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 615 毫秒
1.
Glucagon can interact rapidly with multilamellar vesicles of dimyristoyl glycerophosphocholine over a narrow temperature range around or above the phase transition temperature of the pure phospholipid. The temperature dependence of the rates arises, in large part, from glucagon-induced alterations in the phase transition properties of the phospholipid. Similar effects are observed with dilaury glycerophosphocholine but the rate of reaction of glucagon with multilamellar dipalmitoyl glycerophosphocholine is too slow to measure.The rate of reaction of glucagon with equimolar mixtures of two phospholipid molecules has also been studied. Mixtures of dilauryl glycerophosphocholine and distearoyl glycerophosphocholine are known to exhibit lateral phase separation in the gel state. The presence of distearoyl glycerophosphocholine has no effect on the rate of reaction with glucagon, despite the increased number of phase boundaries present. In the case of mixtures of dilauryl glycerophosphocholine and dimyristoyl glycerophosphocholine, glucagon appears to induce some lateral phase separation. This is demonstrated by the ability of glucagon to react rapidly with this lipid mixture, even at temperatures well below the phase transition temperature of the mixture and by differential scanning calorimetry.The thermodynamics of the binding of glucagon to dimyristoyl glycerophosphocholine and dilauryl glycerophosphocholine were analyzed with Scatchard plots calculated from measurements of the fluorescence enhancement caused by lipids. Equilibrium binding constants of glucagon to dimyristoyl glycerophosphocholine and dilauryl glycerophosphocholine are 1·105 and 5·104 M?1, respectively. These values are relatively insensitive to temperature, indicating that the equilibrium being measured is between lipid-bound glucagon and free lipid which has had its phase transition properties altered. The number of moles of lipid bound per mole of glucagon decreases markedly above the phase transition temperature. In the water-soluble complex formed between glucagon and dimyristoyl glycerophosphocholine, the peptide binds directly to only 40% of the lipid molecules but, nevertheless, is able to modify the phase transition properties of all of the lipid in the particle.  相似文献   

2.
We have studied the influence of changes in lipid organization on the planar distribution of two classes of membrane proteins: integral proteins which have amino groups exposed to labelling at the membrane surface by the biotin-avidin-ferritin procedure, and those proteins which penetrate the lipid bilayer sufficiently to be seen as intramembranous particles by freeze-fracture electron-microscopy.When the membranes are examined at temperatures below the lipid phase transition, the first class is dispersed and the second patched. At temperatures in the middle of the transition range, both classes are patched. At temperatures just above the phase transition the first class is dispersed and the second patched, and at temperatures well above the transition both classes are dispersed. Freeze-etch studies of avidin-ferritin-labeled membranes confirmed that the distribution seen by the labeling and the freeze-fracture techniques coexist in single membranes. Thus, there exist two distinct classes of membrane proteins with differential organizational responses to the lipid state.  相似文献   

3.
(1) Rat liver microsomes were studied by freeze-fracture electron microscopy. The distribution of membrane-associated particles indicated the right-side-out orientation of microsomal vesicles. Studies at different temperatures were performed. At 30 degrees C membrane-associated particles are randomly distributed on membrane A-faces, while aggregations of particles are observed at 4 degrees C. (2) Aggregation is dependent on the cooling rates. It can be prevented by shock-freezing. (3) Particle aggregation is also prevented by cholesterol, added to the microsomes in equal molar ratio to the microsomal phospholid content. (4) These findings suggest that particle aggregation is caused by a partial freezing-out of phospholipid molecules during the phase transition from the liquid-crystalline to the gel state. (5) The results are discussed with respect to an observed increase in activation energy of microsomal drug monooxygenation at lower temperature.  相似文献   

4.
Human high‐density lipoprotein (HDL) plays a key role in the reverse cholesterol transport pathway that delivers excess cholesterol back to the liver for clearance. In vivo, HDL particles vary in size, shape and biological function. The discoidal HDL is a 140–240 kDa, disk‐shaped intermediate of mature HDL. During mature spherical HDL formation, discoidal HDLs play a key role in loading cholesterol ester onto the HDL particles by activating the enzyme, lecithin:cholesterol acyltransferase (LCAT). One of the major problems for high‐resolution structural studies of discoidal HDL is the difficulty in obtaining pure and, foremost, homogenous sample. We demonstrate here that the commonly used cholate dialysis method for discoidal HDL preparation usually contains 5–10% lipid‐poor apoAI that significantly interferes with the high‐resolution structural analysis of discoidal HDL using biophysical methods. Using an ultracentrifugation method, we quickly removed lipid‐poor apoAI. We also purified discoidal reconstituted HDL (rHDL) into two pure discoidal HDL species of different sizes that are amendable for high‐resolution structural studies. A small rHDL has a diameter of 7.6 nm, and a large rHDL has a diameter of 9.8 nm. We show that these two different sizes of discoidal HDL particles display different stability and phospholipid‐binding activity. Interestingly, these property/functional differences are independent from the apoAI α‐helical secondary structure, but are determined by the tertiary structural difference of apoAI on different discoidal rHDL particles, as evidenced by two‐dimensional NMR and negative stain electron microscopy data. Our result further provides the first high‐resolution NMR data, demonstrating a promise of structural determination of discoidal HDL at atomic resolution using a combination of NMR and other biophysical techniques.  相似文献   

5.
E A Bonomo  J B Swaney 《Biochemistry》1990,29(21):5094-5103
Plasma high density lipoproteins (HDL) are synthesized in intestinal mucosal cells and hepatocytes and are secreted into the blood. Factors influencing the structure and function of these HDL, such as lipid and protein composition, are poorly understood. It appears, however, that intracellular, discoidal HDL are enriched, relative to plasma HDL, in phosphatidylethanolamine (PE), a phospholipid known to generate unusual, nonbilayer structures of putative physiological significance. Although incubation of dimyristoylphosphatidylcholine (DMPC) with apolipoprotein A-I at the gel-liquid crystalline phase transition temperature results in the spontaneous formation of lipid-protein complexes, the presence of proportionately small amounts of PE prevents the formation of such complexes, suggesting that PE profoundly alters the phase properties of the phospholipid bilayers. However, by using a detergent-mediated method for the formation of PE-rich model nascent HDL from phospholipids and apolipoprotein A-I, lipid-protein complexes containing as much as 75% DLPE could be formed, thus demonstrating that the presence of PE causes a kinetic, rather than a thermodynamic, barrier to spontaneous complex formation. The products contained a DLPE:DMPC molar ratio similar to that of the initial incubation mixture; however, as the mole percentage of DLPE increased, the products became less heterogeneous, the buoyant density of the products increased, and the Stokes diameter of the products decreased. Similar results were obtained when dimyristoylphosphatidylethanolamine (DMPE) and dipalmitoylphosphatidylethanolamine (DPPE) were employed in lieu of DLPE. Electron microscopy of complexes containing DLPE and DMPC at a 1:1 molar ratio showed that these particles possessed a discoidal, bilayer morphology similar to that seen with complexes containing only phosphatidylcholine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
We have previously described the reconstitution of functional membrane vesicles with lipid content similar to that of the normal sarcoplasmic reticulum membrane (approximately 1.0 mumol of phospholipid/mg of protein). The present study describes methodology to prepare reconstituted membrane vesicles with defined phospholipid to protein ratio, both lower and higher than that of the original membrane. The Ca2+ loading rate and efficiency are greatest in the membranes of highest protein content (0.38 mumol of phospholipid/mg of protein), decline slowly as the lipid content is quadrupled, and decrease markedly as the lipid content is quadrupled again. Such membranes of defined composition can be used to study lipid-protein interaction and to correlate membrane structure with composition. The number of particles observed by freeze-fracture electron microscopy can be correlated with protein content, whereas the percentage of smooth domain is proportional to the lipid content of the reconstituted membrane. Since 90% or more of the protein of the reconstituted membrane is the calcium pump protein, the number of particles observed by freeze-fracture is directly proportional to the amount of calcium pump protein in the membrane. The number of pump molecules calculated to be in the membrane is greater by a factor of two than the number of particles which we observed. This multiplicity ratio could be greater depending upon the assumptions made regarding the width of the membrane (see "Appendix"). Thus, it would appear that the particles consist of two or more molecules of pump protein. The change in protein concentration of the membrane is reflected also in thin sections and by negative staining. In thin sections, the broad inner and outer 70 A bands become discontinuous and patchy and, in the limit, approach a symmetrical 20,20,20 A trilayer as the protein content of the membrane becomes small. In an analogous fashion, the concentration of particles at the surface of the membrane, observed by negative staining, decreases with increasing lipid concentration in the membrane. Thus, the correlation of composition with structure can be observed by each of the three methods of sample preparation for electron microscopic analysis.  相似文献   

7.
Two biologically active, 34 amino acid fragments of parathyroid hormone interact with dimyristoylphosphatidylcholine to form lipoprotein particles. In the lipid-bound form these parathyroid hormone peptides exhibit an increased amount of folded secondary structure and the tryptophan residue of [Nle8, Nle18, Tyr34] b PTH (1-34) amide appears to become buried in a more hydrophobic environment. The lipoprotein particle which is formed has dimensions of approximately 65 X 7 nm but aggregates to larger structures with increasing temperature. Above the phase transition of the phospholipid the peptides no longer affect the morphology of the lipid and the spectral properties of the peptide are not perturbed by the lipid. This is similar to the behavior of glucagon with dimyristoylphatidylcholine. The results indicate that several nonhomologous peptide hormones have common features which allow them to fold into an amphipathic helix and solubilize phospholipid.  相似文献   

8.
A purified protein fraction from the proteolipids of human brain myelin was recombined with different lipids either in aqueous buffer or in a chloroform-methanol-water (10:5:1, v/v/v) mixture. It was found that under both conditions it binds strongly to phospholipids irrespective of surface charge, the presence of cholesterol or double bonds on the fatty acyl chains. The buoyant density of the resulting lipoprotein membranes is intermediate to that of pure lipids, and proteins. The lipoproteins formed by either of these methods were observed by either freeze-fracture or negative stain electron-microscopy. The overall morphology was similar to that of pure phospholipids, showing large closed multilamellar vesicles. The presence of the protein was detected by the appearance of intramembrane particles in freeze-fracture. The addition of the N-2 protein generally increases the permeability vesicles to 22-Na-+ by 2-3 orders of magnitude depending on the concentration. The presence of calcium in the aqueous medium further increases the Na-+ efflux through negatively charged vesicles. Changes in lipid composition, surface charge, cholesterol, etc., have no appreciable influence on the effect of the protein. Differential scanning calorimetry indicates that the presence of small amounts of N-2 have no effect on the lipid phase transition from solid to liquid crystalline. As the amount of protein bound to the phospholipid increases, the enthalpy of the transition decreases, the main endothermic peak broadens, but there is no change on the midpoint temperature. Membranes containing 50% by weight of protein still show a transition with an enthalpy approximately one half that of the original lipid.  相似文献   

9.
J L Soulages  E L Arrese 《Biochemistry》2001,40(47):14279-14290
Quenching of tryptophan fluorescence by nitroxide-labeled phospholipids and nitroxide-labeled fatty acids was used to investigate the lipid-binding domains of apolipophorin III. The location of the Trp residues relative to the lipid bilayer was investigated in discoidal lipoprotein particles made with 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine and five different single-Trp mutants of apoLp-III. A comparison of the quenching efficiencies of phospholipids containing nitroxide groups at the polar head, and at positions 5 and 16 of the sn-2 acyl chain, indicated that the protein is interacting with the acyl chains of the phospholipid along the periphery of the bilayer of the discoidal lipoprotein. N-Bromosuccinimide readily abolished 100% of the fluorescence of all Trp residues in the lipid-bound state. Larger quenching rates were observed for the Trp residues in helices 1, 4, and 5 than for those located in helices 2 and 3, suggesting differences between the interaction of these two groups of helices. However, the extent of Trp fluorescence quenching observed in lipoproteins made with any of the mutants was comparable to that reported for deeply embedded Trp residues, suggesting that all Trp residues interact with the phospholipid acyl chains. This study provides the first experimental evidence of a massive interaction of the alpha-helices of apoLp-III with the phospholipid acyl chains in discoidal lipoproteins. The extent of interaction deduced is consistent with the apolipoprotein adopting a highly extended conformation.  相似文献   

10.
Two fluorescent probes-cis- and trans-parinaric acids were used to study the dimensions, lipid dynamics and apolipoprotein location in the reconstituted discoidal high density lipoproteins (rHDL). The rHDL particles made from apolipoprotein A-I (apoA-I), dipalmitoylphosphatidylcholine (DPPC), with or without cholesterol (Chol) were compared with the analogous particles with two other apolipoproteins-apoE and apoA-II. The data obtained for apoA-I-containing rHDL were as follows: (1) the inclusion of 8 mol.% of cholesterol did not significantly change the particle dimensions (13+/-1 nm) or the mean distance between apoA-I and the disc axis; (2) the phospholipid domains-boundary lipid region in the close vicinity to apoA-I molecule and the remaining part of the bilayer-existed at temperatures both lower and above DPPC transition temperature T(t); (3) at T相似文献   

11.
The structure of discoidal apo A-I-phospholipid complexes, representing the metabolic precursors of mature high-density lipoprotein particles, was studied by a combination of both a theoretical and an experimental approach. The secondary structure of the complex was determined by circular dichroic measurements, while the relative orientation of the apo A-I helical segments and of the phospholipid acyl chains was determined by ATR infrared measurements. Fluorescence energy transfer between the tryptophan residues of apo A-I and fluorescent phospholipid probes yielded an estimation of the relative topography of the lipid and apolipoprotein components in discoidal and spherical particles. The theoretical approach consisted of the identification of the helical segments in various apo A-I species. These segments were then oriented at a lipid/water interface by minimization of their hydrophobic and hydrophilic transfer energies. The calculation of the hydrophobicity profiles along the axis of the helices leads to the identification of specific interactions between pairs of helices. The helices were further assembled together with the phospholipids by computer modelling, enabling an estimation of the dimensions of the complex. The combination of the experimental and theoretical results yielded a model for discoidal apolipoprotein-phospholipid complexes, in which the amphipathic helical segments are oriented along the edges of the discs. Such a model can be extended to the conversion of these complexes into mature spherical HDL, through the formation of a cholesteryl ester core.  相似文献   

12.
For several decades, the standard model for high density lipoprotein (HDL) particles reconstituted from apolipoprotein A-I (apoA-I) and phospholipid (apoA-I/HDL) has been a discoidal particle ∼100 Å in diameter and the thickness of a phospholipid bilayer. Recently, Wu et al. (Wu, Z., Gogonea, V., Lee, X., Wagner, M. A., Li, X. M., Huang, Y., Undurti, A., May, R. P., Haertlein, M., Moulin, M., Gutsche, I., Zaccai, G., Didonato, J. A., and Hazen, S. L. (2009) J. Biol. Chem. 284, 36605–36619) used small angle neutron scattering to develop a new model they termed double superhelix (DSH) apoA-I that is dramatically different from the standard model. Their model possesses an open helical shape that wraps around a prolate ellipsoidal type I hexagonal lyotropic liquid crystalline phase. Here, we used three independent approaches, molecular dynamics, EM tomography, and fluorescence resonance energy transfer spectroscopy (FRET) to assess the validity of the DSH model. (i) By using molecular dynamics, two different approaches, all-atom simulated annealing and coarse-grained simulation, show that initial ellipsoidal DSH particles rapidly collapse to discoidal bilayer structures. These results suggest that, compatible with current knowledge of lipid phase diagrams, apoA-I cannot stabilize hexagonal I phase particles of phospholipid. (ii) By using EM, two different approaches, negative stain and cryo-EM tomography, show that reconstituted apoA-I/HDL particles are discoidal in shape. (iii) By using FRET, reconstituted apoA-I/HDL particles show a 28–34-Å intermolecular separation between terminal domain residues 40 and 240, a distance that is incompatible with the dimensions of the DSH model. Therefore, we suggest that, although novel, the DSH model is energetically unfavorable and not likely to be correct. Rather, we conclude that all evidence supports the likelihood that reconstituted apoA-I/HDL particles, in general, are discoidal in shape.  相似文献   

13.
Apolipophorin III (ApoLp-III) from the Sphinx moth, Manduca sexta, is an 18kDa protein that binds reversibly to hydrophobic surfaces generated on metabolizing lipoprotein particles. It is comprised of amphipathic alpha-helices (H1-H5) organized in an up-and-down topology forming a helix bundle in the lipid-free state. Upon interaction with lipids, apoLp-III has been proposed to undergo a dramatic conformational change, involving helix bundle opening about putative hinge loops such that H1, H2 and H5 move away from H3 and H4. In the present study, we examine the relative spatial disposition of H1 and H5 on discoidal phospholipid complexes and spherical lipoproteins. Cysteine residues were engineered at position 8 in H1 and/or at position 138 in H5 in apoLp-III (which otherwise lacks Cys) yielding A8C-, A138C- and A8C/A138C-apoLp-III. Tethering of H1 and H5 by a disulfide bond between A8C and A138C abolished the ability of apoLp-III to transform phospholipid vesicles to discoidal particles, or to interact with lipoproteins, demonstrating that these helices are required to reposition during lipid interaction. Site-specific labeling of A8C/A138C-apoLp-III with N-(1-pyrene)maleimide in the lipid-free state resulted in intramolecular pyrene "excimer" fluorescence emission indicative of spatial proximity between these sites. Upon association with dimyristoylphosphatidylcholine (DMPC) discoidal complexes, the intramolecular excimer was replaced by intermolecular excimer fluorescence due to proximity between pyrene moieties on A8C and A138C in neighboring apoLp-III molecules on the discoidal particle. No excimer emission was observed in the case of pyrene-A8C-apoLp-III/DMPC or pyrene-A138C-apoLp-III/DMPC complexes. However, equimolar mixing of the two labeled single-cysteine mutants prior to disc formation resulted in excimer emission. In addition, intramolecular pyrene excimer formation was diminished upon binding of pyrene-A8C/A138C-apoLp-III to spherical lipoproteins. The data are consistent with repositioning of H1 away from H5 upon encountering a lipid surface, resulting in an extended conformation of apoLp-III that circumscribes the discoidal bilayer particle.  相似文献   

14.
In the presence of either egg or bovine brain sphingomyelin, the spectral properties of glucagon undergo changes which are similar to those which occur in the presence of synthetic phosphatidylcholines. The fluorescence emission spectra are blue shifted about 10 nm in the presence of lipid and the peptide acquires an increased helical content, determined by circular dichroism. As with phosphatidylcholines, the changes in spectral properties do not occur above the phase transition temperature of the glucagon-lipid mixture. Freeze-fracture electron microscopy indicates that glucagon forms an ellipsoidal complex with bovine brain sphingomyelin, similar to the glucagon-dimyristoylphosphatidylcholine complex. However, the sphingomyelin complexes break down to vesicular structures both above and below the region of the phase transition. These results indicate that the dissociation of glucagon from the lipid at higher temperatures results from changes in the phase of the lipid rather than from a thermal denaturation of glucagon. The effect of glucagon on the phase transition behaviour of palmitoyl sphingosine phosphorylcholine was measured by differential scanning calorimetry. The major effect of glucagon on both this lipid and on dimyristoylphosphatidylcholine is to broaden the phase transition and to shift it to higher temperatures. Similar results are obtained for the effects of glucagon on an equimolar mixture of dimyristoylphosphatidylcholine and palmitoyl sphingosine phosphorylcholine. Glucagon is able to solubilize mixtures of bovine brain sphingomyelin with either dimyristoylphosphatidylcholine or egg lecithin. The lipid composition of the solubilized material is similar to that of the starting lipid film. These results together with those from the differential scanning calorimetry on the synthetic mixtures indicate that glucagon can bind to sphingomyelin-phosphatidylcholine mixtures and that it does not induce extensive lateral phase separation between the components. The maximal stability of the glucagon-lipid complex at the phase transition of the lipids indicates that the glucagon-lipid interaction is highly dependent on the structural organization of the lipid.  相似文献   

15.
Glucagon is found to interact with dimyristoyl glycerophosphocholine both above and below the phase transition temperature of the lipid. Above the phase transition temperature the interaction is manifested by an increase in the rate of vesicle aggregation and by an increased permeability of unilamellar vesicles to Eu3+ and to Fe(CN)63−. However, no stable lipoprotein complex can be detected by gel filtration. Below the phase transition glucagon can form stable complexes with dimyristoyl glycerophosphocholine vesicles but cannot rapidly rearrange these vesicles to disk-shaped particles until the phase transition temperature is approached. The energy of activation for the dissociation of glucagon from the disk-shaped lipoprotein particle is 29 kcal/mol at temperatures above 36°C but increases markedly at lower temperatures, as the region of the lipid phase transition is approached. This increase in energy of activation at lower temperatures is most probably due to the larger amount of energy required to rearrange gel-state lipid in the transition state and provides an explanation for the unusual kinetic stability of the glucagon-dimyristoyl glycerophosphocholine lipoprotein complex only at temperatures below the phase transition of the lipid.  相似文献   

16.
随着各种诱导膜融合的因子相继发现,人们建立了各种膜融合的模型.我们通过对聚乙二醇PEG诱导脂质体融合的分析,认为膜融合的关键在于脱去膜表面的结合水,而其它作用诸如膜脂缺陷.膜脂分相以及脂多型性等尽管是不同膜体系中直接观察到的膜融合形式,都是膜脱去结合水带来的必然结果.膜表面结合水的排除是前因,本文着重讨论脱水及脱水后膜脂结构的一系列变化.  相似文献   

17.
Since spheroidal HDL particles (sHDL) are highly dynamic, molecular dynamics (MD) simulations are useful for obtaining structural models. Here we use MD to simulate sHDL with stoichiometries of reconstituted and circulating particles. The hydrophobic effect during simulations rapidly remodels discoidal HDL containing mixed lipids to sHDL containing a cholesteryl ester/triglyceride (CE/TG) core. We compare the results of simulations of previously characterized reconstituted sHDL particles containing two or three apoA-I created in the absence of phospholipid transfer protein (PLTP) with simulations of circulating human HDL containing two or three apoA-I without apoA-II. We find that circulating sHDL compared with reconstituted sHDL with the same number of apoA-I per particle contain approximately equal volumes of core lipid but significantly less surface lipid monolayers. We conclude that in vitro reconstituted sHDL particles contain kinetically trapped excess phospholipid and are less than ideal models for circulating sHDL particles. In the circulation, phospholipid transfer via PLTP decreases the ratio of phospholipid to apolipoprotein for all sHDL particles. Further, sHDL containing two or three apoA-I adapt to changes in surface area by condensation of common conformational motifs. These results represent an important step toward resolving the complicated issue of the protein and lipid stoichiometry of circulating HDL.  相似文献   

18.
In circulation the phospholipid transfer protein (PLTP) facilitates the transfer of phospholipid-rich surface components from postlipolytic chylomicrons and very low density lipoproteins (VLDL) to HDL and thereby regulates plasma HDL levels. To study the molecular mechanisms involved in PLTP-mediated lipid transfer, we studied the interfacial properties of PLTP using Langmuir phospholipid monolayers and asymmetrical flow field-flow fractionation (AsFlFFF) to follow the transfer of 14C-labeled phospholipids and [35S]PLTP between lipid vesicles and HDL particles. The AsFlFFF method was also used to determine the sizes of spherical and discoidal HDL particles and small unilamellar lipid vesicles. In Langmuir monolayer studies high-activity (HA) and low-activity (LA) forms of PLTP associated with fluid phosphatidylcholine monolayers spread at the air/buffer interphase. Both forms also mediated desorption of [14C]dipalmitoylphosphatidylcholine (DPPC) from the phospholipid monolayer into the buffer phase, even when it contained no physiological acceptor such as HDL. After the addition of HDL3 to the buffer, HA-PLTP caused enhanced lipid transfer to them. The particle diameter of HA-PLTP was approximately 6 nm and that of HDL3 approximately 8 nm as determined by AsFlFFF analysis. Using this method, it could be demonstrated that in the presence of HA-PLTP, but not LA-PLTP, [14C]DPPC was transferred from small unilamellar vesicles (SUV) to acceptor HDL3 molecules. Concomitantly, [35S]-HA-PLTP was transferred from the donor to acceptor, and this transfer was not observed for its low-activity counterpart. These observations suggest that HA-PLTP is capable of transferring lipids by a shuttle mechanism and that formation of a ternary complex between PLTP, acceptor, and donor particles is not necessary for phospholipid transfer.  相似文献   

19.
The polypeptide gramicidin A in a dimeric form is considered to form a helical structure which spans the hydrocarbon region of lipid bilayers. In the present investigation it is used as a model for the interactions of the polypeptide segments of transmembrane proteins within the hydrocarbon region of the lipid bilayers of biomembrane structures. A variety of physical techniques (X-ray diffraction, differential scanning calorimetry, optical and electron microscopy, Raman and electron spin resonance spectroscopy) are applied to a study of the interactions of this polypeptide within the phospholipid bilayers of dimyristoyl and dipalmitoyl lecithins in water, at temperatures both above and below the main endothermic phase transition of the pure lipids.Above the transition temperature of the lipid, the Raman studies show that the polypeptide perturbs the fluid lipid environment and causes a marked decrease in the number of gauche isomers of the lipid hydrocarbon chains, even at quite low relative molar concentrations of the polypeptide to lipid (1:150). At concentrations of phospholipid to polypeptide of less than 5:1, the electron spin resonance studies show the existence of two lipid regions within the bilayer. One region corresponds to the relatively fluid lipid region normally observed at these temperatures and the other to a relatively rigid lipid region. The latter is considered to arise from clusters of the polypeptide in which some of the lipid is entrapped.Below the lipid phase transition temperature, the pretransition endotherm observed with pure lipid-water systems is removed by small molar concentrations of the polypeptide (1:50) and the rippled appearance observed in freeze-fracture electron micrographs with pure dimyristoyl lecithin-water dispersions is replaced by a smooth appearance.The main lipid phase transition becomes broadened by the presence of increasing amounts of the polypeptide within the lipid bilayer as indicated by calorimetry, and electron spin resonance spectroscopy. The enthalpy of the lipid transition decreases linearly with increasing amounts of the polypeptide until, with dipalmitoyl lecithin, a concentration of approximately 20 lipids per polypeptide is reached. This is considered to correspond to the onset of an aggregation process which produces localised polypeptide-lipid clusters within the plane of the membrane.At concentrations of polypeptide less than five lipids per polypeptide, freezefracture electron microscopy shows the presence of liposomes with smooth fracture faces. At higher polypeptide concentrations, sheet-like structures are observed with smooth fracture faces.When a mixed lipid-water system (dilauroyl and dipalmitoyl lecithin) containing low concentrations of the polypeptide is slowly cooled, the calorimetric evidence shows that the polypeptide moves preferentially into the lower melting region of the bilayer, whereas at higher polypcptide eoncentrations a mixing of the two lipids takes place.The various results are discussed to provide insight pertinent to the organisation, interactions, aggregation properties, boundary layer and packing arrangements of helical polypeptides and proteins in reconstituted systems and natural biomembranes.  相似文献   

20.
The interaction between glucagon and dicaprylphosphatidylcholine (DCPC) was studied by fluorescence, circular dichroism and calorimetry, as well as by 1H- and 31P-nuclear magnetic resonance. The water-soluble lipid-protein complex was also characterized by gel filtration and ultracentrifugation. The complex appeared to be monodisperse by sedimentation equilibrium measurements, with a molecular weight of (4.55 ± 0.57)·104. This complex contained approximately 7 molecules of glucagon and 35 molecules of phospholipid. Proton-decoupled 31P-NMR spectra of the phospholipid in the lipid-protein complex display narrower resonances than those of sonicated vesicles of DCPC, and 1H-31P coupling could be detected in proton coupled spectra. These NMR results, together with gel-filtration results, suggest that glucagon ‘solubilizes’ phospholipid aggregates, forming a lipid-protein complex which is smaller than sonicated preparations of DCPC. 1H-NMR resonance of both the methionine methyl group (met-27) and the aromatic envelope of glucagon are broadened by the phospolipid, indicating that the C-terminal region and the aromatic residues are involved in the interaction with the phospholipid. Nuclear magnetic resonance titrations of the imidazole ring C(2) and C(4) protons of the histidine residue of glucagon show that DCPC lowers the pK of the imidazole. The alterations caused by the phospholipid in the far and near ultraviolet CD spectra of glucagon reflect, respectively, the increased helix content of the hormone and the fact that the aromatic residues are located in a more structured environment. The phospholipid also alters the fluorescence properties of glucagon, shifting the fluorescence emission maximum of the hormone to shorter wavelength, and enhancing its relative intensity. This suggests that the fluorophore is experiencing a more hydrophobic environment in the presence of the lipid. Binding of glucagon to the phospholipid was analysed by Scatchard plots of the enhancement of fluorescence caused by the phospholipid and showed that the equilibrium binding constants of glucagon to DCPC are (4.4 ± 0.5)·104M?1 and (7.5±0.5)·104M?1, at 15°C and 25°C, respectively. The average number of moles of phospholipid bound per mole of glucagon is 4.4±0.6. The isothermal enthalpy of reaction of glucagon with DCPC is ?20.5 kcal/mol of glucagon at 25°C and ?32.5 kcal/mol of glucagon at 15°C. The observed enthalpies can arise from glucagon-induced cyrstallization of the phospholipid, from the non-covalent interactions between the peptide and lipid as well as from the lipid-induced conformational change in the protein. These results demonstrate that, unlike the complexes formed between glucagon and phospholipids which form more stable bilayers, the complex formed between glucagon and DCPC is stable over a wide range of temperatures, including temperatures well above the phase transition.  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