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 31 毫秒
1.
Liposomes which have entrapped the metallochromic dye, arsenazo III, constitute a sensitive assay system for ionophoresis of divalent cations. By this means we have compared known calcium ionophores (A23187, ionomycin) with membrane phospholipids, fatty acids, prostanoids, and retinoids. Added at micromolar concentrations to preformed multilamellar liposomes (phosphatidylcholine 7:dicetyl phosphate 2: cholesterol 1) both A23187 and ionomycin, as well as phosphatidic acid and products derived from linoleic acid, linolenic acid, and two eicosatrienoic acids provoked Ca influx (e.g. phosphatidic acid: 0.13 mol of Ca2+/mol of membrane lipid/5 min). A variety of other phospholipids (e.g. phosphatidylinositol), fatty acids (e.g. arachidonic acid), prostanoids (e.g. PGE1) retinoids (e.g. retinoic acid), and glyceryl ether phosphorylcholines ("platelet-activating factors") were without effect. Phosphatidic acid and oxidized fatty acids translocated divalent cations selectively, demonstrating the same rank order as A23187 or ionomycin: Mn greater than Ca greater than Sr much greater than Mg. Membrane lysis did not contribute to the perceived translocation; the liposomes remained impermeable to EDTA, EGTA, arsenazo III, or Mg. Liposomes with phosphatidic acid or oxidized trienoic acids preincorporated at 1-5 mole % of total lipids also permitted translocation of Ca but not Mg. Reduction of ionophoretic fatty acids or ionomycin with stannous chloride abolished their ionophoretic activity. Release of Ca from liposomes which had entrapped arsenazo III-Ca complexes into a medium rich in EGTA permitted calculation of efflux induced by ionophores, whether these were added to the outside of liposomes or preincorporated. Data suggest that phosphatidic acid and oxidized di- and trienoic fatty acids, which act as calcium ionophores in model bilayers, could serve as "endogenous ionophores" in cells.  相似文献   

2.
The antibiotic ionophore ionomycin translocates Ca from an aqueous medium into or across an organic immiscible phase. At pH 8.0, ionomycin translocates less Ca than A23187, the effects of these ionophores being additive to one another. The capacity of ionomycin to translocate Ca across the organic phase is dramatically decreased when the pH of the aqueous media is reduced from 8.0 to 7.5 or lower values. Ionomycin also mediates Ca exchange-diffusion in liposomes, the magnitude of such a process being greater in fluid than in rigid liposomes. At a physiological pH (7.4), ionomycin is unexpectedly as potent as A23187 in mediating Ca transport in fluid liposomes. These findings suggest that the capacity of ionophores to translocate Ca across model membranes depends on both the transverse and lateral mobility of the ionophoretic molecules. The relative importance of the latter phenomenon itself largely depends on the stoichiometry of the Ca-ionophore complex.  相似文献   

3.
We have studied the effect of the polyamines (spermine, spermidine, and putrescine) on the aggregation and fusion of large (approximately 100 nm in diameter) unilamellar liposomes in the presence of 100 mM NaCl, pH 7.4. Liposome fusion was monitored by the Tb/dipicolinic acid fluorescence assay for the intermixing of internal aqueous contents, and the release of contents was followed by carboxyfluorescein fluorescence. Spermine and spermidine at physiological concentrations aggregated liposomes composed of pure phosphatidylserine (PS) or phosphatidate (PA) and mixtures of PA with phosphatidylcholine (PC) but did not induce any fusion. However, liposomes composed of mixtures of acidic phospholipids, cholesterol, and a high mole fraction of phosphatidylethanolamine could be induced to fuse by spermine and spermidine in the absence of divalent cations. Putrescine alone in the physiological concentration range was ineffective for both aggregation and fusion of these liposomes. Liposomes made of pure PC did not aggregate in the presence of polyamines. Addition of aggregating concentrations of spermine caused a drastic increase in the rate of Ca(2+)-induced fusion of PA liposomes and a large decrease in the threshold Ca(2+) concentration required for fusion. This effect was less pronounced in the case of PS or PA/PC vesicles. Preincubation of PA vesicles with spermine before the addition of Ca(2+) resulted in a 30-fold increase in the initial rate of fusion. We propose that polyamines may be involved in the regulation of membrane fusion phenomena accompanying cell growth, cell division, exocytosis, and fertilization.  相似文献   

4.
Bilayer membrane destabilization induced by dolichylphosphate   总被引:1,自引:0,他引:1  
Small vesicles containing the fluorescent probe calcein were used to investigate the effect of dolichyl phosphate (Dol-P) on phospholipid bilayer stability. In the absence of Dol-P, phospholipid vesicles retained the fluorescent probe upon the addition of divalent cations. Small vesicles containing Dol-P, however, exhibited calcein leakage when incubated in the presence of divalent cations. This effect was observed in liposomes composed of a mixture of phosphatidylethanolamine (PE), phosphatidylcholine (PC) and Dol-P, but not in PC/Dol-P liposomes. The rate of calcein leakage was proportional to divalent cation concentration and to temperature, but was independent of vesicle concentration. These results demonstrate that Dol-P has significant effects on the stability of PE containing phospholipid bilayers. Vesicle leakage was also promoted by the addition of rat liver Dol-P-mannose synthase (EC 2.4.1.83) to intact PE/PC/Dol-P vesicles. Enzyme induced leakage from phospholipid vesicles required the presence of both unsaturated PE and Dol-P. The phospholipid composition of leaky vesicles could be correlated with the lipid matrix required for maximal transferase activity of the rat liver synthase. The destabilizing effects of Dol-P on phospholipid bilayers may therefore be involved in the translocation of activated sugars across biological membranes.  相似文献   

5.
Rates of phosphatidic acid- (PA-) mediated Ca2+-traversal are maximal in 'passive bilayers' void of lipid CO and OH groups: dietherphosphatidylcholine (diether-PC) or OH-blocked cholesterol liposomes. Phosphatidylcholine (PC) as bilayer matrix causes 99% inhibition, while 45 mol% cholesterol in passive bilayers inhibits by about 70%. Possibly, the absence of CO and OH groups causes a dehydration of the 'hydrogen belts', i.e., the membrane strata occupied by hydrogen bond acceptors (CO of phospholipids) and donors (OH of cholesterol, sphingosine) and thereby facilitates the formation of dehydrated Ca(PA)2, the ionophoric vehicle; or (our preferred explanation) PC engages in a (non-ionophoric) Ca(PA X PC) complex and thus reduces the concentration of the ionophore, while cholesterol competes with Ca2+ for the CO groups of phosphatidic acid by hydrogen-bonding. The Ca2+-traversal rates realized in bilayers with modified hydrogen belts lend support to the speculation that a Ca(PA)2 ferry may be of physiological importance, e.g., in membranes (such as myelin) containing much ether phospholipid (plasmalogen); and that Ca2+-membrane association and traversal may be controlled by the composition of the hydrogen belts.  相似文献   

6.
I Graham  J Gagné  J R Silvius 《Biochemistry》1985,24(25):7123-7131
The effects of calcium on the mixing of synthetic diacylphosphatidylcholines (PC's) and diacylphosphatidylethanolamines (PE's) with the corresponding phosphatidic acids (PA's) have been examined by high-sensitivity differential scanning calorimetry and by measurements of the fluorescence of labeled PA or PC species in PA-PC bilayers. Calorimetrically derived phase diagrams for dimyristoyl- and dielaidoyl-substituted PA-PC and PA-PE mixtures indicate that these species are readily miscible in the absence of calcium but phase-separate very extensively in the presence of high levels of calcium (30 mM). The limiting solubilities of PA (Ca2+) in liquid-crystalline PC or PE bilayers are less than or equal to 10 and approximately 5 mol %, respectively, while approximately 20 mol % of PC or PE can be introduced into the "cochleate" phase of PA (Ca2+) before a distinct PC-rich (or PE-rich) phase appears. The kinetics of calcium-induced lateral phase separations were examined for dioleoyl- and dielaidoyl-substituted PA-PC unilamellar vesicles labeled with fluorescent (C12-NBD-acyl) PA or PC, whose fluorescence becomes partially quenched upon phase separation. Our results indicate that, for the PA-PC system, lateral phase separation is very rapid (approximately less than 1 s) after calcium addition and develops partially (possibly in only one face of the bilayer) when calcium is present only on one side of the bilayer. Moreover, phase separations can develop at a rate faster than that of vesicle diffusion when calcium is added to dilute suspensions of vesicles, suggesting that interbilayer contacts are not essential to promote phase separations.  相似文献   

7.
N Oku  S Shibamoto  F Ito  H Gondo  M Nango 《Biochemistry》1987,26(25):8145-8150
For the purpose of cytoplasmic delivery of aqueous content in liposomes through endosomes, we synthesized a pH-sensitive polymer, cetylacetyl(imidazol-4-ylmethyl)polyethylenimine (CAIPEI), which generates polycations at acidic pH. CAIPEI in its aqueous phase caused aggregation of sonicated vesicles composed of phosphatidylserine (PS) and phosphatidylcholine (PC) (molar ratio 1:4) when the pH of the solution was lowered. The polymer also induced membrane intermixing as measured by resonance energy transfer between vesicles containing N-(7-nitro-2,1,3-benz[d]oxadiazol-4-yl)phosphatidylethanolamine and those containing N-Rhodamine phosphatidylethanolamine at pH 4-5, while the addition of CAIPEI caused neither aggregation of PC vesicles nor the intermixing of liposomal membranes between PC and PC/PS vesicles at any pH. The CAIPEI-induced membrane intermixing was dependent on the polymer/vesicle ratio rather than on the polymer concentration. Then the polymer was incorporated into the bilayers of PC vesicles. These CAIPEI vesicles also caused membrane intermixing with liposomes containing PS under acidic conditions. The reconstituted CAIPEI did not reduce the trapping efficiency of vesicles or increase their permeability to glucose even at low pH. The vesicles caused the low pH induced aggregation and membrane intermixing with other negatively charged liposomes containing phosphatidic acid or phosphatidylglycerol. These results suggest that the protonation of the polymer at acidic pH endows the CAIPEI vesicles with the activity to fuse with negatively charged liposomes.  相似文献   

8.
Abstract

The interactions of a water-soluble nonmembrane protein aprotinin with multilamellar vesicles (MLV) and small unilamellar vesicles (SUV) from soybean phospholipids were studied using Sephadex G-75 gel chromatography combined with different methods of the analysis of the eluate fractions (fluorescence, light-scattering, turbidity; 31P NMR spectroscopy). The composition of the liposomes mainly containing soybean phosphatidylcholine (PC) was varied by the addition of phosphatidylethanolamine (PE), phosphatidylinositol (PI) and lyso-phosphatidylcholine (lyso-PC). To evaluate the lipid-protein interactions, the amount of aprotinin in the MLV–aprotinin complexes was determined. Lipid–protein interactions were found to strongly depend on the liposome composition, medium pH and ionic strength. These dependencies point to the electrostatic nature of the aprotinin-lipid interactions. 31P NMR spectroscopy of the MLV–aprotinin complexes indicated that aprotinin influences the phospholipid structure in MLV at pH 3.0. In the case of PC:PE:PI and PC:PE:PI:lyso-PC vesicles, aprotinin induced liposome aggregation and a lamellar-to-isotropic phase transition of the phospholipids.  相似文献   

9.
The interactions of a water-soluble nonmembrane protein aprotinin with multilamellar vesicles (MLV) and small unilamellar vesicles (SUV) from soybean phospholipids were studied using Sephadex G-75 gel chromatography combined with different methods of the analysis of the eluate fractions (fluorescence, light-scattering, turbidity; 31P NMR spectroscopy). The composition of the liposomes mainly containing soybean phosphatidylcholine (PC) was varied by the addition of phosphatidylethanolamine (PE), phosphatidylinositol (PI) and lyso-phosphatidylcholine (lyso-PC). To evaluate the lipid-protein interactions, the amount of aprotinin in the MLV-aprotinin complexes was determined. Lipid-protein interactions were found to strongly depend on the liposome composition, medium pH and ionic strength. These dependencies point to the electrostatic nature of the aprotinin-lipid interactions. 31P NMR spectroscopy of the MLV-aprotinin complexes indicated that aprotinin influences the phospholipid structure in MLV at pH 3.0. In the case of PC:PE:PI and PC:PE:PI:lyso-PC vesicles, aprotinin induced liposome aggregation and a lamellar-to-isotropic phase transition of the phospholipids.  相似文献   

10.
J R Silvius 《Biochemistry》1990,29(12):2930-2938
A novel method that uses a carbazole-labeled fluorescent phosphatidylcholine, which partitions preferentially into liquid-crystalline lipid domains, to monitor the kinetics and the extents of thermotropic and ionotropic lateral phase separations in vesicles combining brominated and nonbrominated phosphatidylcholines (PCs), phosphatidic acids (PAs), and phosphatidylserines (PSs) is described. The calcium-induced segregation of several nonbrominated PA species in liquid-crystalline brominated PC bilayers behaves as a well-defined lateral phase separation; the residual solubility of the PA component in the PC-rich phase in the presence of calcium can vary severalfold depending on the PA acyl chain composition. PC/PS mixtures show a pronounced tendency to form metastable solutions in the presence of calcium, particularly when they contain less than equimolar proportions of PS. This metastability is not readily relaxed by repeated freeze-thawing of vesicles in the presence of calcium, by avidin-mediated contacts between PC/PS vesicles containing biotinylated lipids, or by calcium-induced lateral segregation of PA in the same vesicles. Different PS species exhibit different apparent residual solubilities in liquid-crystalline PC bilayers, ranging from less than 10 mol % for dimyristoyl-PS to ca. 45 mol% for dioleoyl-PS, after prolonged incubations of PC/PS multilamellar vesicles with excess calcium. Results are presented, obtained by using the above lipid-segregation assay and parallel assays of intervesicle lipid mixing, that raise questions concerning the relevance of the equilibrium behavior of calcium-treated PS/PC mixtures to the relatively rapid interactions (fusion and lipid mixing) of PC/PS vesicles that follow initial exposure to calcium.  相似文献   

11.
The phase and colloidal properties of phosphatidylcholine/fatty acid (PC/FA) mixed vesicles have been investigated by optical methods, acid-base titration, and theoretically as a function of temperature (5-80 degrees C), molar lipid ratio (0-1), lipid chain length (C14-C18), headgroup ionization (1.5 less than or equal to pH less than or equal to 10), vesicle concentration (0.05-32 mumol vesicle.dm-3, and ionic strength (0.005 less than or equal to J less than or equal to 0.25). Increasing the fatty acid concentration in PC bilayers causes the phase transition temperatures (at 4 less than or equal to pH less than or equal to 5) to rise until, for more than 2 FA molecules per PC molecule, the sample turbidity exhibits only two transitions corresponding to the chain-melting of the 1:2 stoichiometric complexes of PC/FA, and pure fatty acid. The former transition is into a nonlamellar phase and is accompanied by extremely rapid vesicle aggregation (with association rates on the order of Ca approximately 10(7) dm3.mol-1.s-1) and massive lipid precipitation. Fluid-phase vesicles with less than 2 FA per PC associate much more slowly (Ca approximately 10(3) dm3.mol-1.s-1), their aggregation being comparable to that of the ordered-phase liposomes. Under no conditions was the relation between the fatty acid concentration and the vesicle association rate for the fluid-phase vesicles linear. In contrast to the X-ray diffraction data, optical measurements reveal a 'pretransitional region' between the chain-melting temperature of the PC component and the temperature at which the gross transformation into a nonlamellar phase sets in. This is seen for all lipid mixtures investigated. On the relative temperature scale, lipids with different chain lengths behave qualitatively similarly; however, the effective association constants determined for samples of constant lipid concentration seem to decrease somewhat with the number of CH2 groups per chain. Fatty acid protonation, which yields electrically neutral bilayers, invariably increases the rate of vesicle association; we have measured, for example, Ca approximately 10(2) at pH approximately 7 and Ca approximately 10(7) dm3.mol-1.s-1 at pH approximately 4). Protonation of the phosphatidylcholine phosphate groups, which causes a net positive charge to accumulate on the lipid vesicles, initially increases (Ca approximately 10(8) dm3.mol-1.s-1) but ultimately decreases (Ca approximately 10(7) dm3.mol-1.s-1) the rate of association between PC/FA (1:2) mixed vesicles.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
We examined the relationship between phosphatidylcholine (PC) hydrolysis, phosphoinositide hydrolysis, and diacylglycerol (DAG) formation in response to muscarinic acetylcholine receptor (mAChR) stimulation in 1321N1 astrocytoma cells. Carbachol increases the release of [3H]choline and [3H]phosphorylcholine ([3H]Pchol) from cells containing [3H]choline-labeled PC. The production of Pchol is rapid and transient, while choline production continues for at least 30 min. mAChR-stimulated release of Pchol is reduced in cells that have been depleted of intracellular Ca2+ stores by ionomycin pretreatment, whereas choline release is unaffected by this pretreatment. Phorbol 12-myristate 13-acetate (PMA) increases the release of choline, but not Pchol, from 1321N1 cells, and down-regulation of protein kinase C blocks the ability of carbachol to stimulate choline production. Taken together, these results suggest that Ca2+ mobilization is involved in mAChR-mediated hydrolysis of PC by a phospholipase C, whereas protein kinase C activation is required for mAChR-stimulated hydrolysis of PC by a phospholipase D. Both carbachol and PMA rapidly increase the formation of [3H]phosphatidic acid ([3H]PA) in cells containing [3H]myristate-labeled PC. [3H]Diacylglycerol ([3H]DAG) levels increase more slowly, suggesting that the predominant pathway for PC hydrolysis is via phospholipase D. When cells are labeled with [3H]myristate and [14C]arachidonate such that there is a much greater 3H/14C ratio in PC compared with the phosphoinositides, the 3H/14C ratio in DAG and PA increases with PMA treatment but decreases in response to carbachol. By analyzing the increase in 3H versus 14C in DAG, we estimate that the DAG that is formed in response to PMA arises largely from PC. Muscarinic receptor activation also causes formation of DAG from PC, but approximately 20% of carbachol-stimulated DAG appears to arise from hydrolysis of the phosphoinositides.  相似文献   

13.
Phospholipase D 2 (PLD2) is the major PLD isozyme associated with the cardiac sarcolemmal (SL) membrane. Hydrolysis of SL phosphatidylcholine (PC) by PLD2 produces phosphatidic acid (PA), which is then converted to 1,2 diacylglycerol (DAG) by the action of phosphatidate phosphohydrolase type 2 (PAP2). In view of the role of both PA and DAG in the regulation of Ca(2+) movements and the association of abnormal Ca(2+) homeostasis with congestive heart failure (CHF), we examined the status of both PLD2 and PAP2 in SL membranes in the infarcted heart upon occluding the left coronary artery in rats for 1, 2, 4, 8 and 16 weeks. A time-dependent increase in both SL PLD2 and PAP2 activities was observed in the non-infarcted left ventricular tissue following myocardial infarction (MI); however, the increase in PAP2 activity was greater than that in PLD2 activity. Furthermore, the contents of both PA and PC were reduced, whereas that of DAG was increased in the failing heart SL membrane. Treatment of the CHF animals with imidapril, an angiotensin-converting enzyme (ACE) inhibitor, attenuated the observed changes in heart function, SL PLD2 and PAP2 activities, as well as SL PA, PC and DAG contents. The results suggest that heart failure is associated with increased activities of both PLD2 and PAP2 in the SL membrane and the beneficial effect of imidapril on heart function may be due to its ability to prevent these changes in the phospholipid signaling molecules in the cardiac SL membrane.  相似文献   

14.
Sodium and calcium movements in dog red blood cells   总被引:7,自引:5,他引:2       下载免费PDF全文
Determinants of 45Ca influx, 45Ca efflux, and 22Na efflux were examined in dog red blood cells. 45Ca influx is strongly influenced by the Na concentration on either side of the membrane, being stimulated by intracellular Na and inhibited by extracellular Na. A saturation curve is obtained when Ca influx is plotted as a function of medium Ca concentration. The maximum Ca influx is a function of pH (increasing with greater alkalinity) and cell volume (increasing with cell swelling). Quinidine strongly inhibits Ca influx. Efflux of 45Ca is stimulated by increasing concentrations of extracellular Na. 22Na efflux is stimulated by either Ca or Na in the medium, and the effects of the two ions are mutually exclusive rather than additive. Quinidine inhibits Ca-activated 22Na efflux. The results are considered in terms of a model for Ca-Na exchange, and it is concluded that the system shows many features of such a coupled ion transport system. However, the stoichiometric ratio between Ca influx and Ca-dependent Na efflux is highly variable under different experimental conditions. Because the Ca fluxes may reflect a combination of ATP-dependent, outward transport and Na-linked passive movements, the true stoichiometry of an exchanger may not be ascertainable in the absence of a specific Ca pump inhibitor. The meaning of these observations for Ca-dependent volume regulation by dog red blood cells is discussed.  相似文献   

15.
In electrically excitable cells, membrane depolarization opens voltage-dependent Ca(2+) channels eliciting Ca(2+) influx, which plays an important role for the activation of protein kinase C (PKC). However, we do not know whether Ca(2+) influx alone can activate PKC. The present study was conducted to investigate the Ca(2+) influx-induced activation mechanisms for two classes of PKC, conventional PKC (cPKC; PKCalpha) and novel PKC (nPKC; PKCtheta), in insulin-secreting cells. We have demonstrated simultaneous translocation of both DsRed-tagged PKCalpha to the plasma membrane and green fluorescent protein (GFP)-tagged myristoylated alanine-rich C kinase substrate to the cytosol as a dual marker of PKC activity in response to depolarization-evoked Ca(2+) influx in the DsRed-tagged PKCalpha and GFP-tagged myristoylated alanine-rich C kinase substrate co-expressing cells. The result indicates that Ca(2+) influx can generate diacylglycerol (DAG), because cPKC is activated by Ca(2+) and DAG. We showed this in three different ways by demonstrating: 1) Ca(2+) influx-induced translocation of GFP-tagged C1 domain of PKCgamma, 2) Ca(2+) influx-induced translocation of GFP-tagged pleckstrin homology domain, and 3) Ca(2+) influx-induced translocation of GFP-tagged PKCtheta, as a marker of DAG production and/or nPKC activity. Thus, Ca(2+) influx alone via voltage-dependent Ca(2+) channels can generate DAG, thereby activating cPKC and nPKC, whose activation is structurally independent of Ca(2+).  相似文献   

16.
Summary The purpose of this study was to examine (1) the association of tumor extract proteins with phospholipid vesicles of varying physiochemical properties, and (2) the adjuvant and carrier properties of liposome-borne tumor antigens in the in vivo induction of an antitumor immune response. Cell surface antigens of the 3-methylcholanthrene-induced fibrosarcoma of C3H/HeJ mice, MCA-F, were extracted using 2.5% 1-butanol. Crude and electrofocused antigen preparations capable of eliciting a protective antitumor immune response were used to prepare liposome vaccines. The incorporation of extract proteins into liposomes formed by butanol dialysis (BVD) was three- to five-fold greater than the encapsulation of protein into the aqueous compartment of multilamellar vesicles (MLV). The electrochemical properties of the BDV had a significant effect on the induction of an antitumor response: Antigens borne on negatively charged, but not uncharged, liposomes were effective in protecting hosts against supralethal tumor challenge, and displayed a specific activity 20- to 50-fold greater than soluble antigen. Antigens carried by MLV were not effective in generating an immunoprotective response. The lipophilic characteristics of butanol-extracted antigens allowed (1) the passive adsorption of immunoprotective tumor antigen onto the surface of preformed vesicles, and (2) adsorption of MCA-F antigen onto the surface of an antigenically distinct tumor MCA-D. In the latter experiment, adsorption of MCA-F-specific antigen onto MCA-D cells resulted in a change in the membrane antigen phenotype as measured by indirect immunofluorescence. Although butanol released a lipophilic moiety from cells which spontaneously reassociated with phospholipid bilayers, no evidence for a lipoidal antigen was obtained when tumor-derived lipids were used as immunogens. This study demonstrates that butanol-extracted tumor antigen is lipophilic without being a lipid, and that negatively charged liposomes can be effective as carriers and adjuvants for tumor antigens in the induction of an antitumor immune response. Abbreviations used: BDV, butanol dialysis vesicles; CBE-F, crude butanol extract from the MCA-F tumor; Chol, cholesterol; FI, fluorescense index; PBS, Dulbecco's calcium- and magnesium-free phosphate-buffered saline, pH 7.4; PC, phosphatidylcholine; PG, phosphatidylglycerol; pIEF, preparative isoelectric focusing; MLV, multilamellar vesicles; REV, reverse-evaporation vesicles  相似文献   

17.
The kinetics and thermodynamics of the transmembrane movement (flip-flop) of fluorescent analogs of phosphatidic acid (PA), phosphatidylglycerol (PG), phosphatidylcholine (PC), and phosphatidylethanolamine (PE) were investigated to determine the contributions of headgroup composition and acyl chain length to phospholipid flip-flop. The phospholipid derivatives containing n-octanoic, n-decanoic or n-dodecanoic acid in the sn-1 position and 9-(1-pyrenyl)nonanoic acid in the sn-2 position were incorporated at 3 mol% into sonicated single-bilayer vesicles of 1-palmitoyl-2-oleoyl-sn-glycerol-3-phosphocholine (POPC). The kinetics of diffusion of the pyrene-labeled phospholipids from the outer and inner monolayers of the host vesicles to a large pool of POPC acceptor vesicles were monitored by the time-dependent decrease of pyrene excimer fluorescence. The observed kinetics of transfer were biexponential, with a fast component due to the spontaneous transfer of pyrenyl phospholipids in the outer monolayer of labeled vesicles and a slower component due to diffusion of pyrenyl phospholipid from the inner monolayer of the same vesicles. Intervesicular transfer rates decreased approx. 8-fold for every two carbons added to the first acyl chain. Correspondingly, the free energy of activation for transfer increased approx. 1.3 kcal/mol. With the exception of PE, the intervesicular transfer rates for the different headgroups within a homologous series were nearly the same, with the PC derivative being the fastest. Transfer rates for the PE derivatives were 5-to 7-fold slower than the rates observed for PC. Phospholipid flip-flop, in contrast, was strongly dependent on headgroup composition with a smaller dependence on acyl chain length. At pH 7.4, flip-flop rates increased in the order PC less than PG less than PA less than PE, where the rates for PE were at least 10-times greater than those of the homologous PC derivative. Activation energies for flip-flop were large, and ranged from 38 kcal/mol for the longest acyl chain derivative of PC to 25 kcal/mol for the PE derivatives. Titration of the PA headgroup at pH 4.0 produced an approx. 500-fold increase in the flip-flop rate of PA, while the activation energy decreased 10 kcal/mol. Increasing acyl chain length reduced phospholipid flip-flop rates, with the greatest change observed for the PC analogs, which exhibited an approx. 2-fold decrease in flip-flop rate for every two methylene carbons added to the acyl chain at the sn-1 position.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
Initial rates of ionophore-mediated Ca2+ transport across egg phosphatidylcholine bilayers of large unilamellar vesicles were measured using the absorbance change of arsenazo III at 650 nm as an indicator of Ca2+ translocation. A23187 induced the movement of Ca2+ in a 2:1 ionophore: Ca2+ complex, whereas its methyl ester (CH3A23187) and X537A mediated Ca2+ movement in a 1:1 ionophore: Ca2+ complex. The relative potencies of these ionophores in transporting Ca2+ across lipid membranes were A23187 much greater than X537A greater than CH3A23187.  相似文献   

19.
In mouse neuroblastoma N18TG2 cells prelabeled with [3H]arachidonic acid ([3H]AA) the biosynthesis of 2-arachidonoylglycerol (2-AG) is induced by ionomycin in a fashion sensitive to an inhibitor of diacylglycerol (DAG) lipase, RHC 80267, but not to four different phospholipase C (PLC) blockers. Pulse experiments with [3H]AA showed that ionomycin stimulation leads to the sequential formation of [3H]phosphatidic acid ([3H]PA), [3H]DAG, and [3H]2-AG. [3H]2-AG biosynthesis in N18TG2 cells prelabeled with [3H]AA was counteracted by propranolol and N-ethylmaleimide, two inhibitors of the Mg2+/Ca2(+)-dependent brain PA phosphohydrolase. Pretreatment of cells with exogenous phospholipase D (PLD) led to a strong potentiation of ionomycin-induced [3H]2-AG formation. These data indicate that DAG precursors for 2-AG in intact N18TG2 cells are obtained from the hydrolysis of PA and not through the activation of PLC. The presence of 2% ethanol during ionomycin stimulation failed to elicit the synthesis of [3H]phosphatidylethanol and did not counteract the formation of [3H]PA, thus arguing against the activation of PLD by the Ca2+ ionophore. Selective inhibitors of secretory phospholipase A2 and the acyl-CoA acylase inhibitor thimerosal significantly reduced [3H]2-AG biosynthesis. The implications of these latter findings, and of the PA-dependent pathways of 2-AG formation described here, are discussed.  相似文献   

20.
The Streptomyces chromofuscus phospholipase D (PLD) cleavage of phosphatidylcholine in bilayers can be enhanced by the addition of the product phosphatidic acid (PA). Other anionic lipids such as phosphatidylinositol, oleic acid, or phosphatidylmethanol do not activate this PLD. This allosteric activation by PA could involve a conformational change in the enzyme that alters PLD binding to phospholipid surfaces. To test this, the binding of intact PLD and proteolytically cleaved isoforms to styrene divinylbenzene beads coated with a phospholipid monolayer and to unilamellar vesicles was examined. The results indicate that intact PLD has a very high affinity for PA bilayers at pH >/= 7 in the presence of EGTA that is weakened as Ca(2+) or Ba(2+) are added to the system. Proteolytically clipped PLD also binds tightly to PA in the absence of metal ions. However, the isolated catalytic fragment has a considerably weaker affinity for PA surfaces. In contrast to PA surfaces, all PLD forms exhibited very low affinity for PC interfaces with an increased binding when Ba(2+) was added. All PLD forms also bound tightly to other anionic phospholipid surfaces (e.g. phosphatidylserine, phosphatidylinositol, and phosphatidylmethanol). However, this binding was not modulated in the same way by divalent cations. Chemical cross-linking studies suggested that a major effect of PLD binding to PA.Ca(2+) surfaces is aggregation of the enzyme. These results indicate that PLD partitioning to phospholipid surfaces and kinetic activation are two separate events and suggest that the Ca(2+) modulation of PA.PLD binding involves protein aggregation that may be the critical interaction for activation.  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