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.
Here we study the anti-nucleating mechanism of apolipoprotein A-I (apo A-I) on model biliary vesicles in the presence of phospholipase C (PLC) utilizing dynamic light scattering (DLS), steady-state fluorescence spectroscopy, cryogenic transmission electron microscopy (cryo-TEM), and UV/Vis spectroscopy. PLC induces aggregation of cholesterol-free lecithin vesicles from an initial, average size of 100 nm to a maximal size of 600 nm. The presence of apo A-I likely inhibits vesicle aggregation by shielding the PLC-generated hydrophobic moieties, which results in vesicles of an average size of 200 nm. A similar phenomenon is observed in cholesterol-enriched lecithin vesicles. Whereas PLC alone produces aggregates of 300 nm, no aggregation is observed when apo A-I is present along with PLC. However, the ability of apo A-I to inhibit aggregation is temporary, and after 8 h, a broad particle size distribution with sizes as high as 800 nm is observed. Apo A-I possibly induces the formation of small apo A-I/lecithin/cholesterol complexes of about 5-20 nm similar to the discoidal pre-HDL complexes found in blood when it can no longer effectively shield all the DAG molecules. Concomitant with formation of complexes, DAG molecules coalesce into large oil droplets, which account for the large particles observed by light scattering. Thus, apo A-I acts as an anti-nucleating agent by two mechanisms, anti-aggregation and microstructural transition. The mode of protection is dependent on the cholesterol content and the relative amounts of DAG and apo A-I present. This study supports the possibility of apo A-I solubilizing lipids in bile in a similar fashion as it does in blood and also delineates the mechanism of formation of the complexes.  相似文献   

2.
The segment corresponding to residues 121 to 164 of human plasma apolipoprotein A-I (apo-A-I) has been synthesized by the Merrifield solid phase method. The peptide binds to unilamellar phospholipid vesicles and to phospholipid-cholesterol mixed vesicles. The surface affinity of the peptide measured in this way indicated that the mechanism of binding is the same as that of apo A-I (144-165) and apo A-I itself. The peptide appears to be a globular monomer in a aqueous solution, with 17% alpha helix content. The peptide bound to vesicles activates lecithin:cholesterol acyltransferase: compared to apo A-I, the peptide is about 30% as efficient in the activation of cholesterol esterification and of phospholipid hydrolysis when the surface is saturated by the activator. For a variety of amphiphilic peptides and for apo A-I, the lecithin: cholesterol acyltransferase-activating ability correlates well with their alpha helix contents in 50% trifluoroethanol.  相似文献   

3.
We explored the influence of several compositional factors considered capable of influencing the nucleation time of model biles supersaturated in cholesterol. In addition to the classical techniques, e.g., electron microscopy and quasielastic light scattering, employed for size measurement and structural assessment, we employed a novel technique, i.e., video-enhanced microscopy, for particle evaluation in these polydisperse systems which often may simultaneously contain isolated small vesicles, their complex aggregates, and small cholesterol monohydrate crystals. The factors we studied included dilution, degree of cholesterol supersaturation, bile salt/lecithin molar ratio, and Ca2+ concentration. Dilution markedly raised the degree of cholesterol saturation, prolonged nucleation time for cholesterol monohydrate crystals, and favored formation of metastable small unilamellar vesicles. Increasing the degree of cholesterol supersaturation as an independent variable in more concentrated systems both shortened the nucleation time and favored spontaneous formation of a relatively small number of isolated vesicles. A decrease in bile salt/lecithin molar ratio within the physiologically relevant range was accompanied by a prolonged nucleation time and favored spontaneous vesicle formation. Large numbers of small unilamellar vesicles were observed even in concentrated model bile solutions (total lipids: 20 g/dl) when the bile salt/lecithin molar ratio was 1.9 or less. At physiological concentrations, Ca2+ promoted nucleation of cholesterol monohydrate crystals only in vesicle-containing solutions. Taken together, the following conclusions can be drawn. First, spontaneous vesicle formation in dilute systems prolongs solid cholesterol crystal nucleation. It can thus provide a supplementary non-micellar mode of cholesterol transport in micellar systems of supersaturated human bile. Second, dilution, degree of cholesterol supersaturation, and a decrease in bile salt/lecithin ratio prolong cholesterol crystal nucleation time and favor spontaneous vesicle formation. With increasing calcium concentrations, opposite effects are observed. Third, the presence of vesicles may help to account for the frequently observed and otherwise unexplained remarkable degree of metastable supersaturation and prolonged metastability (delayed nucleation time) for cholesterol in human bile.  相似文献   

4.
Various combinations of incorporation and addition of apolipoprotein A-I (apo A-I) and apolipoprotein A-II (apo A-II) individually or together to a defined lecithin-cholesterol (250/12.5 molar ratio) liposome prepared by the cholate dialysis procedure were used to study the effect of apo A-II on lecithin:cholesterol acyltransferase (LCAT, EC 2.3.1.43) activity of both purified enzyme preparations and plasma. When apo A-I (0.1-3.0 nmol/assay) alone was incorporated or added to the liposome, apo A-I effectively activated the enzyme. By contrast, when apo A-II (0.1-3.0 nmol/assay) alone was incorporated into or added to the liposome, apo A-II exhibited minimal activation of LCAT activity, approximately 1% of the activity obtained by an equal amount of apo A-I. Addition of apo A-II (0.1-3.0 nmol/assay) together with apo A-I (0.8 nmol/assay) to the liposome reduced the LCAT activity to approximately 30% of the level obtained with addition of apo A-I alone. On the other hand, addition of apo A-II (0.1-3.0 nmol/assay) or addition of lecithin-cholesterol liposome containing apo A-II (0.1-3.0 nmol/assay) to lecithin-cholesterol liposome containing apo A-I (0.8 nmol/assay) did not significantly alter apo A-I activation of LCAT activity. However, when the same amounts (0.1-3.0 nmol/assay) of apo A-II were incorporated together with apo A-I (0.8 nmol/assay) into the liposome, apo A-II significantly stimulated LCAT activity as compared to activity obtained with incorporation of apo A-I alone. The maximal stimulation was obtained with 0.4 nmol apo A-II/assay for both purified and plasma enzyme. At this apo A-II concentration, approximately 4-fold and 1.8-fold stimulation was observed for purified enzyme and plasma enzyme, respectively. These results indicated that apo A-II must be incorporated together with apo A-I into lecithin-cholesterol liposomes to exert its stimulatory effect on LCAT activity and that apo A-II in high-density lipoprotein may play an important role in the regulation of LCAT activity.  相似文献   

5.
The amino acid sequence of rabbit apolipoprotein A-I (apo A-I) has been determined by degradation and alignment of two overlapping sets of peptides obtained from tryptic and staphylococcal digestions. All of the peptides of rabbit apo A-I resulting from digestion by staphylococcal protease were isolated and sequenced except residues 33-37. A digestion with trypsin was employed to find overlapping and missing peptides. The N-terminus of rabbit apo A-I was confirmed by sequencing the intact protein up to 20 residues while the C-terminus was identified through its homology with human apo A-I. The protein contains 241 residues in its single chain. Its primary structure is highly homologous to the reported canine apo A-I (80%) and human apo A-I (78%), but exhibits less similarity with rat apo A-I (60%). Like human apo A-I, rabbit apo A-I contains very little histidine (2) and methionine (1); it does however have two residues of isoleucine. Based on a comparison of the hydrophobic-hydrophilic character of apo A-I residues with that of the two synthetic peptides that activated lecithin: cholesterol acyltransferase (Pownall et al. and Yokoyama et al.), we found that the five segments with the highest corresponding homologies on the protein are located within the N-terminal half. This suggests that the N-terminal half of apo A-I contains the major portion of regions activating lecithin: cholesterol acyltransferase.  相似文献   

6.
Complexes of egg yolk phosphatidylcholine and apolipoprotein A-I were prepared by a detergent (sodium cholate)-dialysis method and characterized by gradient gel electrophoresis, gel filtration, electron microscopy and chemical analysis. Multicomponent electrophoretic patterns were obtained indicating formation of at least eight classes of discoidal complexes. The relative contribution of the different classes to the electrophoretic pattern was a function of the molar ratio of phosphatidylcholine:apolipoprotein A-I in the interaction mixture. Molar ratios of phosphatidylcholine:apolipoprotein A-I in isolated complexes were strongly and positively correlated with disc diameter obtained by electron microscopy. Incorporation of unesterified cholesterol into phosphatidylcholine/apolipoprotein A-I interaction mixtures also resulted in formation of unique complexes but with considerably different particle size distributions relative to those observed in the absence of cholesterol. One common consequence of cholesterol incorporation into interaction mixtures of 87.5:1 and 150:1 molar ratio of phosphatidylcholine:apolipoprotein A-I was the disappearance of a major complex class with diameter of 10.8 nm and the appearance of a major component with diameter of approximately 8.8 nm. Electrophoretic patterns of cholesterol-containing complexes showed a strong similarity to patterns recently published for high density lipoproteins from plasma of lecithin:cholesterol acyltransferase-deficient subjects, suggesting that the complexes formed in vitro by the detergent-dialysis method may serve as appropriate models for investigation of the origins of the HDL particle size distribution.  相似文献   

7.
Complexes formed between apolipoprotein A-I (apo A-I) and dimyristoylphosphatidylcholine (DMPC) or egg phosphatidylcholine have been studied by high-field 1H NMR, nondenaturing gradient gel electrophoresis, electron microscopy, and gel filtration chromatography. Emphasis has been placed on an analysis of the particle size distribution within the micellar complexes produced at lipid/protein molar ratios of 40-700. As determined by electron microscopy and gel filtration of DMPC/apo A-I complexes, the size of the discoidal micelles produced appears to increase uniformly with an increasing lipid/protein ratio. By electron microscopy, the diameters of isolated DMPC/apo A-I discoidal micelles range from approximately 89 A at a 40 molar ratio to 205 A at a 700 molar ratio. Analysis of the micellar complexes by 1H NMR shows that concomitant with the increase in size is the progressive downfield shift of the choline N-methyl proton resonance of the complex which is observed from 3.245 to 3.267 ppm over the above molar ratio range. The relationship between chemical shift and micelle size is most simply interpreted as arising from a weighted averaging of two lipid environments--lipid-lipid and lipid-protein. In contrast to the above interpretation of the gel filtration experiments on DMPC/apo A-I complexes, nondenaturing gradient gel electrophoresis analysis of particle size distribution leads to an unexpected observation: as the DMPC/apo A-I ratio increases, discrete complexes of increasing size are formed in an apparently quantized manner.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The preparation of discoidal, recombinant HDL (r-HDL) containing various phospholipids, apolipoproteins and a range of concentrations of unesterified cholesterol has been reported by several investigators. The present study describes the preparation of r-HDL containing both apolipoprotein (apo) A-I and apo A-II. r-HDL with 100:1 (mol:mol) egg PC.apo A-I and 0 (Series I), 5 (Series II) or 10 (Series III) mol% unesterified cholesterol were prepared by the cholate dialysis method. The resulting complexes had a Stokes' radius of 4.7 nm and contained two molecules of apo A-I per particle. When the r-HDL (2.0 mg apo A-I) were supplemented with 1.0 mg of apo A-II, one of the apo A-I molecules was replaced by two molecules of apo A-II. This modification was not accompanied by a loss of phospholipid, nor by major change in particle size. The addition of 2.5 or 4.0 mg of apo A-II resulted in the displacement of both apo A-I molecules from a proportion of the r-HDL and the formation of smaller particles (Stokes' radius 3.9 nm), which contained half the original number of egg PC molecules and three molecules of apo A-II. The amount of apo A-I displaced was dependent on the concentration of unesterified cholesterol in the r-HDL: when 2.5 mg of apo A-II was added to the Series I, II and III r-HDL, 44, 60 and 70%, respectively, of the apo A-I was displaced. Addition of 4.0 mg of apo A-II did not promote further displacement of apo A-I from any of the r-HDL. By contrast, the association of apo A-II with r-HDL was independent of the concentration of unesterified cholesterol and was a linear function of the amount of apo A-II which had been added. It is concluded that (1), the structural integrity of egg PC.unesterified cholesterol.apo A-I r-HDL, which contain two molecules of apo A-I, is not affected when one of the apo A-I molecules is replaced by two molecules of apo A-II; (2), when both apo A-I molecules are replaced by apo A-II, small particles which contain three molecules of apo A-II are formed; and (3), the displacement of apo A-I from r-HDL is facilitated by the presence of unesterified cholesterol in the particles.  相似文献   

9.
We studied the effect of bacterial lipopolysaccharide (LPS)-apolipoprotein A-I (apo A-I) interaction on the structure and function of this protein. The micellization process of dimirystoil phosphatidylcholine liposomes (MLV-DMPC) by apo A-I in the presence of LPS was characterized. Apo A-I may interact with MLV-DMPC at the lipid transition temperature, forming micellar complexes. The kinetics of MLV-DMPC micellization was studied by turbidimetry. In the absence of LPS, a monoexponential decrease in turbidity is observed. Preincubation of apo A-I with LPS impairs the micellization reaction, resulting in biphasic kinetics. The amplitude of the fast phase decreases with increasing concentrations of LPS. In the absence or in the presence of low amounts of LPS (1∶0.1 protein:LPS weight ratio), two major micellization products-containing two and three apo A-I molecules per particle-were observed. However, in the presence of higher amounts of LPS (1∶1 protein:LPS weight ratio), particles mainly contained two apo A-I molecules. In contrast, a decrease in intrinsic fluorescence intensity of the protein was observed in the presence of an increasing LPS concentration. Finally, we studied the effect of LPS on the transition temperature (Tt) of MLV-DMPC without detecting changes in Tt. In conclusion, the changes found in the micellization process are likely to be mainly caused by changes in the apo A-I conformation by LPS interaction in solution.  相似文献   

10.
The interaction between apolipoprotein A-I and small unilamellar vesicles of dipalmitoylphosphatidylcholine at the lipid phase transition resulted in complete release of vesicle contents at molar ratios of lipid to protein from 4000:1 down to 50:1. This indicated the existence of two types of stable complexes: a vesicular apo-A-I complex with a maximum of two to three apo-A-Is/vesicle, and a micellar complex (disc) with a stoichiometry of about 50 phosphatidylcholines/apo-A-I (mol/mol). We characterized the complexes by density gradient centrifugation, by gel filtration, and by immunoprecipitation using an anti-apo-A-I antibody. The morphology of the discs was similar to that of previously reported discs. Apo-A-I-induced release of vesicle contents was monitored by the relief of self-quenching of vesicle-encapsulated carboxyfluorescein. Using this assay we characterized the nature of the interaction between apo-A-I and phospholipid vesicles. The formation of complexes between vesicles and apo-A-I followed a two-step process; below or above the lipid phase transition temperature (Tc), apo-A-I bound to phosphatidylcholine vesicles but caused little leakage of contents. Kinetic analysis of the interaction between apo-A-I and dipalmitoylphosphatidylcholine vesicles below Tc indicated that about 1 in 500 collisions leads to a stable apo-A-I-vesicle complex. The second step involved passage of those complexes through Tc, which resulted in a very rapid transition into discs or vesicular complexes. Vesicular complexes contain apo-A-I which was no longer capable of interacting with pure lipid. Discs, on the other hand, interacted with vesicles at their phase transition.  相似文献   

11.
Y S Bae  H Kim 《Journal of biochemistry》1989,106(6):1019-1025
The interactions of human apolipoprotein A-I (apo A-I) with dipalmitoylphosphatidylcholine (DPPC) in vesicular complexes at low protein concentrations and in micellar complexes at high protein concentrations are compared. The C-terminal segment of this protein, with a relative molecular weight (Mr) of about 11,000, is protected on trypsin treatment of apo A-I-vesicle complexes. A segment within the sequence from Leu-189 to Arg-215 of apo A-I penetrates the hydrophobic interior of the membrane, as found in a hydrophobic labeling experiment involving 3-(trifluoromethyl)-3-(m-[125I]iodophenyl)-diazirine ([125I]TID). No appreciable stretch of apo A-I in micellar complexes was found to be protected from the tryptic digestion. This indicates that the interactions of apo A-I with lipids in the vesicular and micellar complexes are different. The binding equilibrium of apo A-I as to DPPC vesicles at low protein concentrations, as studied by hydrophobic labeling of the bilayer-penetrating segment, is reached within about 1 h, while the formation of micellar complexes at high protein concentrations takes about 24 h at 42 degrees C. Time-dependent labeling studies involving photoreactive phosphatidylcholine (PC) with high apo A-I concentrations suggested an initial interaction with the head group region of the bilayer followed by interaction with the tail ends of the acyl chains of the lipid. A possible mechanism for the micellization process is discussed.  相似文献   

12.
Two monoclonal antibodies, A17 and A30, were raised against human apolipoprotein A-I (apo A-I). They were studied by competitive inhibition of 125I-labeled HDL3 with HDL subfractions, delipidated apo A-I, and complexes of dimyristoylphosphatidylcholine (DMPC) containing apo A-I and apo A-II. Immunoblotting located the A17 antibody on CNBr fragment 4 of apo A-I and the A30 antibody on CNBr fragment 1. The A17 antigenic determinant was expressed identically in all HDL subclasses, on delipidated apo A-I as well as all on the DMPC-apo A-I and DMPC-apo A-I/apo A-II complexes. In contrast, the apparent affinity constant of the A30 antibody for delipidated apo A-I was about 30-times less than for HDL3 or for apo A-I/apo A-II-phospholipid complexes. These data suggest that the association of apo A-I with phospholipids improves the reactivity of the A30 monoclonal antibody towards apo A-I, and that this antigenic determinant has a different conformation in delipidated apo A-I compared to apo A-I complexed with phospholipids. Turbidimetric and fluorescence experiments monitoring the phospholipid-apo A-I association in the presence and in the absence of the A17 and A30 antibodies were consistent with the competition experiments carried out by solid phase radioimmunoassay (RIA). After reaction of apo A-I with the A30 antibody, we observed an enhancement of the degradation kinetics of large multilamellar vesicles (LMV), while the A17 antibody did not have a significant effect. Calcein leakage experiments carried out below the transition temperature of DPPC showed an enhancement of the degradation kinetics with both monoclonal antibodies, while the phase-transition release was independent of the reaction of apo A-I with the monoclonal antibodies. These data therefore suggest the existence of at least two different types of epitope on apo A-I, which might account for the differences in immunological reactivity of apo A-I that is either delipidated or present on HDL.  相似文献   

13.
Using a pyrene as a fluorescent probe, we investigated the influence of native and oxidized apolipoprotein A-I (apo A-I) and their complexes with tetrahydrocortisol (THC) on the microviscosity of the erythrocyte plasma membrane. The addition of THC to isolated membranes led to a 17% increase in the membrane microviscosity. In contrast, native apo A-I reduced the microviscosity (i.e., increased the fluidity) of the membranes by 15%. A more pronounced increase (by 25%) in the membrane fluidity was found in the presence of the complex of apo A-I with THC. Unlike native apo A-I, oxidized apo A-I and its complex with THC did not change the membrane viscosity. In view of the fact that apo A-I plays an important role in the binding of membrane cholesterol we suggest that the observed increase in the membrane fluidity under the influence of the native apo A-I is associated with the cholesterol efflux from plasma membrane. Oxidative modification of apo A-I likely disturbs the mechanisms of the cholesterol efflux and prevents the decrease in the membrane microviscosity.  相似文献   

14.
The four peptide analogs of the amphipathic helix whose interactions with dimyristoyl phosphatidylcholine were described in the preceding paper were compared with apolipoproteins (apo) A-I and A-II in ability to displace native apolipoprotein from high density lipoprotein (HDL) and in ability to activate lecithin:cholesterol acyltransferase. The rank order of the ability of the four peptide analogs to displace apo-A-I from intact HDL was 18A-Pro-18A greater than 18A greater than des-Val10-18A greater than reverse-18A, the same order suggested in the preceding paper for relative lipid affinities. Modified HDL from which 40% of the apo-A-I had been displaced by 18A was indistinguishable from unmodified HDL in its ability to act as a lecithin:cholesterol acyltransferase substrate. This suggests that the easily displaced apo-A-I molecules in polydisperse HDL are relatively ineffectual as lecithin:cholesterol acyltransferase activators and/or 18A replaces the lecithin:cholesterol acyltransferase activity lost. The peptide analog 18A-Pro-18A was found to be a powerful activator of lecithin:cholesterol acyltransferase when incubated with unilamellar egg phosphatidylcholine (PC) vesicles, reaching 140% of the activity of apo-A-I at a 1:1.75 peptide-to-egg PC ratio. In another experiment, it was found that discoidal egg PC complexes of 18A-Pro-18A, 18A, and des-Val10-18A, formed by cholate dialysis, had 30-45% of the activity of apo-A-I/egg PC discoidal complexes, also formed by cholate dialysis, at the same peptide/lipid weight ratio. Examination of the structures formed when the 18A-Pro-18A peptide was incubated with unilamellar egg PC vesicles indicated that the ability of 18A-Pro-18A to exceed apo-A-I in lecithin:cholesterol acyltransferase activating ability is due to the spontaneous conversion by 18A-Pro-18A of egg PC vesicles to small protein annulus-bilayer disc structures. Apo-A-I, apo-A-II, nor any of the other three peptide analogs of the amphipathic helix studied were able to convert a significant fraction of egg PC unilamellar vesicles to discoidal structures.  相似文献   

15.
Reconstitution of apolipoprotein A-I was found to occur readily with bovine brain sphingomyelin (BBSM), with a maximum rate occurring at a temperature of 28 degrees C, a temperature approximating the phase transition temperature for this naturally occurring phospholipid. At BBSM:A-I weight ratios of 7.5:1 or less, a single recombinant product was observed which contained three A-I molecules per particle, which had a BBSM:A-I molar ratio of 360 to 1 and which appeared in the electron microscope as a discoidal complex with a thickness of 68 A and a diameter of 217 A. By these criteria, as well as by gel filtration, this product appears very similar to that obtained by recombination of A-I with phosphatidylcholine at elevated ratios of phospholipid/protein. No evidence was found for the existence of any BBSM:A-I complexes comparable to the smaller lecithin:A-I complex containing 200-250 mol of phospholipid and two A-I molecules per complex which has been previously reported. At BBSM:A-I ratios of 15:1 (w/w), a new type of complex was observed which was discoidal by electron microscopy but possessed a larger diameter (390 A) and higher phospholipid:protein molar ratio (535:1) than has been observed previously for recombinant complexes. The BBSM:A-I complexes were found to be significantly more resistant to denaturation by guanidine hydrochloride than the dimyristoyl phosphatidylcholine:A-I recombinant complexes. It is concluded that the mechanisms of interaction between apolipoprotein A-I and either bovine brain sphingomyelin or phosphatidylcholines are similar, but that the nature of the protein-lipid interactions with BBSM are such as to produce larger and more stable complexes than are observed with the phosphatidylcholines.  相似文献   

16.
Chinese hamster ovary cells transfected with the human apolipoprotein A-I gene linked to the human metallothionein gene promoter region secrete large quantities of apolipoprotein A-I (7.1 +/- 0.4% total secreted protein) in the presence of zinc. Approx. 16% of the secreted apolipoprotein A-I is complexed with lipid and can be isolated ultracentrifugally at d less than or equal to 1.21 g/ml. The latter complexes are composed of discs and vesicles as judged by electron microscopy and can be further separated by column chromatography into three fractions: fraction I, mostly vesicles (60-260 nm) and large discs (18-20 nm diameter); fraction II, discs 14.2 +/- 2.6 nm diameter; and fraction III, nonresolvable by electron microscopy. The latter fraction is extremely lipid-poor (94% protein, 6% phospholipid); in contrast, the protein, phospholipid and unesterified cholesterol content for the other fractions are 43, 33 and 24%, respectively, for fraction I and 53, 33 and 14%, respectively, for fraction II. Fraction II particles contain three and four apolipoprotein A-Is per particle as determined by protein crosslinking while large structures in fraction I contain primarily six to seven apolipoprotein A-Is per particle. Following incubation with purified lecithin: cholesterol acyltransferase, discoidal particles were transformed into apparent spherical particles 12.9 +/- 3.4 nm diameter; this transformation coincided with 19-21% conversion of unesterified cholesterol to esterified cholesterol. The apolipoprotein A-I-lipid complexes isolated from Chinese hamster ovary cell media are similar to nascent HDL found in plasma of lecithin:cholesterol acyltransferase-deficient patients and those secreted by the human hepatoma line, Hep G2. The ability of the Chinese hamster ovary cell nascent HDL-like particles to undergo transformation in the presence of purified lecithin:cholesterol acyltransferase indicates that they are functional particles.  相似文献   

17.
Fluorescence lifetime and intensity quenching studies of human plasma apolipoprotein A-I (apo A-I) in aqueous solution and in recombinant lipoprotein complexes with dimyristoylphosphatidylcholine (DMPC) indicate differences in conformational dynamics. In aqueous solution, the bimolecular quenching constants (k*) for lipid-free apo A-I fluorescence quenching by oxygen and acrylamide are 2.4 X 10(9) and 0.38 X 10(9) M-1 s-1, respectively. These values are independent of the oligomeric form of the protein. There is no correlation between the relatively small k* for apo A-I, which reflects rapid, low-amplitude protein fluctuations, and the labile conformational changes of apo A-I folding reactions, like denaturation, which occur on a slower time scale. In recombinant DMPC/apo A-I complexes (100:1 molar ratio) the protein increases in amphiphilic alpha-helical structure as it blankets the lipid matrix. The apparent k* for oxygen quenching of apo A-I fluorescence in the complex is large and increases in a temperature-dependent manner. We have introduced a two-compartment model, which discriminates the source of quencher molecules as aqueous or lipid, to describe oxygen quenching of DMPC/apo A-I fluorescence. The magnitude and temperature dependence of the apparent k* predominantly reflect the partitioning of oxygen between the two phases rather than being a probe of the lipid physical state. Calculations of the helical hydrophobic moment in apo A-I indicate that tryptophan residues 8 and 72 occur at the lipid-protein interface of amphiphilic alpha-helices, whereas the other two tryptophan residues (50, 108) lie on the nonpolar faces of amphiphilic helices.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Reassembly experiments, involving isolated human apoproteins A-I and A-II and (dimyristoylglycerophosphocholine)-cholesterol vesicles were performed with apoprotein mixtures at apoprotein A-I/A-II molar ratios varying between 0 and 3. The apoproteins were incubated at 24 degrees C. 28 degrees C and 32 degrees C with either pure dimyristoyl-glycerophosphocholine vesicles or with dimyristoylglycerophosphocholine cholesterol vesicles containing 2, 5, 10, 15 mol/100 mol cholesterol. The kinetics of association were followed by measuring the increase of the fluorescence polarization ratio after labeling the lipids with diphenyl hexatriene. The complexes were separated from the free protein by gradient ultracentrifugation. Total protein was assayed and the apoproteins A-I and A-II were quantified separately by immunonephelometry. The content of apoprotein A-I was also monitored by measuring the intrinsic tryptophan fluorescence. The results suggest that apoprotein A-II has a greater affinity than apoprotein A-I for the phospholipid-cholesterol vesicles and that apoprotein A-II is able to quantitatively displace apoprotein A-I from the lipid-protein complexes. The content of apoprotein A-II in the complexes increases proportionally to the concentration of apoprotein A-II in the incubation mixture until saturation is reached. At saturation the dimyristoylglycerophosphocholine/apoprotein A-II ratio in the complex is dependent upon the cholesterol content of the original vesicles and increases from 60 to 275 mol/mol between 0 and 15 mol/100 mol cholesterol. From these experiments one can calculate that 1 mol human apoprotein A-I is displaced by 2 mol human apoprotein A-II.  相似文献   

19.
Apolipoprotein A-I (apoA-I) interaction with specific cell lipid domains was suggested to trigger cholesterol and phospholipid efflux. We analyzed here apoA-I interaction with dimyristoylphosphatidylcholine/distearoylphosphatidylcholine (DMPC/DSPC) bilayers at a temperature showing phase coexistence. Solid and liquid-crystalline domains were visualized by two-photon fluorescence microscopy on giant unilamellar vesicles (GUVs) labeled with 6-dodecanoyl-2-dimethyl-amino-naphthalene (Laurdan). A decrease of vesicle size was detected as long as they were incubated with lipid-free apoA-I, together with a shape deformation and a relative enrichment in DSPC. Selective lipid removal mediated by apoA-I from different domains was followed in real time by changes in the Laurdan generalized polarization. The data show a selective interaction of apoA-I with liquid-crystalline domains, from which it removes lipids, at a molar ratio similar to the domain compositions. Next, apoA-I was incubated with DMPC/DSPC small unilamellar vesicles, and products were isolated and quantified. Protein solubilized both lipids but formed complexes relatively enriched in the liquid component. We also show changes in the GUV morphology when cooling down. Our results suggest that the most efficient reaction between apoA-I and DMPC/DSPC occurs in particular bilayer conditions, probably when small fluid domains are nucleated within a continuous gel phase and interfacial packing defects are maximal.  相似文献   

20.
The reversibility of the binding of human apolipoprotein A-I (apo A-I) to phospholipid has been monitored through the influence of guanidine hydrochloride (Gdn-HCl) on the isothermal denaturation and renaturation of apo A-1/dimyristoylphosphatidylcholine (DMPC) complexes at 24 degree C. Denaturation was studied by incubating discoidal 1:100 and vesicular 1:500 mol/mol apo A-I/DMPC complexes with up to 7 M Gdn-HCl for up to 72 h. Unfolding of apo A-I molecules was observed from circular dichroism spectra while the distribution of protein between free and lipid-associated states was monitored by density gradient ultracentrifugation. The ability of apo A-I to combine with DMPC in the presence of Gdn-HCl at 24 degrees C was also investigated by similar procedures. In both the denaturation and renaturation of 1:100 and 1:500 complexes, the final values of the molar ellipticity and the ratio of free to bound apo A-I at various concentrations of Gdn-HCl are dependent on the initial state of the lipid and protein; apo A-I is more resistant to denaturation when Gdn-HCl is added to existing complexes than to a mixture of apo A-I and DMPC. There is an intermediate state in the denaturation pathway of apo A-I/DMPC complexes which is not present in the renaturation; the intermediate comprises partially unfold apo A-I molecules still associated with the complex by some of their apolar residues. Complete unfolding of the alpha helix and subsequent desorption of the apo A-I molecules from the lipid/water interface involve cooperative exposure of these apolar residues to the aqueous phase. The energy barrier associated with this desorption step makes the binding of apo A-I to DMPC a thermodynamically irreversible process. Consequently, binding constants of apo A-I and PC cannot be calculated simply from equilibrium thermodynamic treatments of the partitioning of protein between free and bound states. Apo A-I molecules do not exchange freely between the lipid-free and lipid-bound states, and extra work is required to drive protein molecules off the surface. The required increased in surface pressure can be achieved by a net mass transfer of protein to the surface; in vivo, increases in the surface pressure of lipoproteins by lipolysis can cause protein desorption.  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