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 796 毫秒
1.
The effect of glucocerebroside (GlcCer) on the structure and thermotropic phase behavior of aqueous dispersions of palmitoyloleoylphosphatidylethanolamine (POPE) has been examined using simultaneous small-angle and wide-angle X-ray diffraction methods. Binary mixtures of GlcCer:POPE in molar ratios of 2:100, 5:100, 10:100, 20:100, 30:100, and 40:100 were examined in the temperature range 20-90 degrees C. Cubic phase has been observed in binary mixtures comprised of molar ratios greater than 5:100 in the temperature range of 60-90 degrees C upon heating at a rate of 2 degrees C/min. The cubic phase is relatively stable and coexists with inverted hexagonal or lamellar phases. It persists in the codispersions throughout subsequent cooling scans to 30 degrees C. The space group of the cubic phase is determined to be Pn3m or Pn3. The lattice constant of the Pn3m cubic phase was found to be almost constant when it coexists with lamellar liquid-crystal phase. Marked temperature-dependent changes were observed when cubic phase coexists with hexagonal phase or lamellar-gel phases. This is the first report of cubic phases formed by codispersions of glycosphingolipids and phospholipids. The mechanism of cubic phase formation and the interaction between GlcCer and POPE is discussed in terms of the putative biological functions of glycolipids.  相似文献   

2.
Domain formation in bacteria-mimetic membranes due to cationic peptide binding was recently proposed based on calorimetric data. We now use 2H solid-state NMR to critically examine the presence and absence of domains in bacterial membranes containing zwitterionic 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylethanolamine (POPE) and anionic 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylglycerol (POPG) lipids. Chain-perdeuterated POPE and POPG are used in single-component membranes, binary POPE/POPG (3:1) membranes, and membranes containing one of four cationic peptides: two antimicrobial peptides (AMPs) of the β-hairpin family of protegrin-1 (PG-1), and two cell-penetrating peptides (CPPs), HIV TAT and penetratin. 2H quadrupolar couplings were measured to determine the motional amplitudes of POPE and POPG acyl chains as a function of temperature. Homogeneously mixed POPE/POPG membranes should give the same quadrupolar couplings for the two lipids, whereas the presence of membrane domains enriched in one of the two lipids should cause distinct 2H quadrupolar couplings that reflect different chain disorder. At physiological temperature (308 K), we observed no or only small coupling differences between POPE and POPG in the presence of any of the cationic peptides. However, around ambient temperature (293 K), at which gel- and liquid-crystalline phases coexist in the peptide-free POPE/POPG membrane, the peptides caused distinct quadrupolar couplings for the two lipids, indicating domain formation. The broad-spectrum antimicrobial peptide PG-1 ordered ∼40% of the POPE lipids while disordering POPG. The Gram-negative selective PG-1 mutant, IB549, caused even larger differences in the POPE and POPG disorder: ∼80% of POPE partitioned into the ordered phase, whereas all of the POPG remained in the disordered phase. In comparison, TAT rigidified POPE and POPG similarly in the binary membrane at ambient temperature, indicating that TAT does not cause dynamic heterogeneity but interacts with the membrane with a different mechanism. Penetratin maintained the POPE order but disordered POPG, suggesting moderate domain separation. These results provide insight into the extent of domain formation in bacterial membranes and the possible peptide structural requirements for this phenomenon.  相似文献   

3.
Domain formation in bacteria-mimetic membranes due to cationic peptide binding was recently proposed based on calorimetric data. We now use 2H solid-state NMR to critically examine the presence and absence of domains in bacterial membranes containing zwitterionic 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylethanolamine (POPE) and anionic 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylglycerol (POPG) lipids. Chain-perdeuterated POPE and POPG are used in single-component membranes, binary POPE/POPG (3:1) membranes, and membranes containing one of four cationic peptides: two antimicrobial peptides (AMPs) of the β-hairpin family of protegrin-1 (PG-1), and two cell-penetrating peptides (CPPs), HIV TAT and penetratin. 2H quadrupolar couplings were measured to determine the motional amplitudes of POPE and POPG acyl chains as a function of temperature. Homogeneously mixed POPE/POPG membranes should give the same quadrupolar couplings for the two lipids, whereas the presence of membrane domains enriched in one of the two lipids should cause distinct 2H quadrupolar couplings that reflect different chain disorder. At physiological temperature (308 K), we observed no or only small coupling differences between POPE and POPG in the presence of any of the cationic peptides. However, around ambient temperature (293 K), at which gel- and liquid-crystalline phases coexist in the peptide-free POPE/POPG membrane, the peptides caused distinct quadrupolar couplings for the two lipids, indicating domain formation. The broad-spectrum antimicrobial peptide PG-1 ordered ∼40% of the POPE lipids while disordering POPG. The Gram-negative selective PG-1 mutant, IB549, caused even larger differences in the POPE and POPG disorder: ∼80% of POPE partitioned into the ordered phase, whereas all of the POPG remained in the disordered phase. In comparison, TAT rigidified POPE and POPG similarly in the binary membrane at ambient temperature, indicating that TAT does not cause dynamic heterogeneity but interacts with the membrane with a different mechanism. Penetratin maintained the POPE order but disordered POPG, suggesting moderate domain separation. These results provide insight into the extent of domain formation in bacterial membranes and the possible peptide structural requirements for this phenomenon.  相似文献   

4.
Nisin, a peptide used as a food preservative, is shown, by 31P-nuclear magnetic resonance and infrared spectroscopy, to perturb the structure of membranes formed of unsaturated phosphatidylethanolamine (PE) and to induce the formation of inverted non-lamellar phases. In the case of dioleoyl-sn-glycero-3-phosphatidylethanolamine (DOPE), nisin promotes the formation of inverted hexagonal phase. Similarly, the peptide induces the formation of an isotropic phase, most likely a cubic phase, with 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylethanolamine (POPE). It is proposed that the insertion of the peptide in the bilayer shifts the amphiphilic balance by increasing the hydrophobic contribution and is at the origin of the changes in the polymorphic propensities of PE. This is supported by the fact that the presence of cholesterol in the PE bilayer inhibits the power of nisin to perturb the membrane structure, most likely because the peptide insertion is difficult in the fluid ordered phase. This finding provides insight into possible antibacterial mechanisms of nisin.  相似文献   

5.
Docosahexaenoic acid (DHA), the longest and most unsaturated fatty acid commonly found in biological membranes, is known to affect various membrane properties. In a variety of cell membranes, DHA is primarily incorporated in phosphatidylethanolamines, where its function remains poorly understood. In order to understand the role of DHA in influencing membrane structure, we utilize (31)P NMR spectroscopy to study the phase behavior of 1-stearoyl-2-docosahexaenoyl-sn-glycerophosphoethanolamine (SDPE) in comparison to 1-palmitoyl-2-oleoyl-sn-glycerophosphoethanolamine (POPE) from 20 to 50 degrees C. Spectra of SDPE phospholipids show the formation of inverted hexagonal phase (H(II)) from 20 to 50 degrees C; in contrast, POPE mutilamellar dispersions exist in a lamellar liquid-crystalline phase (L(alpha)) at the same temperatures. The ability of SDPE to adopt nonbilayer phases at a physiological temperature may indicate its role in imparting negative curvature stress upon the membrane and may affect local molecular organization including the formation of lipid microdomains within biological membranes.  相似文献   

6.
WALP peptides consist of repeating alanine-leucine sequences of different lengths, flanked with tryptophan "anchors" at each end. They form membrane-spanning alpha-helices in lipid membranes, and mimic protein transmembrane domains. WALP peptides of increasing length, from 19 to 31 amino acids, were incorporated into N-monomethylated dioleoylphosphatidylethanolamine (DOPE-Me) at concentrations up to 0.5 mol % peptide. When pure DOPE-Me is heated slowly, the lamellar liquid crystalline (L(alpha)) phase first forms an inverted cubic (Q(II)) phase, and the inverted hexagonal (H(II)) phase at higher temperatures. Using time-resolved x-ray diffraction and slow temperature scans (1.5 degrees C/h), WALP peptides were shown to decrease the temperatures of Q(II) and H(II) phase formation (T(Q) and T(H), respectively) as a function of peptide concentration. The shortest and longest peptides reduced T(Q) the most, whereas intermediate lengths had weaker effects. These findings are relevant to membrane fusion because the first step in the L(alpha)/Q(II) phase transition is believed to be the formation of fusion pores between pure lipid membranes. These results imply that physiologically relevant concentrations of these peptides could increase the susceptibility of biomembrane lipids to fusion through an effect on lipid phase behavior, and may explain one role of the membrane-spanning domains in the proteins that mediate membrane fusion.  相似文献   

7.
The structural effects of the fusion peptide of feline leukemia virus (FeLV) on the lipid polymorphism of N-methylated dioleoylphosphatidylethanolamine were studied using a temperature ramp with sequential X-ray diffraction. This peptide, the hydrophobic amino-terminus of p15E, has been proven to be fusogenic and to promote the formation of highly curved, intermediate structures on the lamellar liquid-crystal to inverse hexagonal phase transition pathway. The FeLV peptide produces marked effects on the thermotropic mesomorphic behaviour of MeDOPE, a phospholipid with an intermediate spontaneous radius of curvature. The peptide is shown to reduce the lamellar repeat distance of the membrane prior to the onset of an inverted cubic phase. This suggests that membrane thinning may play a role in peptide-induced membrane fusion and strengthens the link between the fusion pathway and inverted cubic phase formation. The results of this study are interpreted in relation to models of the membrane fusion mechanism.  相似文献   

8.
C Czeslik  R Winter  G Rapp    K Bartels 《Biophysical journal》1995,68(4):1423-1429
We used x-ray and neutron diffraction to study the temperature- and pressure-dependent structure and phase behavior of the monoacylglycerides 1-monoelaidin (ME) and 1-monoolein (MO) in excess water. The monoacylglycerides were chosen for investigation of their phase behavior because they exhibit mesomorphic phases with one-, two-, and three-dimensional periodicity, such as lamellar, an inverted hexagonal and bicontinuous cubic phases, in a rather easily accessible temperature and pressure range. We studied the structure, stability, and transformations of the different phases over a wide temperature and pressure range, explored the epitaxial relations that exist between different phases, and established a relationship between the chemical structure of the lipid molecules and their phase behavior. For both systems, a temperature-pressure phase diagram has been determined in the temperature range from 0 to 100 degrees C at pressures from ambient up to 1400 bar, and drastic differences in phase behavior are found for the two systems. In MO-water dispersions, the cubic phase Pn3m extends over a large phase field in the T,p-plane. At temperatures above 95 degrees C, the inverted hexagonal phase is found. In the lower temperature region, a crystalline lamellar phase is induced at higher pressures. The phases found in ME-water include the lamellar crystalline Lc phase, the L beta gel phase, the L alpha liquid-crystalline phase, and two cubic phases belonging to the crystallographic space groups Im3m and Pn3m. In addition, the existence of metastable phases has been exploited. Between coexisting metastable cubic structures, a metric relationship has been found that is predicted theoretically on the basis of the curvature elastic energy approximation only.  相似文献   

9.
In this work, we present the first characterization of the cell lysing mechanism of MSI-78, an antimicrobial peptide. MSI-78 is an amphipathic alpha-helical peptide designed by Genaera Corporation as a synthetic analog to peptides from the magainin family. (31)P-NMR of mechanically aligned samples and differential scanning calorimetry (DSC) were used to study peptide-containing lipid bilayers. DSC showed that MSI-78 increased the fluid lamellar to inverted hexagonal phase transition temperature of 1,2-dipalmitoleoyl-phosphatidylethanolamine indicating the peptide induces positive curvature strain in lipid bilayers. (31)P-NMR of lipid bilayers composed of MSI-78 and 1-palmitoyl-2-oleoyl-phosphatidylethanolamine demonstrated that the peptide inhibited the fluid lamellar to inverted hexagonal phase transition of 1-palmitoyl-2-oleoyl-phosphatidylethanolamine, supporting the DSC results, and the peptide did not induce the formation of nonlamellar phases, even at very high peptide concentrations (15 mol %). (31)P-NMR of samples containing 1-palmitoyl-2-oleoyl-phosphatidylcholine and MSI-78 revealed that MSI-78 induces significant changes in the bilayer structure, particularly at high peptide concentrations. At lower concentrations (1-5%), the peptide altered the morphology of the bilayer in a way consistent with the formation of a toroidal pore. Higher concentrations of peptide (10-15%) led to the formation of a mixture of normal hexagonal phase and lamellar phase lipids. This work shows that MSI-78 induces significant changes in lipid bilayers via positive curvature strain and presents a model consistent with both the observed spectral changes and previously published work.  相似文献   

10.
M Caffrey 《Biochemistry》1987,26(20):6349-6363
A study of the dynamics and mechanism of the various thermotropic phase transitions undergone by the hydrated monoacylglycerides monoolein and monoelaidin, in the temperature range of 20-120 degrees C and from 0 to 5 M NaCl, has been undertaken. Measurements were made by using time-resolved X-ray diffraction at the Cornell High-Energy Synchrotron Source. The lamellar chain order/disorder, lamellar/cubic (body centered, space group No. 8), cubic (body centered, No. 8)/cubic (primitive, No. 4), cubic (body centered, No. 12)/cubic (primitive, No. 4), cubic (primitive, No. 4)/fluid isotropic, cubic (body centered, No. 12)/inverted hexagonal, cubic (primitive, No. 4)/inverted hexagonal, and hexagonal/fluid isotropic transitions were examined under active heating and passive cooling by using a jump in temperature to effect phase transformation. All of the transitions with the exception of the cubic (body centered, No. 8)/cubic (primitive, No. 4) and the cubic (body centered, No. 12)/cubic (primitive, No. 4) cooling transitions were found (1) to be repeatable, (2) to be reversible, and (3) to have an upper bound on the transit time (time required to complete the transition) of less than or equal to 3 s. The shortest transit times recorded for the various phase changes in the heating direction were less than or equal to 1.9 (lamellar chain melting), less than or equal to 1.7 [lamellar liquid crystal/cubic (body (body centered, No. 8)], less than or equal to 0.5 [cubic (body centered, No. 8)/cubic (primitive, No. 4)], less than or equal to 0.9 [cubic (primitive, No. 4)/hexagonal], less than or equal to 1.3 [cubic (body centered, No. 12)/cubic (primitive, No. 4) and cubic (body centered, No. 12)/hexagonal], and less than or equal to 0.6 s (hexagonal/fluid isotropic). For the exceptions noted above, the transitions were slow with transit times ranging from 0.5 to 30 min and displayed pronounced hysteresis and/or undercooling. Regardless of the direction of the transitions, all but one appear to be two state to within the sensitivity limits of the time-resolved method. In the case of the lamellar liquid crystal/cubic (body centered, No. 8) transition a stable intermediate of unknown identity was apparent. In addition to the time-resolved measurements, data were obtained on the stability of the various phases in the temperature range of 20-120 degrees C and from 0 to 5 M NaCl. In the case of fully hydrated monoolein, high salt strongly favors the hexagonal over the cubic (body centered, No. 8) phase and slightly elevates the hexagonal/fluid isotropic transition temperature.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
M D King  D Marsh 《Biochemistry》1989,28(13):5643-5647
The polymorphic phase behavior of 1-palmitoyl-2-lyso-sn-glycero-3-phosphocholine dispersions in excess water has been studied as a function of temperature and poly(ethylene glycol) (PEG) concentration, using proton dipolar-decoupled 31P NMR spectroscopy and turbidity measurements. The phase behavior was found to depend on both lipid concentration and PEG concentration, and most of the NMR experiments were conducted at a lipid concentration of 15 mg/mL. At low PEG concentrations (0-12 wt %), a thermotropic transition occurs at 3-5 degrees C with increasing temperature, from an interdigitated lamellar gel (L beta i) phase to a normal micellar phase. At intermediate PEG concentrations (12-20 wt %), thermotropic transitions take place with increasing temperature, first from the lamellar gel phase to a fluid cubic (Q alpha) phase and then at higher temperatures from the cubic phase to the micellar phase. At intermediate PEG concentrations above the former range (20-30 wt %), thermotropic transitions take place with increasing temperature, first from the lamellar gel phase to the cubic phase, then from the cubic phase to a normal hexagonal (HI) phase, and finally from the hexagonal phase to the micellar phase. At high PEG concentrations (greater than 30 wt %), a thermotropic transition takes place with increasing temperature from the lamellar gel phase directly to the fluid hexagonal phase. At these high PEG concentrations, the micellar phase is not attained within the accessible temperature range (less than or equal to 90 degrees C).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Salt-induced fluid lamellar (L alpha) to inverted hexagonal (HII) phase transitions have been studied in diphosphatidylglycerols (cardiolipins) with different acyl chain compositions, using 31P nuclear magnetic resonance (NMR) spectroscopy. Cardiolipins with four myristoyl chains, tetramyristoyl cardiolipin (TMCL), and with four oleoyl chains, tetraoleoyl cardiolipin (TOCL), were synthesized chemically. TMCL was found to undergo a thermotropic lamellar gel to lamellar liquid-crystalline phase transition at 33-35 degrees C. This lipid exhibited an axially symmetric 31P-NMR spectrum corresponding to a lamellar phase at all NaCl concentrations between 0 and 6 M. In the case of TOCL, formation of an HII phase was induced by salt concentrations of 3.5 M NaCl or greater. These observations, taken together with earlier findings that bovine heart cardiolipin aqueous dispersions adopt an HII phase at salt concentrations of 1.5 M NaCl or greater, indicate that increasing unsaturation and length of the acyl chains favour formation of the HII phase in diphosphatidylglycerols.  相似文献   

13.
The mechanism of the lamellar/inverted cubic (QII) phase transition is related to that of membrane fusion in lipid systems. N-Monomethylated dioleoylphosphatidylethanolamine (DOPE-Me) exhibits this transition and is commonly used to investigate the effects of exogenous substances, such as viral fusion peptides, on the mechanism of membrane fusion. We studied DOPE-Me phase behavior as a first step in evaluating the effects of membrane-spanning peptides on inverted phase formation and membrane fusion. These measurements show that: a) the onset temperatures for QII and inverted hexagonal (HII) phase formation both are temperature scan rate-dependent; b) longer pre-incubation times at low temperature and lower temperature scan rates favor formation of the QII phase; and c) in temperature-jump experiments between 61 and 65°C, the meta-stable HII phase forms initially, and disappears slowly while the QII phase develops. These observations are rationalized in the context of a mechanism for both the lamellar/non-lamellar phase transition and the related process of membrane fusion. Current address for D.P.S.: Givaudan, Cincinnati, OH 45216 Data Deposition: Relevant transition temperatures in this paper have been deposited in the LIPIDAT ( )  相似文献   

14.
Peptidyl-glycine-leucine-carboxyamide (PGLa), isolated from granular skin glands of Xenopus laevis, is practically devoid of secondary structure in aqueous solution and in the presence of zwitterionic phospholipids, when added exogenously, but adopts an α-helix in the presence of anionic lipids. The peptide was shown to exhibit antifungal activity and to have antimicrobial activity towards both Gram-negative and Gram-positive bacteria. As a broad variety of peptides is found in the secretions of amphibian skin combinatorial treatment of PGLa and magainin 2 was studied showing enhanced activity by a heterodimer formation. Thus production of mutually recognizing peptides seems to be an effective way in nature to increase selective membrane activity. Biophysical studies on membrane mimics demonstrated that PGLa can discriminate between different lipid species, preferentially interacting with negatively charged lipids, which are major components of bacterial but not mammalian cell membranes. This emphasizes the role of electrostatic interactions as a major determinant to trigger the affinity of antimicrobial peptides towards bacterial membranes. PGLa induced the formation of a quasi-interdigitated phase in phosphatidylglycerol bilayers below their chain melting transition, which is due to the creation of voids below the peptide being aligned parallel to the membrane surface. In the fluid phase of phosphatidylglycerol the peptide inserts perpendicularly into the bilayer above a threshold concentration, which results in a hydrophobic mismatch of the peptide length and bilayer core for lipids ≤ C16. This mismatch is compensated by stretching of the acyl chains and in turn thickening of the bilayer demonstrating that membrane thinning cannot be taken generally as the hallmark of pore formation by antimicrobial peptides. Furthermore, PGLa was shown to affect membrane curvature strain of phosphatidylethanolamine, another main lipid component of bacterial membranes, where a cubic phase coexists with the fluid bilayer phase. Investigations on living Escherichia coli showed distinct changes in cell envelope morphology, when treated with the peptide. In a first stage loss of surface stiffness and consequently of topographic features was observed, followed in a second stage by permeabilization of the outer membrane and rupture of the inner (cytoplasmic) membrane supposedly by the mechanism(s) derived from model studies.  相似文献   

15.
We have investigated the effects of the model alpha-helical transmembrane peptide Ac-K(2)L(24)K(2)-amide (L(24)) on the thermotropic phase behavior of aqueous dispersions of 1,2-dielaidoylphosphatidylethanolamine (DEPE) to understand better the interactions between lipid bilayers and the membrane-spanning segments of integral membrane proteins. We studied in particular the effect of L(24) and three derivatives thereof on the liquid-crystalline lamellar (L(alpha))-reversed hexagonal (H(II)) phase transition of DEPE model membranes by differential scanning calorimetry and (31)P nuclear magnetic resonance spectroscopy. We found that the incorporation of L(24) progressively decreases the temperature, enthalpy, and cooperativity of the L(alpha)-H(II) phase transition, as well as induces the formation of an inverted cubic phase, indicating that this transmembrane peptide promotes the formation of inverted nonlamellar phases, despite the fact that the hydrophobic length of this peptide exceeds the hydrophobic thickness of the host lipid bilayer. These characteristic effects are not altered by truncation of the side chains of the terminal lysine residues or by replacing each of the leucine residues at the end of the polyleucine core of L(24) with a tryptophan residue. Thus, the characteristic effects of these transmembrane peptides on DEPE thermotropic phase behavior are independent of their detailed chemical structure. Importantly, significantly shortening the polyleucine core of L(24) results in a smaller decrease in the L(alpha)-H(II) phase transition temperature of the DEPE matrix into which it is incorporated, and reducing the thickness of the host phosphatidylethanolamine bilayer results in a larger reduction in the L(alpha)-H(II) phase transition temperature. These results are not those predicted by hydrophobic mismatch considerations or reported in previous studies of other transmembrane alpha-helical peptides containing a core of an alternating sequence of leucine and alanine residues. We thus conclude that the hydrophobicity and conformational flexibility of transmembrane peptides can affect their propensity to induce the formation of inverted nonlamellar phases by mechanisms not primarily dependent on lipid-peptide hydrophobic mismatch.  相似文献   

16.
The effect of alpha-tocopherol on the thermotropic phase transition behaviour of aqueous dispersions of dimyristoylphosphatidylethanolamine was examined using synchrotron X-ray diffraction methods. The temperature of gel to liquid-crystalline (Lbeta-->Lalpha) phase transition decreases from 49.5 to 44.5 degrees C and temperature range where gel and liquid-crystalline phases coexist increases from 4 to 8 degrees C with increasing concentration of alpha-tocopherol up to 20 mol%. Codispersion of dimyristoylphosphatidylethanolamine containing 2.5 mol% alpha-tocopherol gives similar lamellar diffraction patterns as those of the pure phospholipid both in heating and cooling scans. With 5 mol% alpha-tocopherol in the phospholipid, however, an inverted hexagonal phase is induced which coexists with the lamellar gel phase at temperatures just before transition to liquid-crystalline lamellar phase. The presence of 10 mol% alpha-tocopherol shows a more pronounced inverted hexagonal phase in the lamellar gel phase but, in addition, another non-lamellar phase appears with the lamellar liquid-crystalline phase at higher temperature. This non-lamellar phase coexists with the lamellar liquid-crystalline phase of the pure phospholipid and can be indexed by six diffraction orders to a cubic phase of Pn3m or Pn3 space groups and with a lattice constant of 12.52+/-0.01 nm at 84 degrees C. In mixed aqueous dispersions containing 20 mol% alpha-tocopherol, only inverted hexagonal phase and lamellar phase were observed. The only change seen in the wide-angle scattering region was a transition from sharp symmetrical diffraction peak at 0.43 nm, typical of gel phases, to broad peaks centred at 0.47 nm signifying disordered hydrocarbon chains in all the mixtures examined. Electron density calculations through the lamellar repeat of the gel phase using six orders of reflection indicated no difference in bilayer thickness due to the presence of 10 mol% alpha-tocopherol. The results were interpreted to indicate that alpha-tocopherol is not randomly distributed throughout the phospholipid molecules oriented in bilayer configuration, but it exists either as domains coexisting with gel phase bilayers of pure phospholipid at temperatures lower than Tm or, at higher temperatures, as inverted hexagonal phase consisting of a defined stoichiometry of phospholipid and alpha-tocopherol molecules.  相似文献   

17.
Su Y  Waring AJ  Ruchala P  Hong M 《Biochemistry》2011,50(12):2072-2083
The structural basis for the gram selectivity of two disulfide-bonded β-hairpin antimicrobial peptides (AMPs) is investigated using solid-state nuclear magnetic resonance (NMR) spectroscopy. The hexa-arginine PG-1 exhibits potent activities against both gram-positive and gram-negative bacteria, while a mutant of PG-1 with only three cationic residues maintains gram-positive activity but is 30-fold less active against gram-negative bacteria. We determined the topological structure and lipid interactions of these two peptides in a lipopolysaccharide (LPS)-rich membrane that mimics the outer membrane of gram-negative bacteria and in the POPE/POPG membrane, which mimics the membrane of gram-positive bacteria. (31)P NMR line shapes indicate that both peptides cause less orientational disorder in the LPS-rich membrane than in the POPE/POPG membrane. (13)C chemical shifts and (13)C-(1)H dipolar couplings show that both peptides maintain their β-hairpin conformation in these membranes and are largely immobilized, but the mutant exhibits noticeable intermediate-time scale motion in the LPS membrane at physiological temperature, suggesting shallow insertion. Indeed, (1)H spin diffusion from lipid chains to the peptides shows that PG-1 fully inserts into the LPS-rich membrane whereas the mutant does not. The (13)C-(31)P distances between the most hydrophobically embedded Arg of PG-1 and the lipid (31)P are significantly longer in the LPS membrane than in the POPE/POPG membrane, indicating that PG-1 does not cause toroidal pore defects in the LPS membrane, in contrast to its behavior in the POPE/POPG membrane. Taken together, these data indicate that PG-1 causes transmembrane pores of the barrel-stave type in the LPS membrane, thus allowing further translocation of the peptide into the inner membrane of gram-negative bacteria to kill the cells. In comparison, the less cationic mutant cannot fully cross the LPS membrane because of weaker electrostatic attractions, thus causing weaker antimicrobial activities. Therefore, strong electrostatic attraction between the peptide and the membrane surface, ensured by having a sufficient number of Arg residues, is essential for potent antimicrobial activities against gram-negative bacteria. The data provide a rational basis for controlling gram selectivity of AMPs by adjusting the charge densities.  相似文献   

18.
用小角 X 射线散射(SAXS)法研究了苯、丙酮对磷脂酰乙醇胺(PE)液晶态结构的影响.实验结果表明:苯能使 PE 液晶态先从片层相解束变成立方相.然后再诱导立方相变成六角形 H相,最后促使六角形 H相解束变成液态.在解束相变中出现了红移现象.丙酮也有使 PE 液晶态先从片层相解束变成立方相,在解束相变中也出现红移现象,但再诱导立方相变成六角形 H相之后,不是促使其变成液相,而是将其稳定在六角形 H相.  相似文献   

19.
H Takahashi  I Hatta    P J Quinn 《Biophysical journal》1996,70(3):1407-1411
The structures of fully hydrated 1:1 and 1:2 (mol/mol) dipalmitoylphosphatidylcholine (DPPC)-dipalmitoylglycerol (DPG) mixtures were studied by means of small-angle x-ray diffraction. The x-ray diffraction pattern of the 1:1 (mol/mol) DPPC-DPG mixture at 65 degrees C contains three reflections with spacings in the ratio of 1:1/ square root of 2:1/ square root of 3 in addition to reflections of an inverted hexagonal (H11) phase. A possible interpretation of this result is that a cubic phase of the body-centered space group lm3m, with a lattice constant of 23.1 +/- 0.6 nm, is formed. This cubic phase appears at intermediate temperatures between the lamellar and the H11 phases. The 1:2 (mol/mol) DPPC-DPG mixture gives an x-ray diffraction pattern at temperatures higher than the lamellar-to-H11 transition containing a number of reflections that index a cubic phase structure. The space group of the cubic phase was assigned a face-centered group Fd3m with a lattice constant of 16.3 +/- 0.1 nm at 82 degrees C. The possible role of cubic phases in membrane phenomena such as transmembrane signal transduction and fusion is discussed.  相似文献   

20.
We report a molecular dynamics simulation of the phase transition of monoolein from an inverted cubic phase to an inverted hexagonal phase. The transition proceeds via an intermediate structure consisting of water channels in a cubic geometry, in agreement with the predictions of the modified stalk theory (Siegel, 1999). Two mechanisms are identified by which the topology changes during the transition. Bilayer fusion proceeds via the formation of trans-monolayer contacts, whereas bilayer rupture is observed as a gradual thinning of each monolayer.  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