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 734 毫秒
1.
Ion binding and lipid ionization of the acidic phospholipid 1,2-dimyristoyl-sn-glycero-3-phosphoglycerol (DMPG) in monolayers was studied by measuring the lateral pressure Pi as a function of the molecular area A at the air/water interface at different temperatures. The pH of the subphase (pH 2 and 7) and the ionic strength (NaCl) was varied. In addition, different divalent cations (1mM MgCl2, CaCl2 and SrCl2, pH 7) were added. DMPG is partly protonated on pure water at pH 7. An increase in the NaCl concentration in the subphase leads to film expansion. This effect is caused by an ionization of the headgroup of DMPG, i.e. a shift of the apparent pK. More condensed films are obtained on pure water at pH 2, due to the reduction of electrostatic repulsion by headgroup protonation and the possibility for the formation of a hydrogen bonding network. The divalent cations Mg2+, Ca2+ and Sr2+ interact differently with a DMPG monolayer in pure water at pH 7. In the presence of 1mM CaCl2 a condensation of the DMPG film is induced, whereas an expansion of the monolayer is observed in the presence of Mg2+ and Sr2+. Two counteracting effects are operative: (a) ionization of the headgroup due to electrostatic screening leads to film expansion and (b) binding of the divalent cations to the lipid headgroups leads to condensation. The latter effect is more pronounced in the case of Ca2+, whereas the binding of Mg2+ and Sr2+ to DMPG is weaker. Site-specific cation binding has to be assumed in addition to electrostatic effects.  相似文献   

2.
1,2-Didodecanoyl-sn-glycero-3-phosphoryl-1'-(3'-O-L-alanyl)-sn-glycerol (Ala-PG) has been synthesized. Its ionic properties have been studied at the air-water interface through film compressions and surface potential measurements as a function of subphase pH and ionic content (NaCl, Na2MoO4, CaCl2). The existence of the polar head in a loop conformation allowing for interactions between phosphate and amino groups is suggested. Ionic properties of Ala-PG clearly depended on subphase ionic strength but no specific interactions between either cations or anions in the subphase and phosphate or amino groups in the film could be detected. Results are interpreted in terms of ion-pair interactions at the interface between these two groups and anions and cations from the subphase. Occurrence of charge separation between these two groups, induced by increasing subphase ionic strength, is postulated. Since the molecular packing appeared independent of the subphase ionic content over a large domain of pH (3–8) and surface pressure (π > 5 dyne/cm) and since the lipid can be considered as zwitterionic or slightly positive below pH 5–6, it is suggested that in the parent bacteria, grown under acidic conditions, Ala-PG could play a role in maintaining the membrane intergrity and in preventing the passive diffusion of protons.  相似文献   

3.
The interaction of the low molecular weight group of surfactant-associated proteins, SP 5-18, with the major phospholipids of pulmonary surfactant was studied by fluorescence measurements of liposomal permeability and fusion, morphological studies, and surface activity measurements. The ability of SP 5-18 to increase the permeability of large unilamellar lipid vesicles was enhanced by the presence of negatively charged phospholipid. The permeability of these vesicles increased as the protein concentration was raised and the pH was lowered. SP 5-18 also induced leakage from liposomes made both from a synthetic surfactant lipid mixture and from lipids separated from SP 5-18 during its purification from canine sources. When SP 5-18 was added to egg phosphatidylglycerol liposomes, the population of liposomes which became permeable leaked all encapsulated contents, while the remaining liposomes did not leak at all. The extent of leakage was higher in the presence of 3 mM calcium. SP 5-18 also induced lipid mixing between two populations of egg phosphatidylglycerol liposomes in the presence of 3 mM calcium, as monitored by resonance energy transfer between two different fluorescent lipid probes, N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine and N-(lissamine rhodamine B sulfonyl)phosphatidylethanolamine. Negative-staining electron microscopy showed that the addition of SP 5-18 and 3 mM calcium produced vesicles twice the size of control egg phosphatidylglycerol liposomes. In addition, surface balance measurements revealed that the adsorption of liposomal lipids to an air/water interface was enhanced by the presence of SP 5-18, negatively charged phospholipids, and 3 mM calcium. These observations suggest a similar lipid dependence for the interactions observed in the fluorescence and adsorption experiments.  相似文献   

4.
The relationship between area per molecule and surface pressure of monolayers of phosphatidylglycerol phosphate from extreme halophile Halobacterium cutrirubrum and its deoxy analogue, deoxyphosphatidylglycerol phosphate, spread at an air/water interface was examined. The effect of ionization of the primary and secondary acidic functions of the phosphate groups of the two lipids on surface characteristics of compression isotherms was determined by spreading monolayers on subphases with pH values ranging from below the apparent pKa of the primary ionization (pH 0) to greater than that of secondary ionization (pH 10.9). The limiting molecular area increases with decreasing pH below 2. Ionization of the primary phosphate functions of both phospholipids (with bulk pK1 values close to 4) is associated with a marked expansion of the films, as judged by values of limiting molecular area. Ionization of the secondary phosphate functions causes further expansion of the films, with the apparent pK2 of deoxyphosphatidylglycerol phosphate slightly less than that indicated for phosphatidylglycerol phosphate. Values of surface-compressibility modulus calculated from the surface characteristics of the phosphatidylglcerol phosphate monolayers showed that films spread on subphases with a pH of about the apparent pK1 of the primary phosphate functions were the least compressible. Increasing or decreasing subphase pH caused an increase in compressibility; this effect on compressibility was much less with monolayers of deoxyphosphatidylglycerol phosphate at high pH. The effect of inorganic counter-ions on monolayer characteristics of phosphatidylglycerol phosphate was examined by using subphases of NaCl concentrations varying from 0.01 to 1 M. The limiting molecular area was found to increase exponentially with respect to the subphase NaCl concentration.  相似文献   

5.
Previous studies have demonstrated that lung-specific proteins are associated with surfactant lipids, particularly the highly surface active subfraction known as tubular myelin. We have isolated a surfactant-associated protein complex with molecular weight components of 36 000, 32 000, and 28 000 and reassembled it with protein-free lung surfactant lipids prepared as small unilamellar liposomes. The effects of divalent cations on the structure and surface activity of this protein-lipid mixture were investigated by following (1) the state of lipid dispersion by changes in turbidity and by electron microscopy and (2) the ability of the surfactant lipids to form a surface film from an aqueous subphase at 37 degrees C. The protein complex markedly increased the rate of Ca2+-induced surfactant-lipid aggregation. Electron microscopy demonstrated transformation of the small unilamellar liposomes (median diameter 440 A) into large aggregates. The threshold Ca2+ concentration required for rapid lipid aggregation was reduced from 13 to 0.5 mM by the protein complex. This protein-facilitated lipid aggregation did not occur if Mg2+ was the only divalent cation present. Similarly, 5 mM Ca2+ but not 5 mM Mg2+ improved the ability of the protein-lipid mixture to form a surface film at 37 degrees C. Extensive aggregation of the surfactant lipids without protein by 20 mM Ca2+ or 20 mM Mg2+ did not promote rapid surface film formation. These results add to the growing evidence that specific Ca2+-protein-lipid interactions are important in determining both the structure and function of extracellular lung surfactant fractions.  相似文献   

6.
The glycerolipid composition of pea (Pisum sativum L.) root plastids and their capacity to synthesize glycerolipids from [UL-14C]glycerol-3-phosphate were determined. Pea root plastids primarily consist of monogalactosyldiacylglycerol, triacylglycerol, phosphatidylcholine, digalactosyldiacylglycerol, and diacylglycerol. Maximum rates of total glycerolipid biosynthesis were obtained in the presence of 2.4 mM glycerol-3-phosphate, 15 mM KHCO3, 0.2 mM sodium-acetate, 0.5 mM each of NADH and NADPH, 0.05 mM coenzyme A, 2 mM MgCl2, 1 mM ATP, 0.1 M Bis-Tris propane (pH 7.5), and 0.31 M sorbitol. Glycerolipid biosynthesis was completely dependent on exogenously supplied ATP, coenzyme A, and a divalent cation, whereas the remaining cofactors improved their activity from 1.3- to 2.4-fold. Radioactivity from glycerol-3-phosphate was recovered predominantly in phosphatidic acid, phosphatidylglycerol, diacylglycerol, and triacylglycerol with lesser amounts in phosphatidylcholine and monoacylglycerol. The proportions of the various radiolabeled lipids that accumulated were dependent on the pH and the concentration of ATP and glycerol-3-phosphate. The data presented indicate that pea root plastids can synthesize almost all of their component glycerolipids and that glycerolipid biosynthesis is tightly coupled to de novo fatty acid biosynthesis. pH and the availability of ATP may have important roles in the regulation of lipid biosynthesis at the levels of phosphatidic acid phosphatase and in the reactions that are involved in phosphatidylglycerol and triacylglycerol biosynthesis.  相似文献   

7.
Net K movements in reconstituted human red cell ghosts and the resealing of ghosts to cations after osmotic hemolysis of red cells have been studied as functions of the free Ca ion concentration. The Ca-dependent specific increase in K permeability was shown to be mediated by a site close to the internal surface of the membrane with an apparent dissociation constant ap pH 7.2 for Ca (K'p1) of 3-5 X 10(-7) M, for Sr of 7 X 10(-6) M. Ba and Mg did not increase the K-permeability of the membrane but inhibited the Ca-mediated permeability changes. K'D1 decreased in a nonlinear fashion when the pH was increased from 6.0 to 8.5. Two different pK' values of this membrane site were found at pH 8.3 and 6.3. The Ca-activated net K efflux into a K-free medium was almost completely inhibited by an increase in intracellular Na from 4 to 70mM. Extracellular K antagonized this Na effect. Changes in the extracellular Na (0.1-140 mM) or K (0.1-6 mM) concentrations had little effect and did not change K'p1. The Ca-stimulated recovery of a low cation permeability in ghost cells appeared to be mediated by a second membrane site which was accessible to divalent cations only during the process of hemolysis in media of low ionic strength. The apparent dissociation constant for Ca at this site (K'p2) varied between 6 X 10(-7) and 4 X 10(-6) M at pH 7.2 Mg, Sr, and Ba could replace Ca functionally. The selectivity sequence was Ca greater than Sr greater than Ba greater than Mg. K'p2 was independent on the pH value in the range between 6.0 and 8.0 Hill coefficients of 2 were observed for the interaction of Ca with both membrane sites suggesting that more than one Ca ion is bound per site. The Hill cofficients were affected neither by the ion composition nor by the Ph values of the intra-and extracellular media. It is concluded that two different pathways for the permeation of cations across the membrane are controlled by membrane sites with high affinities for Ca: One specific for K, one unspecific with respect to cations. The K-specific "channel" has properties similar to the K channel in excitable tissues.  相似文献   

8.
Essentially chlorophyll-free mitochondria were isolated from green leaves of spinach (Spinacia oleracea L. cv. Viking II). Uncoupled oxidation of exogenous NADPH (1 mM) to oxygen had an optimum at pH 6.0, and activity was relatively low at pH 7.0, even in the presence of 1 mM-CaCl2. There was a proportional increase in the apparent Km for NADPH with decreasing H+ concentrations, suggesting that NADPH protonated on the 2'-phosphate group was the true substrate. Exogenous NADH was oxidized by oxygen with an optimum at pH 6.9. Under low-cation conditions, EGTA or EDTA (both 1 mM) had no effect on the Vmax. of NADH oxidation, although the removal of bivalent cations from the membrane surface by the chelators could be observed by use of 9-aminoacridine fluorescence. In contrast, under high-cation conditions, chelators lowered the Vmax. by about 50%, probably due to a better approach of the negatively charged chelators to the negative membrane surface than under low-cation conditions. In a low-cation medium, the Vmax. of NADH oxidation was increased by about 50% by the addition of cations. This was caused by a lowering of the size of the negative surface potential through charge screening. In contrast with other cations, La3+ inhibited NADH oxidation, possibly through binding to lipids essential for NADH oxidation. The apparent Km for NADH varied 6-fold in response to changes in the size of the surface potential, suggesting that the approach of the negatively charged NADH to the active site is hampered by the negative surface potential. The results demonstrate that the spinach leaf cell can regulate the mitochondrial NAD(P)H oxidation through several mechanisms: the pH; the cation concentration in general; and the concentration of Ca2+ in particular. The results also emphasize the importance of electrostatic considerations when investigating the kinetic behaviour of membrane-bound enzymes.  相似文献   

9.
Escherichia coli membrane particulate fraction has been spin-labeled by incubating with sn-glycerol-3-phosphate, CTP, palmitoyl CoA and 12-nitroxide stearoyl CoA. Incorporation of the spin-labeled acyl chain into phosphatidyl-glycerol was confirmed. ESR spectrum of the spin-labeled phosphatidylglycerol in E. coli membrane consisted at least of two components; one due to the labels undergoing rapid anisotropic motions and the other due to strongly immobilized labels (the overall splitting value, approx. 58 G). The relative intensity of the two components was dependent on the concentration of divalent cations. The immobilized component decreased on treatment of the membrane with EDTA and increased on addition of Mg2+ or Ca2+. The spectrum at 1 mM Mg2+ or Ca2+ consisted almost only of the immobilized component. Spin-labeled phosphatidylglycerol in total lipid membrane showed ESR spectrum due to mobile labels and the spectrum was not affected appreciably by the divalent cations. The results suggest the divalent cation-mediated interaction of phosphatidylglycerol with proteins in E. coli membrane. Phosphoenolpyruvate-dependent uptake of methyl-alpha-D-glucoside was markedly accelerated by Mg2+. Ca2+ was not effective for the enhancement. The divalent cation-induced interaction of phosphatidylglycerol with proteins was discussed in relation to the sugar transport.  相似文献   

10.
We have characterized the surface activity of different-sized poly(ethylene-glycols) (PEG; M(r) 200-100,000 Da) in the presence or absence of lipid monolayers and over a wide range of bulk PEG concentrations (10(-8)-10% w/v). Measurements of the surface potential and surface pressure demonstrate that PEGs interact with the air-water and lipid-water interfaces. Without lipid, PEG added either to the subphase or to the air-water interface forms relatively stable monolayers. Except for very low molecular weight polymers (PEGs < 1000 Da), low concentrations of PEG in the subphase (between 10(-5) and 10(-4)% w/v) increase the surface potential from zero (with respect to the potential of a pure air-water interface) to a plateau value of approximately 440 mV. At much higher polymer concentrations, > 10(-1)% (w/v), depending on the molecular weight of the PEG and corresponding to the concentration at which the polymers in solution are likely to overlap, the surface potential decreases. High concentrations of PEG in the subphase cause a similar decrease in the surface potential of densely packed lipid monolayers spread from either diphytanoyl phosphatidylcholine (DPhPC), dipalmitoyl phosphatidylcholine (DPPC), or dioleoyl phosphatidylserine (DOPS). Adding PEG as a monolayer at the air-water interface also affects the surface activity of DPhPC or DPPC monolayers. At low lipid concentration, the surface pressure and potential are determined by the polymer. For intermediate lipid concentrations, the surface pressure-area and surface potential-area isotherms show that the effects due to lipid and PEG are not always additive and that the polymer's effect is distinct for the two lipids. When PEG-lipid-mixed monolayers are compressed to surface pressures greater than the collapse pressure for a PEG monolayer, the surface pressure-area and surface potential-area isotherms approach that of the lipid alone, suggesting that for this experimental condition PEG is expelled from the interface.  相似文献   

11.
We manipulate lipid bilayer surface charge and gauge its influence on gramicidin A channel conductance by two strategies: titration of the lipid charge through bulk solution pH and dilution of a charged lipid by neutral. Using diphytanoyl phosphatidylserine (PS) bilayers with CsCl aqueous solutions, we show that the effects of lipid charge titration on channel conductance are masked 1) by conductance saturation with Cs+ ions in the neutral pH range and 2) by increased proton concentration when the bathing solution pH is less than 3. A smeared charge model permits us to separate different contributions to the channel conductance and to introduce a new method for "bilayer pKa" determination. We use the Gouy-Chapman expression for the charged surface potential to obtain equilibria of protons and cations with lipid charges. To calculate cation concentration at the channel mouth, we compare different models for the ion distribution, exact and linearized forms of the planar Poisson-Boltzmann equation, as well as the construction of a "Gibbs dividing surface" between salt bath and charged membrane. All approximations yield the intrinsic pKain of PS lipid in 0.1 M CsCl to be in the range 2.5-3.0. By diluting PS surface charge at a fixed pH with admixed neutral diphytanoyl phosphatidylcholine (PC), we obtain a conductance decrease in magnitude greater than expected from the electrostatic model. This observation is in accord with the different conductance saturation values for PS and PC lipids reported earlier (, Biochim. Biophys. Acta. 552:369-378) and verified in the present work for solvent-free membranes. In addition to electrostatic effects of surface charge, gramicidin A channel conductance is also influenced by lipid-dependent structural factors.  相似文献   

12.
We investigated the interaction of a major apolipoprotein of pulmonary surfactant with mixtures of lipids analogous to those found in natural surfactant. The apolipoprotein was extracted from canine surfactant and was purified to about 90% homogeneity. The apolipoprotein was mixed with liposomes of lipids in buffers containing 0.1 M sodium chloride and 3 mM calcium chloride at 22°C for 2 h or 37°C for 30 min. Two fractions were separated by centrifugation in sucrose density gradients at 15 000 rev./min. One was comprised of an aggregated, relatively high density recombinant lipoprotein which sedimented to a position toward the bottom of the centrifuge tube; the other remained at the top of the centrifuge tube and was mainly comprised of unbound lipid. The amount of lipid recovered as a sedimenting lipoprotein was dependent upon its composition. Those mixtures of lipids which contained dipalmitoyl phosphatidylglycerol formed sedimenting complexes which comprised 14% to 53% of the recovered lipid; those without phosphatidylglycerol formed such aggregates with less than 13% of the available lipid. Moreover, the lipid-to-protein stoichiometry of the recombinant was also dependent upon phosphatidylglycerol, and lipids containing this phospholipid displayed enhanced binding at a critical concentration of lipid which varied with temperature and composition. Calcium was required to form the sedimenting complex at 37°C. These results suggest that phosphatidylglycerol may be involved in the formation of a micelle-like complex, the stoichiometry of which is regulated over a narrow range of lipid concentration, and the structure of which involves calcium. The physiological advantage of forming this complex has not been determined. We found, however, that lipids containing phosphatidylglycerol absorbed more rapidly to an air/liquid interface than did those without. This rate of adsorption was further increased after interaction with the apolipoprotein.  相似文献   

13.
A crude preparation of PEP carboxylase (EC 4.1.1.31) from the yellow lupin roots exhibits the pH optimum of activity within the range of 7.4-8.6 and the temperature optimum at 32 - 40 degrees C. Its Km for PEP is 0.1 mM, and Km for HCO3- is 0.7 mM. The affinity of the enzyme towards Mg2+ diminishes with the metal ion concentration. At the concentration of Mg2+ below 0.5 mM Km for Mg2+ is 0.07 mM and at the Mg2+ concentration over 1.5 mM it rises to 0.47 mM. The Hill coefficients are 0.37 and 0.88, respectively. Among several compounds affecting the PEP carboxylase activity, such as organic acids, amino acids, and sugar phosphates, at physiological pH (7.0 and 7.8), malate shows the strongest inhibition of a competitive character, its Ki being 2 mM. Also acidic amino acids strongly inhibit the enzyme activity, aspartate being more effective than glutamate. Glucose 6-phosphate and fructose 1,6-diphosphate markedly activate the enzyme. Both the inhibition by malate, aspartate and glutamate, and the activation by sugar phosphates rises considerably when pH is decreased from 7.8 to 7.0. Malonate scarcely affects the enzyme.  相似文献   

14.
Ionic properties of didodecanoylphosphatidylglycerol (C12PG), didodecanolyphosphatidyl-l'-propanol (C12PP), di-(12-methyl, 13-methyl)-pentadecanoylphosphatidylglycerols (C15PG) and dihexadecanoylphosphatidylglycerol (C16PG) have been studied at the air-water interface using titration experiments at constant ionic strength and film expansion experiments at constant pH, with Li+, Na+, K+ and Cs+ in the subphase. For each lipid, the apparent pK in the surface is strongly dependent on the subphase salt concentration and differs from expected intrinsic pK in the bulk. Discrimination between alkaline cations is observed. These results can be accounted for by strong surface potentials, which are satisfactorily calculated by using the Gouy and Chapman theory of the diffuse double layer. The comparison of C12PP and PG expansion data shows the importance of the glycerol residue of PG ionic properties, favouring penetration of cations in the films. Lipids in the liquid-crystalline state, such as C12-and C15PG, do not interact with alkaline cations as does C16PG in the gel phase. In particular, film condensations bring about a clear-cut discrimination between Na+ and K+. Results are discussed with regard to cation penetration and the structure of water at the interface. The importance on membrane functions of these strong surface potentials generated by PG monolayers is suggested.  相似文献   

15.
The effects of phospholipid vesicles and divalent cations in the subphase solution on the surface tension of phospholipid monolayer membranes were studied in order to elucidate the nature of the divalent cation-induced vesicle-membrane interaction. The monolayers were formed at the air/water interface. Various concentrations of unilamellar phospholipid (phosphatidylserine, phosphatidylcholine and their mixtures) vesicles and divalent cations (Mg2+, Ca2+, Mn2+, etc.) were introduced into the subphase solution of the monolayers. The changes of surface tension of monolayers were measured by the Wilhelmy plate (Teflon) method with respect to divalent ion concentrations and time.When a monolayer of phosphatidylserine and vesicles of phosphatidylserine/phosphatidylcholine (1 : 1) were used, there were critical concentrations of divalent cations to produce a large reduction in surface tension of the monolayer. These concentrations were 16 mM for Mg2+, 7 mM for Sr2+, 6 mM for Ca2+, 3.5 mM for Ba2+ and 1.8 mM for Mn2+. On the other hand, for a phosphatidylcholine monolayer and phosphatidylcholine vesicles, there was no change in surface tension of the monolayer up to 25 mM of any divalent ion used. When a phosphatidylserine monolayer and phosphatidylcholine vesicles were used, the order of divalent ions to effect the large reduction of surface tension was Mn2+ > Ca2+ > Mg2+ and their critical concentrations were in between the former two cases. The threshold concentrations also depended upon vesicle concentrations as well as the area/molecule of monolayers. For phosphatidylserine monolayers and phosphatidylserine/phosphatidylcholine (1 : 1) vesicles, above the critical concentrations of Mn2+ and Ca2+, the surface tension decreased to a value close to the equilibrium pressure of the monolayers within 0.5 h.This decrease in surface tension of the monolayers is interpreted partly as the consequence of fusion of the vesicles with the monolayer membranes. The  相似文献   

16.
In the present study the influence of the lipid environment on the organization of the main light-harvesting complex of photosystem II (LHCII) was investigated by 77K fluorescence spectroscopy. Measurements were carried out with a lipid-depleted and highly aggregated LHCII which was supplemented with the different thylakoid membrane lipids. The results show that the thylakoid lipids are able to modulate the spectroscopic properties of the LHCII aggregates and that the extent of the lipid effect depends on both the lipid species and the lipid concentration. Addition of the neutral galactolipids monogalactosyldiacylglycerol (MGDG) and digalactosyldiacylglycerol (DGDG) seems to induce a modification of the disorganized structures of the lipid-depleted LHCII and to support the aggregated state of the complex. In contrast, we found that the anionic lipids sulfoquinovosyldiacylglycerol (SQDG) and phosphatidylglycerol (PG) exert a strong disaggregating effect on the isolated LHCII. LHCII disaggregation was partly suppressed under a high proton concentration and in the presence of cations. The strongest suppression was visible at the lowest pH value (pH 5) and the highest Mg(2+) concentration (40 mM) used in the present study. This suggests that the negative charge of the anionic lipids in conjunction with negatively charged domains of the LHCII proteins is responsible for the disaggregation. Additional measurements by photon correlation spectroscopy and sucrose gradient centrifugation, which were used to gain information about the size and molecular mass of the LHCII aggregates, confirmed the results of the fluorescence spectroscopy. LHCII treated with MGDG and DGDG formed an increased number of aggregates with large particle sizes in the micromm-range, whereas the incubation with anionic lipids led to much smaller LHCII particles (around 40 nm in the case of PG) with a homogeneous distribution.  相似文献   

17.
The structure and surface activity of the extracellular fraction of pulmonary surfactant known as tubular myelin are Ca2+ dependent. Previous studies have demonstrated surfactant-specific proteins with monomeric molecular weights of 28,000-36,000 (SP28-36) are associated with this fraction. In reassembled lipoprotein mixtures, SP28-36 promotes the Ca2+-induced aggregation and surface activity of surfactant lipids, but the detailed interactions between Ca2+, SP28-36, and surfactant lipids have not been established. In this study, we investigated the effect of various cations on the aggregation of surfactant lipid liposomes in the presence of SP28-36. SP28-36 reduced the threshold ion concentration for liposome aggregation from greater than 10 to 0.5 mM for Ca2+, Ba2+, and Sr2+ but not Mg2+ or Mn2+. The liposome aggregation was reversed by ethylenediaminetetraacetic acid and not associated with leakage of carboxyfluorescein. SP28-36 promoted similar liposome aggregation at pH less than 5 in the absence of divalent cations. Surfactant lipids adsorbed slowly to an air-fluid interface in all ionic conditions unless SP28-36 was present. Both Ca2+ and H+ induced rapid lipid adsorption in the presence of SP28-36. The surface activity of native surfactant had a similar ion dependence. Electron micrographs of native surfactant showed typical tubular myelin structures at pH 7.4 only in the presence of Ca2+. At pH 4.4 in the absence of Ca2+, similar but not identical structures were seen. In the reconstituted system, SP28-36 in the presence of Ca2+ induced the formation of larger multilayered structures including parallel bilayers and small areas of squares and triangles with dimensions similar to structures found in the native material.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
1. The interaction between [(14)C]carboxymethylated cytochrome c and monolayers of egg phosphatidylethanolamine at the air/water interface has been investigated by measurements of surface radioactivity, pressure and potential. 2. On adding (14)C-labelled cytochrome c to the subphase under monolayers with a surface pressure below 24dynes/cm. there was an initial surface pressure increment as the protein penetrated, followed by an adsorption that could be detected only by a continued increase in the surface radioactivity. 3. Above film pressures of 24dynes/cm. only adsorption was observed, i.e. an increment in surface radioactivity with none in surface pressure. 4. The changes in surface parameters with penetration of cytochrome c added to the subphase were indirectly proportional to the initial pressure of the monolayer. With hydrogenated phosphatidylethanolamine the constant of proportionality was increased but penetration again ceased at 24dynes/cm. 5. On compressing a phosphatidylethanolamine film containing penetrated cytochrome c to 40dynes/cm. only a proportion of the protein was ejected on a subphase of 10mm-sodium chloride, whereas on a subphase of m-sodium chloride nearly all the protein was lost. 6. With both penetration and adsorption only a small proportion of the added cytochrome c interacted with the phospholipid films, and initially the amount bound was proportional to the added protein concentration. There was no evidence of a stoicheiometric relationship between the protein and phospholipid or the build-up of multilayers. The bonded protein was not released by removing cytochrome c from the subphase. 7. The addition of m-sodium chloride to the subphase delays the rate of protein penetration into low-pressure films, but the final surface-pressure increment is not appreciably decreased. In contrast, m-sodium chloride almost completely stops adsorption on to films at all pressures. 8. When sodium chloride is added to the subphase below cytochrome c adsorbed to monolayers at high pressures, so that the final concentration is 1m, only a proportion of the protein is desorbed and this decreases as the time of the interaction increases. This indicates that adsorption is initially electrostatic, followed by the formation of non-ionic bonds. 9. Alteration of the subphase pH under a high-pressure film leads to a steady increase in adsorption from pH3 to 8.5 followed by a rapid fall to zero adsorption at pH11. 10. The penetration into phospholipid monolayers at 10dynes/cm. shows a rate that is consistent with the relative electrostatic status of the two components of the interaction as the subphase pH is varied between 3 and 10.5. The final equilibrium penetration shows a pronounced peak in the increments of surface pressure at pH9.0 although a similar peak is not observed in the surface radioactivity. This indicates that more residues of the protein are penetrating into the film at about this pH. 11. Determinations were made of the electrophoretic mobilities of phosphatidylethanolamine particles both alone and after interaction with cytochrome c. 12. The electrophoretic mobilities of cytochrome c adsorbed on lipid particles showed an isoelectric point below that of cytochrome c. This and the observations on the monolayers suggest that, with cytochrome c, protein-protein interactions are weak compared with other proteins.  相似文献   

19.
1. The surface potentials of mixed monolayers of synthetic phospholipids with lipids that are fusogenic for hen erythrocytes were investigated. 2. At pH 5.6 and 10, but not at pH2, mixed monolayers of the fusogenic lipid, glycerol mono-oleate, with phosphatidylcholine exhibited negative deviations from the ideality rule in surface potential per molecule which were accompanied by negative deviations in mean molecular area. 3. Interactions of this type were not seen with chemically related but non-fusogenic lipids, nor were they found in mixed monolayers of any of the lipids with phosphatidylethanolamine. 4. Experiments with dihexadecyl phosphate and hexadecyltrimethyl-ammonium indicated that the complete head group of phosphatidylcholine is required for its observed behaviour with fusogenic lipids. 5. Bivalent cations (Ca2+, UO2(2+) or Zn2+) in the subphase at pH 5.6 significantly modified the behaviour of mixed monolayers of fusogenic lipids with phospholipids; there was a parallel perturbing effect of fusogenic lipids on interactions between monolayers of phospholipids and bivalent cations. 6. Possible molecular interactions of fusogenic lipids with membrane phospholipids, and the role of Ca2+, are discussed which may be relevant to cell fusion in erythrocytes induced by low-melting lipids in the presence of Ca2+.  相似文献   

20.
Through a monolayer investigation (π, ΔV), it is shown that the cationic antibiotic polymyxin B (or E) strongly interacts with films of acidic lipids, namely the didodecanoyl- and dihexadecanoylphosphatidylglycerol. The zwitterionic dihexadecanoylphosphatidylcholine was an unsuitable substrate. Interactions occurred at and above a polymyxin B concentration in the subphase of 2.5 · 10?7 M, bringing about a considerable increase of both π and ΔV. These interactions proceeded in two steps, as revealed by a biphasic change of ΔV with time. They were independent of the film molecular packing (fluid or gel states) and of the initial film pressure.Since it was possible to monitor the relative number of polymyxin B and didodecanoyl- or dihexadecanoylphosphatidylglycerol molecules in the monolayer, it is demonstrated that, at saturation, one polymyxin B molecule is bound to five phosphatidylglycerol molecules, a result which accounts for an exact neutralization of the charges.From competition experiments, it is shown that Na+ is ineffective in removing polymyxin B from the interface. Ca2+ appeared to be a stronger competitor but no complete antibiotic desorption was observed even at a Ca2+ concentration of 100 mM.As a working hypothesis, the antibiotic/lipid (15) system was assumed to constitute by itself one molecular species. The mixing of the polymyxin B/didodecanoylphosphatidylglycerol (15) system with an excess of lipid molecules in the monolayer was found to be ideal both in terms of π and ΔV. With dihexadecanoylphosphatidylglycerol, a small condensing effect could be detected only at intermediate surface pressures, in a region where the lipid phase transition occurred.The molecular area of polymyxin B interacting with didodecanoylphosphatidylglycerol can be calculated to be 1.23 ± 0.05 nm2. It is proposed that the whole antibiotic molecule penetrates the film, the five bound lipid molecules being distributed around the peptide structure, at given positions imposed by the five 2,4-diaminobutyric acid residues.  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