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 effect of cardiotoxin IIa from Naja mossambica mossambica, a small basic protein extracted from snake venom, on dimyristoylphosphatidic acid (DMPA) and on equimolar mixtures of DMPA and dimyristoylphosphatidylcholine (DMPC) has been studied by Fourier transform infrared spectroscopy. The interaction of cardiotoxin with DMPA dispersions decreases both the cooperativity of the phase transition of the lipid and the molecular order of the lipid acyl chains in the gel phase. This effect increases with the proportion of the toxin in the complexes and leads to the total abolition of the phase transition of DMPA at a lipid-to-protein molar ratio of 5. Small-angle X-ray results demonstrate that the structure of the lipid-protein complexes is poorly ordered and gives rise to broad diffusion peaks rather than to well-resolved diffraction patterns. Infrared spectra of oriented cardiotoxin-DMPA films show that the protein is not homogeneously oriented with respect to the bilayer surface. The destabilization of the gel-phase structure of DMPA by cardiotoxin also results in a deeper water penetration in the interfacial region of the lipid since more carbonyl ester groups appear to be hydrogen bonded in the presence of the toxin. The infrared results on the phosphate group vibrations also indicate clearly that the basic residues of cardiotoxin interact strongly with the phosphate group of DMPA that becomes partly ionized at a pH as low as 6.5. The results obtained on the interaction of cardiotoxin with an equimolar mixture of DMPA and DMPC clearly demonstrate the ability of this toxin to induce lateral phase separation in this mixture with one phase containing DMPA-rich domains perturbed by cardiotoxin while the second phase is composed of regions enriched in DMPC. Comparison of the results of the current study with those obtained on other basic proteins and polypeptides suggests that charge-induced phase separation occurs only when the charge density on certain regions of the protein structure is high enough to lead to efficient electrostatic interactions with anionic phospholipids. This condition occurs only when the conformation of the protein or polypeptide is well-ordered at the lipid interface.  相似文献   

2.
The cardiotoxin (cytotoxin II, or CTII) isolated from cobra snake (Naja oxiana) venom is a 60-residue basic membrane-active protein featuring three-finger beta sheet fold. To assess possible modes of CTII/membrane interaction 31P- and 1H-NMR spectroscopy was used to study binding of the toxin and its effect onto multilamellar vesicles (MLV) composed of either zwitterionic or anionic phospholipid, dipalmitoylglycerophosphocholine (Pam2Gro-PCho) or dipalmitoylglycerophosphoglycerol (Pam2Gro-PGro), respectively. The analysis of 1H-NMR linewidths of the toxin and 31P-NMR spectral lineshapes of the phospholipid as a function of temperature, lipid-to-protein ratios, and pH values showed that at least three distinct modes of CTII interaction with membranes exist: (a) nonpenetrating mode; in the gel state of the negatively charged MLV the toxin is bound to the surface electrostatically; the binding to Pam2Gro-PCho membranes was not observed; (b) penetrating mode; hydrophobic interactions develop due to penetration of the toxin into Pam2Gro-PGro membranes in the liquid-crystalline state; it is presumed that in this mode CTII is located at the membrane/water interface deepening the side-chains of hydrophobic residues at the tips of the loops 1-3 down to the boundary between the glycerol and acyl regions of the bilayer; (c) the penetrating mode gives way to isotropic phase, stoichiometrically well-defined CTII/phospholipid complexes at CTII/lipid ratio exceeding a threshold value which was found to depend at physiological pH values upon ionization of the imidazole ring of His31. Biological implications of the observed modes of the toxin-membrane interactions are discussed.  相似文献   

3.
It has recently been shown that cardiotoxin II from Naja mossambica mossambica specifically interacts with negatively charged phospholipids (Dufourcq, J. and Faucon, J.F. (1978) Biochemistry 17, 1170–1176). In order to investigate whether or not short neurotoxins give rise to similar interactions, four techniques have been used, namely intrinsic fluorescence, fluorescence polarization of 1,6-diphenylhexatriene, turbidity measurements and release of 6-carboxyfluorescein trapped inside single shelled vesicles.Neurotoxin III from Naja mossambica mossambica and neurotoxin I from the venom of the scorpion Androctonus australis Hector, specifically interact with negatively charged phospholipids leading to changes in tryptophan fluorescence and to a decrease of the fluidity of the bilayer. Cardiotoxin II from the same snake venom gives similar results. On the other hand, it seems that either a very weak or no interaction at all occurs in the case of neurotoxin I from the same Naja venom.There are important differences in the behaviour of cardiotoxin and neurotoxins: (i) neurotoxins lead to only weak release of 6-carboxyfluorescein from lipid vesicles, whereas cardiotoxin II induces fast and quantitative escape of the dye and then a general breakdown of the vesicular structure; (ii) binding of neurotoxins can be easily reversed by 100–200 mM NaCl or less than 1 mM Ca2+ and so it is essentially electrostatic, whereas binding of cardiotoxin II seems to involve some hydrophobic contribution.The short neurotoxins and cardiotoxins from snake venom having a great homology in sequence, their differences on binding properties are discussed in terms of changes in a particular area of the sequence.  相似文献   

4.
The effect of hydrophobic peptides on the lipid phase behavior of an aqueous dispersion of dioleoylphosphatidylethanolamine and dioleoylphosphatidylglycerol (7:3 molar ratio) was studied by (31)P NMR spectroscopy. The peptides (WALPn peptides, where n is the total number of amino acid residues) are designed as models for transmembrane parts of integral membrane proteins and consist of a hydrophobic sequence of alternating leucines and alanines, of variable length, that is flanked on both ends by tryptophans. The pure lipid dispersion was shown to undergo a lamellar-to-isotropic phase transition at approximately 60 degrees C. Small-angle x-ray scattering showed that at a lower water content a cubic phase belonging to the space group Pn3m is formed, suggesting also that the isotropic phase in the lipid dispersion represents a cubic liquid crystalline phase. It was found that the WALP peptides very efficiently promote formation of nonlamellar phases in this lipid system. At a peptide-to-lipid (P/L) molar ratio of 1:1000, the shortest peptide used, WALP16, lowered the lamellar-to-isotropic phase transition by approximately 15 degrees C. This effect was less for longer peptides. For all of the WALP peptides used, an increase in peptide concentration led to a further lowering of the phase transition temperature. At the highest P/L ratio (1:25) studied, WALP16 induced a reversed hexagonal liquid crystalline (H(II)) phase, while the longer peptides still promoted the formation of an isotropic phase. Peptides with a hydrophobic length larger than the bilayer thickness were found to be unable to inhibit formation of the isotropic phase. The results are discussed in terms of mismatch between the hydrophobic length of the peptide and the hydrophobic thickness of the lipid bilayer and its consequences for lipid-protein interactions in membranes.  相似文献   

5.
The interactions of the antibiotic polymixin B, a polycationic cyclic polypeptide containing a branched acyl side chain, with dimyristoylphosphatidylcholine (DMPC) and dimyristoylphosphatidic acid (DMPA) bilayers were investigated by Raman spectroscopy for a wide range of lipid/polypeptide mole fractions. Temperature profiles, constructed from peak height intensity ratios derived from the lipid methylene C-H stretching and acyl chain C-C stretching mode regions, reflected changes originating from lateral chain packing effects and intrachain trans / gauche rotamer formation, respectively. For DMPC/polymyxin B bilayers the temperature dependent curves indicate a broadening of the gel-liquid crystalline phase transition accompanied by an approx. 3 C deg. increase in the phase transition temperature from 22.8°C for the pure bilayer to 26°C for the polypeptide complex. For a 10:1 lipid/polypeptide mole ratio the temperature profile derived from the C-C mode spectral parameters displays a second order/disorder transition, at approx. 35.5°C, associated with the melting behavior of approximately three bilayer lipids immobilized by the antibiotic's charged cyclic headgroup and hydrophobic side chain. For the 10:1 mole ratio DMPA/polypeptide liposomes, the temperature profiles indicate three order/disorder transitions at 46, 36 and 24°C. Pure DMPA bilayers display a sharp lamellar-micellar phase transition at 51°C.  相似文献   

6.
The interaction of cardiotoxin II of Naja mossambica mossambica with cardiolipin model membranes was investigated by binding, fluorescence, resonance energy transfer, fluorescence quenching, 31P NMR, freeze-fracture, and small-angle X-ray experiments. An initially electrostatic binding appeared to be accompanied by a deep penetration, most likely into the acyl chain region of the phospholipids, indicating a hydrophobic contribution to the strong interaction (KD congruent to 5 X 10(-8) M). This binding results in a fusion of unilamellar vesicles as indicated by a fluorescence-based fusion assay, freeze-fracture, and X-ray diffraction. In these fused structures freeze-fracture electron microscopy reveals the appearance of particles, which is accompanied by the induction of an isotropic component in 31P NMR. The well-defined particles are interpreted as inverted micelles, and the localization of the cardiotoxin molecule in these structures is discussed.  相似文献   

7.
The effect of solubilized hydrophobic peptides on the phase behavior of dioleoylphosphatidylcholine (DOPC)/water system was studied by 2H- and 31P-NMR spectroscopy and by x-ray diffraction, and partial phase diagrams were constructed. The utilized peptides were HCO-AWW(LA)5WWA-NHCH2CH2OH (WALP16), which is an artificial peptide designed to resemble a transmembrane part of a membrane protein; and VEYAGIALFFVAAVLTLWSMLQYLSAAR (Pgs peptide E), a peptide that is identical to one of the putative transmembrane segments of the membrane-associated protein phosphatidylglycerophosphate synthase (Pgs) in Escherichia coli. Circular dichroism spectroscopy suggests that both peptides are mostly alpha-helical in DOPC vesicles. The most striking features in the phase diagram of the WALP16/DOPC/water system are 1) a single lamellar liquid crystalline (L alpha) phase forms only at very low peptide concentrations. 2) At low water content and above a peptide/lipid molar ratio of approximately 1:75 a reversed hexagonal liquid crystalline (H[II]) phase coexists with an L alpha phase, while in excess water this phase forms at a peptide/lipid molar ratio of approximately 1:25. 3) At peptide/lipid ratios > or =1:6 a single H(II) phase is stable. Also, the Pgs peptide E strongly affects the phase behavior, and a single L alpha phase is only found at low peptide concentrations (peptide/lipid molar ratios <1:50), and water concentrations <45% (w/w). Higher peptide content results in coexistence of L alpha and isotropic phases. Generally, the fraction of the isotropic phase increases with increasing temperature and water concentration, and at 80% (w/w) water content only a single isotropic phase is stable at 55 degrees C. Thus, both peptides were found to be able to induce nonlamellar phases, although different in structure, in the DOPC/water system. The phase transitions, the extensions of the one-phase regions, and the phase structures observed for the two systems are discussed in terms of the molecular structure of the two peptides and the matching between the hydrophobic lengths of the peptides and the bilayer thickness of DOPC.  相似文献   

8.
Cardiotoxin isolated from Naja mossambica mossambica selectively deactivates the sodium-potassium activated adenosine triphosphatase of axonal membranes. Tetrodotoxin binding and acetylcholinesterase activities are unaffected by cardiotoxin treatment. The details of association of cardiotoxin with the axonal membrane were studied by following the deactivation of the sodium-potassium activated adenosine triphosphatase and by direct binding measurements with a tritiated derivative of the native cardiotoxin. The maximal binding capacity of the membrane is 42-50 nmol of cardiotoxin/mg of membrane protein. The high amount of binding suggests association of the toxin with the lipid phase of the membrane. It has been shown that cardiotoxin first associates rapidly and reversibly to membrane lipids, then, in a second step, it induces a rearrangement of the membrane structure which produces and irreversible deactivation of the sodium-potassium activated adenosine triphosphatase. Solubilization of the membrane-bound ATPase with Lubrol WX gives an active enzyme species that is resistant to cardiotoxin-induced deactivation. Cardiotoxin binding to the membrane is prevented by high concentrations of Ca 2+ and dibucaine. Although cardiotoxins and neurotoxins of cobra venom have large sequence homologies, their mode of action on membranes is very different. The cardiotoxin seems to bind to the lipid phase of the axonal membrane and inhibits the sodium-potassium activated adenosine triphosphatase, whereas the neurotoxin associates with a protein receptor in the post-synaptic membrane and blocks acetylcholine transmission.  相似文献   

9.
We investigated the interaction between α-lactalbumin and sonicated dimyristoylphosphatidylcholine at pH 4 and different temperatures. (1) At 23°C and lipid-to-protein molar ratios below 170, the interaction results in a disruption of the original vesicles to form smaller complex particles. By the sedimentation velocity method we determined for this particle a molar mass of (1.05 ± 0.16) · 106 g·mol?1. The lipid-to-protein molar ratio within the complex particle is 70/1, as earlier estimated. It follows that there are approximately 1200 lipid and 17 α-lactalbumin molecules per particle. At molar ratios above 170, α-lactalbumin strongly associates with the vesicles. In this case the vesicle entity remains. The ability of α-lactalbumin to break up the vesicles at this temperature is determined by the number of protein molecules which are required in the complex particle. (2) By means of fluorescence polarization of the lipophilic probe 1,6-diphenyl-1,3,5-hexatriene and energy transfer of the tryptophan groups of the protein to 1,3-(1,1′-dipyrenyl)propane located in the hydrocarbon region of the vesicles, it is shown that with increasing temperature above 25°C, complexes of decreasing internal lipid-to-protein molar ratio are formed. However, by electron microscopy we show that the overall size of these complexes remains approximately the same, i.e., bars with dimensions 70 × 220 A?. A temperature-reversible transformation occurs between these complexes, which cannot be isolated by gel chromatography. In contrast, the complex of molar ratio 70/1 remains stable at lower temperatures.  相似文献   

10.
The molecular orientation in a lipid membrane of the peptide fragment VEYAGIALFFVAAVLTLWSMLQYLSAAR (phosphatidylglycerophosphate synthase (Pgs) peptide E) of an integral membrane protein, Pgs, in Escherichia coli has been investigated by solid-state 15N nuclear magnetic resonance (NMR) on macroscopically aligned lipid bilayers. The secondary structure of the peptide in lipid vesicles was determined by circular dichroism spectroscopy. Furthermore, the phase behaviour of the Pgs peptide E/dierucoylphosphatidylcholine (DEruPC)/water system was determined by (2)H, (31)P and 15N solid-state NMR spectroscopy. The phase behaviour obtained was then compared to that of the Pgs peptide E solubilised in dioleoylphosphatidylcholine and water that was previously studied by Morein et al. [Biophys. J. 73 (1997) 3078-3088]. This was aimed to answer the question whether a difference in the length of the hydrophobic part of this peptide and the hydrophobic thickness of the lipid bilayer (hydrophobic mismatch) will affect the phase behaviour. The peptide mostly has a transmembrane orientation and is in an alpha-helical conformation. An isotropic phase is formed in DEruPC with high peptide content (peptide/lipid molar ratio (p/l) > or =1:15) and high water content (> or =50%, w/w) at 35 degrees C. At 55 and 65 degrees C an isotropic phase is induced at high water content (> or =50%, w/w) at all peptide contents studied (no isotropic phase forms in the lipid/water system under the conditions in this study). At high peptide contents (p/l> or =1:15) an isotropic phase forms at 20 and 40% (w/w) of water at 55 and 65 degrees C. A comparison of the phase behaviour of the two homologous lipid systems reveals striking similarities, although the thicknesses of the two lipid bilayers differ by 7 A. This suggests that the rationalisation of the phase behaviour in terms of the hydrophobic mismatch is not applicable to these systems. The C-terminus of Pgs peptide E is amphiphilic and a considerable part of the peptide is situated outside the hydrophobic part of the bilayer, a property of the peptide that to a large extent will affect the lipid/peptide phase behaviour.  相似文献   

11.
The vibrational Raman spectra of both pure 1-alpha-dimyristoylphosphatidic acid (DMPA) liposomes and DMPA multilayers reconstituted with ferricytochrome c at pH 7 and pH 4, with either sodium or calcium as the cation, are reported as a function of temperature. Multilayers composed of a 1:1 mol ratio DMPA and dimyristoylphosphatidylcholine with perdeuterated acyl chains (DMPC-d54) have also been reconstituted with approximately 10(-4) M ferricytochrome c for Raman spectroscopic observation. Total integrated band intensities and relative peak height intensity ratios, two spectral Raman scattering parameters used to characterize bilayer properties, are sensitive to the presence of both ferricytochrome c and the cation in the reconstituted liposomes. Temperature profiles, derived from the various Raman intensity parameters for the 3,100-2,800 cm-1 lipid acyl chain C-H stretching mode region specifically reflect bilayer perturbations due to the interactions of ferricytochrome c. At pH 4 the calcium DMPA multilamellar gel to liquid crystalline phase transition temperatures Tm, defined by either the C-H stretching mode I2850/I2880 and I2935/I2880 peak height intensity ratios, are 58.5 +/- 0.5 degrees C and 60.0 +/- 0.3 degrees C, respectively. This difference in Tm's resolves the phase transition process into first an expansion of the lipid lattice and then a melting of the lipid acyl chains. At pH 7 the calcium DMPA liposomes show no distinct phase transition characteristics below 75 degrees C. For sodium DMPA liposomes reconstituted with ferricytochrome c at either pH 4.0 or pH 7.0, spontaneous Raman spectra show altered lipid structures at temperatures above 40 degrees C. Resonance Raman spectra indicate that ferricytochrome c reconstituted in either calcium or sodium DMPA liposomes changes irreversibly above Tm. For either the binary lipid or ternary lipid-protein systems reconstituted with DMPC-d54, linewidth parameters of the DMPC-d54 acyl chain CD2 symmetric stretching modes at 2,103 cm-1 provide a sensitive measure of the conformational and dynamic properties of the perdeuterated lipid component, while the 3,000 cm-1 C-H spectral region reflects the bilayer characteristics of the DMPA species in the complex. Although calcium clearly induces a lateral phase separation in the DMPA/DMPC-d54 system at pH 7.5 (Kouaouci, R., J.R. Silvius, I. Grah, and M. Pezolet. 1985. Biochemistry. 24:7132-7140), no distinct lateral segregation of the lipid components is observed in the mixed DMPA/DMPC-d54 lipid system in the presence of either ferricytochrome c or the sodium and calcium cations at pH 4.0.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
The influence of melittin on two DMPA membrane systems at pH 4.2 and 8.2 has been investigated by solid-state 31P and 2H NMR, as a function of temperature and peptide concentration. Melittin promotes greater morphological changes for both systems in the fluid phase, the effect being larger at pH 4.2. Close inspection of fatty acyl chain dynamics suggests that some parallels can be drawn between the DMPA/melittin at pH 8.2 and PC/melittin systems. In addition, at pH 8.2 a direct neutralization at the interface of one of the lipid negative charges by a positive charge of the peptide occurs, as can be monitored by 31P NMR at the molecular level. For the system at pH 4.2 and at high temperature, a lipid-to-peptide molar ratio of 30 is sufficient to transform the whole system into an isotropic phase, proposed to be inverted micelles. When the system is cooled down towards the gel phase one observes an intermediate hexagonal phase in a narrow range of temperature.  相似文献   

13.
The interaction of melittin, a polypeptide consisting of 26 amino acid residues, with dimyristoyl phosphatidylcholine bilayers was investigated by vibrational Raman spectroscopy. Spectral peak height intensity ratios, involving vibrational transitions in both the 3000 cm?1 acyl chain methylene carbon-hydrogen stretching mode region and the 1100 cm?1 acyl chain carbon-carbon skeletal stretching mode interval, served as temperature profile indices for monitoring the bilayer order-disorder processes. For a lipid : melittin molar ratio of 14 : 1 two order-disorder transitions were observed. In comparison to a gel to liquid crystalline phase transition of 22.5°C for the pure lipid, the lower transition, exhibiting a 2°C width, is centered at 17°C and is associated with a depression of the main lipid phase transition of dimyristoyl phosphatidylcholine. The second thermal transition, displaying a 7°C interval, occurs at approx. 29°C and is associated with the melting behavior of approximately seven immobilized boundary lipids which surround the inserted hydrophobic segment of the polypeptide. For a lipid : melittin molar ratio of 10 : 1 two thermal transitions are also observed at 11 and 30°C. As before, they represent, respectively, the main gel to liquid crystalline phase transition and the melting behavior of approximately four boundary lipids attached to melittin. From these data alternative schemes are suggested for disposing the immobilized lipids around the hydrophobic portion of the polypeptide within the bilayer.  相似文献   

14.
The interaction of the basic polypeptide poly-L-lysine with the negatively charged phospholipid dipalmitoylphosphatidylglycerol was studied using Raman spectroscopy. The nature of the interaction appeared to depend on the molar ratio of the constituents. At up to one lysine group per lipid molecule, the bilayer was stabilized by the polypeptide that underwent a conformational transition toward an ordered alpha-helical structure, in which the electrostatic interactions were probably maximal. The stabilization of the bilayer was detected by an increase in both the temperature of the thermotropic transition of the lipid and the interchain vibrational coupling of the methylene C-H vibrations. At higher poly-L-lysine concentration, hydrophobic interactions must have been involved to explain the binding of excess polypeptide. There seemed to be a penetration of poly-L-lysine in the bilayer that increased with the polypeptide concentration. Under these conditions, the chain-packing lattice gradually changed from hexagonal to either orthorhombic or monoclinic symmetry. We believe that this change of structure is associated with the interdigitation of the acyl chains.  相似文献   

15.
The interactions of three polypeptide antibiotics (polymyxin B, gramicidin S, and valinomycin) with artificial lecithin membranes were studied by nuclear magnetic resonance (NMR). Combination of 31P and 2H NMR allowed observation of perturbations of the bilayer membrane structure induced by each of the antibiotics in the regions of the polar headgroups and acyl side chains of the phospholipids. The comparative study of the effects of these membrane-active antibiotics and the lipid bilayer structure demonstrated distinct types of antibiotic-membrane interactions in each case. Thus, the results showed the absence of interaction of polymyxin B with the dimyristoyllecithin membranes. In contrast, gramicidin S exhibited strong interaction with the lipid above the gel to liquid-crystalline phase transition temperature: disordering of the acyl side chains was evident. Increasing the concentration of gramicidin S led to disintegration of the bilayer membrane structure. At a molar ratio of 1:16 of gramicidin S to lecithin, the results are consistent with coexistence of gel and liquid-crystalline phases of the phospholipids near the phase transition temperature. Valinomycin decreased the phase transition temperature of the lipids and increased the order parameters of the lipid side chains. Such behavior is consistent with penetration of the valinomycin molecule into the interior of the lipid bilayers.  相似文献   

16.
The cooperative binding process between the antibiotic peptide polymyxin-B and negatively-charged phosphatidic acid bilayers was investigated by differential thermal analysis and completed by fluorescence polarization measurements. The sigmoidal binding curves were analyzed in terms of the interaction energy within a domain formed by polymyxin and phosphatidic acid molecules. The formation of such a heterogeneous domain structure was favoured by high concentration of external monovalent ions. The cooperativity of the binding increased while a charge-induced decrease in the phase transition temperature of the pure lipid phase was observed with increasing ion concentration at a given pH. The reduced lateral coupling within the lipid bilayer in the presence of salt ions, as demonstrated by an increase in the lipid phase transition enthalpy, was considered to facilitate the cooperative domain formation. Moreover, an increase in the cooperativity of the polymyxin binding could be observed if phosphatidic acids of smaller chain length and thus of a lowered phase transition temperature were used. By the use of chemically-modified polymyxin we were able to demonstrate the effect of electrostatic and hydrophobic interaction. Acetylated polymyxin with a reduced positive charge was used to demonstrate the pure hydrophobic effect of polymyxin binding leading to a decrease in the phosphatidic acid phase transition temperature by about 20°C. The cooperativity of the binding was strongly reduced. Cleavage of the hydrophobic polymyxin tail yielded a colistinnonapeptide which caused an electrostatically-induced increase in the phosphatidic acid phase transition temperature. With unmodified polymyxin we observed the combined effects of electrostatic as well as hydrophobic interaction making this model system interesting for the understanding of lipid-protein interactions. Evidence is presented that the formation of the polymyxin-phosphatidic acid complex is a lateral phase separation phenomenon.  相似文献   

17.
The cooperative binding process between the antibiotic peptide polymyxin-B and negatively-charged phosphatidic acid bilayers was investigated by differential thermal analysis and completed by fluorescence polarization measurements. The sigmoidal binding curves were analyzed in terms of the interaction energy within a domain formed by polymyxin and phosphatidic acid molecules. The formation of such a heterogeneous domain structure was favoured by high concentration of external monovalent ions. The cooperativity of the binding increased while a charge-induced decrease in the phase transition temperature of the pure lipid phase was observed with increasing ion concentration at a given pH. The reduced lateral coupling within the lipid bilayer in the presence of salt ions, as demonstrated by an increase in the lipid phase transition enthalpy, was considered to facilitate the cooperative domain formation. Moreover, an increase in the cooperativity of the polymyxin binding could be observed if phosphatidic acids of smaller chain length and thus of a lowered phase transition temperature were used. By the use of chemically-modified polymyxin we were able to demonstrate the effect of electrostatic and hydrophobic interaction. Acetylated polymyxin with a reduced positive charge was used to demonstrate the pure hydrophobic effect of polymyxin binding leading to a decrease in the phosphatidic acid phase transition temperature by about 20 degrees C. The cooperativity of the binding was strongly reduced. Cleavage of the hydrophobic polymyxin tail yielded a colistinnonapeptide which caused an electrostatically-induced increase in the phosphatidic acid phase transition temperature. With unmodified polymyxin we observed the combined effects of electrostatic as well as hydrophobic interaction making this model system interesting for the understanding of lipid-protein interactions. Evidence is presented that the formation of the polymyxin-phosphatidic acid complex is a lateral phase separation phenomenon.  相似文献   

18.
Hydrophobic matching, in which transmembrane proteins cause the surrounding lipid bilayer to adjust its hydrocarbon thickness to match the length of the hydrophobic surface of the protein, is a commonly accepted idea in membrane biophysics. To test this idea, gramicidin (gD) was embedded in 1, 2-dilauroyl-sn-glycero-3-phosphocholine (DLPC) and 1, 2-myristoyl-sn-glycero-3-phosphocholine (DMPC) bilayers at the peptide/lipid molar ratio of 1:10. Circular dichroism (CD) was measured to ensure that the gramicidin was in the beta6.3 helix form. The bilayer thickness (the phosphate-to-phosphate distance, or PtP) was measured by x-ray lamellar diffraction. In the Lalpha phase near full hydration, PtP is 30.8 A for pure DLPC, 32.1 A for the DLPC/gD mixture, 35.3 A for pure DMPC, and 32.7 A for the DMPC/gD mixture. Gramicidin apparently stretches DLPC and thins DMPC toward a common thickness as expected by hydrophobic matching. Concurrently, gramicidin-gramicidin correlations were measured by x-ray in-plane scattering. In the fluid phase, the gramicidin-gramicidin nearest-neighbor separation is 26.8 A in DLPC, but shortens to 23.3 A in DMPC. These experiments confirm the conjecture that when proteins are embedded in a membrane, hydrophobic matching creates a strain field in the lipid bilayer that in turn gives rise to a membrane-mediated attractive potential between proteins.  相似文献   

19.
The interaction of three vitamin A derivatives or retinoids: all-trans-retinoic acid, 13-cis-retinoic acid and retinol with multilamellar phospholipid bilayers was studied using a combination of 2H- and 31P-NMR measurements. The following model membrane systems were used: (1) dipalmitoylphosphatidylcholine (DPPC) bilayers; (2) bilayers composed of a mixture of DPPC and bovine heart phosphatidylcholine (PC); (3) mixed PC/phosphatidylethanolamine (PE) bilayers. Only a weak interaction was observed between 13-cis-retinoic acid and DPPC membranes. Addition of all-trans-retinoic acid at a molar ratio of 1:2 to the lipid causes a small decrease (5 C degrees) in the gel to liquid crystalline phase-transition temperature of DPPC, a small increase in the order parameters of the lipid side-chains of single component bilayers and no measurable effect in the other lipid systems studied. Considerably larger perturbation in the lipid bilayer structure is introduced by addition of retinol which, at a molar ratio of 1:2 to the lipid, lowered the gel to liquid crystalline phase-transition temperature of DPPC by 21 C degrees and caused a decrease of order parameters of the lipid side-chains in all three lipid bilayer systems. These effects are consistent with intercalation of retinol molecules into the bilayer interior. The results for the mixed PC/PE bilayers indicate that the presence of retinol caused lateral separation of PE- and retinol-enriched regions.  相似文献   

20.
The interaction of phosphatidylcholine bilayers with Triton X-100   总被引:1,自引:0,他引:1  
The interaction of multilamellar phosphatidylcholine vesicles with the non-ionic detergent Triton X-100 has been studied under equilibrium conditions, specially in the sub-lytic range of surfactant concentrations. Equilibrium was achieved in less than 24 h. Estimations of detergent binding to bilayers, using [3H]Triton X-100, indicate that the amphiphile is incorporated even at very low concentrations (below its critical micellar concentration); a dramatic increase in the amount of bound Triton X-100 occurs at detergent concentrations just below those producing membrane solubilization. Solubilization occurs at phospholipid/detergent molar ratios near 0.65 irrespective of lipid concentration. The perturbation produced by the surfactant in the phospholipid bilayer has been studied by differential scanning calorimetry, NMR and Fourier-transform infrared spectroscopy. At low detergent concentration (lipid/detergent molar ratios above 3), a reduction in 2H-NMR quadrupolar splitting occurs, suggesting a decrease in the static order of the acyl chains; the same effect is detected by Fourier-transform infrared spectroscopy in the form of blue shifts of the methylene stretching vibration bands. Simultaneously, the enthalpy variation of the main phospholipid phase transition is decreased by about a third with respect to its value in the pure lipid/water system. For phospholipid/detergent molar ratios between 3 and 1, the decrease in lipid static order does not proceed any further; rather an increase in fluidity is observed, characterized by a marked decrease in the midpoint transition temperature of the gel-to-fluid phospholipid transition. At the same time an isotropic component is apparent in both 31P-NMR and 2H-NMR spectra, and a new low-temperature endotherm is detected in differential scanning calorimetric traces. When phospholipid and Triton X-100 are present at equimolar ratios some bilayer structure persists, as judged from calorimetric observations, but NMR reveals only one-component isotropic signals. At lipid/detergent molar ratios below unity, the NMR lines become narrower, the main (lamellar) calorimetric endotherm tends to vanish and solubilization occurs.  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