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 484 毫秒
1.
The pH-dependence of the phase transition of dimyristoyl phosphatidic acid and dihexadecyl phosphatidic acid has been investigated using differential scanning calorimetry. Varying the pH induces different degrees of ionization of the polar head group. The changes in transition temperature with pH as observed by calorimetry are in good agreement with those obtained by measuring the changes in light scattering, whereas the transition temperatures reported by the fluorescent probe N-phenylnaphthylamine do not always coincide with those determined from calorimetry [1]. The observed maximum of the transition temperature at pH 3.5 corresponds to a minimum in the transition enthalpy vs. pH diagram. At this pH a particular stable bilayer phase is formed. Full protonation of phosphatidic acids leads to suspensions of mycrocrystals. The transition enthalpy approaches the value of the melting enthalpy of crystalline anhydrous phosphatidic acid. The decrease in the transition enthalpy at high pH values is due to a change in the hydrocarbon chain interactions induced by the doubly charged head groups. The cooperativity of the transition varies with the degree of ionization of the head group, being lower for doubly charged phosphatidic acids.  相似文献   

2.
A complete titration of phosphatidic acid bilayer membranes was possible for the first time by the introduction of a new anaologue, 1,2-dihexadecyl-sn-glycerol-3-phosphoric acid, which has the advantage of a high chemical stability at extreme pH values. The synthesis of the phosphatidic acid is described and the phase transition behaviour in aqueous dispersions is compared with that of three ester phosphatidic acids; 1,2-dimyristoyl-sn-glycerol-3-phosphoric acid, 1,3-dimyristoylglycerol-2-phosphoric acid and 1,2-dipalmitoyl-sn-glycerol-3-phosphoric acid. The phase transition temperatures (Tt) of aqueous phosphatidic acid dispersions at different degrees of dissociation were measured using fluorescence spectroscopy and 90 degrees light scattering. The Tt values are comparable to the melting points of the solid phosphatidic acids in the fully protonated states, but large differences exist for the charged states. The Tt vs. pH diagrams of the four phosphatidic acids are quite similar and of a characteristic shape. Increasing ionisation results in a maximum value for the transition temperatures at pH 3.5 (pK1). The regions between the first and the second pK of the phosphatidic acids are characterised by only small variations in the transition temperatures (extended plateau) in spite of the large changes occurring in the surface charge of the membranes. The slope of the plateau is very shallow with increasing ionisation. A further decrease in the H+ concentration results in an abrupt change of the transition temperature. The slope of the Tt vs. pH diagram beyond pK2 becomes very steep. This is the result of reduced hydrocarbon interaction energy, which was demonstrated by differential scanning calorimetry (Blume, A. and Eibl, H., unpublished data).  相似文献   

3.
A complete titration of phosphatidic acid bilayer membranes was possible for the first time by the introduction of a new anaologue, 1,2-dihexadecyl-sn-glycerol-3-phosphoric acid, which has the advantage of a high chemical stability at extreme pH values. The synthesis of this phosphatidic acid is described and the phase transition behaviour in aqueous dispersions is compared with that of three ester phosphatidic acids; 1,2-dimyristoyl-sn-glycerol-3-phosphoric acid, 1,3-dimyristoylglycerol-2-phosphoric acid and 1,2-dipalmitoyl-sn-glycerol-3-phosphoric acid.The phase transition temperatures (Tt) of aqueous phosphatidic acid dispersions at different degrees of dissociation were measured using fluorescence spectroscopy and 90° light scattering. The Tt values are comparable to the melting points of the solid phosphatidic acids in the fully protonated states, but large differences exist for the charged states.The Tt vs. pH diagrams of the four phosphatidic acids are quite similar and of a characteristic shape. Increasing ionisation results in a maximum value for the transition temperatures at pH 3.5 (pK1). The regions between the first and the second pK of the phosphatidic acids are characterised by only small variations in the transition temperatures (extended plateau) in spite of the large changes occurring in the surface charge of the membranes. The slope of the plateau is very shallow with increasing ionisation. A further decrease in the H+ concentration results in an abrupt change of the transition temperature. The slope of the Tt vs. pH diagram beyond pK2 becomes very steep. This is the  相似文献   

4.
Fourier transform infrared spectroscopy is used to characterize specific interactions between negatively charged lipids, such as phosphatidic acid, and the purified nicotinic acetylcholine receptor from Torpedo californica. The specific interaction of phosphatidic acid with acetylcholine receptor is demonstrated by the receptor-induced perturbation of the lipid ionization state, which is monitored using Fourier transform infrared bands arising from the phosphate head group. The acetylcholine receptor shifts the pKa of phosphatidic acid molecules adjacent to the receptor to a lower value by almost 2 pH units from 8.5 to 6.6. Decreased pH also leads to changes in ion channel function and to changes in the secondary structure of the acetylcholine receptor in membranes containing ionizable phospholipids. Phospholipase D restores functional activity of acetylcholine receptor reconstituted in an unfavorable environment containing phosphatidylcholine by generating phosphatidic acid. Lipids such as phosphatidic acid may serve as allosteric effectors for membrane protein function and the lipid-protein interface could be a site for activity-dependent changes that lead to modulation of synaptic efficacy.  相似文献   

5.
The monolayer and thermal behaviour of different phosphatidic acids are presented. At neutral pH and 22°C dilauroylphosphatidic acid and unsaturated phosphatidic acids form liquid-expanded monolayers, while dipalmitoyl- and distearoylphosphatidic acid form condensed monolayers. Dimyristoylphosphatidic acid undergoes a transition from the liquid-expanded to the condensed state. With long-chain saturated and unsaturated phosphatidic acids little change in molecular area is observed between pH 2 and 7. In contrast, the short chain saturated phosphatidic acids, dilauroyl- and dimyristoylphosphatidic acids, undergo a condensation in the pH range 2 to 7. This is so in spite of the fact that the phosphoric acid group dissociates and the phosphatidic acid molecule attains one negative charge over this pH range. This finding is interpreted to indicate that the electrostatic repulsion between negatively charged phosphatidic acid molecules is compensated for or even outweighed by other intermolecular forces. Hydrogen bonding at the lipid/water interface is supposed to play a major role. All phosphatidates studied exhibit a significant expansion in the pH range 7 to 12. The second apparent pK of the primary phosphate group of phosphatidic acids is 8.6 and the expansion observed in this pH range is therefore due to electrostatic repulsion. At neutral pH the ether analogues of saturated phosphatidic acids have monolayer properties similar to those of the ester compounds. Considering the total pH range of 2 to 12 studied the force-area curves of the ether analogues are more condensed compared to the ester compounds. Synthetic phosphatidates and their ether analogues give reversible sharp crystal(gel)-to-liquid crystal transitions while the naturally occurring egg phosphatidate gives a broad, asymmetric one. The transition temperature Tm of saturated phosphatidates increases with increasing hydrocarbon chain length and at a given chain length Tm decreases markedly with unsaturation. The Tm values of the ether analogues are about 10°C higher and the ΔH values are 10–15% lower than those of the corresponding esters.  相似文献   

6.
(1) The thermotropic behaviour of dimyristoyl phosphatidylglycerol, phosphatidylserine, phosphatidic acid and phosphatidylcholine was investigated by differential scanning calorimetry and freeze-fracture electron microscopy as a function of pH and of Ca2+ concentration. (2) From the thermotropic behaviour as a function of pH, profiles could be constructed from which apparent pK values of the charged groups of the lipids could be determined. (3) Excess Ca2+ induced a shift of the total phase transition in 14 : 0/14 : 0-glycerophosphocholine and 14 : 0/14 : 0-glycerophosphoglycerol mixtures. In 14 : 0/14 : 0-glycerophosphocholine bilayers containing 16 : 0/16 : 0-glycerophosphoglycerol lateral phase separation was induced by Ca2+. (4) Up to molar ratios of 1 : 2 of 14 : 0/14 : 0-glycerophosphoserine to 14 : 0/14: 0-glycerophosphocholine, excess Ca2+ induced lateral phase separation. Addition to mixtures of higher molar ratios caused segregation into different structures: the liposome organization and the stacked lamellae/cylindrical organization. (5) Addition of excess Ca2+ to mixtures of 14 : 0/14 : 0-glycerophosphocholine and 14 : 0/14 : 0-phosphatidic acid caused, independent of the molar ratio, separation into two structural different organizations. (6) The nature of Ca2+-induced changes in bilayers containing negatively charged phospholipids is strongly dependent on the character of the polar headgroup of the negatively charged phospholipid involved.  相似文献   

7.
Phospholipids interact on Triton X-100 solubilized GDP-fucose: asialofetuin fucosyltransferase (EC 2.4.1.68) isolated from sheep brain. This enzymatic activity is modulated by charged phospholipids. In particular, phosphatidic acid and analogues markedly inhibit the transfer of fucose from GDP-[14C]fucose. Kinetic studies show that phosphatidic acid interacts as a mixed inhibitor: the velocity and affinity of fucosyltransferase for the GDP-fucose and asialofetuin substrates are strongly decreased. However, this inhibitory effect is not related to stereospecificity, and the different parameters involved in the enzymatic reaction of glycosylation are not modified. The nature of fatty acids and chemical bond (ester or ether) occurring in the carbohydrate chain does not modify the behaviour of phosphatidic acid with respect to fucosyltransferase activity. Further, the physical state of phosphatidic acid (gel phase or liquid crystalline phase) has no influence. However, as the inhibition is closely pH-dependent, these data suggest that phosphatidic acid might directly interact with the active site of the enzyme and induce a conformational change.  相似文献   

8.
The effect of salt and pH titration on the selectivity of spin-labeled analogues of phosphatidic acid, phosphatidylserine, phosphatidylcholine, and stearic acid for the nicotinic acetylcholine receptor (nAcChoR) reconstituted into dioleoylphosphatidylcholine was examined at 0 degrees C using electron spin resonance spectroscopy. The order of selectivity at pH 7.4 and 0 mM NaCl was phosphatidylserine > stearic acid > phosphatidic acid > phosphatidylcholine. The addition up to 2 M NaCl or titration of pH from 5.0 to > 9.0 did not alter the selectivity of the phospholipids for the nAcChoR. For stearic acid, conversely, titration of pH from 5.0 to 9.0 at 0 mM NaCl and titration of NaCl from 0 to 2 M at pH 9.0 both increased selectivity for the nAcChoR. It is concluded that electrostatic interactions do not account for the selectivity of the negatively charged phospholipids, phosphatidylserine, and phosphatidic acid for the nAcChoR. This is consistent with the known orientation of the transmembrane sequences M1 and M4, which predicts a balance in the number of negative and positive charges in the lipid-protein interface and suggests that the two positive charges on each M3 helix are not exposed to the lipid-protein interface.  相似文献   

9.
Differential scanning calorimetry (DSC) and fluorescence polarization of embedded probe molecules were used to detect phase behavior of various phospholipids. The techniques were directly compared for detecting the transition of dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidic acid (DPPA) dispersed in aqueous salt solutions. Excellent agreement occurred in the case of phosphatidylcholine; however, in the case of phosphatidic acid, at pH 6.5, transitions detected by fluorescence polarization using the disc-like perylene molecule occurred about 10 degrees lower than those detected by DSC. Discrepancy between fluorescence and DSC methods is eliminated by using a rod-like molecule, diphenylhexatriene (DPH). Both techniques show that doubly ionizing the phosphate group reduces the Tc by about 9 degrees. Direct pH titration of fluidity can be accomplished and this effect is most dramatic when membranes are in their transition temperature range (ca. 50 degrees). Phosphatidic acid transitions occur at higher temperatures, and have appreciably lower transition enthalpies and entropies than phosphatidylcholine. These effect could not be explained simply on the basis of double layer electrostatics and several other factors were discussed in an attempt to rationalize the results. Addition of monovalent cations (0.01-0.5 M) is shown to increase the Tc of dipalmitoylphosphatidylglycerol by less than 3 degrees. However, addition of (1 x 10-3 M) Ca2+ abolishes the phase transition of both phosphatidyglycerol and phosphatidylserine in the range 0-70 degrees. Preliminary X-ray evidence indicates the phosphatidylserine-Ca2+ bilayers are in a crystalline state at 24 degrees. In contrast, 5 x 10-3 M Mg2+ only broadens the transition and increases the Tc indicating a considerable difference between the effects of Ca2+ and Mg2+. Neutralization of PS increases the Tc from 6 degrees (at pH 7.4) to 20-26 degrees (at pH 2.5-3.0) but does not abolish the transition, suggesting the Ca2+ effect involves more than charge neutralization. Addition of Ca2+ to mixed phosphatidylserine-phosphatidylcholine dispersions, induces a phase separation of the dipalmitoyl- (and also distearoyl-) phosphatidylcholine as seen by the appearance of a new endothermic peak at 41 degrees (58 degrees). Similarly, in mixed (dipalmitoyl) phosphatidic acid-phosphatidylcholine (2:1) dispersions, Ca2+ again can separate the phosphatidylcholine component.  相似文献   

10.
The influence of pH on the structure of 1,2-(ditetradecyl)-phosphatidic acid was investigated by differential scanning calorimetry and freeze-fracture electron microscopy. At pH 13.5--14 (2.6 M K+), where phosphatidic acid has two negative charges, calorimetric scans show a small transition (pretransition) below the main phase transition temperature. Freeze-fracture studies of the same dispersions reveal regular band patterns (so-called ripples) in the plane of the bilayers, when the lipid is quenched from below the main phase transition temperature. This rippled structure is similar to the well-known rippled structure of phosphatidylcholines.  相似文献   

11.
The influence of pH on the structure of 1,2-(ditetradecyl)-phosphatidic acid was investigated by differential scanning calorimetry and freeze-fracture electron microscopy. At pH 13.5–14 (2.6 M K+), where phosphatidic acid has two negative charges, calorimetric scans show a small transition (pretransition) below the main phase transition temperature. Freeze-fracture studies of the same dispersions reveal regular band patterns (so-called ripples) in the plane of the bilayers, when the lipid is quenched from below the main phase transition temperature. This rippled structure is similar to the well-known rippled structure of phosphatidylcholines.  相似文献   

12.
Monolayers of phospholipids spread at the air/water interface were studied by means of fluorescence measurements. Using linearly polarised incident light and following the behaviour of a fluorescent covalently-labelled phospholipid (dansylphosphatidylethanolamine) embedded in the monolayer, it was possible to obtain information about the orientation changes at the glycerol level of the phospholipid.When using dipalmitoylphosphatidylcholine as phospholipid, the main orientation of the probe appears unchanged during the phase transition process. On the other hand, the standard deviation of the distribution function of orientations is larger in the liquid-expanded state relative to the condensed state.When using phosphatidic acid or a mixture of dipalmitoylphosphatidylcholine and phosphatidic acid as phospholipids, the same orientation of the probe is observed as with pure dipalmitoylphosphatidylcholine. This orientation stays the same whatever the ionic content of the subphase, i.e. the ionisation state of the polar head.The glycerol region of a phospholipid spread in monolayer at the air/water interface appears unaffected structurally either by the nature of the polar moiety, by its ionisation state, or by the physical state of the hydrocarbon chains.  相似文献   

13.
The C1 domains of novel PKCs mediate the diacylglycerol-dependent translocation of these enzymes. The four different C1B domains of novel PKCs (δ, ε, θ and η) were studied, together with different lipid mixtures containing acidic phospholipids and diacylglycerol or phorbol ester. The results show that either in the presence or in the absence of diacylglycerol, C1Bε and C1Bη exhibit a substantially higher propensity to bind to vesicles containing negatively charged phospholipids than C1Bδ and C1Bθ. The observed differences between the C1B domains of novel PKCs (in two groups of two each) were also evident in RBL-2H3 cells and it was found that, as with model membranes, in which C1Bε and C1Bη could be translocated to membranes by the addition of a soluble phosphatidic acid without diacylglycerol or phorbol ester, C1Bδ and C1Bθ were not translocated when soluble phosphatidic acid was added, and diacylglycerol was required to achieve a detectable binding to cell membranes. It is concluded that two different subfamilies of novel PKCs can be established with respect to their propensity to bind to the cell membrane and that these peculiarities in recognizing lipids may explain why these isoenzymes are specialized in responding to different triggering signals and bind to different cell membranes.  相似文献   

14.
The interactions of salmon cacitonin with a number of phospholipids are studied by electron microscopy, circular dichroism and the leakage of carboxyfluorescein. At room temperature, calcitonin reacts strongly with dimyristoylphosphatidylglycerol and egg phosphatidic acid, while only moderate or no interaction is observed with several other phospholipids. The interaction is judged by the dissolution of the phospholipid dispersion and by electron microscopic observation and is in general concomitant with an increase in the helical content of the peptide. The electrostatic charge and the transition temperature of each of the phospholipids are important factors in determining the extent of reaction with salmon calcitonin. An exception is the sulphatide from bovine brain. The resulting morphology of the complex formed between salmon calcitonin and phosphatidic acid is quite different from that formed with phosphatidylglycerol. In the case of phosphatidylglycerol and most other negatively charged phospholipids, disc-shaped complexes are observed under the electron microscope by negative staining. The calcitonin-DMPG complexes are about 7 nm thick and their diameter increases with an increasing lipid-to-peptide ratio. In contrast, phosphatidic acids form spherical complexes with salmon calcitonin causing large multilamellar structures to spontaneously break-up into smaller particles of about 10 to 20 nm in diameter independent of the lipid-to-peptide ratio. The contrasting effects of salmon calcitonin on the morphology of these two phospholipids is explicable by consideration of the size of the lipid headgroup. Phosphatidic acid can accommodate the peptide without rupture of the bilayer, while the larger headgroup of phosphatidylglycerol requires the bilayer to rupture. This model is supported by studies of calcitonin-induced leakage of carboxyfluorescein from sonicated vesicles of 75% egg phosphatidylcholine and 25% either egg phosphatidic acid, egg phosphatidylglycerol or dimyristoylphosphatidylglycerol. There was a much greater increase in carboxyfluorescein leakage from phosphatidylglycerol-containing vesicles induced by salmon calcitonin demonstrating the greater ability of the peptide to rupture bilayers containing this phospholipid.  相似文献   

15.
The interactions of salmon calcitonin with a number of phospholipids are studied by electron microscopy, circular dichroism and the leakage of carboxyfluorescein. At room temperature, calcitonin reacts strongly with dimyristoylphosphatidylglycerol and egg phosphatidic acid, while only moderate or no interaction is observed with several other phospholipids. The interaction is judged by the dissolution of the phospholipid dispersion and by electron microscopic observation and is in general concomitant with an increase in the helical content of the peptide. The electrostatic charge and the transition temperature of each of the phospholipids are important factors in determining the extent of reaction with salmon calcitonin. An exception is the sulphatide from bovine brain. The resulting morphology of the complex formed between salmon calcitonin and phosphatidic acid is quite different from that formed with phosphatidylglycerol. In the case of phosphatidylglycerol and most other negatively charged phospholipids, disc-shaped complexes are observed under the electron microscope by negative staining. The calcitonin- DMPG complexes are about 7 nm thick and their diameter increases with an increasing lipid-to-peptide ratio. In contrast, phosphatidic acids form spherical complexes with salmon calcitonin causing large multilamellar structures to spontaneously break-up into smaller particles of about 10 to 20 nm in diameter independent of the lipid-to-peptide ratio. The contrasting effects of salmon calcitonin on the morphology of these two phospholipids is explicable by consideration of the size of the lipid headgroup. Phosphatidic acid can accommodate the peptide without rupture of the bilayer, while the larger headgroup of phosphatidylglycerol requires the bilayer to rupture. This model is supported by studies of calcitonin-induced leakage of carboxyfluorescein from sonicated vesicles of 75% egg phosphatidylcholine and 25% either egg phosphatidic acid, egg phosphatidylglycerol or dimyristoylphosphatidylglycerol . There was a much greater increase in carboxyfluorescein leakage from phosphatidylglycerol-containing vesicles induced by salmon calcitonin demonstrating the greater ability of the peptide to rupture bilayers containing this phospholipid.  相似文献   

16.
The interaction between a positively charged peptide (poly-L-lysine) and model membranes containing charged lipids has been investigated. Conformational changes of the polypeptide as well as changes in the membrane lipid distribution were observed upon lipid-protein agglutination: 1. The strong binding of polylysine is shown directly by the use of spinlabelled polypeptide. Upon binding to phosphatidic acid a shift in the hyperfine coupling constant from 16.5 to 14.6 Oe is observed. The spectrum of the lipid-bound peptide is superimposed on the spectrum of polylysine in solution. Half of the lysine groups are bound to the charged membranes. A change in the conformation of polylysine from a random coil to a partially ordered configuration is suggested. 2. Spin labelling of the lipid component gives evidence concerning the molecular organization of a lipid mixture containing charged phosphatitid acid. Addition of polylysine induces the formation of crystalline patches of bound phosphatidic acid. 3. Excimer forming pyrene decanoic acid has been employed. Addition of positively charged polylysine (pH 9.0) to phosphatidic acid membranes increases the transition temperature of the lipid from Tt = 50 to Tt = 62 degrees C. Thus, a lipid segregation of lipid into regions of phosphatidic acid bound to the peptide which differ in their microviscosity from the surrounding membrane is induced. One lysine group binds one phosphatidic acid molecule, but only half of the phosphatidic acid is bound. 4. Direct evidence for charge induced domain formation in lipid mixtures containing phosphatidic acid is given by electron microscopy. Addition of polylysine leads to a change in the surface curvature of the bound charged lipid. The domain size is estimated from the electron micrographs. The number of domains present is dependent on both the ratio of charged to uncharged lipids as well as on the amount of polylysine added to the vesicles. The size of the domains is not dependent on membrane composition. However, the size seems to increase in a stepwise manner that is correlated with a multiple of the area covered by one polylysine molecule.  相似文献   

17.
Marsh D 《Biophysical journal》2012,102(5):1079-1087
Negatively charged phospholipids are an important component of biological membranes. The thermodynamic parameters governing self-assembly of anionic phospholipids are deduced here from isothermal titration calorimetry. Heats of demicellization were determined for dioctanoyl phosphatidylglycerol (PG) and phosphatidylserine (PS) at different ionic strengths, and for dioctanoyl phosphatidic acid at different pH values. The large heat capacity (ΔCoP ∼ −400 J.mol−1 K−1 for PG and PS), and zero enthalpy at a characteristic temperature near the physiological range (T ∼ 300 K for PG and PS), demonstrate that the driving force for self-assembly is the hydrophobic effect. The pH and ionic-strength dependences indicate that the principal electrostatic contribution to self-assembly comes from the entropy associated with the electrostatic double layer, in agreement with theoretical predictions. These measurements help define the thermodynamic effects of anionic lipids on biomembrane stability.  相似文献   

18.
The finding that negatively charged phospholipids activate the plasma-membrane (Ca2+ + Mg2+)-ATPase and that polycations counteract this stimulation suggest that negative charges in the environment of the ATPase protein could be important for its function. The aim of the present work was to investigate whether changing the charges on the ATPase protein itself by modifying the pH within the physiological range affects the activity of the purified plasma-membrane Ca2+ pump from stomach smooth muscle. Increasing the pH from 6.9 to 7.4 and using 1,2-bis(o-aminophenoxy)ethane-N,N,N',N'-tetra-acetic acid (BAPTA) as a Ca2+ buffer, doubled the ATPase activity at 0.3 microM-Ca2+ in the presence of 100% phosphatidylcholine (PC) or after substituting 20% of the PC by negatively charged phospholipids PtdIns, PtdIns4P, phosphatidylserine and phosphatidic acid. This stimulatory effect was due to an increased affinity of the enzyme for Ca2+, while the Vmax. remained unaffected. In the case of PtdIns(4,5)P2, a stimulatory effect upon alkalinization was only observed at a PtdIns(4,5)P2 concentration of 10%. When a concentration of 20% was used, alkalinization decreased the Vmax. and no stimulatory effect on the ATPase at 0.3 microM-Ca2+ could be observed. Alkalinization not only stimulated the purified Ca2+ pump, but it also increased the activity of the enzyme in a plasma-membrane-enriched fraction from stomach smooth muscle by a factor of 2.06. The ionophore A23187-induced Ca2+ uptake in closed inside-out vesicles also increased by a factor of 2.54 if the pH was changed from 6.9 to 7.4. This finding indicates that the effect of pH is most likely to be exerted at the cytoplasmic site of the Ca2+ pump protein.  相似文献   

19.
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   

20.
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   

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