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 46 毫秒
1.
The mechanism of the all-or-none release of the contents of phospholipid vesicles induced by the antimicrobial peptide cecropin A was investigated. A detailed experimental study of the kinetics of dye release showed that the rate of release increases with the ratio of peptide bound per vesicle and, at constant concentration, with the fraction of the anionic lipid phosphatidylglycerol in neutral, phosphatidylcholine membranes. Direct measurement of the kinetics of peptide binding and dissociation from vesicles revealed that the on-rate is almost independent of vesicle composition, whereas the off-rate decreases by orders of magnitude with increasing content of anionic lipid. A simple, exact model fits all experimental kinetic data quantitatively. This is the first time that an exact kinetic model is implemented for a peptide with an all-or-none mechanism. In this model, cecropin A binds reversibly to vesicles, which at a certain point enter an unstable state. In this state, a pore probably opens and all vesicle contents are released, consistent with the all-or-none mechanism. This pore state is a state of the whole vesicle, but does not necessarily involve all peptides on that vesicle. No peptide oligomerization on the vesicle is involved; alternative models that assume oligomerization are inconsistent with the experimental data. Thus, formation of well-defined, peptide-lined pores is unlikely.  相似文献   

2.
The antibacterial peptide PGLa exerts its activity by permeabilizing bacterial membranes whereas eukaryotic membranes are not affected. To provide insight into the selectivity and the permeabilization mechanism, the binding of PGLa to neutral and negatively charged model membranes was studied with high-sensitivity isothermal titration calorimetry (ITC), circular dichroism (CD), and solid-state deuterium nuclear magnetic resonance ((2)H NMR). The binding of PGLa to negatively charged phosphatidylcholine (PC)/phosphatidylglycerol (PG) (3:1) vesicles was by a factor of approximately 50 larger than that to neutral PC vesicles. The negatively charged membrane accumulates the cationic peptide at the lipid-water interface, thus facilitating the binding to the membrane. However, if bulk concentrations are replaced by surface concentrations, very similar binding constants are obtained for neutral and charged membranes (K approximately 800-1500 M(-)(1)). Membrane selectivity is thus caused almost exclusively by electrostatic attraction to the membrane surface and not by hydrophobic insertion. Membrane insertion is driven by an exothermic enthalpy (DeltaH approximately -11 to -15 kcal/mol) but opposed by entropy. An important contribution to the binding process is the membrane-induced random coil --> alpha-helix transition of PGLa. The peptide is random coil in solution but adopts an approximately 80% alpha-helical conformation when bound to the membrane. Helix formation is an exothermic process, contributing approximately 70% to the binding enthalpy and approximately 30% to the free energy of binding. The (2)H NMR measurements with selectively deuterated lipids revealed small structural changes in the lipid headgroups and in the hydrocarbon interior upon peptide binding which were continuous over the whole concentration range. In contrast, isothermal titration calorimetry of PGLa solutions with PC/PG(3:1) vesicles gave rise to two processes: (i) an exothermic binding of PGLa to the membrane followed by (ii) a slower endothermic process. The latter is only detected at peptide-to-lipid ratios >17 mmol/mol and is paralleled by the induction of membrane leakiness. Dye efflux measurements are consistent with the critical limit derived from ITC measurements. The endothermic process is assigned to peptide pore formation and/or lipid perturbation. The enthalpy of pore formation is 9.7 kcal/mol of peptide. If the same excess enthalpy is assigned to the lipid phase, the lipid perturbation enthalpy is 180 cal/mol of lipid. The functional synergism between PGLa and magainin 2 amide could also be followed by ITC and dye release experiments and is traced back to an enhanced pore formation activity of a peptide mixture.  相似文献   

3.
We previously proposed three hypotheses relating the mechanism of antimicrobial and cytolytic peptides in model membranes to the Gibbs free energies of binding and insertion into the membrane [Almeida, P. F., and Pokorny, A. (2009) Biochemistry 48, 8083-8093]. Two sets of peptides were designed to test those hypotheses, by mutating of the sequences of δ-lysin, cecropin A, and magainin 2. Peptide binding and activity were measured on phosphatidylcholine membranes. In the first set, the peptide charge was changed by mutating basic to acidic residues or vice versa, but the amino acid sequence was not altered much otherwise. The type of dye release changed from graded to all-or-none according to prediction. However, location of charged residues in the sequence with the correct spacing to form salt bridges failed to improve binding. In the second set, the charged and other key residues were kept in the same positions, whereas most of the sequence was significantly but conservatively simplified, maintaining the same hydrophobicity and amphipathicity. This set behaved completely different from predicted. The type of release, which was expected to be maintained, changed dramatically from all-or-none to graded in the mutants of cecropin and magainin. Finally, contrary to the hypotheses, the results indicate that the Gibbs energy of binding to the membrane, not the Gibbs energy of insertion, is the primary determinant of peptide activity.  相似文献   

4.
Aminoacylated phosphatidylglycerols are common lipids in bacterial cytoplasmic membranes. Their presence in Staphylococcus aureus has been linked to increased resistance to a number of antibacterial agents, including antimicrobial peptides. Most commonly, the phosphatidylglycerol headgroup is esterified to lysine, which converts anionic phosphatidylglycerol into a cationic lipid with a considerably increased headgroup size. In the present work, we investigated the interactions of two well-studied antimicrobial peptides, cecropin A and mastoparan X, with lipid vesicles composed of 1-palmitoyl-2-oleoyl-phosphatidylcholine (POPC) and 1-palmitoyl-2-oleoyl-phosphatidylglycerol (POPG), containing varying fractions of an aminoacylated phosphatidylethanolamine, a stable analog of the corresponding phosphatidylglycerol-derivative. To differentiate between the effects of headgroup size and charge on peptide–lipid interactions, we synthesized two different derivatives. In one, the headgroup was modified by the addition of lysine, and in the other, by glutamine. The modification by glutamine results in a phospholipid with a headgroup size comparable to that of the lysylated version. However, whereas lysyl-phosphatidylethanolamine (Lys-PE) is cationic, glutaminyl-phosphatidylethanolamine (Gln-PE) is zwitterionic. We found that binding of mastoparan X and cecropin A was not significantly altered if the content of aminoacylated phosphatidylethanolamines did not exceed 20 mol.%, which is the concentration found in bacterial membranes. However, a lysyl-phosphatidylethanolamine content of 20 mol% significantly inhibits dye release from lipid vesicles, to a degree that depends on the peptide. In the case of mastoparan X, dye release is essentially abolished at 20 mol.% lysyl-phosphatidylethanolamine, whereas cecropin A is less sensitive to the presence of lysyl-phosphatidylethanolamine. These observations are understood through the complex interplay between peptide binding and membrane stabilization as a function of the aminoacylated lipid content. This article is part of a Special Issue entitled: Interfacially Active Peptides and Proteins. Guest Editors: William C. Wimley and Kalina Hristova.  相似文献   

5.
Melittin, an amphiphathic peptide, affects the permeability of vesicles. This can be demonstrated using the dye release technique. Calcein, a fluorescent marker, is trapped in large unilamellar 1-palmitoyl-2-oleoyl-phosphatidylcholine (POPC) vesicles and melittin-induced leakage of the dye can be monitored directly by increasing fluorescence intensity. First, we characterized the effect of increasing cholesterol content in the membrane on melittin-induced leakage and our results reveal that cholesterol inhibits the lytic activity of the peptide. Using intrinsic fluorescence of the single tryptophan of melittin and 2H-NMR of headgroup deuterated phosphatidylcholine, we demonstrated that the affinity of melittin for phosphatidylcholine vesicles is reduced in the presence of cholesterol; this is associated with the tighter lipid packing of the cholesterol-containing bilayer. This reduced binding is responsible for the reduced melittin-induced leakage from cholesterol-containing membranes. The pathway of release was determined to be an all-or-none mechanism. Finally, we investigated the possibility of achieving specific membrane targeting with melittin, when vesicles of different lipid composition are simultaneously present. Melittin incubated together with vesicles made of pure POPC and POPC containing 30(mol)% cholesterol can empty nearly all the cholesterol-free vesicles while the cholesterol-containing vesicles remain almost intact. Owing to the preferential interaction of melittin with the pure POPC vesicles, we were able to achieve controlled release of encapsulated material from a specific vesicle population. Received: 8 May 1996 / Accepted: 12 September 1996  相似文献   

6.
Surface pressure measurements, external reflection-Fourier transform infrared spectroscopy, and neutron reflectivity have been used to investigate the lipid-binding behavior of three antimicrobial peptides: melittin, magainin II, and cecropin P1. As expected, all three cationic peptides were shown to interact more strongly with the anionic lipid, 1,2 dihexadecanoyl-sn-glycerol-3-(phosphor-rac-(1-glycerol)) (DPPG), compared to the zwitterionic lipid, 1,2 dihexadecanoyl-sn-glycerol-3-phosphocholine (DPPC). All three peptides have been shown to penetrate DPPC lipid layers by surface pressure, and this was confirmed for the melittin-DPPC interaction by neutron reflectivity measurements. Adsorption of peptide was, however, minimal, with a maximum of 0.4 mg m(-2) seen for melittin adsorption compared to 2.1 mg m(-2) for adsorption to DPPG (from 0.7 microM solution). The mode of binding to DPPG was shown to depend on the distribution of basic residues within the peptide alpha-helix, although in all cases adsorption below the lipid layer was shown to dominate over insertion within the layer. Melittin adsorption to DPPG altered the lipid layer structure observed through changes in the external reflection-Fourier transform infrared lipid spectra and neutron reflectivity. This lipid disruption was not observed for magainin or cecropin. In addition, melittin binding to both lipids was shown to be 50% greater than for either magainin or cecropin. Adsorption to the bare air-water interface was also investigated and surface activity followed the trend melittin>magainin>cecropin. External reflection-Fourier transform infrared amide spectra revealed that melittin adopted a helical structure only in the presence of lipid, whereas magainin and cecropin adopted helical structure also at an air-water interface. This behavior has been related to the different charge distributions on the peptide amino acid sequences.  相似文献   

7.
The cecropin-melittin hybrid antimicrobial peptide BP100 (H-KKLFKKILKYL-NH2) is selective for Gram-negative bacteria, negatively charged membranes, and weakly hemolytic. We studied BP100 conformational and functional properties upon interaction with large unilamellar vesicles, LUVs, and giant unilamellar vesicles, GUVs, containing variable proportions of phosphatidylcholine (PC) and negatively charged phosphatidylglycerol (PG). CD and NMR spectra showed that upon binding to PG-containing LUVs BP100 acquires α-helical conformation, the helix spanning residues 3–11. Theoretical analyses indicated that the helix is amphipathic and surface-seeking. CD and dynamic light scattering data evinced peptide and/or vesicle aggregation, modulated by peptide:lipid ratio and PG content. BP100 decreased the absolute value of the zeta potential (ζ) of LUVs with low PG contents; for higher PG, binding was analyzed as an ion-exchange process. At high salt, BP100-induced LUVS leakage requires higher peptide concentration, indicating that both electrostatic and hydrophobic interactions contribute to peptide binding. While a gradual release took place at low peptide:lipid ratios, instantaneous loss occurred at high ratios, suggesting vesicle disruption. Optical microscopy of GUVs confirmed BP100-promoted disruption of negatively charged membranes. The mechanism of action of BP100 is determined by both peptide:lipid ratio and negatively charged lipid content. While gradual release results from membrane perturbation by a small number of peptide molecules giving rise to changes in acyl chain packing, lipid clustering (leading to membrane defects), and/or membrane thinning, membrane disruption results from a sequence of events – large-scale peptide and lipid clustering, giving rise to peptide-lipid patches that eventually would leave the membrane in a carpet-like mechanism.  相似文献   

8.
Magainins and other antimicrobial peptides increase ion flux across the membrane. They may do this by forming some type of pore or by perturbing lipid organization due to peptide lying on the bilayer surface. In order to determine if magainins perturb the lipid sufficiently to permeabilize the bilayer, their effect on the motion of fatty acid and lipid spin labels in phosphatidylcholine/phosphatidylglycerol (PC/PG) lipid vesicles was determined. Their effect was compared to two synthetic peptides, 18L and Ac-18A-NH(2), designed to mimic the naturally occurring classes of lytic (class L) and apolipoprotein (class A) amphipathic helices, respectively. We show that although magainins and 18L both had significant effects on lipid chain order, much greater than Ac-18A-NH(2), there was no correlation between these effects and the relative ability of these three peptide classes to permeabilize PC/PG vesicles in the order magainins=Ac-18A-NH(2) > 18L. This suggests that the perturbing effects of magainins on lipid chain order at permeabilizing concentrations are not directly responsible for the increased leakage of vesicle contents. The greater ability of the magainins to permeabilize PC/PG vesicles relative to 18L is thus more likely due to formation of some type of pore by magainins. The greater ability of Ac-18A-NH(2) relative to 18L to permeabilize PC/PG vesicles despite its lack of disordering effect must be due to its ability to cause membrane fragmentation. Effects of these peptides on other lipids indicated that the mechanism by which they permeabilize lipid bilayers depends both on the peptide and on the lipid composition of the vesicles.  相似文献   

9.
Tamba Y  Yamazaki M 《Biochemistry》2005,44(48):15823-15833
It is thought that magainin 2, an antimicrobial peptide, acts by binding to lipid membranes. Recent studies using a suspension of large unilamellar vesicles (LUVs) indicate that magainin 2 causes gradual leakage from LUVs containing negatively charged lipids. However, the details of the characteristics of the membrane permeability and the mechanism of pore formation remain unclear. In this report, we investigated the interaction of magainin 2 with single giant unilamellar vesicles (GUVs) composed of a dioleoylphosphatidylcholine and dioleoylphosphatidylglycerol mixture (50% DOPG/50% DOPC GUVs) containing the fluorescent dye, calcein, by phase contrast, fluorescence microscopy using the single GUV method. Low concentrations (3-10 microM) of magainin 2 caused the rapid leakage of calcein from single GUVs but did not disrupt the liposomes or change the membrane structure, showing directly that magainin 2 forms membrane pores through which calcein leaked. The rapid leakage of calcein from a GUV started stochastically, and once it began, the complete leakage occurred rapidly (6-60 s). The fraction of completely leaked GUV, P(L), increased with time and also with an increase in magainin 2 concentration. Shape changes in these GUVs occurred prior to the pore formation and also at lower concentrations of magainin 2, which could not induce the pore formation. Their analysis indicates that binding of magainin 2 to the external monolayer of the GUV increases its membrane area, thereby raising its surface pressure. The addition of lysophosphatidylcholine into the external monolayer of GUVs increased P(L). On the basis of these results, we propose the two-state transition model for the pore formation.  相似文献   

10.
The magainins, peptide antibiotics secreted by the frog Xenopus laevis, have previously been shown to permeabilize phospholipid vesicles. To elucidate the mechanism of permeabilization, we have conducted detailed kinetic studies of magainin 2 amide (mgn2a)-induced release of 6-carboxyfluorescein from vesicles of phosphatidylserine. The results show that dye release occurs in (at least) two stages--an initial rapid phase, with t1/2 approximately 3 s, followed by a much slower phase that approaches zero leakage rate before all the dye is released. Light-scattering studies showed that mgn2a does not cause gross changes in vesicle structure. The peptide was found to rapidly equilibrate between vesicles; this was demonstrated by determining a binding isotherm for the peptide-lipid interaction, and by showing that addition of unloaded vesicles rapidly quenches peptide-induced leakage from loaded vesicles. Transient dye release in the presence of an equilibrating peptide can be explained in two ways: (1) the peptide exists only transiently in an active form; (2) the vesicles are only transiently leaky. Preincubation of mgn2a at assay concentrations in buffer alone or with unloaded vesicles did not inactivate the peptide. Therefore, rapid leakage is probably due to transient destabilization of the vesicle upon addition of mgn2a.  相似文献   

11.
Pardaxin is a 33-amino-acid neurotoxin from the Red Sea Moses sole Pardachirus marmoratus, whose mode of action shows remarkable sensitivity to lipid chain length and charge, although the effect of pH is unclear. Here we combine optical spectroscopy and dye release experiments with laser scanning confocal microscopy and natural abundance 13C solid-state nuclear magnetic resonance to provide a more complete picture of how pardaxin interacts with lipids. The kinetics and efficiency of release of entrapped calcein is highly sensitive to pH. In vesicles containing zwitterionic lipids (PC), release occurs most rapidly at low pH, whereas in vesicles containing 20% anionic lipid (PG), release occurs most rapidly at high pH. Pardaxin forms stable or transient pores in PC vesicles that allow release of contents without loss of vesicle integrity, whereas the inclusion of PG promotes total vesicle collapse. In agreement with this, solid-state nuclear magnetic resonance reveals that pardaxin takes up a trans-membrane orientation in 14-O-PC/6-O-PC bicelles, whereas the inclusion of 14-0-PG restricts it to contacts with lipid headgroups, promoting membrane lysis. Pore formation in zwitterionic vesicles is more efficient than lysis of anionic vesicles, suggesting that electrostatic interactions may trap pardaxin in several suboptimal interconverting conformations on the membrane surface.  相似文献   

12.
Sticholysin I and II (St I and St II), two basic cytolysins purified from the Caribbean sea anemone Stichodactyla helianthus, efficiently permeabilize lipid vesicles by forming pores in their membranes. A general characteristic of these toxins is their preference for membranes containing sphingomyelin (SM). As a consequence, vesicles formed by equimolar mixtures of SM with phosphatidylcholine (PC) are very good targets for St I and II. To better characterize the lipid dependence of the cytolysin-membrane interaction, we have now evaluated the effect of including different lipids in the composition of the vesicles. We observed that at low doses of either St I or St II vesicles composed of SM and phosphatidic acid (PA) were permeabilized faster and to a higher extent than vesicles of PC and SM. As in the case of PC/SM mixtures, permeabilization was optimal when the molar ratio of PA/SM was ~1. The preference for membranes containing PA was confirmed by inhibition experiments in which the hemolytic activity of St I was diminished by pre-incubation with vesicles of different composition. The inclusion of even small proportions of PA into PC/SM LUVs led to a marked increase in calcein release caused by both St I and St II, reaching maximal effect at ~5 mol % of PA. Inclusion of other negatively charged lipids (phosphatidylserine (PS), phosphatidylglycerol (PG), phosphatidylinositol (PI), or cardiolipin (CL)), all at 5 mol %, also elicited an increase in calcein release, the potency being in the order CL approximately PA > PG approximately PI approximately PS. However, some boosting effect was also obtained, including the zwitterionic lipid phosphatidylethanolamine (PE) or even, albeit to a lesser extent, the positively charged lipid stearylamine (SA). This indicated that the effect was not mediated by electrostatic interactions between the cytolysin and the negative surface of the vesicles. In fact, increasing the ionic strength of the medium had only a small inhibitory effect on the interaction, but this was actually larger with uncharged vesicles than with negatively charged vesicles. A study of the fluidity of the different vesicles, probed by the environment-sensitive fluorescent dye diphenylhexatriene (DPH), showed that toxin activity was also not correlated to the average membrane fluidity. It is suggested that the insertion of the toxin channel could imply the formation in the bilayer of a nonlamellar structure, a toroidal lipid pore. In this case, the presence of lipids favoring a nonlamellar phase, in particular PA and CL, strong inducers of negative curvature in the bilayer, could help in the formation of the pore. This possibility is confirmed by the fact that the formation of toxin pores strongly promotes the rate of transbilayer movement of lipid molecules, which indicates local disruption of the lamellar structure.  相似文献   

13.
D Rapaport  R Peled  S Nir    Y Shai 《Biophysical journal》1996,70(6):2502-2512
The mechanism of leakage induced by surface active peptides is not yet fully understood. To gain insight into the molecular events underlying this process, the leakage induced by the peptide pardaxin from phosphatidylcholine/ phosphatidylserine/cholesterol large unilamellar vesicles was studied by monitoring the rate and extent of dye release and by theoretical modeling. The leakage occurred by an all-or-none mechanism: vesicles either leaked or retained all of their contents. We further developed a mathematical model that includes the assumption that certain peptides become incorporated into the vesicle bilayer and aggregate to form a pore. The current experimental results can be explained by the model only if the surface aggregation of the peptide is reversible. Considering this reversibility, the model can explain the final extents of calcein leakage for lipid/peptide ratios of > 2000:1 to 25:1 by assuming that only a fraction of the bound peptide forms pores consisting of M = 6 +/- 3 peptides. Interestingly, less leakage occurred at 43 degrees C, than at 30 degrees C, although peptide partitioning into the bilayer was enhanced upon elevation of the temperature. We deduced that the increased leakage at 30 degrees C was due to an increase in the extent of reversible surface aggregation at the lower temperature. Experiments employing fluorescein-labeled pardaxin demonstrated reversible aggregation of the peptide in suspension and within the membrane, and exchange of the peptide between liposomes. In summary, our experimental and theoretical results support reversible surface aggregation as the mechanism of pore formation by pardaxin.  相似文献   

14.
Understanding the mechanisms of antimicrobial, cytolytic and cell-penetrating peptides is important for the design of new peptides to be used as cargo-delivery systems or antimicrobials. But these peptides should not be hemolytic. Recently, we designed a series of such membrane-active peptides and tested several hypotheses about their mechanisms on model membranes. To that end, the Gibbs free energy of binding to 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) vesicles was determined experimentally. Because the main lipid components of the outermost monolayer of erythrocyte membranes are zwitterionic, like POPC, we hypothesized that the Gibbs free energy of binding of these peptides to POPC would also be a good indicator of their hemolytic activity. Now, the hemolytic activity of those synthetic peptides was examined by measuring the lysis of sheep erythrocyte suspensions after peptide addition. Indeed, the Gibbs free energy of binding was in good correlation with the hemolytic activity, which was represented by the concentration of peptide in solution that produced 50 % hemolysis. Furthermore, with two exceptions, those peptides that caused graded dye release from POPC vesicles were also hemolytic, while most of those that caused all-or-none release were not.  相似文献   

15.
16.
Antimicrobial, cytolytic, and cell-penetrating peptides induce pores or perturbations in phospholipid membranes that result in fluxes of dyes into or out of lipid vesicles. Here we examine the fluxes induced by four of these membrane-active peptides in giant unilamellar vesicles. The type of flux is determined from the modality of the distributions of vesicles as a function of their dye content using the statistical Hartigan dip test. Graded and all-or-none fluxes correspond to unimodal and bimodal distributions, respectively. To understand how these distributions arise, we perform Monte Carlo simulations of peptide-induced dye flux into vesicles using a very simple model. The modality of the distributions depends on the rate constants of pore opening and closing, and dye flux. If the rate constants of pore opening and closing are both much smaller than that of dye flux through the pore, all-or-none influx occurs. However, if one of them, especially the rate constant for pore opening, increases significantly relative to the flux rate constant, the process becomes graded. In the experiments, we find that the flux type is the same in giant and large vesicles, for all peptides except one. But this one exception indicates that the flux type cannot be used to unambiguously predict the mechanism of membrane permeabilization by the peptides.  相似文献   

17.
Dipalmitoyl-3-aza-dehydroxy-lysylphosphatidylglycerol (DP3adLPG), is a chemically stable synthetic analogue of the bacterial lipid lysylphosphatidylglycerol (LPG), designed as a substitute for the notoriously labile native lipid in biophysical investigations. In Staphylococcus aureus, LPG is known to play a role in resistance to antibiotics by altering membrane charge properties in response to environmental stress, but little is known about how LPG influences other bilayer physicochemical properties or lateral organisation, through the formation of complexes with lipids such as phosphatidylglycerol (PG). In this study we have investigated the different phases formed by biomimetic mixtures of 3adLPG and PG in different thermotropic states, using neutron diffraction and electron microscopy. In a DPPG/DP3adLPG 70:30 mol% mixture, two distinct lamellar phases were observed below the lipid melting transition: Lβ′ 1 and Lβ′ 2 with respective periodicities of 82 and 62 Å. Increasing the proportion of DP3adLPG to mimic the effects of environmental stress led to the disappearance of the Lβ′ 1 phase and the formation of an inverse hexagonal phase. The compositions of these different phases were identified by investigating the thermotropic properties of the two mixtures, and probing their interaction with the antimicrobial peptide magainin 2 F5W. We propose that the observed polymorphism results from the preferential formation of either triplet PG-3adLPG-PG, or paired PG-3adLPG complexes, dependent upon the mixing proportions of the two lipids. The relevance of these findings to the role native LPG in S. aureus, are discussed with respect to their influence on antibiotic resistance and lateral membrane organisation.  相似文献   

18.
Most antimicrobial peptides (AMPs) damage the cell membrane of bacterial cells and induce rapid leakage of the internal cell contents, which is a main cause of their bactericidal activity. One of the AMPs, magainin 2 (Mag), forms nanopores in giant unilamellar vesicles (GUVs) comprising phosphatidylcholine (PC) and phosphatidylglycerol (PG), inducing leakage of fluorescent probes. In this study, to elucidate the Mag-induced pore formation in lipid bilayer region in E. coli cell membrane, we examined the interaction of Mag with single GUVs comprising E. coli polar lipids (E. coli-lipid-GUVs). First, we investigated the Mag-induced leakage of a fluorescent probe AF488 from single E. coli-lipid-GUVs, and found that Mag caused rupture of GUVs, inducing rapid AF488 leakage. The rate constant of Mag-induced GUV rupture increased with the Mag concentration. Using fluorescence microscopy with a time resolution of 5 ms, we revealed the GUV rupture process: first, a small micropore was observed in the GUV membrane, then the pore radius increased within 50 ms without changing the GUV diameter, the thickness of the membrane at the pore rim concomitantly increased, and eventually membrane aggregates were formed. Mag bound to only the outer monolayer of the GUV before GUV rupture, which increased the area of the GUV bilayer. We also examined the physical properties of E. coli-lipid-GUVs themselves. We found that the rate constant of the constant tension-induced rupture of E. coli-lipid-GUVs was higher than that of PG/PC-GUVs. Based on these results, we discussed the Mag-induced rupture of E. coli-lipid-GUVs and its mechanism.  相似文献   

19.
Haque ME  Lentz BR 《Biochemistry》2002,41(35):10866-10876
The fusion peptide of the HIV fusion protein gp41 is required for viral fusion and entry into a host cell, but it is unclear whether this 23-residue peptide can fuse model membranes. We address this question for model membrane vesicles in the presence and absence of aggregating concentrations of poly(ethylene glycol) (PEG). PEG had no effect on the physical properties of peptide bound to membranes or free in solution. We tested for fusion of both highly curved and uncurved PC/PE/SM/CH (35:30:15:20 mol %) vesicles and highly curved PC/PE/CH (1:1:1) vesicles treated with peptide in the presence and absence of PEG. Fusion was never observed in the absence of PEG, although high peptide concentrations led to aggregation and rupture, especially in unstable PC/PE/CH (1:1:1) vesicles. When 5 wt % PEG was present to aggregate vesicles, peptide enhanced the rate of lipid mixing between curved PC/PE/SM/CH vesicles in proportion to the peptide concentration, with this effect leveling off at peptide/lipid (P/L) ratios approximately 1:200. Peptide produced an even larger effect on the rate of contents mixing but inhibited contents mixing at P/L ratios >1:200. No fusion enhancement was seen with uncurved vesicles. The rate of fusion was also enhanced by the presence of hexadecane, and peptide-induced rate enhancement was not observed in the presence of hexadecane. We conclude that gp41 fusion peptide does not induce vesicle fusion at subrupturing concentrations but can enhance fusion between highly curved vesicles induced to fuse by PEG. The different effects of peptide on the rates of lipid mixing and fusion pore formation suggest that, while gp41 fusion peptide does affect hemifusion, it mainly affects pore formation.  相似文献   

20.
Interactions of two antimicrobial peptides, magainin 2 and indolicidin, with three different model biomembranes, namely, monolayers, large unilamellar vesicles (LUVs), and giant liposomes, were studied. Insertion of both peptides into lipid monolayers was progressively enhanced when the content of an acidic phospholipid, 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol (POPG) in a film of 1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine (SOPC) was increased. Indolicidin and magainin 2 penetrated also into lipid monolayers containing cholesterol (mole fraction, X = 0.1). Membrane association of magainin 2 attenuated lipid lateral diffusion in POPG-containing LUVs as revealed by the decrease in the excimer/monomer fluorescence ratio I(e)/I(m) for the pyrene fatty-acid-containing phospholipid derivative 1-palmitoyl-2-[10-(pyren-1-yl) decanoyl]-sn-glycero-3-phospho-rac-glycerol (PPDPG). Likewise, an increase in steady-state fluorescence anisotropy of the membrane-incorporated diphenylhexatriene (DPH) was observed, revealing magainin 2 to increase acyl chain order and induce segregation of acidic phospholipids. Similar effects were observed for indolicidin. The topological effects of magainin 2 and indolicidin on phospholipid membranes were investigated using optical microscopy of giant vesicles. Magainin 2 had essentially no influence on either SOPC or SOPC:cholesterol (X = 0.1) giant liposomes. However, effective vesiculation was observed when acidic phospholipid (X(PG) = 0.1) was included in the giant vesicles. Indolicidin caused only a minor shrinkage of giant SOPC vesicles whereas the formation of endocytotic vesicles was observed when the giant liposome contained POPG (X(PG) = 0.1). Interestingly, for indolicidin, vesiculation was also observed for giant vesicles composed of SOPC/cholesterol (X(chol) = 0.1). Possible mechanisms of membrane transformation induced by these two peptides are discussed.  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