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 78 毫秒
1.
The mechanism of folding and membrane insertion of integral membrane proteins, including helix bundle and beta-barrel proteins is not well understood. A key question is whether folding and insertion are coupled or separable processes. We have used the beta-barrel outer membrane protein A (OmpA) of Escherichia coli as a model to study the kinetics of folding and insertion into dioleoylphosphatidylcholine (DOPC) bilayers, as a function of temperature by gel electrophoresis, protease digestion, and fluorescence spectroscopy. OmpA was unfolded in 8 M urea solution (without detergent), and refolding and membrane insertion was initiated by rapid dilution of the urea concentration in the presence of phospholipid vesicles. In addition to the kinetically unresolved hydrophobic collapse in water, the time course of refolding of OmpA into DOPC bilayers exhibited three kinetic phases over a large temperature range. The first step was fast (k1 = 0.16 min-1) and not very dependent on temperature. The second step was up to two orders of magnitude slower at low temperatures (2 degrees C), but approached the rate of the first step at higher temperatures (40 degrees C). The activation energy for this process was 46 +/- 4 kJ/mol. A third slow process (k3 = 0.9 x 10(-2) min-1 at 40 degrees C) was observed at the higher temperatures. These results suggest that at least two membrane-bound intermediates exist when OmpA folds and inserts into lipid bilayers. We also show that both membrane-bound intermediates can be stabilized in fluid lipid bilayers at low temperatures. These intermediates share many properties with the adsorbed/partially inserted form of OmpA that was previously characterized in gel phase lipid bilayers [Rodionova et al. (1995) Biochemistry 34, 1921-1929]. Temperature jump experiments demonstrate, that the low-temperature intermediates can be rapidly converted to fully inserted native OmpA. On the basis of these and previous results, we present a simple folding model for beta-barrel membrane proteins, in which folding and membrane insertion are coupled processes which involve at least four kinetically distinguishable steps.  相似文献   

2.
Deuterium (2H) nuclear magnetic resonance (NMR) quadrupole splittings and relaxation times have been measured for a variety of specifically deuterated lipids intercalated in lamellar-multibilayer dispersions and single-bilayer vesicles of egg lecithin and lecithin-cholesterol mixtures. The deduced order parameters and relaxation times vary with position of deuteration, acyl chain length, unsaturation, and temperature. The order parameters and spinlattice relaxation times T1 indicate rapid intramolecular motions of restricted amplitude in both the choline head group and hydrocarbon chains. The ordering profile for the acyl chains is similar to that predicted by statistical-mechanical theory. The order parameters yield estimates of the bilayer thickness and linear coefficient of expansion in close agreement with the x-ray determinations. A comparison of the deuterium and electron spin resonance spinprobe order parameters demonstrates the perturbation of the bilayer by the bulky nitroxide probe. The transverse relaxation time T2 for single-bilayer vesicles is quantitatively accounted for by a simple modification of classical relaxation theory which takes into account the modulation of the static quadrupole interaction by rapid local molecular motions and the modulation of the residual quadrupole interaction by the slower overall tumbling of the vesicle. It is unambiguously demonstrated that molecular motion and order in single-bilayer vesicles are very similar to those in lamellar multibilayers. Significant differences occur only for a few segments near the terminal methyl groups of the acyl chains, where the order parameters for vesicles are 10-30% smaller than those found for lamellae. The incorporation of cholesterol in lecithin bilayers is shown to increase the degree of orientational order in vesicles and lamellae, and to increase the hydrodynamic radius of vesicles. Thus, single-bilayer vesicles and multilamellar dispersions of phospholipids are equally useful models for biological membranes. They yield equivalent information about the internal organization and mobility of lipid bilayers, when the spectral manifestations of overall vesicle motion are correctly taken into account.  相似文献   

3.
A cationic amphiphilic peptide made of 10 leucine and 10 lysine residues, and four of its fluorescent derivatives in which leucines were substituted by Trp residues at different locations on the primary sequence have been synthesized. The interactions of these five peptides with neutral anionic or cationic vesicles were investigated using circular dichroism, steady state and time-resolved fluorescence with a combination of Trp quenching by brominated lipid probes, monolayers, modeling with minimization and simulated annealing procedures. We show that all the five peptides interact with neutral and anionic DMPC, DMPG, DOPC or egg yolk PC vesicles. The binding takes place whatever the peptide conformation in solution is. In the case of DMPC bilayers the binding free energy DeltaG is estimated at -8 kcal mole-1 and the number of phospholipid molecules involved is about 20-25 per peptide molecule. Peptides are bound as single-stranded alpha helices orientated parallel to the bilayer surface. In the anchoring of phospholipid head groups around the peptides, the lipid molecules are not smeared out in a plane parallel to the membrane surface but are organized around the hydrophilic face of the alpha helices like 'wheat grains around an ear' and protrude outside the bilayer towards the solvent. We suggest that such a lipid arrangement generates transient structural defects responsible for the membrane permeability enhancement. When an electrical potential is applied, the axis of the peptide helices remains parallel to the membrane surface and does not reorient to give rise to a bundle of helix monomers that forms transmembrane channels via a 'barrel stave' mechanism. The penetration depth of alpha helices in relation to the position of phosphorus atoms in the unperturbed lipid leaflet is estimated at 3.2 A.  相似文献   

4.
Processes occurring in dispersions of dimyristoyl phosphatidylcholine containing myristic acid have been studied by light scattering of dilute dispersions (concn. less than or equal to 1 mg/ml) at temperatures above and below the phase transition temperatures of these dispersions. The transition temperatures increase with increasing mol fraction of myristic acid. Above these temperatures, vesicles with different mol fractions of myristic acid exchange lipid molecules. The exchange process leads to vesicles having phase transition temperatures and radii, which ar both intermediate between the initial transitions and radii, respectively. In contrast with the observations above the phase transitions, it was found that when dimyristoyl phosphatidylcholine/myristic acid vesicles were cooled to a few degrees below the phase transition, larger particles were formed. These observations are consistent with a mechanism consisting of vesicle aggregation followed by fustion of the aggregated vesicles. The aggregation process is of second order in the vesicle concentration, and its rate increases with increasing mol fraction of myristic acid.  相似文献   

5.
We studied the mechanism of membrane permeabilization by the 18L model peptide (GIKKFLGSIWKFIKAFVG), which features the consensus class L sequence averaged from the number of naturally occurring lytic peptides. Two aspects of membrane lipid composition significantly affected peptide-membrane interactions: the presence of acidic lipids and, in zwitterionic membranes, and the presence of nonbilayer forming lipids. In zwitterionic membranes, 18L peptide destabilizes the membrane, leading to a transient formation of large defects in the membrane which result generally in contents leakage, but in the presence of bilayer-bilayer contact can alternatively lead to vesicle fusion. In membranes containing acidic lipids (DOPC:DOPG, DOPG), 18L caused leakage but not fusion, probably due to mutual repulsion of acidic vesicles. While the extent of contents leakage was approximately the same as for zwitterionic membranes, the kinetics of leakage could be resolved only by using stopped-flow, leakage being essentially complete within the first minute. Previously, we reported that apolipoprotein (class A) and lytic (class L) peptide analogs have opposing effects on some properties of biological membranes. This reciprocal effect of 18L and Ac-18A-NH2, class A model peptide, is restricted to membranes with a high propensity for nonbilayer phase formation (DOPE, Me-DOPE, DOPC:DOPE, DOPC:Me-DOPE). The decrease in the content of nonbilayer phase forming lipid or the addition of acidic lipids reduces or eliminates the reciprocal effects. This suggests the importance of nonbilayer phase propensity for certain functions of biological membranes.  相似文献   

6.
The construction of a mutant Escherichia coli strain which cannot synthesize phosphatidylethanolamine provides a tool to study the involvement of non-bilayer lipids in membrane function. This strain produces phosphatidylglycerol and cardiolipin (CL) as major membrane constituents and requires millimolar concentrations of divalent cations for growth. In this strain, the lipid phase behaviour is tightly regulated by adjustment of the level of CL which favours a nonbilayer organization in the presence of specific divalent cations. We have used an in vitro system of inverted membrane vesicles to study the involvement of non-bilayer lipids in protein translocation in the secretion pathway. In this system, protein translocation is very low in the absence of divalent cations but can be enhanced by inclusion of Mg2+, Ca2+ or Sr2+ but not by Ba2+ which is unable to sustain growth of the mutant strain and cannot induce a non-bilayer phase in E. coli CL dispersions. Alternatively, translocation in cation depleted vesicles could be increased by incorporation of the non-bilayer lipid DOPE (18:1) but not by DMPE (14:0) or DOPC (18:1), both of which are bilayer lipids under physiological conditions. We conclude that non-bilayer lipids are essential for efficient protein transport across the plasma membrane of E. coli.  相似文献   

7.
Free thin liquid films (foam films) formed from aqueous dispersions of dimyristoylphosphatidylcholine (DMPC) and dipalmitoylphosphatidylethanolamine with covalently bound poly-(ethylene glycol) of molecular weight 2000 (DPPE-PEG-2000) were studied by the thin liquid film microinterferometric technique of Scheludko and Exerowa in the temperature range 14-36 degrees C. The surface tension kinetics of the dispersions were studied in order to ensure equilibration of the foam films. These measurements showed that the rate of surface coverage depends slightly on the temperature and does not reach equilibrium values within reasonable time intervals for the dispersions containing only one amphiphile (DPPE-PEG-2000). The destruction of the vesicles at the air/(aqueous dispersion) interface was much faster for the dispersions containing DMPC/DPPE-PEG-2000 mixtures above 23 degrees C, the temperature of the chain-melting phase transition of the main lipid component (DMPC). The dependence of the equilibrium thickness of the foam films on the electrolyte concentration was measured for 1 and 9 mol% DPPE-PEG-2000 at 28 degrees C in the range 10-3 to 0.5 M NaCl. These results indicate that at the low electrolyte concentrations the electrostatic and van der Waals interactions are dominant similar to the foam films stabilized with DMPC alone. At the high electrolyte concentrations the steric repulsion of the PEG layers becomes dominant. The temperature-composition dependence of the bilayer thickness was measured for the foam bilayers at 0.14 M NaCl. The data for the foam bilayer thickness and the comparison with the phase diagrams of PC/PE-PEG dispersions, show that the DMPC/DPPE-PEG-2000 foam bilayers are able to exist in two phase states characterised by different conformations (mushroom and brush) of the grafted polymer.  相似文献   

8.
When subject to an osmotic 'up-shock', water flows outward from bacterial cytoplasm of the bacterium. Lipid bilayers can shrink very little in area and therefore must wrinkle to accommodate the smaller volume. The usual consequence is that all the layers of the cell envelope must become wrinkled together because they adhere to each other and must now cover a smaller surface. Plasmolysis spaces are formed if the cytoplasmic membrane (CM) separates from the other components of the wall. However, because the CM bilayer is essentially an incompressible two-dimensional liquid, this constraint restricts the location and shape of plasmolysis spaces. With mild up-shocks they form at the pole and around constricting regions in the cell. Elsewhere their creation requires the formation of endocytotic or exocytotic vesicles. The formation of endocytotic vesicles occurs in animal and plant cells as well as in bacterial cells. With stronger up-shocks tubular structures (Bayer adhesion sites), or other special geometric shapes (e.g., Scheie structures) allow the bilayer to surround an irregular shaped cytoplast. Periosmotic agents, that is, those that extract water from the periplasm as well as the cytoplasm, are molecules such as poly-vinyl-pyrrolidone and alpha-cyclodextrin that are too large to pass through the porins in the outer membrane. They were found to significantly inhibit the formation of plasmolysis spaces. Presumably, they inhibit the plasmolysis process, which requires that extracellular fluid enter between the CM and the outer membrane (OM). In the extreme case, with the dehydrating action of both osmotic agents and periosmotic agents, periplasmic space formation tends to be prevented and a new kind of space develops within the cytoplasm. We have designated these as 'cytoplasmic voids'. These novel structures are not bounded by lipid bilayers, in contrast to the endocytotic vesicles. These new spaces appear to result from the negative turgor pressure generated by the application of the combination of osmotic and periosmotic agents causing bubble formation. Several ideas in the literature about the wall biology (periseptal annuli, leading edge, osmotic pressure in the periplasm) are presented and critiqued. The basic criticism of these is that much of the phenomena can be explained because of the physics of the phospholipid bilayers and osmotic forces and thus does not imply the existence of a special control mechanism to regulate growth and division.  相似文献   

9.
We present the phenomenology of transformations in lipid bilayers that are excited by laser tweezers. A variety of dynamic instabilities and shape transformations are observed, including the pearling instability, expulsion of vesicles, and more exotic ones, such as the formation of passages. Our physical picture of the laser-membrane interaction is based on the generation of tension in the bilayer and loss of surface area. Although tension is the origin of the pearling instability, it does not suffice to explain expulsion of vesicles, where we observe opening of giant pores and creeping motion of bilayers. We present a quantitative theoretical framework to understand most of the observed phenomenology. The main hypothesis is that lipid is pulled into the optical trap by the familiar dielectric effect, is disrupted, and finally is repackaged into an optically unresolvable suspension of colloidal particles. This suspension, in turn, can produce osmotic pressure and depletion forces, driving the observed transformations.  相似文献   

10.
Recombinant human cystic fibrosis transmembrane conductance regulator (CFTR) has been produced in a Saccharomyces cerevisiae expression system used previously to produce transport ATPases with high yields. The arrangement of the bases in the region immediately upstream from the ATG start codon of the CFTR is extremely important for high expression levels. The maximal CFTR expression level is about 5-10% of that in Sf9 insect cells as judged by comparison of immunoblots. Upon sucrose gradient centrifugation, the majority of the CFTR is found in a light vesicle fraction separated from the yeast plasma membrane in a heavier fraction. It thus appears that most of expressed CFTR is not directed to the plasma membrane in this system. CFTR expressed in yeast has the same mobility (ca. 140 kDa) as recombinant CFTR produced in Sf9 cells in a high resolution SDS-PAGE gel before and after N-glycosidase F treatment, suggesting that it is not glycosylated. The channel function of the expressed CFTR was measured by an isotope flux assay in isolated yeast membrane vesicles and single channel recording following reconstitution into planar lipid bilayers. In the isotope flux assay, protein kinase A (PKA) increased the rate of 125I- uptake by about 30% in membrane vesicles containing the CFTR, but not in control membranes. The single channel recordings showed that a PKA-activated small conductance anion channel (8 pS) with a linear I-V relationship was present in the CFTR membranes, but not in control membranes. These results show that the human CFTR has been expressed in functional form in yeast. With the reasonably high yield and the ability to grow massive quantities of yeast at low cost, this CFTR expression system may provide a valuable new source of starting material for purification of large quantities of the CFTR for biochemical studies.  相似文献   

11.
We have examined the fractional sterol concentration dependence of dehydroergosterol (DHE) fluorescence in DHE/cholesterol/dimyristoyl-L-alpha-phosphatidylcholine (DMPC), DHE/ergosterol/DMPC and DHE/cholesterol/dipalmitoyl-L-alpha-phosphatidylcholine (DPPC) liquid-crystalline bilayers. Fluorescence intensity and lifetime exhibit local minima (dips) whenever the total sterol mole fraction, irrespective of the DHE content, is near the critical mole fractions predicted for sterols being regularly distributed in hexagonal superlattices. This result provides evidence that all three of these naturally occurring sterols (e.g., cholesterol, ergosterol, and DHE) can be regularly distributed in the membrane and that the bulky tetracyclic ring of the sterols is the cause of regular distribution. Moreover, at the critical sterol mole fractions, the steady-state anisotropy of DHE fluorescence and the calculated rotational relaxation times exhibit distinct peaks, suggesting that membrane free volume reaches a local minimum at critical sterol mole fractions. This, combined with the well-known sterol condensing effect on lipid acyl chains, provides a new understanding of how variations in membrane sterol content change membrane free volume. In addition to the fluorescence dips/peaks corresponding to hexagonal superlattices, we have observed intermediate fluorescence dips/peaks at concentrations predicted by the centered rectangular superlattice model. However, the 22.2 mol% dip for centered rectangular superlattices in DHE/ergosterol/DMPC mixtures becomes diminished after long incubation (4 weeks), whereas on the same time frame the 22.2 mol% dip in DHE/cholesterol/DMPC mixtures remains discernible, suggesting that although all three of these sterols can be regularly distributed, subtle differences in sterol structure cause changes in lateral sterol organization in the membrane.  相似文献   

12.
A kinetic model for pore-mediated and perturbation-mediated flip-flop is presented and used to characterize the mechanism of peptide-induced phospholipid flip-flop in bilayers. The model assumes that certain peptides can bind to and aggregate within the membrane. When the aggregate attains a critical size (M peptides), a channel is created that results in a fast flip-flop of phospholipids. In addition, certain peptides induce flip-flop through perturbation of the membrane without forming a pore. Donor phospholipid vesicles with an asymmetrical distribution of the fluorescent phospholipid 1-oleoyl-2-[12-[(7-nitro-1,2,3-benzoxadiazol-4- yl)amino]dodecanoyl]phosphatidylcholine (NBD-PC) were used to measure the extent of flip-flop by quantitating the decrease in fluorescence as the NBD-PC exchanged from the donor vesicles to acceptor vesicles that contained a quencher of the NBD fluorescence. Flip-flop curves generated at lipid/peptide ratios ranging from 30/1 to 300000/1 could be well-simulated by the model. Pore-forming peptides, such as melittin or the synthetic peptide GALA (WEAALAEALAEALAEHLAEALAEALEALAA), induce rapid phospholipid flip-flop with half-times for flip-flop of seconds at low peptide/vesicle ratios. The deduced pore sizes are M = 10 +/- 2 for GALA and M = 2 - 4 for melittin. The synthetic peptide LAGA (WEAALAEAEALALAEHEALALAEAELALAA) can catalyze flip-flop via bilayer perturbation. In contrast, hydrophobic peptides such as gramicidin A and valinomycin intercalate into the membrane, but induce little flip-flop. Modeling of the kinetics of phospholipid translocation supports pore formation as the key factor in accelerating phospholipid flip-flop. Thus, amphipathic segments from membrane proteins may account for non-energy-dependent phospholipid flip-flop in biological membranes.  相似文献   

13.
High concentrations (> or = 20 wt %) of poly(ethylene glycol) (PEG) induce large, unilamellar, dipalmitoylphosphatidylcholine model membrane vesicles to fuse when the bilayers contain small amounts of amphipathic peturbant molecules. In addition to fusion, similar concentrations of PEG induce these vesicles to leak their contents. In this paper, we have asked if fusion could occur independently of leakage or if fusion might be described as local bilayer rupture followed by resealing. By following the release of MW 10,000 fluoresceinated dextran trapped inside vesicles, it was determined that PEG-induced leakage was the result of major membrane disruption and not small-pore formation. Fusion of vesicles containing 0.5 mol % palmitic acid was clearly observed at 20 wt % PEG, while 25 wt % was needed to cause rupture. On the other hand, vesicles containing 0.5 mol % lysophosphatidylcholine ruptured at roughly the same concentration needed to induce rupture. Two methods were developed for removing PEG so that fusion products could be characterized. Quasi-elastic light scattering demonstrated that fusing vesicles grew in size and that nonfusing vesicles did not. Moreover, PEG concentrations that induced rupture led to the appearance of species with mean diameters much larger than those of fused vesicles. High-resolution nuclear magnetic resonance showed that the population of large vesicles that correlated with rupture was composed of multilamellar vesicles while the population resulting from fusion alone remained unilamellar. We conclude that, upon incubation with and subsequent removal of PEG, vesicles were either unaffected, or fused to form larger, unilamellar vesicles, or ruptured to form larger, nonunilamellar species.  相似文献   

14.
Preparation of inside-out vesicles of pig lymphocyte plasma membrane   总被引:3,自引:0,他引:3  
Between 30 and 50% of pig lymphocyte plasma membrane vesicles were not bound by concanavalin A (Con A)-Sepharose. Various results suggest that the Con A-unretarded fraction represents "inside-out" membrane vesicles. First, an alternative cell surface ligand, anti-lymphocytic serum, gave a similar fractionation to Con A. Second, lack of binding by Con A was not due to lack of carbohydrate or to masking of carbohydrate by extraneous protein, because the unfractionated membrane and the unretarded fraction had similar carbohydrate and polypeptide compositions. Third although the carbohydrate of the unretarded membrane vesicles was accessible to 125I-labelled Con A and to release by soluble trypsin, it was not accessible to ferritin-Con A or trypsin-Sepharose. Fourth, antisera against the external surface of the Con A-unretarded vesicles strongly agglutinated the unretarded membrane, but caused negligible agglutination of whole lymphocytes. When attached to Sepharose these antisera bound all of the Con A-unretarded fraction, but failed to bind the membrane that adhered to Con A-Sepharose.  相似文献   

15.
A new and sensitive molecular probe, 2-(2'-hydroxyphenyl)imidazo[1, 2-a]pyridine (HPIP), for monitoring structural changes in lipid bilayers is presented. Migration of HPIP from water into vesicles involves rupture of hydrogen (H) bonds with water and formation of an internal H bond once the probe is inside the vesicle. These structural changes of the dye allow the occurrence of a photoinduced intramolecular proton-transfer reaction and a subsequent twisting/rotational process upon electronic excitation of the probe. The resulting large Stokes-shifted fluorescence band depends on the twisting motion of the zwitterionic phototautomer and is characterized in vesicles of dimyristoyl-phosphatidylcholine and in dipalmitoyl-phosphatidylcholine at the temperature range of interest and in the presence of cholesterol. Because the fluorescence of aqueous HPIP does not interfere in the emission of the probe within the vesicles, HPIP proton-transfer/twisting motion fluorescence directly allows us to monitor and quantify structural changes within bilayers. The static and dynamic fluorescence parameters are sensitive enough to such changes to suggest this photostable dye as a potential molecular probe of the physical properties of lipid bilayers.  相似文献   

16.
The formation and annealing of structural defects in lipid bilayer vesicles   总被引:2,自引:0,他引:2  
It is shown that sonication of phospholipid-water dispersions below the crystalline leads to liquid crystalline phase transition temperature (Tc) produces bilayer vesicles with structural defects within the bilayer membrane, which permit rapid permeation of ions and catalyze vesicle-vesicle fusion. These structural defects are annihilated simply by annealing the vesicle suspension above Tc. The rate of annealing was found to be slow, of the order of an hour for T = 3 degrees C above Tc, but annealing is complete within 10 min for T = 10 degrees C above Tc. It is proposed that these structural defects are fault-dislocations in the bilayer structure, which arise from a population defect in the distribution of the lipid molecules between the outer and inner monolayers, when small bilayer fragments reassemble to form the small bilayer vesicles during the sonication procedure. Such a population defect can only be remedied by lipid transport via the inside in equilibrium outside flip-flop mechanism, which would account for the slow kinetics of annealing observed even at 3 degrees C above the phase transition.  相似文献   

17.
The structure of softly supported polymer-cushioned lipid bilayers, prepared in two different ways at the quartz-solution interface, were determined using neutron reflectometry. The polymer cushion consisted of a thin layer of branched, cationic polyethyleneimine (PEI), and the bilayers were formed by adsorption of small unilamellar dimyristoylphosphatidylcholine (DMPC) vesicles. When vesicles were first allowed to adsorb to a bare quartz substrate, an almost perfect bilayer formed. When the polymer was then added to the aqueous solution, it appeared to diffuse beneath this bilayer, effectively lifting it from the substrate. In contrast, if the polymer layer is adsorbed first to the bare quartz substrate followed by addition of vesicles to the solution, there is very little interaction of the vesicles with the polymer layer, and the result is a complex structure most likely consisting of patchy multilayers or adsorbed vesicles.  相似文献   

18.
Studies of asymmetric membrane assembly   总被引:1,自引:0,他引:1  
The major capsid protein of M13 bacteriophage is incorporated at each stage of infection into the host plasma membrane with its amino terminus exposed on the outer surface. Purified M13 coat protein is incorporated with the same asymmetry into synthetic phosphatidylcholine vesicles formed near the Tm of the lipid by a cholate dilution technique. We now report that the lipid in the pre-dilution mixture exists as mixed micelles of uniform size. Prior to dilution, the coat protein is present in at least two states of aggregation, both of which behave similarly in the model membrane assembly reaction. No detectable lipid-protein interaction occurs prior to dilution. Upon dilution there is rapid production of small closed vesicles and coat protein is converted to a chymotrypsin-resistant form, presumably reflecting its incorporation into these vesicle bilayers. Formation of large (greater than 6000 A diameter) vesicles occurs slowly with preservation of coat protein asymmetry and internal volume. A model for this assembly reaction is proposed.  相似文献   

19.
Anthracyclines are potent anticancer agents. Their use is limited by the problem of multidrug resistance (MDR) associated with a decreased intracellular accumulation of drug correlated with the presence, in the membrane of resistant cells, of the P-glycoprotein responsible for an active efflux of the drug. The activity of a drug depends upon its intracellular concentration which itself depends on the kinetics (a) of passive influx (b) of passive efflux and (c) of the P-glycoprotein-mediated efflux of drug across the cell membrane. The ability of an anthracycline to overcome MDR depends largely on the first point. The passive drug uptake is governed by their incorporation into the lipid matrix and both electrostatic and hydrophobic forces seem necessary for the stabilization of anthracyclines into lipid bilayers. The aim of the present study was to determine the relative importance of these two interactions. Using microspectrofluorometry and the observation that the fluorescence of anthracycline is enhanced when the dihydroanthraquinone part is embedded within the lipid bilayer, we have determined the partition coefficient (alternatively, the binding constant) of 12 anthracycline derivatives in large unilamellar vesicles. The anthracyclines were (a) doxorubicin, daunorubicin and idarubicin which, at pH 7.2, bear a single positive charge at the level of the amino group on the sugar, (b) their corresponding neutral 3'-hydroxy derivatives where the amino group in the sugar has been replaced by a hydroxyl, (c) the three 13-hydroxy derivatives, doxorubicinol, daunorubicinol and idarubicinol, (d) pirarubicin and (e) two permanently positively charged derivatives. The large unilamellar vesicles contained phosphatidylcholine with various amounts of phosphatidic acid which is negatively charged and of cholesterol. We came to the conclusion that the efficiency of drug incorporation in the bilayers depends neither on the presence of a positive charge on the drug nor on the presence of anionic phospholipid but on the hydrophobicity of the molecule: the neutral and the positively charged form have the same ability to partition into the bilayer. However, the percentage of each form present should depend on the electrostatic parameters.  相似文献   

20.
Nisin is a 34 residue long peptide belonging to the group A lantibiotics with antimicrobial activity against Gram-positive bacteria. The antimicrobial activity is based on pore formation in the cytoplasmic membrane of target organisms. The mechanism which leads to pore formation remains to be clarified. We studied the orientation of nisin via site-directed tryptophan fluorescence spectroscopy. Therefore, we engineered three nisin Z variants with unique tryptophan residues at positions 1, 17, and 32, respectively. The activity of the tryptophan mutants against Gram-positive bacteria and in model membrane systems composed of DOPC or DOPG was established to be similar to that of wild type nisin Z. The tryptophan fluorescence emission maximum showed an increasing blue-shift upon interaction with vesicles containing increased amounts of DOPG, with the largest effect for the 1W peptide. Studies with the aqueous quencher acrylamide showed that all tryptophans became inaccessible from the aqueous phase in the presence of negatively charged lipids in the vesicles. From these results it is concluded that anionic lipids mediate insertion of the tryptophan residues in at least three positions of the molecule into the lipid bilayer. The depth of insertion of the tryptophan residues was determined via quenching of the tryptophan fluorescence by spin-labeled lipids. The results showed that the depth of insertion was dependent on the amount of negatively charged lipids. In membranes containing 50% DOPG, the distances from the bilayer center were determined to be 15.7, 15.0, and 18.4 A for the tryptophan at position 1, 17, and 32, respectively. In membranes containing 90% DOPG, these distances were calculated to be 10.8, 11.5, and 13.1 A, respectively. These results suggest an overall parallel average orientation of nisin in the membrane, with respect to the membrane surface, with the N-terminus more deeply inserted than the C-terminus. These data were used to model the orientation of nisin in the membrane.  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