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 328 毫秒
1.
The cationic antimicrobial peptide PGLa is electrostatically attracted to bacterial membranes, binds as an amphiphilic α-helix, and is thus able to permeabilize the lipid bilayer. Using solid state 2H-NMR of non-perturbing Ala-d3 labels on the peptide, we have characterized the helix alignment under a range of different conditions. Even at a very high peptide-to-lipid ratio (1:20) and in the presence of negatively charged lipids, there was no indication of a toroidal wormhole structure. Instead, PGLa re-aligns from a surface-bound S-state to an obliquely tilted T-state, which is presumably dimeric. An intermediate structure half-way between the S- and T-state was observed in fully hydrated multilamellar DMPC vesicles at 1:50, suggesting a fast exchange between the two states on the time scale of >50 kHz. We demonstrate that this equilibrium is shifted from the S- towards the T-state either upon (i) increasing the peptide concentration, (ii) adding negatively charged DMPG, or (iii) decreasing the level of hydration. The threshold concentration for re-alignment in DMPC is found to be between 1:200 and 1:100 in oriented samples at 96% humidity. In fully hydrated multilamellar DMPC vesicles, it shifts to an effective peptide-to-lipid ratio of 1:50 as some peptides are able to escape into the bulk water phase.  相似文献   

2.
The interaction of the synthetic antimicrobial peptide P5 (KWKKLLKKPLLKKLLKKL-NH2) with model phospholipid membranes was studied using solid-state NMR and circular dichroism (CD) spectroscopy. P5 peptide had little secondary structure in buffer, but addition of large unilamellar vesicles (LUV) composed of dimyristoylphosphatidylcholine (DMPC) increased the β-sheet content to ~20%. Addition of negatively charged LUV, DMPC–dimyristoylphosphatidylglycerol (DMPG) 2:1, led to a substantial (~40%) increase of the α-helical conformation. The peptide structure did not change significantly above and below the phospholipid phase transition temperature. P5 peptide interacted differently with DMPC bilayers with deuterated acyl chains (d54-DMPC) and mixed d54-DMPC–DMPG bilayers, used to mimic eukaryotic and prokaryotic membranes, respectively. In DMPC vesicles, P5 peptide had no significant interaction apart from slightly perturbing the upper region of the lipid acyl chain with minimum effect at the terminal methyl groups. By contrast, in the DMPC–DMPG vesicles the peptide increased disorder throughout the entire acyl chain of DMPC in the mixed bilayer. P5 promoted disordering of the headgroup of neutral membranes, observed by 31P NMR. However, no perturbations in the T 1 relaxation nor the T 2- values were observed at 30°C, although a slight change in the dynamics of the headgroup at 20°C was noticeable compared with peptide-free vesicles. However, the P5 peptide caused similar perturbations of the headgroup of negatively charged vesicles at both temperatures. These data correlate with the non-haemolytic activity of the P5 peptide against red blood cells (neutral membranes) while inhibiting bacterial growth (negatively charged membranes).  相似文献   

3.
We have studied the effects of the antimicrobial peptide gramicidin S (GS) on the thermotropic phase behavior of large multilamellar vesicles of dimyristoylphosphatidylcholine (DMPC), dimyristoylphosphatidylethanolamine (DMPE) and dimyristoyl phosphatidylglycerol (DMPG) by high-sensitivity differential scanning calorimetry. We find that the effect of GS on the lamellar gel to liquid-crystalline phase transition of these phospholipids varies markedly with the structure and charge of their polar headgroups. Specifically, the presence of even large quantities of GS has essentially no effect on the main phase transition of zwitterionic DMPE vesicles, even after repeating cycling through the phase transition, unless these vesicles are exposed to high temperatures, after which a small reduction in the temperature, enthalpy and cooperativity of the gel to liquid-crystalline phase transitions is observed. Similarly, even large amounts of GS produce similar modest decreases in the temperature, enthalpy and cooperativity of the main phase transition of DMPC vesicles, although the pretransition is abolished at low peptide concentrations. However, exposure to high temperatures is not required for these effects of GS on DMPC bilayers to be manifested. In contrast, GS has a much greater effect on the thermotropic phase behavior of anionic DMPG vesicles, substantially reducing the temperature, enthalpy and cooperativity of the main phase transition at higher peptide concentrations, and abolishing the pretransition at lower peptide concentrations as compared to DMPC. Moreover, the relatively larger effects of GS on the thermotropic phase behavior of DMPG vesicles are also manifest without cycling through the phase transition or exposure to high temperatures. Furthermore, the addition of GS to DMPG vesicles protects the phospholipid molecules from the chemical hydrolysis induced by their repeated exposure to high temperatures. These results indicate that GS interacts more strongly with anionic than with zwitterionic phospholipid bilayers, probably because of the more favorable net attractive electrostatic interactions between the positively charged peptide and the negatively charged polar headgroup in such systems. Moreover, at comparable reduced temperatures, GS appears to interact more strongly with zwitterionic DMPC than with zwitterionic DMPE bilayers, probably because of the more fluid character of the former system. In addition, the general effects of GS on the thermotropic phase behavior of zwitterionic and anionic phospholipids suggest that it is located at the polar/apolar interface of liquid-crystalline bilayers, where it interacts primarily with the polar headgroup and glycerol-backbone regions of the phospholipid molecules and only secondarily with the lipid hydrocarbon chains. Finally, the considerable lipid specificity of GS interactions with phospholipid bilayers may prove useful in the design of peptide analogs with stronger interactions with microbial as opposed to eucaryotic membrane lipids.  相似文献   

4.
The energetics and partition of two hybrid peptides of cecropin A and melittin (CA(1-8)M(1-18) and CA(1-7)M(2-9)) with liposomes of different composition were studied by time-resolved fluorescence spectroscopy, isothermal titration calorimetry, and surface plasmon resonance. The study was carried out with large unilamellar vesicles of three different lipid compositions: 1,2-dimyristoil-sn-glycero-3-phosphocholine (DMPC), 1,2-dimyristoyl-sn-glycero-3-phospho-rac-(1-glycerol) (DMPG), and a 3:1 binary mixture of DMPC/DMPG in a wide range of peptide/lipid ratios. The results are compatible with a model involving a strong electrostatic surface interaction between the peptides and the negatively charged liposomes, giving rise to aggregation and precipitation. A correlation is observed in the calorimetric experiments between the observed events and charge neutralization for negatively charged and mixed membranes. In the case of zwitterionic membranes, a very interesting case study was obtained with the smaller peptide, CA(1-7)M(2-9). The calorimetric results obtained for this peptide in a large range of peptide/lipid ratios can be interpreted on the basis of an initial and progressive surface coverage until a threshold concentration, where the orientation changes from parallel to perpendicular to the membrane, followed by pore formation and eventually membrane disruption. The importance of negatively charged lipids on the discrimination between bacterial and eukaryotic membranes is emphasized.  相似文献   

5.
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   

6.
Bak is a pro-apoptotic protein widely distributed in different cell types that is associated with the mitochondrial outer membrane, apparently through a C-terminal hydrophobic domain. We used infrared spectroscopy to study the secondary structure of a synthetic peptide (+3HN-188ILNVLVVLGVVLLGQFVVRRFFKS211-COO-) with the same sequence as the C-terminal domain of Bak. The spectrum of this peptide in D2O buffer shows an amide I′ band with a maximum at 1636 cm−1, which clearly indicates the predominance of an extended β-structure in aqueous solvent. However, the peptide incorporated in multilamellar dimyristoylphosphatidylcholine (DMPC) membranes shows a different amide I′ band spectrum, with a maximum at 1658 cm−1, indicating a predominantly α-helical structure induced by its interaction with the membrane. It was observed that through differential scanning calorimetry the transition of the phospholipid model membrane was broadened in the presence of the peptide. Fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene (DPH) in fluid DMPC vesicles showed that increasing concentrations of the peptide produced increased polarization values, which is compatible with the peptide being inserted into the membrane. High concentrations of the peptide considerably broaden the phase transition of DMPC multilamellar vesicles, and DPH polarization increased, especially at temperatures above the Tc transition temperature of the pure phospholipid. The addition of peptide destabilized unilamellar vesicles and released encapsulated carboxyfluorescein. These results indicate that this domain is able to insert itself into membranes, where it adopts an α-helical structure and considerably perturbs the physical properties of the membrane.  相似文献   

7.
Molecular dynamics simulation of the hydrated dimyristoylphosphatidylcholine (DMPC) bilayer membrane in the liquid-crystalline phase was carried out for 5 ns to study the interaction among DMPC headgroups in the membrane/water interface region. The phosphatidylcholine headgroup contains a positively charged choline group and negatively charged phosphate and carbonyl groups, although it is a neutral molecule as a whole. Our previous study (Pasenkiewicz-Gierula, M., Y. Takaoka, H. Miyagawa, K. Kitamura, and A. Kusumi. 1997. J. Phys. Chem. 101:3677-3691) showed the formation of water cross-bridges between negatively charged groups in which a water molecule is simultaneously hydrogen bonded to two DMPC molecules. Water bridges link 76% of DMPC molecules in the membrane. In the present study we show that relatively stable charge associations (charge pairs) are formed between the positively and negatively charged groups of two DMPC molecules. Charge pairs link 93% of DMPC molecules in the membrane. Water bridges and charge pairs together form an extended network of interactions among DMPC headgroups linking 98% of all membrane phospholipids. The average lifetimes of DMPC-DMPC associations via charge pairs, water bridges and both, are at least 730, 1400, and over 1500 ps, respectively. However, these associations are dynamic states and they break and re-form several times during their lifetime.  相似文献   

8.
The secondary structure of amyloid betaAP(25-35) peptide was studied in pure form and in the presence of different phospholipid vesicles, by using Fourier transform infrared spectroscopy (FT-IR). Pure peptide aggregated with time, forming fibrils with beta-structure. Phospholipid vesicles formed by negatively charged phospholipids such as 1,2-dimyristoyl-sn-glycerol-3-phospho-L-serine (Myr2PtdSer), 1,2-dimyristoyl-sn-glycerol-3-phospho-rac-1-glycerol (Myr2PtdGro) and 1,2-dimyristoyl-sn-glycerol 3-phosphate (Myr2PtdH), greatly accelerated the aggregation of the peptide. However, the presence of vesicles formed by the zwitterionic phospholipid, 1, 2-dimyristoyl-sn-glycerol-3-phosphocholine (Myr2PtdCho), slowed down the aggregation process. Differential scanning calorimetry (DSC) measurements showed that the effect of betaAP(25-35) on the gel to crystal liquid phase transition was small at neutral pH for negatively charged phospholipids and practically nil for Myr2PtdCho. In the case of Myr2PtdSer the effect was also zero at pH 9 but the effect was large at pH 3. The effect on Myr2PtdH was not, however, very dependent on pH. These results were fully confirmed by the observation through FT-IR of the change with temperature of the CH2 antisymmetric stretching vibration. The case of Myr2PtdGro was special as this phospholipid presents polymorphism giving solid quasicrystalline phases when it is not sufficiently hydrated, and it is remarkable that betaAP(25-35) was able to induce the formation of crystalline phases in samples prepared through a method which ensure a good hydration of phospholipid. These results show that the interaction of amyloid betaAP(25-35) peptide with phospholipids is based on electrostatic interactions, that these interactions favour the aggregation of the peptides, and that the presence of the aggregates may disturb the lipid-water interphase of the membrane.  相似文献   

9.
We investigated the effects of various concentrations of staphylococcal delta-lysin on the thermotropic phase behavior of large multilamellar dimyristoylphosphatidylcholine (DMPC) vesicles by differential scanning calorimetry (DSC), 31P nuclear magnetic resonance (NMR) and Fourier transform infrared (FTIR) spectroscopy, and X-ray diffraction. The DSC studies revealed that at all concentrations, the addition of delta-lysin progressively decreases the enthalpy of the pretransition of DMPC bilayers without significantly affecting its temperature or cooperativity. Similarly, the addition of smaller quantities of peptide has little effect on the temperature of the main phase transition of DMPC bilayers but does reduce the cooperativity and enthalpy of this transition somewhat. However, at higher peptide concentrations, a second phase transition with a slightly increased temperature and a markedly reduced cooperativity and enthalpy is also induced, and this latter phase transition resolves itself into two components at the highest peptide concentrations that are tested. Moreover, our 31P NMR spectroscopic studies reveal that at relatively low delta-lysin concentrations, essentially all of the phospholipid molecules produce spectra characteristic of the lamellar phase, whereas at the higher peptide concentrations, an increasing proportion exhibit an isotropic signal. Also, at the highest delta-lysin concentrations that are studied, the isotropic component of the 31P NMR spectrum also resolves itself into two components. At the highest peptide concentration that was tested, we are also able to effect a macroscopic separation of our sample into two fractions by centrifugation, a pellet containing relatively smaller amounts of delta-lysin and a supernatant containing larger amounts of peptide relative to the amount of lipid present. We are also able to show that the more cooperative phase transition detected calorimetrically, and the lamellar phase 31P NMR signal, arise from the pelleted material, while the less cooperative phase transition and the isotropic 31P NMR signal arise from the supernatant. In addition, we demonstrate by X-ray diffraction that the pelleted material corresponds to delta-lysin-containing large multilamellar vesicles and the supernatant to a mixture of delta-lysin-containing small unilamellar vesicles and discoidal particles. We also show by FTIR spectroscopy that delta-lysin exists predominantly in the alpha-helical conformation in aqueous solution or when interacting with DMPC, and that a large fraction of the peptide bonds undergo H-D exchange in D(2)O. However, upon interaction with DMPC, the fraction of exchangeable amide protons decreases. We also demonstrate by this technique that both of the phase transitions detected by DSC correspond to phospholipid hydrocarbon chain-melting phase transitions. Finally, we show by several techniques that the absolute concentrations of delta-lysin and the thermal history, as well as the lipid:peptide ratio, can affect the thermotropic phase behavior and morphology of peptide-lipid aggregates.  相似文献   

10.
The membrane-disruptive antimicrobial peptide PGLa is found to change its orientation in a dimyristoyl-phosphatidylcholine bilayer when its concentration is increased to biologically active levels. The alignment of the alpha-helix was determined by highly sensitive solid-state NMR measurements of (19)F dipolar couplings on CF(3)-labeled side chains, and supported by a nonperturbing (15)N label. At a low peptide/lipid ratio of 1:200 the amphiphilic peptide resides on the membrane surface in the so-called S-state, as expected. However, at high peptide concentration (>/=1:50 molar ratio) the helix axis changes its tilt angle from approximately 90 degrees to approximately 120 degrees , with the C-terminus pointing toward the bilayer interior. This tilted "T-state" represents a novel feature of antimicrobial peptides, which is distinct from a membrane-inserted I-state. At intermediate concentration, PGLa is in exchange between the S- and T-state in the timescale of the NMR experiment. In both states the peptide molecules undergo fast rotation around the membrane normal in liquid crystalline bilayers; hence, large peptide aggregates do not form. Very likely the obliquely tilted T-state represents an antiparallel dimer of PGLa that is formed in the membrane at increasing concentration.  相似文献   

11.
The amphiphilic alpha-helical peptide (KIAGKIA)3-NH2 (MSI-103) is a designer-made antibiotic, based on the natural sequence of PGLa from Xenopus laevis. Here, we have characterized the concentration-dependent alignment and dynamic behavior of MSI-103 in lipid membranes by solid-state 2H and 19F NMR, using orientational constraints from seven Ala-d3-labeled analogues and five 4-CF3-phenylglycine labels. As previously found for PGLa, MSI-103, too, assumes a flat surface-bound S-state alignment at low peptide concentrations, and it also realigns to a tilted T-state at higher concentrations. For PGLa, the stability of the T-state had been attributed to the specific assembly of antiparallel dimers; hence, it is remarkable that the artificial KIAGKIA repeat sequence can also dimerize in the same way in liquid crystalline lipid bilayers. Oriented circular dichroism analysis shows that for MSI-103 the threshold for realignment from the S-state to the T-state is approximately 3-fold lower than for PGLa (at a peptide-to-lipid ratio of 1:240 in dimyristoylphosphatidylcholine, compared to 1:80). Furthermore, MSI-103 becomes laterally immobilized in the lipid bilayer at a concentration ratio of 1:50, which occurs for PGLa only above 1:20. The superior antimicrobial activity of MSI-103 over PGLa thus appears to correlate with its stronger tendency to realign and self-assemble. The hemolytic activities of MSI-103 and its analogues, on the other hand, are shown here to correlate purely with the respective changes in hydrophobicity.  相似文献   

12.
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   

13.
The binding of dextran sulfates (DSs) with varying chain lengths to phosphatidylcholine multilamellar vesicles was investigated as a function of polyelectrolyte, NaCl, and Ca2+ concentration. Attractive forces between negatively charged polyelectrolytes and zwitterionic phospholipids arise from the assembly of calcium bridges. The formation of calcium bridges between the sulfate groups on the dextran sulfate and the phosphate group of the lipid results in increased calcium binding in mixtures of DS and 1, 2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC). At high NaCl concentration, the plateau adsorption of DS 500 is increased. The strength of dextran sulfate binding to DMPC is reflected in the changes of the 2H NMR quadrupolar splittings of the headgroup methylenes. Association forces increase with the number of calcium bridges formed. Low-molecular-weight DS does not bind to DMPC surfaces whereas longer-chain DSs strongly influence headgroup structure as a result of strong association. DS binding increases with increasing concentration; however, further association of the polyelectrolyte can be promoted only if negative charges are sufficiently screened. DS binding to lipid bilayers is a complicated balance of calcium bridging and charge screening. From our data we postulate that the structure of the adsorbed layer resembles a lattice of DS strands sandwiched between the bilayer lamellae.  相似文献   

14.
The interaction of water-soluble nonmembraneous proteins (trypsin and the basic pancreatic trypsin inhibitor (BPTI)) with soybean phospholipids was studied using multilamellar vesicles. Multilamellar vesicles were obtained from soybean lipid extracts and mixtures of individual phospholipids based on phosphatidylcholine. These mixtures contain different phospholipids: "bilayer", "nonbilayer", and negatively charged. It was shown that the content of both proteins in the complex depends on pH and the presence of negatively charged components. On the basis of this finding, the conclusion about the electrostatic nature of lipid-protein interaction was made. The structural organization of soybean phospholipids in multilamellar vesicles was studied in the presence and absence of the proteins using broad-line 31P-NMR spectroscopy. It was found that, in mixtures of phospholipids of complex composition, different types of phases coexist, and phospholipids of different classes can compensate the effects of each other. Trypsin and BPTI affect the structure of phospholipids in a similar way, inducing considerable structural changes in multilamellar vesicles of preparations containing negatively charged components in whose structure there coexisted primordially the bilayer and isotropic phases.  相似文献   

15.
Recombinant human interleukin-2 (rhIL-2) was incorporated in liposomes for potential therapeutic applications using a novel process. In this process, rhIL-2 caused the formation of large, unique multilamellar vesicles (MLVs) from small unilamellar vesicles (SUVs) of dimyristoylphosphatidylcholine (DMPC). Vesicle coalescence occurred most rapidly at 19 degrees C, between the pre- and main phase transition temperatures of DMPC, and showed a dependence upon pH (pH <5.5), ionic strength (>50 mM) and the initial size of the unilamellar vesicles (相似文献   

16.
In order to investigate structural and dynamical properties of local anesthetic articaine in a model lipid bilayer, a series of molecular dynamics simulations have been performed. Simulations were carried out for neutral and charged (protonated) forms of articaine inserted in fully hydrated dimyristoylphosphatidylcholine (DMPC) lipid bilayer. For comparison purpose, a fully hydrated DMPC bilayer without articaine was also simulated. The length of each simulation was 200 ns. Various properties of the lipid bilayer systems in the presence of both charged and uncharged forms of articaine taken at two different concentrations have been examined: membrane area per lipid, mass density distributions, order parameters, radial distribution functions, head group tilt, diffusion coefficients, electrostatic potential, etc, and compared with results of previous simulations of DMPC bilayer in the presence of lidocaine. It was shown that addition of both charged and neutral forms of articaine causes increase of the dipole electrostatic potential in the membrane interior.  相似文献   

17.
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   

18.
We have studied the interaction between three compounds which accumulate in organisms under hydration stress--proline, betaine, and trehalose--and the membrane phospholipids dimyristoylphosphatidylcholine (DMPC), palmitoyloleoylphosphatidylcholine (POPC), and dimyristoylphosphatidylethanolamine in bulk solution. Film balance studies reveal that these compounds increase the area/molecule of these lipids. Differential scanning calorimetry has been employed to investigate the effect these agents have on the gel-to-liquid crystalline phase transition of multilamellar and small unilamellar vesicles of DMPC, dipalmitoylphosphatidylcholine, and POPC:phosphatidylserine (90:10 mole ratio) in bulk solution. In the presence of 1 M proline, trehalose, or betaine, the midtransition temperature in small unilamellar vesicles is reduced (up to 7 degrees C in 1 M trehalose), and the transition broadened. In contrast, multilamellar vesicles of similar lipid composition show an increased transition temperature in the presence of the same concentration of these compounds. This result suggests that the inner lamellae in multilamellar vesicles may be dehydrated with only a few outer lamellae exposed to the protective compound. Finally, we have used stereomodels of phosphatidylcholine to investigate the mechanism of action of these agents. Hydrogen bonding of trehalose to the head group region results in an increase in the distance between head groups of 6.9 A. This amount of spreading compares well with data from the monolayer experiments which indicate that maximal spreading of DMPC monolayers by trehalose is 6.5 A. Molecular models of proline and betaine have also been constructed, and these models suggest potential interactions between these compounds and phosphatidylcholines. For the amphipath proline, this interaction may involve intercalation between phospholipid head groups.  相似文献   

19.
The acetylated and amidated hexapeptide FRWWHR (combi-2), previously identified by combinatorial chemistry methods, shows strong antimicrobial activity. The binding of the peptide to 1-palmitoyl-2-oleoyl-sn-glycero-3-[(phospho-rac-(1-glycerol)] (POPG) and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) vesicles was studied using fluorescence spectroscopy and isothermal titration calorimetry (ITC). Differential scanning calorimetry (DSC) with dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylglycerol (DPPG) multilamellar vesicles was performed to determine changes in the lipid phase behaviour upon binding the peptide. Two-dimensional proton nuclear magnetic resonance (NMR) spectroscopy, to solve the bound peptide structure, was performed in the presence of dodecylphosphatidylcholine (DPC) and sodium dodecyl sulphate (SDS) micelles. The fluorescence, ITC and DSC studies indicate that the peptide interacts preferentially with lipid vesicles containing negatively charged head groups. Conformational information determined using NMR indicate that the combi-2 peptide adopts a coiled amphipathic conformation when bound to SDS and DPC micelles. Leakage assays indicate that the peptide is not very efficient at causing leakage from calcein-filled large unilamellar vesicles comprised of POPG/POPC (1 : 1). The rapid passage of either the fluorescent-tagged peptides combi-2 or the previously studied peptide Ac-RRWWRF-NH(2) (combi-1) into Escherichia coli and Staphylococcus aureus suggests that instead of membrane disruption, the main bactericidal site of action of these peptides might be located inside bacteria.  相似文献   

20.
The amphipathic helical theory of Segrest and colleagues (FEBS Lett.:38:247-253, 1974) proposes that the lipid-binding segments of serum apolipoproteins are in an alpha helical conformation. Furthermore the helices have a hydrophobic face and a hydrophilic face with a specific distribution of positively and negatively charged residues. The importance of the pattern of the charged residues in the lipid binding and lecithin:cholesterol acyltransferase (LCAT) activation by the segments is still debated. We designed a 30-residue peptide, GALA, which in the alpha helical conformation has a hydrophilic face composed of glutamic acid residues (Sabbarao et al.: Biochemistry 26:2964-2972, 1987). GALA behaves like the serum apolipoproteins in its interaction with dimyristoylphosphatidylcholine (DMPC) at neutral pH; the amino terminal tryptophan of GALA undergoes a blue shift in its fluorescence emission spectrum, and the circular dichroism (CD) spectrum indicates that GALA acquires alpha helical structure in the presence of DMPC. A DMPC-GALA:19/1 (molar ratio) complex can be isolated by gel-permeation chromatography. This complex has a discoidal structure with the approximate dimensions of 44-A edge thickness and a 170- to 350-A diameter. GALA activates LCAT with DMPC but not with unsaturated phospholipids as the substrate. The apparent partition coefficient of GALA into DMPC vesicles is 100-fold larger than into egg phosphatidylcholine vesicles. The interaction of GALA with unsaturated lipids at neutral pH is so weak that no detectable change in the spectroscopic properties of GALA or the structure of the liposomes can be detected under the conditions used here. The sequence of GALA differs from previously studied model Apo A1 peptides by the absence of positively charged residues on the hydrophilic face. This indicates that positive charges in Apo A1-like peptides are not required in order to form discoidal structures with saturated phospholipids or to activate LCAT with such lipid substrates.  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