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 617 毫秒
1.
1. Amphoteric electrolytes form salts with both acids and alkalies. It is shown for two amphoteric electrolytes, Al(OH)3 and gelatin, that in the presence of an acid salt water diffuses through a collodion membrane into a solution of these substances as if its particles were negatively charged, while water diffuses into solutions of these electrolytes, when they exist as monovalent or bivalent metal salts, as if the particles of water were positively charged. The turning point for the sign of the electrification of water seems to be near or to coincide with the isoelectric point of these two ampholytes which is a hydrogen ion concentration of about 2 x 10–5 N for gelatin and about 10–7 for Al(OH)3. 2. In conformity with the rules given in a preceding paper the apparently positively charged water diffuses with less rapidity through a collodion membrane into a solution of Ca and Ba gelatinate than into a solution of Li, Na, K, or NH4 gelatinate of the same concentration of gelatin and of hydrogen ions. Apparently negatively charged water diffuses also with less rapidity through a collodion membrane into a solution of gelatin sulfate than into a solution of gelatin chloride or nitrate of the same concentration of gelatin and of hydrogen ions. 3. If we define osmotic pressure as that additional pressure upon the solution required to cause as many molecules of water to diffuse from solution to the pure water as diffuse simultaneously in the opposite direction through the membrane, it follows that the osmotic pressure cannot depend only on the concentration of the solute but must depend also on the electrostatic effects of the ions present and that the influence of ions on the osmotic pressure must be the same as that on the initial velocity of diffusion. This assumption was put to a test in experiments with gelatin salts for which a collodion membrane is strictly semipermeable and the tests confirmed the expectation.  相似文献   

2.
The effects of negatively charged and neutral lipids on the function of the reconstituted nicotinic acetylcholine receptor from Torpedo californica were determined with two assays using acetylcholine receptor-containing vesicles: the ion flux response and the affinity-state transition. The receptor was reconstituted into three different lipid environments, with and without neutral lipids: (1) phosphatidylcholine/phosphatidylserine; (2) phosphatidylcholine/phosphatidic acid; and (3) phosphatidylcholine/cardiolipin. Analysis of the ion flux responses showed that: (1) all three negatively charged lipid environments gave fully functional acetylcholine receptor ion channels, provided neutral lipids were added; (2) in each lipid environment, the neutral lipids tested were functionally equivalent to cholesterol; and (3) the rate of receptor desensitization depends upon the type of neutral lipid and negatively charged phospholipid reconstituted with the receptor. The functional effects of neutral and negatively charged lipids on the acetylcholine receptor are discussed in terms of protein-lipid interactions and stabilization of protein structure by lipids.  相似文献   

3.
The pores formed by Bacillus thuringiensis insecticidal toxins have been shown to allow the diffusion of a variety of monovalent cations and anions and neutral solutes. To further characterize their ion selectivity, membrane permeability induced by Cry1Aa and Cry1Ac to amino acids (Asp, Glu, Ser, Leu, His, Lys and Arg) and to divalent cations (Mg2+, Ca2+ and Ba2+) and anions (SO42− and phosphate) was analyzed at pH 7.5 and 10.5 with midgut brush border membrane vesicles isolated from Manduca sexta and an osmotic swelling assay. Shifting pH from 7.5 to 10.5 increases the proportion of the more negatively charged species of amino acids and phosphate ions. All amino acids diffused well across the toxin-induced pores, but, except for aspartate and glutamate, amino acid permeability was lower at the higher pH. In the presence of either toxin, membrane permeability was higher for the chloride salts of divalent cations than for the potassium salts of divalent anions. These results clearly indicate that the pores are cation-selective.  相似文献   

4.
1. The osmotic effects on phospholipid vesicles in the presence and absence of bound glycoprotein and protein were used to determine the permeability to some ions and neutral molecules. 2. The permeation times were registered as a function of membrane surface charge and glycoprotein protein molar ratio in the vesicles. 3. The permeation times for ions varied considerably with the change of glycoprotein protein molar ratio. 4. When neutral molecules passed through the membrane, the permeation times did not change varying glycoprotein-protein molar ratio. 5. In all cases the permeation times did not depend on the surface charge. 6. Our results suggest that topographical distribution of hydrophobic and hydrophilic regions of glycoprotein molecule would play a substantial role and influence the permeability in this case.  相似文献   

5.
The kinetics of the partitioning of lipid vesicles containing acidic phospholipids in aqueous two-phase polymer systems are dependent upon the vesicle size; the larger the vesicles, the more readily they absorb to the interfaces between the two polymer phases and hence are cleared from the top phase as phase separation proceeds. The partitioning of neutral lipid vesicles is principally to the bulk interface and is the same in phase systems of both low and high electrostatic potential difference between the two phases (delta psi). The incorporation of negatively charged lipids has two effects upon partition. First, vesicles with negatively charged lipids exhibit increased bottom phase partitioning in phases of low delta psi due to an enhanced wetting of the charged lipids by the lower phase. Second, the presence of a negatively charged group on the vesicle surface results in increased partition to the interface and top phase in phase systems of high delta psi. Differences observed in the partition of vesicles containing various species of negatively charged lipid thus reflect a competition between these two opposing factors.  相似文献   

6.
The protein-mediated phospholipid exchange between small unilamellar vesicles was investigated by fluorescence polarization measurements with diphenylhexatriene as optical probe. Thermotropic phase-transition measurements were taken after mixing two vesicle preparations of distinct and different phase-transition temperatures or having different states of charge. From the heights of each phase-transition step, we were able to follow the lipid-exchange process in the presence, as well as in the absence (natural exchange), of so-called transfer protein isolated from beef liver. A strong enhancement of the lipid transfer was observed at the corresponding lipid-phase-transition temperature, which is explained by the presence of fluctuating fluid and ordered domains co-existing at the lipid-phase-transition temperature. A unidirectional lipid transfer of the neutral component was observed between negatively charged phosphatidic acid and neutral phosphatidylcholine vesicles. Fluorescence polarization measurements showed the disappearance of the phosphatidylcholine phase transition, whereas the phosphatidic acid phase transition broadened and its phase transition temperature became lower.  相似文献   

7.
To better understand bilayer property dependency on lipid electrostatics and headgroup size, we use atomistic molecular dynamics simulations to study negatively charged and neutral lipid membranes. We compare the negatively charged phosphatidic acid (PA), which at physiological pH and salt concentration has a negative spontaneous curvature, with the negatively charged phosphatidylglycerol (PG) and neutrally charged phosphatidylcholine (PC), both of which have zero spontaneous curvature. The PA lipids are simulated using two different sets of partial charges for the headgroup and the varied charge distribution between the two PA systems results in significantly different locations for the Na+ ions relative to the water/membrane interface. For one PA system, the Na+ ions are localized around the phosphate group. In the second PA system, the Na+ ions are located near the ester carbonyl atoms, which coincides with the preferred location site for the PG Na+ ions. We find that the Na+ ion location has a larger effect on bilayer fluidity properties than lipid headgroup size, where the Alipid and acyl chain order parameter values are more similar between the PA and PG bilayers that have Na+ ions located near the ester groups than between the two PA bilayers.  相似文献   

8.
N Poklar  J Fritz  P Macek  G Vesnaver  T V Chalikian 《Biochemistry》1999,38(45):14999-15008
The interactions of equinatoxin II (EqTxII) with zwitterionic (DPPC) and anionic (DPPG) phospholipids and an equimolar mixture of the two phospholipids (DPPC/DPPG) have been investigated by differential scanning calorimetry (DSC), CD-spectropolarimetry, intrinsic emission fluorescence spectroscopy, and ultrasonic velocimetry. EqTxII binds to small unilamellar vesicles formed from negatively charged DPPG lipids, causing a marked reduction in the cooperativity and enthalpy of their gel/liquid-crystalline phase transition. This transition is completely abolished at a lipid-to-protein ratio, L/P, of 10. For the mixed DPPC/DPPG vesicles, a 2-fold greater lipid-to-protein ratio (L/P = 20) is required to abolish the phase transition, which corresponds to the same negative charge (-10) of lipid molecules per EqTxII molecule. The disappearance of the phase transition of the lipids apparently corresponds to the precipitation of the lipid-protein complex, as suggested by our sound velocity measurements. Based on the far-UV CD spectra, EqTxII undergoes two structural transitions in the presence of negatively charged vesicles (DPPG). The first transition coincides with the gel/liquid-crystalline phase transition of the lipids, which suggests that the liquid-crystalline form of negatively charged lipids triggers structural changes in EqTxII. The second transition involves the formation of alpha-helical structure. Based on these observations, we propose that, in addition to electrostatic interactions, hydrophobic interactions play an important role in EqTxII-membrane association.  相似文献   

9.
The interactions of the assembly factor P17 of bacteriophage PRD1 with liposomes were investigated by static light scattering, fluorescence spectroscopy, and differential scanning calorimetry. Our data show that P17 binds to positively charged large unilamellar vesicles composed of the zwitterionic 1-palmitoyl-2-oleoyl-phosphatidylcholine and sphingosine, whereas only a weak interaction is evident for 1-palmitoyl-2-oleoyl-phosphatidylcholine vesicles. P17 does not bind to negatively charged membranes composed of 1-palmitoyl-2-oleoyl-phosphatidylglycerol and 1-palmitoyl-2-oleoyl-phosphatidylcholine. Our differential scanning calorimetry results reveal that P17 slightly perturbs the phase behaviour of neutral phosphatidylcholine and negatively charged multilamellar vesicles. In contrast, the phase transition temperature of positively charged dimyristoylphosphatidylcholine/sphingosine multilamellar vesicles (molar ratio 9 : 1, respectively) is increased by approximately 2.4 degrees C and the half width of the enthalpy peak broadened from 1.9 to 5.6 degrees C in the presence of P17 (protein : lipid molar ratio 1 : 47). Moreover, the enthalpy peak is asymmetrical, suggesting that lipid phase separation is induced by P17. Based on the far-UV CD spectra, the alpha-helicity of P17 increases upon binding to positively charged micelles composed of Triton X-100 and sphingosine. We propose that P17 can interact with positively charged lipid membranes and that this binding induces a structural change on P17 to a more tightly packed and ordered structure.  相似文献   

10.
Summary The volume change of sarcoplasmic reticulum vesicles was followed by measuring the light scattering intensity. When the salt concentration of the suspension of sarcoplasmic reticulum vesicles was increased by using a stopped flow apparatus, the light scattering intensity rapidly increased at the beginning and then decreased. The fast increase in the light scattering intensity is caused by the decrease of the volume of sarcoplasmic reticulum vesicles due to the outflow of water. The following decrease in the light scattering intensity is caused by the increase of the volume due to the inflow of the solutes and water. From the former and the latter rates, the permeation times of water and the solutes could be calculated, respectively. According to the same method, permeation times of various salts were determined. The rate of the inflow of the salts was dependent on the movement of the slower ions, that is, ions move as a pair.In the case of potassium salts, an increase in the permeation rate of the salts was observed when valinomycin was added to the membrane suspensions. From these experiments, as a measure of permeability, half permeation times of various ions and molecules were determined. The following are typical results: water 0.1, Li+ 36, Na+ 26, K+ 20, Rb+ 16, Cl 0.4, methanesulfonate 20, phosphate 10.5, oxalate 40 in seconds at room temperature. As a whole, sarcoplasmic reticulum was found to be an anion permeable membrane.  相似文献   

11.
We have compared hydration forces, electrical dipole potentials, and structural parameters of dispersions of dipalmitoylphosphatidylcholine (DPPC) and dihexadecylphosphatidylcholine (DHPC) to evaluate the influence of fatty acid carbonyl groups on phospholipid bilayers. NMR and x-ray investigations performed over a wide range of water concentrations in the samples show, that in the liquid crystalline lamellar phase, the presence of carbonyl groups is not essential for lipid structure and hydration. Within experimental error, the two lipids have identical repulsive hydration forces between their bilayers. The higher transport rate of the negatively charged tetraphenylboron over the positively charged tetraphenylarsonium indicates that the dipole potential is positive inside the membranes of both lipids. However, the lack of fatty acid carbonyl groups in the ether lipid DHPC decreased the potential by (118 +/- 15) mV. By considering the sign of the potential and the orientation of carbonyl groups and headgroups, we conclude that the first layer of water molecules at the lipid water interface makes a major contribution to the dipole potential.  相似文献   

12.
The pores formed by Bacillus thuringiensis insecticidal toxins have been shown to allow the diffusion of a variety of monovalent cations and anions and neutral solutes. To further characterize their ion selectivity, membrane permeability induced by Cry1Aa and Cry1Ac to amino acids (Asp, Glu, Ser, Leu, His, Lys and Arg) and to divalent cations (Mg(2+), Ca(2+) and Ba(2+)) and anions (SO(4)(2-) and phosphate) was analyzed at pH 7.5 and 10.5 with midgut brush border membrane vesicles isolated from Manduca sexta and an osmotic swelling assay. Shifting pH from 7.5 to 10.5 increases the proportion of the more negatively charged species of amino acids and phosphate ions. All amino acids diffused well across the toxin-induced pores, but, except for aspartate and glutamate, amino acid permeability was lower at the higher pH. In the presence of either toxin, membrane permeability was higher for the chloride salts of divalent cations than for the potassium salts of divalent anions. These results clearly indicate that the pores are cation-selective.  相似文献   

13.
The interaction and mixing of membrane components in sonicated unilamellar vesicles and also non-sonicated multilamellar vesicles prepared from highly purified phospholipids suspended in NaCl solutions has been examined. Electron microscopy and differential scanning calorimetry were used to characterize the extent and kinetics of mixing of membrane components between different vesicle populations. No appreciable fusion was detected between populations of non-sonicated phospholipid vesicles incubated in aqueous salt (NaCl) solutions. Mixing of vesicle membrane components via diffusion of phospholipid molecules between vesicles was observed in populations of negatively charged phosphatidylglycerol vesicles but similar exchange diffusion was not detected in populations of neutral phosphatidylcholine vesicles. Incubation of sonicated vesicle populations at temperatures close to or above the phospholipid transition temperature resulted in an increase in vesicle size and mixing of vesicle membrane components as determined by a gradual change in the thermotropic properties of the mixed vesicle population. The interaction of purified phospholipid vesicles was also examined in the presence of myristic acid and lysolecithin. Our results indicate that while these agents enhance mixing of vesicle membrane components, in most cases mixing probably proceeds via diffusion of phospholipid molecules rather than by fusion of entire vesicles. Increased mixing of vesicle membrane components was also produced when vesicles were prepared containing a purified hydrophobic protein (myelin proteolipid apoprotein) or were incubated in the presence of dimethylsulfoxide. In these two systems, however, the evidence suggests that mixing of membrane components results from the fusion of entire vesicles.  相似文献   

14.
M Rafalski  J D Lear  W F DeGrado 《Biochemistry》1990,29(34):7917-7922
Peptides representing the N-terminal 23 residues of the surface protein gp41 of LAV1a and LAVmal strains of the human immunodeficiency virus were synthesized and their interactions with phospholipid vesicles studied. The peptides are surface-active and penetrate lipid monolayers composed of negatively charged but not neutral lipids. Similarly, the peptides induce lipid mixing and solute (6-carboxyfluorescein) leakage of negatively charged, but not neutral, vesicles. Circular dichroism and infrared spectroscopy show that at low peptide:lipid ratios (approximately 1:200), the peptides bind to negatively charged vesicles as alpha-helices. At higher peptide:lipid ratios (1:30), a beta conformation is observed for the LAV1a peptide, accompanied by a large increase in light scattering. The LAVmal peptide showed less beta-structure and induced less light scattering. With neutral vesicles, only the beta conformation and a peptide:lipid ratio-dependent increase in vesicle suspension light scattering were observed for both peptides. We hypothesize that the inserted alpha-helical form causes vesicle membrane disruption whereas the surface-bound beta form induces aggregation.  相似文献   

15.
We have performed molecular dynamics simulations to investigate the structure and dynamics of charged bilayers as well as the distribution of counterions at the bilayer interface. For this, we have considered the negatively charged di-myristoyl-phosphatidyl-glycerol (DMPG) and di-myristoyl-phosphatidyl-serine (DMPS) bilayers as well as a protonated di-myristoyl-phosphatidyl-serine (DMPSH) bilayer. We were particularly interested in calcium ions due to their important role in biological systems. Simulations performed in the presence of calcium ions (DMPG, DMPS) or sodium ions (DMPS) were run for 45-60 ns. Simulation results for DMPG are compared with fluorescence measurements. The average areas per molecule were 47.4 ± 0.5 Å2 (DMPG with calcium), 47.3 ± 0.5 Å2 (DMPS with calcium), 51.3 ± 1.0 Å2 (DMPS with sodium) and 45.3 ± 0.5 Å2 (DMPSH). The structure of the negatively charged lipids is significantly affected by the counterions, where calcium ions have a more pronounced effect than sodium ions. Calcium ions were found to be tightly bound to the anionic groups of the lipid molecules and as such appear to constitute an integral part of the membrane interface on nanoseconds time scales. In contrast to sodium ions, calcium ions are localised in a narrow (∼ 10 Å) band around the phosphate group. The interaction of calcium with the lipid molecules enhances the molecular packing of the PG and PS lipids. This observation is in good agreement with emission spectra of the membrane partitioning probe Laurdan in DMPG multilamellar vesicles that indicate an increase in the ordering of the DMPG bilayer due to the presence of calcium. Our results indicate that calcium ions, which often function as a second messengers in living cells have a pronounced effect on membrane structures, which may have implications during signal transduction events.  相似文献   

16.
The 53-kDa insulin receptor substrate protein (IRSp53) organizes the actin cytoskeleton in response to stimulation of small GTPases, promoting the formation of cell protrusions such as filopodia and lamellipodia. IMD is the N-terminal 250 amino acid domain (IRSp53/MIM Homology Domain) of IRSp53 (also called I-BAR), which can bind to negatively charged lipid molecules. Overexpression of IMD induces filopodia formation in cells and purified IMD assembles finger-like protrusions in reconstituted lipid membranes. IMD was shown by several groups to bundle actin filaments, but other groups showed that it also binds to membranes. IMD binds to negatively charged lipid molecules with preference to clusters of PI(4,5)P2. Here, we performed a range of different in vitro fluorescence experiments to determine the binding properties of the IMD to phospholipids. We used different constructs of large unilamellar vesicles (LUVETs), containing neutral or negatively charged phospholipids. We found that IMD has a stronger binding interaction with negatively charged PI(4,5)P2 or PS lipids than PS/PC or neutral PC lipids. The equilibrium dissociation constant for the IMD–lipid interaction falls into the 78–170 μM range for all the lipids tested. The solvent accessibility of the fluorescence labels on the IMD during its binding to lipids is also reduced as the lipids become more negatively charged. Actin affects the IMD–lipid interaction, depending on its polymerization state. Monomeric actin partially disrupts the binding, while filamentous actin can further stabilize the IMD–lipid interaction.  相似文献   

17.
Electrophoretic mobility data of SR vesicles reconstituted with uncharged and two mixtures of charged and uncharged lipids (Brethes, D., Dulon, D., Johannin, G., Arrio, B., Gulik-Krzywicki, T., Chevallier, J. 1986. Study of the electrokinetic properties of reconstituted sarcoplasmic reticulum vesicles. Arch. Biochem. Biophys. 246:355–356) were analyzed in terms of four models of the membrane-water interface: (I) a smooth, negatively charged surface; (II) a negatively charged surface of lipid bilayer covered with an electrically neutral surface frictional layer; (III) an electrically neutral lipid bilayer covered with a neutral frictional layer containing a sheet of negative charge at some distance above the surface of the bilayer; (IV) an electrically neutral lipid bilayer covered with a homogeneously charged frictional layer. The electrophoretic mobility was predicted from the numerical integration of Poisson-Boltzmann and Navier-Stokes equations. Experimental results were consistent only with predictions based on Model-III with charged sheet about 4 nm above the bilayer and frictional layer about 10 nm thick. Assuming that the charge of the SR membrane is solely due to that on Ca++-ATPase pumps, the dominant SR protein, the mobility data of SR and reconstituted SR vesicles are consistent with 12 electron charges/ATPase. This value compares well to the net charge of the cytoplasmic portion of ATPase estimated from the amino acid sequence (-11e). The position of the charged sheet suggests that the charge on the ATPase is concentrated in the middle of the cytoplasmic portion. The frictional layer of SR can be also assigned to the cytoplasmic portion of Ca++-ATPase. The layer has been characterized with hydrodynamic shielding length of 1.1 nm. Its thickness is comparable to the height of the cytoplasmic portion of Ca++-ATPase. Received: 15 June 1998/Revised: 8 October 1998  相似文献   

18.

Background

Awad et al. [1] reported on the Ca2+-induced transitions of dioleoyl-phosphatidylglycerol (DOPG)/monoolein (MO) vesicles to bicontinuous cubic phases at equilibrium conditions. In the present study, the combination of rapid mixing and time-resolved synchrotron small-angle X-ray scattering (SAXS) was applied for the in-situ investigations of fast structural transitions of diluted DOPG/MO vesicles into well-ordered nanostructures by the addition of low concentrated Ca2+ solutions.

Methodology/Principal Findings

Under static conditions and the in absence of the divalent cations, the DOPG/MO system forms large vesicles composed of weakly correlated bilayers with a d-spacing of ∼140 Å (Lα-phase). The utilization of a stopped-flow apparatus allowed mixing these DOPG/MO vesicles with a solution of Ca2+ ions within 10 milliseconds (ms). In such a way the dynamics of negatively charged PG to divalent cation interactions, and the kinetics of the induced structural transitions were studied. Ca2+ ions have a very strong impact on the lipidic nanostructures. Intriguingly, already at low salt concentrations (DOPG/Ca2+>2), Ca2+ ions trigger the transformation from bilayers to monolayer nanotubes (inverted hexagonal phase, H2). Our results reveal that a binding ratio of 1 Ca2+ per 8 DOPG is sufficient for the formation of the H2 phase. At 50°C a direct transition from the vesicles to the H2 phase was observed, whereas at ambient temperature (20°C) a short lived intermediate phase (possibly the cubic Pn3m phase) coexisting with the H2 phase was detected.

Conclusions/Significance

The strong binding of the divalent cations to the negatively charged DOPG molecules enhances the negative spontaneous curvature of the monolayers and causes a rapid collapsing of the vesicles. The rapid loss of the bilayer stability and the reorganization of the lipid molecules within ms support the argument that the transition mechanism is based on a leaky fusion of the vesicles.  相似文献   

19.
Planar asymmetric bilayer membranes, formed by apposing a monolayer of the neutral lipid glyceroldioleate (GDO) with one of the negatively charged lipid oleyl acid phosphate (OAP), were used to measure the rate of transmembrane OAP migration. The assay for this lipid flip-flop was the interaction of Ca2+ ions with negatively charged lipids which causes membranes to break: when Ca2+ is added to the compartment limited initially by the neutral lipid, flip-flop of the charged lipid eventually results in membrane breakdown. At 22 ± 2°C, in the absence of an externally applied electric field, an upper limit to the half time of OAP flip-flop was measured as 18.7 h, with a tentative lower limit of 14.4 h.  相似文献   

20.
The interaction of Sendai virus with small, unilamellar vesicles, lacking virus receptors and loaded with self-quenched 6-carboxyfluorescein, was studied. Sendai virions induced release of carboxyfluorescein from vesicles composed of negative charged phospholipids, despite the fact that they did not contain virus receptors. Preliminary experiments indicate that the carboxyfluorescein release is accompanied by mixing of the virus and liposome lipids and their entrapped contents, suggesting liposome-virus fusion. No release of carboxyfluorescein was observed with vesicles containing only phosphatidylcholine. The rate of virus-induced carboxyfluorescein release was temperature dependent; the lytic activity of the virus was greatly enhanced above 25 degrees C. This effect was not due to a thermal phase transition of the lipids in either the lipid vesicles or the virions. Virus-induced carboxyfluorescein release was inhibited by the presence of calcium ions in the medium and of cholesterol in the lipid vesicles. It increased with increasing concentrations of either the lipid vesicles or the virions. pretreatment of virions with increasing concentrations of three different proteolytic enzymes (trypsin, chymotrypsin and proteinase) inhibited the virus' ability to cause release of carboxyfluorescein from negatively charged liposomes. Inhibition of the viral lytic activity was also observed after virions were incubated above 56 degrees C.  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