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 15 毫秒
1.
The effects of negatively charged and neutral lipids on the function of the reconstituted nicotinic acetylcholine receptor from Torpedo californica were determined with two assays using acetylcholine receptor-containing vesicles: the ion flux response and the affinity-state transition. The receptor was reconstituted into three different lipid environments, with and without neutral lipids: (1) phosphatidylcholine/phosphatidylserine; (2) phosphatidylcholine/phosphatidic acid; and (3) phosphatidylcholine/cardiolipin. Analysis of the ion flux responses showed that: (1) all three negatively charged lipid environments gave fully functional acetylcholine receptor ion channels, provided neutral lipids were added; (2) in each lipid environment, the neutral lipids tested were functionally equivalent to cholesterol; and (3) the rate of receptor desensitization depends upon the type of neutral lipid and negatively charged phospholipid reconstituted with the receptor. The functional effects of neutral and negatively charged lipids on the acetylcholine receptor are discussed in terms of protein-lipid interactions and stabilization of protein structure by lipids.  相似文献   

2.
Binding of polyclonal and monoclonal antibodies, quantitated by enzyme-linked immunosorbent assay, to porcine SP-B reconstituted in different phospholipid bilayers has been used to assess differences in protein structure due to lipid-protein interactions. SP-B bound significantly more antibodies when it was reconstituted in bilayers made of anionic phospholipids (phosphatidic acid, cardiolipin, phosphatidylglycerol, phosphatidylinositol or phosphatidylserine) than in zwitterionic bilayers (phosphatidylcholine, phosphatidylcholine/cholesterol, or phosphatidylethanolamine) or in fatty acid micelles (made of salts of palmitic or stearic acids). These differences in immunoreactivity can be important in the development of quantitation methods for SP-B in clinical samples based on immunological techniques.  相似文献   

3.
The spontaneous incorporation of the polyene antibiotic amphotericin B from a micellar solution into phospholipid vesicles was examined as a function of the lipid composition of the vesicles and their physical state. Virtually no insertion of the antibiotic into egg phosphatidylcholine vesicles was observed even when cholesterol was also present in the bilayer. In contrast, rapid incorporation occurred into systems containing an anionic phospholipid such as phosphatidylglycerol or phosphatidylserine with the fastest rates observed for lipids containing the saturated dimyristoyl fatty acyl species. Insertion of amphotericin B into vesicles composed of dimyristoylphosphatidylcholine and dimyristoylphosphatidylglycerol (7:3 mole ratio) was rapid either above, below or within the gel-to-liquid-crystalline phase transition temperature (23 degrees C). The ability of amphotericin B to intercalate into lipid vesicles is discussed in relation to their relative bilayer stabilities.  相似文献   

4.
Steady-state and time-resolved fluorescence properties of the 7-nitrobenz-2-oxa-1, 3-diazole-4-yl (NBD) fluorophore attached either to the sn-2 acyl chain of various phospholipids (phosphatidylcholine, phosphatidylethanolamine, phosphatidylserine, and phosphatidic acid) or to the polar headgroup of phosphatidylethanolamine were studied after insertion of these NBD-labeled lipid probes into unilamellar vesicles of phosphatidylcholine, phosphatidylglycerol, phosphatidic acid, and phosphatidylserine. The fluorescence response of the NBD group was observed to strongly depend on the chemical structure and physical state of the host phospholipids and on the chemical structure of the lipid probe itself. Among the various fluorescence parameters studied, i.e., Stokes' shifts, lifetimes, and quantum yields, the quantum yields were by far the most affected by these structural and environmental factors, whereas the Stokes' shifts were practically unaffected. Thus, depending on the phospholipid probe and the host phospholipid, the fluorescence emission of the NBD group was found to vary by a factor of up to 5. Careful analysis of the data shows that for the various couples of probe and host lipid molecules studied, deexcitation of the fluorophore was dominated by nonradiative deactivation processes. This great sensitivity of the NBD group to environmental factors originates from its well-known solvatochromic properties, and comparison of these knr values with those obtained for n-propylamino-NBD in a set of organic solvents covering a large scale of polarity indicates that in phospholipids, the NBD fluorophore experiences a dielectric constant of around 27-41, corresponding to a medium of relatively high polarity. From these epsilon values and on the basis of models of the dielectric transition that characterizes any water-phospholipid interface, it can be inferred that for all of the phospholipid probes and host phospholipids tested, the NBD group is located in the region of the polar headgroups, near the phosphoglycerol moiety of the lipids.  相似文献   

5.
Hydrolysis of lipid mixtures by rat hepatic lipase   总被引:1,自引:0,他引:1  
The hydrolysis of phospholipid mixtures by purified rat hepatic lipase, also known as hepatic triglyceride lipase, was studied in a Triton X-100/lipid mixed micellar system. Column chromatography of the mixed micelles showed elution of Triton X-100 and binary lipid mixtures of phosphatidic acid, phosphatidylcholine, phosphatidylethanolamine and phosphatidylserine as a single peak. This indicated that the mixed micelles were homogenous and contained all components in the designated molar ratios. The molar ratio of Triton X-100 to lipid was kept constant at 4 to 1. Labeling one lipid with 3H and the other lipid with 14C enabled us to determine the hydrolysis of both components of these binary lipid mixed micelles. We found that the hydrolysis of phosphatidylcholine was activated by the inclusion of small amounts of phosphatidic acid (2.5-fold), phosphatidylethanolamine (1.5-fold) or phosphatidylserine (1.4-fold). The maximal activation of phosphatidylcholine hydrolysis was observed when 5 mol% of phosphatidylethanolamine, 7.5 mol% phosphatidic acid or 5 mol% phosphatidylserine was added to Triton X-100 mixed micelles. The hydrolysis of phosphatidic acid was activated 30%, and that of phosphatidylserine was inhibited 30% when the molar proportion of phosphatidylcholine was less than 50 mol%. The hydrolysis of phosphatidylethanolamine was slightly activated when the mol% of phosphatidylcholine was below 5. The hydrolysis of phosphatidylserine was inhibited by phosphatidylethanolamine when the mol% of the latter was 50 or less whereas phosphatidylethanolamine hydrolysis was not affected by phosphatidylserine. Under the conditions used sphingomyelin and cholesterol did not have a significant effect on the hydrolysis of the phospholipids studied. In agreement with our previous study (Kucera et al. (1988) J. Biol. Chem. 263, 1920-1928) these studies show that the phospholipid polar head group is an important factor which influences the action of hepatic lipase and that the interfacial properties of the substrate play a role in the expression of the activity of this enzyme. The molar ratios of phosphatidic acid, phosphatidylethanolamine and phosphatidylserine which activated phosphatidylcholine hydrolysis correspond closely to the molar ratios of these lipids found in the surface lipid film of lipoproteins e.g., high density lipoproteins.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
A recently developed fluorimetric transfer assay (Somerharju, P., Brockerhoff, H. and Wirtz, K.W.A. (1981) Biochim. Biophys. Acta 649, 521–528) has been applied to study the substrate specificity and membrane binding of the phosphatidylinositol-transfer protein from bovine brain. The substrate specificity was investigated by measuring the rate of transfer, either directly or indirectly, for a series of phosphatidylinositol analogues which included phosphatidic acid, phosphatidylglycerol as well as three lipids obtained from yeast phosphatidylinositol by partial periodate oxidation and subsequent borohydride reduction. Phosphatidylglycerol and the oxidation products of phosphatidylinositol were transferred at about one tenth of the rate observed for phosphatidylinositol while phosphatidic acid was not transferred. It is concluded that an intact inositol moiety favours the formation of the putative transfer protein-phosphatidylinositol complex. In addition to phosphatidylinositol, the transfer protein also transfers phosphatidylcholine. In order to obtain information on the possible occurrence of two sites of interaction, vesicles consisting of either pure 1-acyl-2-parinaroylphosphatidylinositol or 1-acyl-2-parinaroylphosphatidylcholine were titrated with the protein. Binding of labeled phospholipid to the protein was represented by an increase of lipid fluorescence and found to be much more efficient for phosphatidylinositol than for phosphatidylcholine. This is interpreted to indicate that the protein contains an endogenous phosphatidylinositol molecule which can be easily replaced by exogenous phosphatidylinositol but not by phosphatidylcholine, a lipid with a lower affinity for this protein. Thus the binding sites for the two phospholipids are mutually exclusive, i.e. phosphatidylinositol and phosphatidylcholine cannot be bound to the protein simultaneously. Finally, the effect of acidic phospholipids on the transfer protein activity was studied either by varying the content of phosphatidic acid in the acceptor vesicles or by adding vesicles of pure acidic phospholipids to the normal assay system. The latter vesicles consisted of either phosphatidic acid, phosphatidylglycerol, phosphatidylserine, phosphatidylinositol or cardiolipin. In both instances the transfer protein activity was inhibited, obviously through the enhanced association of the protein with the negatively charged vesicles. These findings strongly suggest that relatively nonspecific ionic forces rather than specific protein-phospholipid headgroup interactions contribute to the association of the phosphatidylinositol-transfer protein with membranes.  相似文献   

7.
A recently developed fluorimetric transfer assay (Somerharju, P., Brockerhoff, H. and Wirtz, K.W.A. (1981) Biochim. Biophys. Acta 649, 521-528) has been applied to study the substrate specificity and membrane binding of the phosphatidylinositol-transfer protein from bovine brain. The substrate specificity was investigated by measuring the rate of transfer, either directly or indirectly, for a series of phosphatidylinositol analogues which included phosphatidic acid, phosphatidylglycerol as well as three lipids obtained from yeast phosphatidylinositol by partial periodate oxidation and subsequent borohydride reduction. Phosphatidylglycerol and the oxidation products of phosphatidylinositol were transferred at about one tenth of the rate observed for phosphatidylinositol while phosphatidic acid was not transferred. It is concluded that an intact inositol moiety favours the formation of the putative transfer protein-phosphatidylinositol complex. In addition to phosphatidylinositol, the transfer protein also transfers phosphatidylcholine. In order to obtain information on the possible occurrence of two sites of interaction, vesicles consisting of either pure 1-acyl-2-parinaroylphosphatidylinositol or 1-acyl-2-parinaroylphosphatidylcholine were titrated with the protein. Binding of labeled phospholipid to the protein was represented by an increase of lipid fluorescence and found to be much more efficient for phosphatidylinositol than for phosphatidylcholine. This is interpreted to indicate that the protein contains an endogenous phosphatidylinositol molecule which can be easily replaced by exogenous phosphatidylinositol but not by phosphatidylcholine, a lipid with a lower affinity for this protein. Thus the binding sites for the two phospholipids are mutually exclusive, i.e. phosphatidylinositol and phosphatidylcholine cannot be bound to the protein simultaneously. Finally, the effect of acidic phospholipids on the transfer protein activity was studied either by varying the content of phosphatidic acid in the acceptor vesicles or by adding vesicles of pure acidic phospholipids to the normal assay system. The latter vesicles consisted of either phosphatidic acid, phosphatidylglycerol, phosphatidylserine, phosphatidylinositol or cardiolipin. In both instances the transfer protein activity was inhibited, obviously through the enhanced association of the protein with the negatively charged vesicles. These findings strongly suggest that relatively nonspecific ionic forces rather than specific protein-phospholipid headgroup interactions contribute to the association of the phosphatidylinositol-transfer protein with membranes.  相似文献   

8.
The stoichiometry and specificity of lipid-protein interaction, as well as the lipid exchange rates at the protein interface, have been determined from the electron spin resonance spectra of spin-labeled lipids in reconstituted complexes of the mitochondrial ADP-ATP carrier with egg phosphatidylcholine. With the exception of cardiolipin and phosphatidic acid, the lipids studied are found to compete for approximately 50 sites at the intramembranous surface of the protein dimer. This number of first-shell lipid sites is unusually large for a protein of this size. The specificity for the protein is in the order stearic acid approximately phosphatidic acid approximately cardiolipin greater than phosphatidylserine greater than phosphatidylglycerol approximately phosphatidylcholine, with the maximum association constant relative to phosphatidylcholine being approximately 4. The selectivity for anionic lipids was partially screened with increasing ionic strength, but to a lesser extent for cardiolipin and phosphatidic acid than for stearic acid. Only in the case of phosphatidylserine was the selectivity reduced at high ionic strength to a level close to that for phosphatidylcholine. The off rates for lipid exchange at the protein surface were independent of lipid/protein ratio and correlated in a reciprocal fashion with the different lipid selectivities, varying from 5 x 10(6) s-1 for stearic acid at low ionic strength to 2 x 10(7) s-1 for phosphatidylcholine and phosphatidylglycerol. The off rates for cardiolipin were unusually low in comparison with the observed selectivity, and indicated the existence of a special population of sites (ca. 30% of the total) for cardiolipin, at which the exchange rate was very low.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The major coat protein of bacteriophage M13 was incorporated in mixed dimyristoylphosphatidylcholine/dimyristoylphosphatidylglycerol (80/20 w/w) vesicles probed with different spin-labeled phospholipids, labeled on the C-14 atom of the sn-2 chain. The specificity for a series of phospholipids was determined from a motionally restricted component seen in the electron spin resonance (ESR) spectra of vesicles with the coat protein incorporated. At 30 degrees C and pH 8, the fraction of motionally restricted phosphatidic acid spin-label is 0.36, 0.52, and 0.72 for lipid/protein ratios of 18, 14, and 9 mol/mol, respectively. The ESR spectra, analyzed by digital subtraction, resulted in a phospholipid preference following the pattern cardiolipin = phosphatidic acid greater than stearic acid = phosphatidylserine = phosphatidylglycerol greater than phosphatidylcholine = phosphatidylethanolamine. The specificities found are related to the composition of the target Escherichia coli cytoplasmic membrane.  相似文献   

10.
Complement-mediated lysis of reconstituted lipid-myelin basic protein (BP) vesicles and myelin vesicles due to antibody raised against BP and isolated myelin is measured by determination of the amount of a water-soluble spin label, tempocholine chloride, released from the vesicles. The response is shown to be antigen-specific, antibody-dependent, and complement mediated. The relative response to different anti-BP antibody samples is similar to that determined by radioimmunoassay procedures. In contrast to immunoassays with BP in aqueous solution, this method measures immune recognition of the protein in either a synthetic or a natural membranous environment. This is important because this protein has been shown to have a different conformation when bound to lipid bilayers than in aqueous solution and its conformation depends on lipid composition. It is also a more rapid method because no separation of spin label still trapped in the vesicles and that released due to immune lysis is required. In synthetic membranes consisting of sphingomyelin, cholesterol, and an acidic lipid, either phosphatidylglycerol, phosphatidic acid, or phosphatidylserine, the response was greatest when the acidic lipid was phosphatidic acid. The response did not depend significantly on the antigen concentration expressed as molar ratio of BP to sphingomyelin, over the range 0.15:600 to 2:600, although it decreased at molar ratios less than 0.15:600. The antigen density required for immune lysis of vesicles containing this protein antigen is similar to that reported elsewhere for lipid antigens, although the time required for maximal lysis was greater. Both anti-BP and anti-myelin antibodies caused a greater specific complement-mediated response with synthetic vesicles than with myelin vesicles, which may be due to the different lipid and/or protein composition of myelin. Response was also obtained with the myelin vesicles, however, indicating that some determinants of BP can be recognized on the surface of the bilayer in isolated myelin by anti-BP.  相似文献   

11.
The monolayer technique has been used to study the transfer of [14C]phosphatidylinositol from the monolayer to phosphatidylcholine vesicles. An equivalent transfer rate was found for egg phosphatidylcholine, dioleoylphosphatidylcholine, dielaidoylphosphatidylcholine and dipalmitoylphosphatidylcholine. A reduced transfer rate was found for a shorter-chain derivative, dimyristoylphosphatidylcholine, and for species with two polyunsaturated fatty acid chains such as dilinoleoylphosphatidylcholine, diheptadecadienoylphosphatidylcholine, dilinolenoylphosphatidylcholine and diether and dialkyl derivatives. No activity was found for 1,3-dipalmitoylphosphatidylcholine. The presence of up to 5 mol% phosphatidylinositol in egg phosphatidylcholine vesicles had no effect on the transfer rate. Introduction of more than 5 mol% phosphatidylinositol or phosphatidic acid into the phosphatidylcholine vesicles gradually decreased the rate of phosphatidylinositol transfer from the monolayer. 20 mol% acidic phospholipid was nearly completely inhibitory. Transfer experiments between separate monolayers of phosphatidylcholine and phosphatidylinositol showed that the protein-bound phosphatidylcholine is readily exchanged for phosphatidylinositol, but the protein-bound phosphatidylinositol exchange for phosphatidylcholine occurs at a 20-times lower rate. The release of phosphatidylinositol is dependent on the lipid composition and the concentration of charged lipid in the acceptor membrane, but also on the ratio between donor and acceptor membranes. The main transfer protein from bovine brain which transfer phosphatidylinositol and phosphatidylcholine transfers also phosphatidylglycerol, but not phosphatidylserine or phosphatidic acid. The absence of significant changes in the surface pressure indicate that the phosphatidylinositol and phosphatidylcholine transfer is not accompanied by net mass transfer.  相似文献   

12.
We have used assays of lipid probe mixing, contents mixing and contents leakage to monitor the divalent cation-mediated interactions between lipid vesicles containing phosphatidylserine (PS) as a minority component together with mixtures of phosphatidylethanolamine (PE), phosphatidylcholine (PC) or sphingomyelin, and cholesterol in varying proportions. The initial rates of calcium- and magnesium-induced lipid probe quenching between vesicles, which reflect primarily the rates of vesicle aggregation, are strongly reduced as progressively higher proportions of PC or sphingomyelin are incorporated into PE/PS vesicles. The initial rates of divalent cation-induced contents mixing and contents leakage for PE/PS vesicles are also strongly reduced when choline phospholipids are incorporated into the vesicles in even low molar proportions. Sphingomyelin has a more potent inhibitory effect on these processes than does PC at an equal level in the vesicle membranes. The inclusion of cholesterol in these vesicles, at levels up to 1:2 moles sterol/mole phospholipid, has little effect on the rates of calcium- or magnesium-induced vesicle aggregation. However, cholesterol significantly enhances the initial rates of vesicle contents mixing and contents leakage in the presence of divalent cations when the vesicles contain choline as well as amino phospholipids. This effect is substantial only when the level of cholesterol exceeds the level of choline phospholipids in the vesicles. These results may have significance for the fusion of certain cellular membranes in mammalian cells, whose cytoplasmic faces have lipid compositions very similar to those of the vesicles examined in this study.  相似文献   

13.
Purified G-protein from vesicular stomatitis virus was reconstituted into egg phosphatidylcholine vesicles by detergent dialysis of octyl glucoside. A homogeneous population of reconstituted vesicles could be obtained, provided the protein to lipid ratio was high (about 0.3 mol % protein) and the detergent removal was slow. The reconstituted vesicles were assayed for fusion activity using electron microscopy and fluorescence energy transfer. The fusion activity mediated by the viral envelope protein was dependent upon pH, temperature, and target membrane lipid composition. Incubation of reconstituted vesicles at low pH with small unilamellar vesicles containing negatively charged lipids resulted in the appearance of large cochleate structures, as shown by electron microscopy using negative stain. This process did not cause leakage of a vesicle-encapsulated aqueous marker. The rate of fusion was pH-dependent with a pK of about 4 and the apparent energy of activation for the fusion was 16 +/- 1 kcal/mol. G-protein-mediated fusion showed a large preference for target membranes which contain phosphatidylserine or phosphatidic acid. Inclusion of 36% cholesterol in any of the lipid compositions had no effect on the rate of fusion. These reconstituted vesicles provide a system to study the mechanism of pH-dependent fusion induced by a viral spike protein.  相似文献   

14.
Na+-Ca2+ exchange activity in cardiac sarcolemmal vesicles is known to be sensitive to charged, membrane lipid components. To examine the interactions between membrane components and the exchanger in more detail, we have solubilized and reconstituted the Na+-Ca2+ exchanger into membranes of defined lipid composition. Our results indicate that optimal Na+-Ca2+ exchange activity requires the presence of certain anionic phospholipids. In particular, phosphatidylserine (PS), cardiolipin, or phosphatidic acid at 50% by weight results in high Na+-Ca2+ exchange activity, whereas phosphatidylinositol and phosphatidylglycerol provide a poor environment for exchange. In addition, incorporation of cholesterol at 20% by weight greatly facilitates Na+-Ca2+ exchange activity. Thus, for example, an optimal lipid environment for Na+-Ca2+ exchange is phosphatidylcholine (PC, 30%)/PS (50%)/cholesterol (20%). Na+-Ca2+ exchange activity is also high when cardiac sarcolemma is solubilized and then reconstituted into asolectin liposomes. We fractionated the lipids of asolectin into subclasses for further reconstitution studies. When sarcolemma is reconstituted into vesicles formed from the phospholipid component of asolectin, Na+-Ca2+ exchange activity is low. When the neutral lipid fraction of asolectin (including sterols) is also included in the reconstitution medium, Na+-Ca2+ exchange activity is greatly stimulated. This result is consistent with the requirement for cholesterol described above. Proteinase treatment, high pH, intravesicular Ca2+ and dodecyl sulfate all stimulate Na+-Ca2+ exchange in native sarcolemmal vesicles. We examined the effects of these interventions on exchange activity in reconstituted vesicles of varying lipid composition. In general, Na+-Ca2+ exchange could be stimulated only when reconstituted into vesicles of a suboptimal lipid composition. That is, when reconstituted into asolectin or PC/PS/cholesterol (30:50:20), the exchanger is already in an activated state and can no longer be stimulated. The one exception was that the Na+-Ca2+ exchanger responded to altered pH in an identical manner, independent of vesicle lipid composition. The mechanism of action of altered pH on the exchanger thus appears to be different from other interventions.  相似文献   

15.
Insulin receptor kinase, affinity-purified by adsorption and elution from immobilized insulin, is stimulated 2-3-fold by insulin in detergent solution. Reconstitution of the receptor kinase into leaky vesicles containing phosphatidylcholine and phosphatidylethanolamine (1:1, w/w) by detergent removal on Sephadex G-50 results in the complete loss of receptor kinase sensitivity to activation by insulin. Insulin receptors in these vesicles also exhibit an increase in their apparent affinity for 125I-insulin (Kd = 0.12 nM versus 0.76 nM). Inclusion of 8.3-16.7% phosphatidylserine into the reconstituted vesicles restores 40-50% of the insulin-sensitivity to the receptor kinase. An elevated apparent affinity for 125I-insulin of insulin receptors in vesicles containing phosphatidylcholine and phosphatidylethanolamine is also restored to the value observed in detergent solution by the inclusion of phosphatidylserine in the reconstituted system. The effect of phosphatidylserine on insulin receptor kinase appears specific, because cholesterol, phosphatidylinositol and phosphatidic acid are all unable to restore insulin-sensitivity to the receptor kinase. Autophosphorylation sites on the insulin receptor as analysed by h.p.l.c. of tryptic 32P-labelled receptor phosphopeptides are not different for insulin receptors autophosphorylated in detergent solution or for the reconstituted vesicles in the presence or absence of phosphatidylserine. These data indicate that the phospholipid environment of insulin receptors can modulate its binding and kinase activity, and phosphatidylserine acts to restore insulin-sensitivity to the receptor kinase incorporated into phosphatidylcholine/phosphatidylethanolamine vesicles.  相似文献   

16.
The structural phase behavior of phospholipid mixtures consisting of short-chain (dihexanoyl phosphatidylcholine) and long-chain lipids (dimyristoyl phosphatidylcholine and dimyristoyl phosphatidylglycerol), with and without lanthanide ions was investigated by small-angle neutron scattering (SANS). SANS profiles were obtained from 10 degrees C to 55 degrees C using lipid concentrations ranging from 0.0025 g/ml to 0.25 g/ml. The results reveal a wealth of distinct morphologies, including lamellae, multi-lamellar vesicles, unilamellar vesicles, and bicellar disks.  相似文献   

17.
The interaction between amphotericin B and egg yolk phosphatidylcholine, dimyristoyl (DMPC) and dipalmitoyl phosphatidylcholine (DPPC) phospholipid bilayer vesicles has been monitored by the circular dichroism (CD) spectra of amphotericin B at a 1 · 10?5 M concentration. This method has revealed that amphotericin B may be present in a number of different forms depending on the time elapsed after the mixing, the cholesterol content of the vesicles and the vesicles' physical state. Some striking features of these CD detected species are the following: with egg yolk phosphatidylcholine and a molar cholesterol percentage lower than 25, at 25°C several forms are coexistent, their amount is time-dependent; with dipalmitoyl or dimyristoyl phosphatidylcholines without cholesterol or with a cholesterol molar percentage lower than 25, in the gel state, a form different from the former appears very rapidly; with egg yolk phosphatidylcholine, DMPC and DPPC at a molar cholesterol percentage between 25 and 50 a new form is monitored, identical in the three cases and observed in the liquid crystalline state as well as in the gel state. In the case of the three phospholipids without cholesterol a definite interaction with the antibiotic is observed but with different characteristics according to the nature of lipid.With amphotericin B ‘Fungizone’ the same species are monitored but their appearance is much slower.Two explanations are proposed for the origin of the discrepancies between CD and electronic absorption.  相似文献   

18.
Cytochrome P-450 LM2 was reconstituted by the cholate-dialysis method into vesicles containing a mixture of either phosphatidylcholine or phosphatidylethanolamine with up to 50 mol% of phosphatidic acid. Phase transition curves in the presence or absence of cytochrome P-450 were obtained from electron paramagnetic resonance experiments by measuring the partitioning of 2,2,6,6-tetramethylpiperidine-1-oxyl. Protein-free phospholipid vesicles exhibit a phase separation into domains of gel phase enriched in phosphatidic acid in a surrounding fluid matrix containing mainly phosphatidylcholine. The phase transition of the phosphatidic acid domains disappeared following incorporation of cytochrome P-450 into the bilayers. In contrast, in vesicles containing mixtures of egg-phosphatidic acid and dimyristoyl phosphatidylcholine, the phase transition of the domains enriched in dimyristoyl phosphatidylcholine was less sharp than in the corresponding vesicles containing cytochrome P-450. The results of both of these experiments could be explained by a redistribution of the mol fraction of the two phospholipids in the gel phase due to preferential binding of the egg-phosphatidic acid to the cytochrome P-450. For comparison, incorporation of cytochrome P-450 into uncharged vesicles of dimyristoyl phosphatidylcholine and egg-phosphatidylethanolamine did not alter the  相似文献   

19.
The protein-mediated phospholipid exchange between small unilamellar vesicles was investigated by fluorescence polarization measurements with diphenylhexatriene as optical probe. Thermotropic phase-transition measurements were taken after mixing two vesicle preparations of distinct and different phase-transition temperatures or having different states of charge. From the heights of each phase-transition step, we were able to follow the lipid-exchange process in the presence, as well as in the absence (natural exchange), of so-called transfer protein isolated from beef liver. A strong enhancement of the lipid transfer was observed at the corresponding lipid-phase-transition temperature, which is explained by the presence of fluctuating fluid and ordered domains co-existing at the lipid-phase-transition temperature. A unidirectional lipid transfer of the neutral component was observed between negatively charged phosphatidic acid and neutral phosphatidylcholine vesicles. Fluorescence polarization measurements showed the disappearance of the phosphatidylcholine phase transition, whereas the phosphatidic acid phase transition broadened and its phase transition temperature became lower.  相似文献   

20.
Synapsin I, a major neuron-specific phosphoprotein, is localized on the cytoplasmic surface of small synaptic vesicles to which it binds with high affinity. It contains a collagenase-resistant head domain and a collagenase-sensitive elongated tail domain. In the present study, the interaction between synapsin I and phospholipid vesicles has been characterized, and the protein domains involved in these interactions have been identified. When lipid vesicles were prepared from cholesterol and phospholipids using a lipid composition similar to that found in native synaptic vesicle membranes (40% phosphatidylcholine, 32% phosphatidylethanolamine, 12% phosphatidylserine, 5% phosphatidylinositol, 10% cholesterol, wt/wt), synapsin I bound with a dissociation constant of 14 nM and a maximal binding capacity of about 160 fmol of synapsin I/microgram of phospholipid. Increasing the ionic strength decreased the affinity without greatly affecting the maximal amount of synapsin I bound. When vesicles containing cholesterol and either phosphatidylcholine or phosphatidylcholine/phosphatidylethanolamine were tested, no significant binding was detected under any conditions examined. On the other hand, phosphatidylcholine vesicles containing either phosphatidylserine or phosphatidylinositol strongly interacted with synapsin I. The amount of synapsin I maximally bound was directly proportional to the percentage of acidic phospholipids present in the lipid bilayer, whereas the Kd value was not affected by varying the phospholipid composition. A study of synapsin I fragments obtained by cysteine-specific cleavage showed that the collagenase-resistant head domain actively bound to phospholipid vesicles; in contrast, the collagenase-sensitive tail domain, though strongly basic, did not significantly interact. Photolabeling of synapsin I was performed with the phosphatidylcholine analogue 1-palmitoyl-2-[11-[4-[3-(trifluoromethyl)diazirinyl]phenyl] [2-3H]undecanoyl]-sn-glycero-3-phosphocholine; this compound generates a highly reactive carbene that selectively interacts with membrane-embedded domains of membrane proteins. Synapsin I was significantly labeled upon photolysis when incubated with lipid vesicles containing acidic phospholipids and trace amounts of the photoactivatable phospholipid. Proteolytic cleavage of photolabeled synapsin I localized the label to the head domain of the molecule.(ABSTRACT TRUNCATED AT 400 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