bilayer: the measured activation energy for water permeation is 50% above the predicted value. From this it was inferred that the hydrocarbon chains in the lipid bilayer are more ordered than in the bulk hydrocarbon liquid. The present study tests the liquid hydrocarbon model for the monoolein/triolein bilayer, which has been shown to contain very little triolein in the plane of the membrane (Waldbillig, R.C. and Szabo, G. (1979) Biochim. Biophys. Acta 557, 295–305). Measurements of the water permeability coefficient of the bilayer are compared with predictions of the liquid hydrocarbon model based on measurements of the water permeability coefficient of bulk 8-heptadecene. The predicted and measured values agree quite closely over the temperature range studied (15–35°C): the predicted activation energy is 11.1±0.2 kcal/mol, whereas the measured activation energy for the bilayer is 9.8±0.7 kcal/mol. This close agreement is in contrast with the monoolein/n-hexadecane results and suggests that, insofar as water permeation is concerned, the liquid hydrocarbon model quite closely represents the monoolein/triolein bilayer.  相似文献   

9.
We examined the action of porcine pancreatic and bee-venom phospholipase A2 towards bilayers of phosphatidylcholine as a function of several physical characteristics of the lipid-water interface. 1. Unsonicated liposomes of dimyristoyl phosphatidylcholine are degraded by both phospholipases in the temperature region of the phase transition only (cf. Op den Kamp et al. (1974) Biochim. Biophys. Acta 345, 253--256 and Op den Kamp et al. (1975) Biochim. Biophys. Acta 406, 169--177). With sonicates the temperature range in which hydrolysis occurs is much wider. This discrepancy between liposomes and sonicates cannot be ascribed entirely to differences in available substrate surface. 2. Below the phase-transition temperature the phospholipases degrade dimyristoyl phosphatidylcholine single-bilayer vesicles with a strongly curved surface much more effectively than larger single-bilayer vesicles with a relatively low degree of curvature. 3. Vesicles composed of egg phosphatidylcholine can be degraded by pancreatic phospholipase A2 at 37 degrees C, provided that the substrate bilayer is strongly curved. The bee-venom enzyme shows a similar, but less pronounced, preference for small substrate vesicles. 4. In a limited temperature region just above the transition temperature of the substrate the action of both phospholipases initially proceeds with a gradually increasing velocity. This stimulation is presumably due to an increase of the transition temperature, effectuated by the products of the phospholipase action. 5. Structural defects in the substrate bilayer, introduced by sonication below the phase-transition temperature (cf. Lawaczeck et al. (1976) Biochim. Biophys. Acta 443, 313--330) facilitate the action of both phospholipases. The results lead to the general conclusion that structural irregularities in the packing of the substrate molecules facilitate the action of phospholipases A2 on phosphatidylcholine bilayers. Within the phase transition and with bilayers containing structural defects these irregularities represent boundaries between separate lipid domains. The stimulatory effect of strong bilayer curvature can be ascribed to an overall perturbation of the lipid packing as well as to a change in the phase-transition temperature.  相似文献   

10.
The ribose-modified chromophoric and fluorescent analog of ATP 2′,3′-O-(2,4,6-trinitrocyclohexadienylidene) adenosine 5′-triphosphate (TNP-ATP) has been synthesized previously (Hiratsuka, T., and Uchida, K. (1973) Biochim. Biophys. Acta 320, 635–647 and Hiratsuka, T. (1976) Biochim. Biophys. Acta 453, 293–297). In the present study, four TNP-derivatives of ATP, ADP, AMP and adenosine were synthesized and compared for several chemical, spectral and enzymatic properties. Their visible absorption and fluorescent properties were found to be quite similar. Visible absorption and fluorescence spectra of TNP-derivatives were sensitive to solvent polarity. TNP-adenosine and TNP-AMP showed considerable substrate activities with adenosine deaminase and alkaline phosphatase, respectively. TNP-ATP proved to be an excellent substitute for ATP in adenylate kinase and myosin ATPase systems. The results indicate that these analogs are useful as chromophoric and fluorescent probes for hydrophobic regions in adenine nucleoside and nucleotide requiring enzymes.  相似文献   

11.
The fusion of two black lipid membranes results in the formation of peculiar bilayer lipid tubes (‘cylindrical’) membranes (Neher, E. (1974) Biochim. Biophys. Acta 373, 328–336 and Melikyan, G.B., Abidor, L.G., Chernomordik, L.V. and Chailakhyan, L.M. (1983) Biochim. Biophys. Acta 730, 395–398). The mechanical stability of such tubes has been investigated experimentally and theoretically. With increasing hydrostatic pressure on the outside of the tube the radius of its middle part decreases. After this radius has reached a critical value, which constitutes 0.55 of the radius of the tube base, there occurs a collapse of the tube and its disintegration into two planar bilayers (fission). Expressions are obtained which relate the transmembrane difference of the hydrostatic pressure, causing the collapse, to the geometrical characteristics of the tube (its length and the radius of its base) and to the tension of the lipid bilayer. A method for measuring the membrane tension is proposed on the basis of the phenomenon considered.  相似文献   

12.
In an accompanying publication by Duckwitz-Peterlein, Eilenberger and Overath ((1977) Biochim. Biophys. Acta 469,311--325) it is shown that the exchange of lipid molecules between negatively charged vesicles consisting of total phospholipid extracts from Escherichia coli occurs by the transfer of single lipid monomers or small micelles through the water. Here a kinetic interpretation is presented in terms of a rate constant, k--, for the escape of lipid molecules from the vesicle bilayer into the water. The evaluated rate constants are kP- = (0.86 +/- 0.05) - 10(-5) S-1 and ke- = (1.09 +/- 0.13) - 10(-6) s-1 for phospholipid molecules with trans-delta 9-hexadecenoate and trans-delta 9-octadecenoate, respectively, as the predominant acyl chain component. The rate constants are discussed in terms of the acyl chain and polar head group composition of the lipids.  相似文献   

13.
We have previously suggested the existence of two distinctive states of cholesterol in erythrocyte and lymphoma cell membranes as revealed by high- and low-affinity binding sites for theta-toxin of Clostridium perfringens [Ohno-Iwashita, Y., Iwamoto, M., Mitsui, K., Ando, S., & Nagai, Y. (1988) Eur. J. Biochem. 176, 95-101; Ohno-Iwashita, Y., Iwamoto, M., Ando, S., Mitsui, K., & Iwashita, S. (1990) Biochim. Biophys. Acta 1023, 441-448]. To understand factor(s) which determine membrane cholesterol heterogeneity, we analyzed toxin binding to large unilamellar liposomes composed of cholesterol and phospholipids (phosphatidylcholine/phosphatidylglycerol = 82:18, mol/mol). Liposomes containing phospholipids with 18-carbon hydrocarbon chains at both positions 1 and 2 of the glycerol have both high- and low-affinity toxin-binding sites with Kd values similar to those of intact erythrocytes, whereas liposomes with hydrocarbon chains containing 16 or fewer carbons at either position 1 or 2 have only low-affinity toxin-binding sites. The cholesterol/phospholipid ratio, in addition to the length of phospholipid hydrocarbon chain, also determines the number of toxin-binding sites, indicating that at least these two factors determine the topology of membrane cholesterol by creating distinctively different affinity sites for the toxin. Since theta-toxin binding detects specific populations of membrane cholesterol that are not detectable by the measurements of susceptibility to cholesterol oxidase and cholesterol desorption from membranes, the toxin could provide a unique probe for studying the organization of cholesterol in membranes.  相似文献   

14.
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   

15.
P J Spooner  D M Small 《Biochemistry》1987,26(18):5820-5825
Triacylglycerols are the major substrates for lipolytic enzymes that act at the surface of emulsion-like particles such as triglyceride-rich lipoproteins, chylomicrons, and intracellular lipid droplets. This study examines the effect of cholesterol on the solubility of a triacylglycerol, triolein, in phospholipid surfaces. Solubilities of [carbonyl-13C]triolein in phospholipid bilayer vesicles containing between 0 and 50 mol % free cholesterol, prepared by cosonication, were measured by 13C NMR. The carbonyl resonances from bilayer-incorporated triglyceride were shifted downfield in the 13C NMR spectra from those corresponding to excess, nonincorporated material. This enabled solubilities to be determined directly from carbonyl peak intensities at most cholesterol concentrations. The bilayer solubility of triolein was inversely proportional to the cholesterol/phospholipid mole ratio. In pure phospholipid vesicles the triolein solubility was 2.2 mol %. The triglyceride incorporation decreased to 1.1 mol % at a cholesterol/phospholipid mole ratio of 0.5, and at a mole ratio of 1.0 for the bilayer lipids, the triolein solubility was reduced to just 0.15 mol %. The effects of free cholesterol were more pronounced and progressive than observed previously on the bilayer solubility of cholesteryl oleate (Spooner, P. J. R., Hamilton, J. A., Gantz, D. L., & Small, D. M. (1986) Biochim. Biophys. Acta 860, 345-353]. As with cholesteryl oleate, we suggest that cholesterol also displaces solubilized triglyceride to deeper regions of the bilayer.  相似文献   

16.
A mutant of Escherichia coli missing energy-transducing ATPase and known to be defective in a variety of membrane functions from earlier studies (Yamamoto, T. H., Mével-Ninio, M. and Valentine, R. C. (1973) Biochim. Biophys. Acta 314, 267–275; Thipayathasana, P. and Valentine, R. C. (1974) Biochim. Biophys. Acta 347, 464–468; Mével-Ninio, M. and Yamamoto, T. (1974) Biochim. Biophys. Acta 357, 63–66) has been found to be blocked for anaerobic DNA synthesis. The rate of anaerobic DNA synthesis in the mutant, measured as radioactive adenine incorporation into the alkali-resistant fraction of whole cells, is about 1/6 the rate of DNA synthesis in the wild type culture under similar conditions. Addition of NO3- or O2 restores DNA biosynthesis in the mutant. The entry of radioactive adenine is not appreciably affected in the mutant by anaerobiosis. It is concluded that coupling factor plays a role in some step(s) of DNA biosynthesis.  相似文献   

17.
The cholate method originally introduced by Kagawa et al. (J. Biol. Chem. (1973) 248, 676–684) and further developed by Brunner et al. (Biochim. Biophys. Acta (1976) 455, 322–331) has been used to prepare single bilayer vesicles containing 5 mol% lysophosphatidylcholine embedded in a matrix of phosphatidylcholine. The distribution of lysophosphatidylcholine over outer and inner monolayer was found to be highly asymmetric (ratio 9 : 1), as determined by lysophospholipase treatment of the vesicles. This distribution is similar to the value found in sonicated vesicles.Up to 20 mol% cholesterol could be incorporated in the vesicles by the cholate method. The method was succesfully used also for the preparation of single bilayer vesicles from total rat liver microsomal lipids, to which 5 mol% of 1-[1-14C]palmitoyl lysophosphatidylcholine had been added. Surprisingly, almost 100% of lysophosphatidylcholine in the latter vesicles was directly available for hydrolysis by lysophospholipase. In contrast, only 70% of the lysophosphatidylcholine in sonicated vesicles of similar composition could be hydrolyzed by lysophospholipase.  相似文献   

18.
We previously showed [Herbette, L. G., Blasie, J. K., DeFoor, P., Fleischer, S., Bick, R. J., Van Winkle, W. B., Tate, C. A., & Entman, M. L. (1984) Arch. Biochem. Biophys. 234, 235-242; Herbette, L. G., DeFoor, P., Fleischer, S., Pascolini, D., Scarpa, A., & Blasie, J. K. (1985) Biochim. Biophys. Acta 817, 103-122] that the phospholipid head-group distribution in the membrane bilayer of isolated sarcoplasmic reticulum is asymmetric. From these studies, both the total number of phospholipid head groups and the total lipid, as well as the head-group species for these lipids, were found to be different for each monolayer of the membrane bilayer. In this paper, we demonstrate for the first time that there is significant asymmetry in the distribution of unsaturated fatty acids between the two monolayers; i.e., the outer monolayer of the sarcoplasmic reticulum contained more unsaturated and polyunsaturated chains when compared to the inner monolayer. X-ray diffraction measurements demonstrated that the time-averaged fatty acyl chain extension for the outer monolayer was approximately 20% less than for the inner monolayer. This is consistent with the concept that the greater degree of unsaturation in the outer monolayer may provide for a decreased average fatty acyl chain extension for that layer. This architecture for the bilayer may be related to both the "resting" state mass distribution of the calcium pump protein within the membrane bilayer and possible "conformational" states of the calcium pump protein during calcium transport by the sarcoplasmic reticulum.  相似文献   

19.
Cytochrome b5 was found to bind readily to sonicated vesicles containing as much as 0.8 mol cholesterol per mol egg phosphatidylcholine. This observation conflicts with the suggestion of Enomoto and Sato ((1977) Biochim. Biophys. Acta 466, 136–147) that cholesterol prevents binding of this protein to erythrocyte membranes.  相似文献   

20.
Solid core liposomes with encapsulated colloidal gold particles were prepared through four major steps: Preparation of prevesicles with encapsulated solid cores of agarose-gelatin by emulsification of agarose-gelatin sol in organic solvent containing emulsifiers followed by cooling. Extraction of lipophilic components from prevesicles to obtain microspherules of agarose-gelatin. Introducing colloidal gold particles into microspherules and coating with protein molecules. Encapsulation of colloidal gold-bearing microspherules with the modified organic solvent spherule evaporation method for preparation of liposomes (Kim et al. (1983) Biochim. Biophys. Acta 728, 339-348 and Kim et al. (1984) Biochim. Biophys. Acta 812, 793-801). Electron micrographs showed that if liposomes were prepared by using a lipid mixture containing dioleoylphosphatidylcholine/cholesterol/dioleoylphosphatidylglycerol/tri olein (molar ratio 4.5:4.5:1:1), there was only a single continuous bilayer membrane for each solid core liposome. However, if no triolein was added to the lipid mixture, it would cause the formation of multilamellar liposomes. In both cases, there were hundreds to thousands of colloidal gold particles within each solid core liposome.  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