首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The lipid distribution in binary mixed membranes containing charged and uncharged lipids and the effect of Ca2+ and polylysine on the lipid organization was studied by the spin label technique. Dipalmitoyl phosphatidic acid was the charged, and spin labelled dipalmitoyl lecithin was the uncharged (zwitterionic) component. The ESR spectra were analyzed in terms of the spin exchange frequency, Wex. By measuring Wex as a function of the molar percentage of labelled lecithin a distinction between a random and a heterogeneous lipid distribution could be made. It is established that mixed lecithin-phosphatidic acid membranes exhibit lipid segregation (or a miscibility gap) in the fluid state. Comparative experiments with bilayer and monolayer membranes strongly suggest a lateral lipid segregation. At low lecithin concentration, aggregates containing between 25% and 40% lecithin are formed in the fluid phosphatidic acid membrane. This phase separation in membranes containing charged lipids is understandable on the basis of the Gouy-Chapman theory of electric double layers. In dipalmitoyl lecithin and in dimyristoyl phosphatidylethanolamine membranes the labelled lecithin is randomly distributed above the phase transition and has a coefficient of lateral diffusion of D = 2.8-10(-8) cm2/s at 59 degrees C. Addition of Ca2+ dramatically increases the extent of phase separation in lecithin-phosphatidic acid membranes. This chemically (and isothermally) induced phase separation is caused by the formation of crystalline patches of the Ca2+-bound phosphatidic acid. Lecithin is squeezed out from these patches of rigid lipid. The observed dependence of Wex on the Ca2+ concentration could be interpreted quantitatively on the basis of a two-cluster model. At low lecithin and Ca2+ concentration clusters containing about 30 mol % lecithin are formed. At high lecithin or Ca2+ concentrations a second type of precipitation containing 100% lecithin starts to form in addition. A one-to-one binding of divalent ions and phosphatidic acid at pH 9 was assumed. Such a one-to-one binding at pH 9 was established for the case of Mn2+ using ESR spectroscopy. Polylysine leads to the same strong increase in the lecithin segregation as Ca2+. The transition of the phosphatidic acid bound by the polypeptide is shifted from Tt = 47.5 degrees to Tt = 62 degrees C. This finding suggests the possibility of cooperative conformational changes in the lipid matrix and in the surface proteins in biological membranes.  相似文献   

2.
The effect of myotoxin a on the thermotropic phase behavior of aqueous dispersions of dimyristoyl phosphatidylcholine (DMPC) and dimyristoyl phosphatidylserine (DMPS) was examined using differential scanning calorimetry (DSC). Myotoxin a significantly altered the normal phase behavior of DMPC in a concentration dependent fashion. This effect is perturbed by Ca2+ and is sensitive to ionic strength and pH. High concentrations of toxin eliminate the characteristic pretransition associated with the polar head group of DMPC. They also increase the temperature of the main gel-to-liquid crystal transition from 23 degrees C to 32-35 degrees C. At low concentrations of toxin, the first visible effect is upon the pretransition which is split into two components that diminish with time. The main transition is less affected at low toxin concentrations, although the magnitude of the transition is reduced while it is simultaneously shifted to higher temperatures. The main transition is also split into multiple components. The toxin also had pH specific effects on the phase behavior of DMPS. Above physiological pH (8.5) the normal transition of DMPS at 36-38 degrees C was split in the presence of myotoxin a and new components appeared centered at 31 degrees C and 35 degrees C. These observations are consistent with reports that the skeletal muscle membrane system is the major site of the myonecrotic effect of myotoxin a.  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 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   

4.
Ca2+-induced phase separation in phosphatidylserine/phosphatidylethanolamine and phosphatidylserine/phosphatidylethanolamine/phosphatidylcholine model membranes was studied using spin-labeled phosphatidylethanolamine and phosphatidylcholine and compared with that in phosphatidylserine/phosphatidylcholine model membranes studied previously. The phosphatidylethanolamine-containing membranes behaved in qualitatively the same way as did phosphatidylserine/phosphatidylcholine model membranes. There were some quantitative differences between them. The degree of phase separation was higher in the phosphatidylethanolamine-containing membranes. For example, the degree of phase separation in phosphatidylserine/phosphatidylethanolamine membranes containing various mole fractions of phosphatidylserine was 94--100% at 23 degrees C and 84--88% at 40 degrees C, while the corresponding value for phosphatidylserine/phosphatidylcholine membranes was 74--85% at 23 degrees C and 61--79% at 40 degrees C. Ca2+ concentration required for the phase separation was lower for phosphatidylserine/phosphatidylethanolamine than that for phosphatidylserine/phosphatidylcholine membranes; concentration to cause a half-maximal phase separation was 1.4 . 10(-7) M for phosphatidylserine-phosphatidylethanolamine and 1.2 . 10(-6) M for phosphatidylserine/phosphatidylcholine membranes. The phase diagram of phosphatidylserine/phosphatidylethanolamine membranes in the presence of Ca2+ was also qualitatively the same as that of phosphatidylserine/phosphatidylcholine except for the different phase transition temperatures of phosphatidylethanolamine (17 degrees C) and phosphatidylcholine (-15 degrees C). These differences were explained in terms of a greater tendency for phosphatidylethanolamine, compared to phosphatidylcholine, to form its own fluid phase separated from the Ca2+-chelated solid-phase phosphatidylserine domain.  相似文献   

5.
(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   

6.
Seven geometrical or positional isomers of dipalmitoyl cyclopentanophosphoric acid (DPCPA) have been synthesized and studied: 1,3/2-1P (I); 1,2/3-1P (II); 1,2/3-3P (III), 1,2,3/0-1P (IV); 1,2,3/0-2P (V); 1,3/2-2P (VI); 1,2/3-2P (VII). When dispersed in 0.1 M Tris-HCl at pH 7.4, I-VII gave thermal transitions (Tc) of 60.0 degrees, 59.0 degrees, 56.8 degrees, 55.3 degrees, 38.3 degrees, 36.8 degrees and 34.0 degrees C, respectively, as measured by differential scanning calorimetry (DSC). When the lipids were dispersed at pH 9.5 in 0.1 M borate, Tc of I-IV decreased, whereas Tc of V-VII increased. In contrast, at pH 1.5 in 0.1 M HCl/KCl, Tc of I-IV decreased slightly, but Tc of V-VII rose markedly. To determine the effect of head group geometry and substitution pattern on acyl chain motion, EPR spectra of 1-palmitoyl, 2-[16-doxylstearoyl]-glycero-3-phosphoric acid in bilayers of DPCPA isomers were acquired. Abrupt spectral changes occurred at temperatures closely correlating with transition temperatures observed by DSC. These results have led to the conclusions that: (i) isomers I-IV containing vicinal acyl chains form bilayers that exhibit structural transitions at temperatures higher than those at which transitions are exhibited by isomers V-VII which have a polar phosphate group interposed between the two chains; (ii) the effects of differences in backbone structure are transmitted down the entire length of the acyl chains; (iii) the orientation of the cyclopentane ring in the isomers I-IV is significantly different from that in isomers V-VII at pH values where the phosphate group is doubly negatively charged.  相似文献   

7.
Ca2+-translocating activities of phosphatidylinositol, diacylglycerol and phosphatidic acid were investigated in phosphatidylcholine liposomes. Using a fluorescent indicator of Ca2+ concentration, quin-2, release of encapsulated Ca2+ from egg yolk phosphatidylcholine liposomes containing 2 mol% of one of these lipids was measured at 37 degrees C. The rate of Ca2+ translocation across the liposomal membrane mediated by phosphatidic acid was about 3-fold larger than those mediated by phosphatidylinositol and diacylglycerol. The result implies that phosphatidic acid has Ca2+-ionophore activity in the agonist dependent metabolism of inositol phospholipids. The ionophoretic activity depended on the degree of unsaturation of the fatty acyl chains. The Ca2+ translocation rate was smallest in dipalmitoylphosphatidic acid, and it increased in the order of dioleoyl-, dilinoleoyl- and dilinolenoyl-phosphatidic acid. Ca2+ mobilization of a stimulated cell is discussed in the light of Ca2+-ionophore activity of phosphatidic acid converted from inositol phospholipids.  相似文献   

8.
Phase separation in mixed monolayers of phosphatidylcholine (PC) and pyrene-labeled phosphatidic acid (PA) was observed by fluorescence microscopy on an air/water interface as a function of subphase Ca2+ concentration and lateral packing pressure of the film. Below 45 mN m-1 and in the absence of Ca2+ no indications of phase immiscibility were observed. Addition of 1 mM Ca2+ caused extensive phase separation, which was evident immediately after spreading of the film. Further increase in Ca2+ concentration up to 30 mM increased the pyrene excimer intensity of the separated phosphatidic acid enriched domains. In the presence of Ca2+ (1-30 mM) and at surface pressures below 10 mN m-1 phase separation was always evident. However, as surface pressure exceeded 10 mN m-1, mixing of PC and PA occurred. Upon decompression of the film, phase separation reappeared at surface pressures close to 10 mN m-1. The surface textures of the film before and after the compression and subsequent relaxation were different. Inclusion of 30 mol% cholesterol increased the number and decreased the size of the PA domains. In films containing 50 mol% cholesterol no phase separation could be detected at the resolution available.  相似文献   

9.
Two kinds of calcium-dimyristoylphosphatidic acid (DMPA) complexes at acidic and neutral pH conditions were prepared in the following ways. The complex at pH 4 was obtained by adding Ca2+ to DMPA dispersion in pure water. On the other hand, the complex at pH 7.4 was obtained by adding Ca2+ to DMPA dispersion in the presence of NaOH. The stoichiometries of Ca2+ ion to DMPA molecule are 0.5-0.67 and approximately 1 for the complexes at pH 4 and 7.4, respectively. Static x-ray diffraction shows that the hydrocarbon chains of the Ca(2+)-DMPA complex at pH 4 at 20 degrees C are more tightly packed than those of the complex at pH 7.4 at 20 degrees C. Furthermore, the complex at pH 4 at 20 degrees C gives rise to several reflections that might be related to the ordered arrangement of the Ca2+ ions. These results indicate that the structure of the complex at pH 4 is crystalline-like. In the differential scanning calorimetry (DSC) thermogram, the complex at pH 7.4 undergoes no phase transition in a temperature range between 30 and 80 degrees C. On the other hand, in the DSC thermogram for the complex at pH 4, a peak appears at 65.8 degrees C in the first heating scan. In the successive second heating scan, a transition peak appears at 63.5 degrees C. In connection with the DSC results, the structural changes associated with these phase transitions were studied with temperature-scan x-ray diffraction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
We report on the effects of temperature and pressure on the structure, conformation and phase behavior of aqueous dispersions of the model lipid "raft" mixture palmitoyloleoylphosphatidylcholine (POPC)/bovine brain sphingomyelin (SM)/cholesterol (Chol) (1:1:1). We investigated interchain interactions, hydrogen bonding, conformational and structural properties as well as phase transformations of this system using Fourier transform-infrared (FT-IR) spectroscopy, small-angle X-ray scattering (SAXS), differential scanning calorimetry (DSC) coupled with pressure perturbation calorimetry (PPC), and Laurdan fluorescence spectroscopy. The IR spectral parameters in combination with the scattering patterns from the SAXS measurements were used to detect structural and conformational transformations upon changes of pressure up to 7-9 kbar and temperature in the range from 1 to about 80 degrees C. The generalized polarization function (GP) values, obtained from the Laurdan fluorescence spectroscopy studies also reveal temperature and pressure dependent phase changes. DSC and PPC were used to detect thermodynamic properties accompanying the temperature-dependent phase changes. In combination with literature fluorescence spectroscopy and microscopy data, a tentative p,T stability diagram of the mixture has been established. The data reveal a broad liquid-order/solid-ordered (lo+so) two-phase coexistence region below 8+/-2 degrees C at ambient pressure. With increasing temperature, a lo+ld+so three-phase region is formed, which extends up to approximately 27 degrees C, where a liquid-ordered/liquid-disordered (lo+ld) immiscibility region is formed. Finally, above 48+/-2 degrees C, the POPC/SM/Chol (1:1:1) mixture becomes completely fluid-like (liquid-disordered, ld). With increasing pressure, all phase transition lines shift to higher temperatures. Notably, the lo+ld (+so) phase coexistence region, mimicking raft-like lateral phase separation in natural membranes, extends over a rather wide temperature range of about 40 degrees C, and a pressure range, which extends up to about 2 kbar for T=37 degrees C. Interestingly, in this pressure range, ceasing of membrane protein function in natural membrane environments has been observed for a variety of systems.  相似文献   

11.
We studied the effects of calcium ion concentration on the temperature dependence of rheological behavior of human red blood cells (RBCs) and concentrated hemoglobin solutions. Our previous study (G. M. Artmann, C. Kelemen, D. Porst, G. Büldt, and S. Chien, 1998, Biophys. J., 75:3179-3183) showed a critical temperature (Tc) of 36.4 +/- 0.3 degrees C at which the RBCs underwent a transition from non-passage to passage through 1.3 microm micropipettes in response to an aspiration pressure of -2.3 kPa. An increase in intracellular Ca2+ concentration by using the ionophore A23187 reduced the passability of intact RBCs through small micropipettes above T(c); the micropipette diameter needed for >90% passage increased to 1.7 microm. Viscometry of concentrated hemoglobin solutions (45 and 50 g/dl) showed a sudden viscosity transition at 36 +/- 1 degrees C (Tc(eta)) at all calcium concentrations investigated. Below Tc(eta), the viscosity value of the concentrated hemoglobin solution at 1.8 mM Ca(2+) was higher than that at other concentrations (0.2 microM, 9 mM, and 18 mM). Above Tc(eta), the viscosity was almost Ca2+ independent. At 1.8 mM Ca2+ and 36 +/- 1 degrees C, the activation energy calculated from the viscometry data showed a strong dependence on the hemoglobin concentration. We propose that the transition of rheological behavior is attributable to a high-to-low viscosity transition mediated by a partial release of the hemoglobin-bound water.  相似文献   

12.
S Wakabayashi  M Shigekawa 《Biochemistry》1990,29(31):7309-7318
The mechanism for activation of sarcoplasmic reticulum ATPase by Ca2+ was investigated in 2 mM MgCl2 and 0.1 M KCl at pH 6.5 and 11 degrees C by using enzyme preparations in which a specific amino acid residue (Cys-344) was labeled with 4-nitrobenzo-2-oxa-1,3-diazole (NBD) [Wakabayashi, S., Imagawa, T., & Shigekawa, M. (1990) J. Biochem. (Tokyo) 107, 563-571]. We compared the kinetics of binding and dissociation of Ca2+ from the enzyme with those of the accompanying NBD fluorescence changes. The fluorescence rise following addition of Ca2+ proceeded monoexponentially. At 2-100 microM Ca2+ and in the absence of nucleotides, the Ca2(+)-induced fluorescence rise and Ca2+ binding to the enzyme proceeded at similar rates, which were almost independent of the Ca2+ concentration. In contrast, the fluorescence decrease induced by Ca2+ removal was slower than the Ca2+ dissociation, and both of these processes were inhibited markedly by increasing medium Ca2+. ATP by binding at 1 mol/mol of the phosphorylation site markedly accelerated both the Ca2(+)-induced fluorescence rise and Ca2+ binding, ADP and AMPPNP but not GTP also being effective. In contrast, ADP minimally affected the NBD fluorescence decrease and the Ca2+ dissociation. These data are consistent with a reaction model in which binding of Ca2+ occurs after the conformational transition of the free enzyme from a state (E2) having low affinity for Ca2+ to one (E1) having high affinity for Ca2+ and in which ATP bound at the catalytic site of E2, whose affinity for ATP is about 30-fold less than that of E1, accelerates this conformational transition.  相似文献   

13.
We have investigated the temperature dependence of the fusion of phospholipid vesicles composed of pure bovine brain phosphatidylserine (PS) induced by Ca2+ or Mg2+. Aggregation of the vesicles was monitored by 90 degrees light-scattering measurements, fusion by the terbium/dipicolinic acid assay for mixing of internal aqueous volumes, and release of vesicle contents by carboxyfluorescein fluorescence. Membrane fluidity was determined by diphenylhexatriene fluorescence polarization measurements. Small unilamellar vesicles (SUV, diameter 250 A) or large unilamellar vesicles (LUV, diameter 1000 A) were used, and the measurements were done in 0.1 M NaCl at pH 7.4. The following results were obtained: (1) At temperatures (0-5 degrees C) below the phase transition temperature (Tc) of the lipid, LUV (PS) show very little fusion in the presence of Ca2+, although vesicle aggregation is rapid and extensive. With increasing temperature, the initial rate of fusion increases dramatically. Leakage of contents at the higher temperatures remains limited initially, but subsequently complete release occurs as a result of collapse of the internal aqueous space of the fusion products. (2) SUV (PS) are still in the fluid state down to 0 degree C, due to the effect of bilayer curvature, and fuse rapidly in the entire temperature range from 0 to 35 degrees C in the presence of Ca2+. The initial rate of leakage is low relative to the rate of fusion. At higher temperatures (15 degrees C and above), subsequent collapse of the vesicles' internal space causes complete release.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Neutral phospholipase A2 activity, which hydrolyzed phosphatidylcholine and phosphatidylethanolamine with the same efficiency, was identified in the nuclear matrix prepared from purified nuclei of rat ascites hepatoma cells (AH 7974). The enzyme activity was optimal at pH 7.0 and required Ca2+ absolutely. Concentrations of Ca2+ for a maximal and a half-maximal activation were 1.10(-2) and 1.10(-3) M, respectively, and little activity was detected at Ca2+ concentrations lower than 1.10(-5) M. Addition of acidic phospholipids markedly stimulated the enzyme activity, and further, lowered the minimum Ca2+ concentration required for activation. In particular, the polyphosphoionositides phosphatidylinositol 4-monophosphate and 4,5-diphosphate were most effective. These two polyphosphoinositides lowered the Ca2+ concentration required for half-maximal activation to 10(-5) M and dramatically stimulated the activity at that Ca2+ concentration (greater than 30-fold). The neutral phospholipase A2 activity such as characterized in the present study was very low in the other subcellular fractions including mitochondria, microsome, plasma membrane and cytosol.  相似文献   

15.
T F Busby  K C Ingham 《Biochemistry》1987,26(17):5564-5571
Fluorescent probes and other methods have been used to investigate the thermal stability of activated C1r and functionally intact fragments isolated from tryptic digests of the protein. This enzyme exhibits two irreversible transitions that differ with respect to their sensitivity to metal ions. The high-temperature transition occurs with a midpoint near 53 degrees C in 0.02 M tris(hydroxymethyl)aminomethane buffer and 0.15 M NaCl, pH 7.4. It is relatively insensitive to Ca2+ and ionic strength and is accompanied by a loss of catalytic activity. The low-temperature transition is most easily observed in the presence of ethylenediaminetetraacetic acid and is completely abolished by 100 microM Ca2+. Its midpoint varies between 26 degrees C at low ionic strength and 40 degrees C in the presence of 0.5 M NaCl. The low-temperature transition results in extensive polymerization of the protein without loss of the esterolytic activity or the ability to react with C1 inhibitor; however, the ability to reconstitute hemolytically active C1 or even bind to C1s in the presence of Ca2+ is destroyed. A highly purified N-terminal fragment generated by tryptic digestion of C1r in the presence of Ca2+ retained its ability to interact with C1s, disrupting the formation of C1s dimers in the presence of Ca2+. In the absence of Ca2+, this fragment displays only a low-temperature transition that is very similar to the one observed with the whole protein and that destroys its ability to bind to C1s. Addition of Ca2+ stabilizes this fragment, shifting the midpoint of its melting transition upward by more than 20 degrees C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The size and size distribution of unilamellar phospholipid vesicles present in unsonicated phosphatidic acid and mixed phosphatidic acid/phosphatidylcholine dispersions were determined by gel filtration, quasi-elastic light scattering and freeze-fracture electron microscopy. The vesiculation in these dispersions was induced by a transient increase in pH as described previously (Hauser, H. and Gains, N. (1982) Proc. Natl. Acad. Sci. USA 79, 1683–1687). The resulting phospholipid dispersions are heterogeneous consisting of small unilamellar vesicles (average radius r < 50 nm) and large unilamellar vesicles (average r ranging from about 50 to 500 nm). The smallest vesicles with r = 11 ± 2 nm are observed with dispersions of pure phosphatidic acid, the population of these vesicles amounting to about 80% of the total lipid. With increasing phosphatidylcholine content the radius of the small unilamellar vesicles increases and at the same time the population of small unilamellar vesicles decreases. The average radius of small unilamellar vesicles present in phosphatidic acid/phosphatidylcholine dispersions (mole ratio, 1:1) is 17.5 ± 2 nm, the population of these vesicles amounting to about 70% of the total lipid. By a combination of gel filtration, quasi-elastic light scattering and freeze-fracture electron microscopy it was possible to characterize the large unilamellar vesicles. This population is heterogeneous with its mean radius also increasing with increasing phosphatidylcholine content. After separating the large unilamellar vesicles from small unilamellar vesicles on Sepharose 4B it can be shown by quasi-elastic light scattering that in pure phosphatidic acid dispersions 80–90% of the large unilamellar vesicle population consist of vesicles with a mean radius of 170 nm. In mixed phosphatidic acid/phosphatidylcholine dispersions this radius increases to about 265 nm as the phosphatidylcholine content is raised to 90 mol%.  相似文献   

17.
S Wakabayashi  T Ogurusu  M Shigekawa 《Biochemistry》1990,29(47):10613-10620
The binding of Ca2+ to 4-nitro-2,1,3-benzoxadiazole (NBD)-labeled sarcoplasmic reticulum Ca2(+)-ATPase was accelerated markedly when the pH was changed at 11 degrees C from 6.5 to 8.0 at the time of Ca2+ addition. We examined the effect of pH on the enzyme conformational transition by measuring the kinetics of NBD fluorescence rises induced by a pH jump under various ligand conditions. The fast fluorescence rise following a pH jump from 6.0 or 6.5 to various test pHs in the presence and absence of Ca2+ proceeded monoexponentially. The amplitude of this fluorescence rise in the presence of Ca2+ was independent of the test pH, whereas the observed rate constant (kobs) increased markedly as the test pH increased. In contrast, the amplitude of the fast fluorescence rise in the absence of Ca2+ increased with increasing test pH, whereas kobs decreased. MgATP or Mg2+ influenced the pH dependences of these parameters in a complex way except for the amplitudes measured in the presence of Ca2+. These data could be simulated by using a reaction model in which Ca2+ binding is preceded by a rate-limiting enzyme conformational transition from a low to a high NBD fluorescence state and 1 mol each of H+ is liberated before and after this conformational transition. MgATP or Mg2+ appeared to promote this conformational transition by enhancing deprotonation of the enzyme. These results suggest that deprotonation may be the primary event in the activation of the unphosphorylated enzyme by Ca2+.  相似文献   

18.
The interaction of Ca2+ and Mg2+ with phosphatidylserine (PS) vesicles in 0.1 M NaCl aqueous solution was studied by equilibrium dialysis binding, X-ray diffraction, batch microcalorimetry, kinetics of cation-induced vesicle aggregation, release of vesicle contents, and fusion. Addition of either cation causes aggregation of PS vesicles and produces complexes with similar stoichiometry (1:2 cation/PS) at saturating concentrations, although the details of the interactions and the resulting complexes are quite different. Addition of Ca2+ to PS vesicles at T greater than or equal to 25 degrees C induces the formation of an "anhydrous" complex of closely apposed membranes with highly ordered crystalline acyl chains and a very high transition temperature (Tc greater than 100 degrees C). The formation of this complex is accompanied by a release of heat (5.5 kcal/mol), rapid release of vesicle contents, and fusion of the vesicles into larger membranous structures. By contrast, addition of Mg2+ produces a complex with PS which is much more hydrated, has no crystallization of the acyl chains at T greater than or equal to 20 degrees C, and has comparatively little fusion. Studies with both Ca2+ and Mg2+ added simultaneously indicate that there is a synergistic effect between the two cations, which results in an enhancement of the ability of Ca2+ to form its specific complex with PS at lower concentrations. The presence of the erythrocyte protein "spectrin" inhibits this synergism and interferes with the formation of the specific PS/Ca complex. It also inhibits the fusion of PS vesicles. It is proposed that the unique PS/Ca complex, which involves close apposition of vesicle membranes, is an intermembrane "trans" complex. We further propose that such a complex is a key step for the resultant phase transition and fusion of PS vesicles. By contrast, the PS/Mg complex is proposed to be a "cis" complex with respect to each membrane. The results are discussed in terms of the mechanism of membrane fusion.  相似文献   

19.
Human placental anticoagulant protein-I (PAP-I) is a member of the lipocortin/calpactin/annexin family of Ca2+-dependent phospholipid binding proteins. PAP-I was labeled with fluorescein 5-isothiocyanate (1 mol/mol); this derivative had anticoagulant activity identical to the unlabeled protein and could be used to measure Ca2+-dependent binding to phospholipid vesicles through changes in fluorescence quenching. At 1.2 mM Ca2+, 0.50 M ionic strength, pH 7.4, 25 degrees C, fluorescein-labeled PAP-I bound to phospholipid vesicles containing 80% phosphatidylcholine, 20% phosphatidylserine with a Kd of 1.2 +/- 0.2 nM (mean +/- S.D.). At an ionic strength of 0.15 M, the Kd decreased to less than 0.1 nM. Prothrombin and factor Xa both competed with fluorescein-labeled PAP-I for binding to anionic phospholipid vesicles, but with affinities at least 1000-fold weaker than PAP-I. PAP-I bound only weakly (Kd greater than 2 x 10(-5) M) to neutral or anionic phospholipid monomers, and this binding was not calcium-dependent. These results show that the affinity of PAP-I for anionic phospholipid surfaces is sufficient to explain its potency as an in vitro anticoagulant.  相似文献   

20.
Thermal denaturation of porcine pancreatic elastase was studied by difference spectrophotometry. At 293 nm, and pH 8.0, the thermal transition of elastase occurs with a midpoint temperature (Tm) of (58.0 +/- 0.5) degrees C. Mg2+ and Ca2+ stabilize the native form in increasing the midpoint temperature of the transition, Ca2+ being more effective than Mg2+ in the 0-0.02 M concentration range. Furthermore, Ca2+ protects pancreatic elastase against the destabilizing effect of Cu2+. Whatever be the temperature between 40 degrees C and 55 degrees C, Ca2+ protects pancreatic elastase against loss of enzymatic activity.  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