共查询到20条相似文献,搜索用时 15 毫秒
1.
Cationic, O-alkylphosphatidylcholines, recently developed as DNA transfection agents, form bilayers indistinguishable from those of natural phospholipids and undergo fusion with anionic bilayers. Membrane merging (lipid mixing), contents release, and contents mixing between populations of positive vesicles containing O-ethylphosphatidylcholine (EDOPC) and negative vesicles containing dioleolylphosphatidylglycerol (DOPG) have been determined with standard fluorometric vesicle-population assays. Surface-charge densities were varied from zero to full charge. All interactions depended critically on surface-charge density, as expected from the adhesion-condensation mechanism. Membrane mixing ranged from zero to 100%, with significant mixing (>10 <70%) occurring between cationic vesicles that were fully charged and anionic vesicles that had fractional surface charges as low as 0.1. Such mixing with membranes as weakly charged as cell membranes should be relevant to transfection with cationic lipids. Unexpectedly, lipid mixing was higher at high than at low ionic strength when one lipid dispersion was prepared from EDOPC plus DOPG (in different proportions), especially when the other vesicles were of EDOPC; this may somehow be a consequence of the ability of the former mixture to assume non-lamellar phases. Leakage of aqueous contents was also a strong function of charge, with fully charged vesicles releasing essentially all of their contents less than 1 min after mixing. EDOPC was more active in this regard than was DOPG, which probably reflects stronger intermolecular interactions of DOPG. Fusion, as measured by contents mixing, exhibited maximal values of 10% at intermediate surface charge. Reduced fusion at higher charge is attributed to multiple vesicle interactions leading to rupture. The existence of previously published data on individual interactions of vesicles of the same composition made it possible for the first time to compare pairwise with population interactions, confirming the likelihood of population studies to overestimate rupture and hemifusion and underestimate true vesicle fusion. 相似文献
2.
Equilibrium Theory and Geometrical Constraint Equation for Two-Component Lipid Bilayer Vesicles 总被引:1,自引:0,他引:1
This paper aims at the general mathematical framework for the equilibrium theory of two-component lipid bilayer vesicles.
To take into account the influences of the local compositions together with the mean curvature and Gaussian curvature of the
membrane surface, a general potential functional is constructed. We introduce two kinds of virtual displacement modes: the
normal one and the tangential one. By minimizing the potential functional, the equilibrium differential equations and the
boundary conditions of two-component lipid vesicles are derived. Additionally, the geometrical constraint equation and geometrically
permissible condition for the two-component lipid vesicles are presented. The physical, mathematical, and biological meanings
of the equilibrium differential equations and the geometrical constraint equations are discussed. The influences of physical
parameters on the geometrically permissible phase diagrams are predicted. Numerical results can be used to explain recent
experiments. 相似文献
3.
Equilibrium shapes of vesicles composed of a mixture of partially miscible amphiphiles are investigated. To take into account the influences of the composition, a simple phenomenological coupling between the co mposition and the curvatures, including the mean curvature and the Gauss curvature of the membrane surface, is suggested. By minimizing the potential functional, the general shape equation is obtained and solved analytically for vesicles with simple shapes. Besides, the geometrical constraint equation and geometrically permissible condition for the two-component lipid vesicles are put forward. The influences of physical parameters on the geometrically permissible phase diagrams are predicted. The close relations between the predictions and existing experimental phenomena published recently are shown. 相似文献
4.
5.
Rafael G. Oliveira Emanuel Schneck Sergio S. Funari Bruno Maggio 《Biophysical journal》2010,99(5):1500-1509
Purified myelin can be spread as monomolecular films at the air/aqueous interface. These films were visualized by fluorescence and Brewster angle microscopy, showing phase coexistence at low and medium surface pressures (<20-30 mN/m). Beyond this threshold, the film becomes homogeneous or not, depending on the aqueous subphase composition. Pure water as well as sucrose, glycerol, dimethylsulfoxide, and dimethylformamide solutions (20% in water) produced monolayers that become homogeneous at high surface pressures; on the other hand, the presence of salts (NaCl, CaCl2) in Ringer's and physiological solution leads to phase domain microheterogeneity over the whole compression isotherm. These results show that surface heterogeneity is favored by the ionic milieu. The modulation of the phase-mixing behavior in monolayers is paralleled by the behavior of multilamellar vesicles as determined by small-angle and wide-angle x-ray scattering. The correspondence of the behavior of monolayers and multilayers is achieved only at high surface pressures near the equilibrium adsorption surface pressure; at lower surface pressures, the correspondence breaks down. The equilibrium surface tension on all subphases corresponds to that of the air/alkane interface (27 mN/m), independently on the surface tension of the clean subphase. 相似文献
6.
Reihaneh Ramezani Majid Sadeghizadeh Mehrdad Behmanesh Saman Hosseinkhani 《Molecular biotechnology》2013,55(2):120-130
Entrapment of plasmid DNA (pDNA) in an aqueous compartment separated from the bulk external aqueous medium by a phospholipid bilayer resembles a structure similar to a primitive living cell, and interestingly, this phenomenon occurs completely self-assembled. Being inspired by such a structure as well as using the dehydration–rehydration technique, we were able to encapsulate pDNA without using multivalent cations and with high efficiency (98 %) into noncationic lipid bilayer vesicles. These liposomes which were composed of dimyristoyl-sn-glycero-3-phosphocholine unlike cationic liposomes, were nontoxic. The obtained liposome structure was able protect DNA against nuclease and was completely stable, in a way that even after 6 months, it still kept the pDNA in its structure, and there was a small change in its size (100–150 nm) determined by dynamic light scattering. The purpose of this research is to polarize the researchers’ interest toward utilization of neutral liposomes originating from the cell membrane as the most efficient carrier for gene delivery. We indicated that in using such carriers, which are the most similar synthetic structures to viruses, their inability in electrostatic interaction with DNA would not be an obstacle for entrapping nucleic acids. 相似文献
7.
Julie M. Wolf Javier Espadas-Moreno Jose L. Luque-Garcia Arturo Casadevall 《Eukaryotic cell》2014,13(12):1484-1493
Cryptococcus neoformans produces extracellular vesicles containing a variety of cargo, including virulence factors. To become extracellular, these vesicles not only must be released from the plasma membrane but also must pass through the dense matrix of the cell wall. The greatest unknown in the area of fungal vesicles is the mechanism by which these vesicles are released to the extracellular space given the presence of the fungal cell wall. Here we used electron microscopy techniques to image the interactions of vesicles with the cell wall. Our goal was to define the ultrastructural morphology of the process to gain insights into the mechanisms involved. We describe single and multiple vesicle-leaving events, which we hypothesized were due to plasma membrane and multivesicular body vesicle origins, respectively. We further utilized melanized cells to “trap” vesicles and visualize those passing through the cell wall. Vesicle size differed depending on whether vesicles left the cytoplasm in single versus multiple release events. Furthermore, we analyzed different vesicle populations for vesicle dimensions and protein composition. Proteomic analysis tripled the number of proteins known to be associated with vesicles. Despite separation of vesicles into batches differing in size, we did not identify major differences in protein composition. In summary, our results indicate that vesicles are generated by more than one mechanism, that vesicles exit the cell by traversing the cell wall, and that vesicle populations exist as a continuum with regard to size and protein composition. 相似文献
8.
《Journal of liposome research》2013,23(2):311-326
AbstractThe interaction between myoglobin and negatively-charged liposomes composed of phosphatidylcholine/phosphatidylglycerol (1:1) was studied at low ionic strength under acidic conditions. Changes in the absorbance and the fluorescence spectra of myoglobin were recorded upon addition of liposomes to partially unfolded (pH 3.5) and native (pH 4.5 and pH 6.5) myoglobin. Association of myoglobin with liposomes was a relatively fast process at pH 3.5 and pH 4.5. Although at pH 3.5 myoglobin was unfolded partially before the addition of the liposomes while at pH 4.5 before the addition of liposomes myoglobin retained its native form, similar interaction patterns of myoglobin with liposomes were observed. The fluorescence and absorption spectra in the Soret region of myoglobin clearly indicated that at these pH values myoglobin was associated with the liposomes in a (partially) unfolded state. At pH 6.5 the kinetics of myoglobin association with liposomes was much slower than at pH 3.5 and 4.5. The spectroscopic measurements also indicated that the interaction of myoglobin with liposomes at pH 6.5 followed a different pattern and resulted in different protein structures in comparison with pH 3.5/4.5. 相似文献
9.
Background
Protein transport across cellular membranes is an important aspect of toxin biology. Escherichia coli cell killing by nuclease colicins occurs through DNA (DNases) or RNA (RNases) hydrolysis and to this end their cytotoxic domains require transportation across two sets of membranes. In order to begin to unravel the molecular mechanisms underlying the membrane translocation of colicin nuclease domains, we have analysed the membrane association of four DNase domains (E9, a charge reduction E9 mutant, E8, and E7) and one ribosomal RNase domain (E3) using a biomembrane model system.Principal Results
We demonstrate, through the use of large unilamellar vesicles composed of synthetic and E. coli lipids and a membrane surface potential sensor, that the colicin nuclease domains bind anionic membranes only, with micromolar affinity and via a cooperative binding mechanism. The evaluation of the nuclease bilayer insertion depth, through a fluorescence quenching analysis using brominated lipids, indicates that the nucleases locate to differential regions in the bilayer. Colicin DNases target the interfacial region of the lipid bilayer, with the DNase E7 showing the deepest insertion, whereas the ribosomal RNase E3 penetrates into the hydrophobic core region of the bilayer. Furthermore, the membrane association of the DNase E7 and the ribosomal RNase E3 induces vesicle aggregation, lipid mixing and content leakage to a much larger extent than that of the other DNases analysed.Conclusions/Significance
Our results show, for the first time, that after the initial electrostatically driven membrane association, the pleiotropic membrane effects induced by colicin nuclease domains relate to their bilayer insertion depth and may be linked to their in vivo membrane translocation. 相似文献10.
The kinetics of absorbance and fluorescence changes of cytochrome c as induced by an aqueous solution of the anionic surfactant sodium dodecyl sulfate (SDS) or sodium bis(2-ethylhexyl)sulfosuccinate (AOT) are studied. The results are compared with far-UV circular dichroism (CD) spectra. Both surfactants cause similar alterations in the secondary structure of cytochrome c, while their influence on the heme environment of cytochrome c is different. In the presence of AOT below and above critical micellar concentration a conversion of the low-spin native cytochrome c to a denatured low-spin protein not having methionine ligand takes place. In the presence of SDS micelles conversion of the native protein to a denatured mixed-spin form occurs. The changes in the heme group induced by both surfactants occur independently of the alterations in tertiary structure. 相似文献
11.
《Journal of liposome research》2013,23(1):35-62
AbstractThe effects of adriamycin on the organization of dipalmitoylphosphatidylcholine (DPPC) membranes alone or in the presence of 1 mol% cardiolipin (CL) in the form of single and multibilayer vesicles has been studied by spin-labeling, and by high sensitivity differential scanning calorimetry. With sonicated small unilamellar vesicles (SUVs), adriamycin increased the order parameter of 5-doxylstearate spin-labeled vesicles, an effect primarily observed in the gel phase of DPPC. thermal transition profiles, obtained by high-sensitivity differential scanning calorimetry and by 2,2,6,6-tetramethylpiperidinyl-l-oxy (TEMPO) partitioning studies, indicated that the drug induced aggregation and fusion of SUVs, especially at high drug-lipid molar ratios, and this phenomenon was further verified by negative-staining electron microscopy. the presence of J mol% CL in the lecithin bilayer markedly enhanced the effect of adriamycin on membrain order and fusion, especially under conditions of low ionic strength. the ordering effect of the drug was insensitive to the presence of other acidic phospholipids and was only partially inhibited by Ca2+ or Mg2+. In contrast to the small and highly-curved SUVs, however, the phase behavior of fused unilamellar vesicles (FUVs) or multilamellar vesicles (MLVs) was not significantly affected by the drug, suggesting that the bilayer curvature is an important factor in the interaction of the antibiotic with the corresponding bilayers. these findings demonstrate that changes in the bilayer packing configuration, due to differences in the radii of the curvature of the vesicles, must be considered in studies of adriamycin-lipid membrane interactions, as well as the phospholipid composition of the vesicles. 相似文献
12.
13.
Yuri A. Kosinsky Peter V. Dubovskii Dmitry E. Nolde Alexander S. Arseniev Roman G. Efremov 《Molecular simulation》2013,39(4-6):341-349
Abstract Conformation of 20-residue peptide E5, an analog of the fusion peptide of influenza virus hemagglutinin, was explored by Monte-Carlo technique starting with the fully buried in the membrane ideal α-helix. The lipid bilayer (of 30 Å width) together with surrounding water were modeled by the atomic solvation parameters. During the simulation, residues 2–18 of the peptide retained α-helical conformation, and the peptide was found to be partially immersed into the bilayer. In the resulting low-energy conformers, the N-terminus was buried inside the membrane, its position with respect to the bilayer surface (ZNT) being varied from 2.5 to 7.5 Å, and the orientation of the helical axis relative to the membrane plane (Θ) – from 10 to 35°. The low-energy conformers (below -200kcal/mol) were clustered in the space (ZNT, Θ) into 4 groups. To select low-energy states of the peptide compatible with NMR data, we calculated pKa values of E5 ionizable groups and compared them with the experimental values. It was shown that the best correlation coefficient (0.87) and rmsd (0.68 in pH units) were obtained for the group of states which is characterized by Θ = 15–19° and ZNT = 3.5–4.5Å. 相似文献
14.
Phospholipid-DNA complexes were made of the cationic triester derivative of phosphatidylcholine, EDOPC (1,2-dioleoyl- sn-glycero-3-ethylphosphocholine), by varying conditions of complex formation, in particular, the rate and direction of mixing, as well as by changing the mode of dispersing the lipid (extrusion or vortexing). The biological effects of variations in the formulation procedure were assessed by measuring transfection activity and cell association in cultures of BHK cells. Formulation procedures generally had little effect on cell association, but had marked effects on transfection efficiency. Transfection varied from effectively nil to extremely efficient with what appeared to be modest changes in formulation procedure. Formulation procedures also had significant effects on average sizes and size distributions of lipoplexes as determined by dynamic light scattering. Among the four possibilities of rapid or slow mixing combined with the two possible directions of mixing, slow addition of DNA to lipid gave results that differed significantly from the other three modes. In the case of vortexed lipid, the latter procedure was much less satisfactory than the other three, whereas in the case of extruded lipid, it was the only mode that produced satisfactory transfection. The factors that determine the difference in lipoplex properties can be identified as both geometric and physical. The geometric factor has to do with the symmetries of the participating units. There are three physical factors that are critical: the difference in vesicle stability upon interaction with DNA, the time dependence of interdiffusion of the components relative to that of vesicle rupture, and difference in input concentrations. These factors determine lipoplex size and, as already also shown by others, lipoplex size influences transfection efficiency. 相似文献
15.
Myong-Chul Koag Stephan Wilkens Raymond D. Fenton Josh Resnik Evanly Vo Timothy J. Close 《Plant physiology》2009,150(3):1503-1514
Dehydrins (DHNs; late embryogenesis abundant D11 family) are a family of intrinsically unstructured plant proteins that accumulate in the late stages of seed development and in vegetative tissues subjected to water deficit, salinity, low temperature, or abscisic acid treatment. We demonstrated previously that maize (Zea mays) DHNs bind preferentially to anionic phospholipid vesicles; this binding is accompanied by an increase in α-helicity of the protein, and adoption of α-helicity can be induced by sodium dodecyl sulfate. All DHNs contain at least one “K-segment,” a lysine-rich 15-amino acid consensus sequence. The K-segment is predicted to form a class A2 amphipathic α-helix, a structural element known to interact with membranes and proteins. Here, three K-segment deletion proteins of maize DHN1 were produced. Lipid vesicle-binding assays revealed that the K-segment is required for binding to anionic phospholipid vesicles, and adoption of α-helicity of the K-segment accounts for most of the conformational change of DHNs upon binding to anionic phospholipid vesicles or sodium dodecyl sulfate. The adoption of structure may help stabilize cellular components, including membranes, under stress conditions.When plants encounter environmental stresses such as drought or low temperature, various responses take place to adapt to these conditions. Typical responses include increased expression of chaperones, signal transduction pathway and late embryogenesis abundant (LEA) proteins, osmotic adjustment, and induction of degradation and repair systems (Ingram and Bartels, 1996).Dehydrins (DHNs; LEA D11 family) are a subfamily of group 2 LEA proteins that accumulate to high levels during late stages of seed development and in vegetative tissues subjected to water deficit, salinity, low temperature, or abscisic acid (ABA) treatment (Svensson et al., 2002). Some DHNs are expressed constitutively during normal growth (Nylander et al., 2001; Rorat et al., 2004, 2006; Rodriguez et al., 2005). DHNs exist in a wide range of photosynthetic organisms, including angiosperms, gymnosperms, algae, and mosses (Svensson et al., 2002). DHNs are encoded by a dispersed multigene family and are differentially regulated, at least in higher plants. For example, 13 Dhn genes have been identified in barley (Hordeum vulgare), dispersed over seven genetic map locations (Choi et al., 1999; Svensson et al., 2002) and regulated variably by drought, low temperature, and embryo development (Tommasini et al., 2008). DHNs are localized in various subcellular compartments, including cytosol (Roberts et al., 1993), nucleus (Houde et al., 1995), chloroplast (Artus et al., 1996), vacuole (Heyen et al., 2002), and proximal to the plasma membrane and protein bodies (Asghar et al., 1994; Egerton-Warburton et al., 1997; Puhakainen et al., 2004). Elevated expression of Dhn genes generally has been correlated with the acquisition of tolerance to abiotic stresses such as drought (Whitsitt et al., 1997), salt (Godoy et al., 1994; Jayaprakash et al., 1998), chilling (Ismail et al., 1999a), or freezing (Houde et al., 1995; Danyluk et al., 1998; Fowler et al., 2001). The differences in expression and tissue location suggest that individual members of the Dhn multigene family have somewhat distinct biological functions (Close, 1997; Zhu et al., 2000; Nylander et al., 2001). Many studies have observed a positive correlation between the accumulation of DHNs and tolerance to abiotic stresses (Svensson et al., 2002). However, overexpression of a single DHN protein has not, in general, been sufficient to confer stress tolerance (Puhakainen et al., 2004).DHNs are subclassified by sequence motifs referred to as the K-segment (Lys-rich consensus sequence), the Y-segment (N-terminal conserved sequence), the S-segment (a tract of Ser residues), and the φ-segment (Close, 1996). Because of high hydrophilicity, high content of Gly (>20%), and the lack of a defined three-dimensional structure in the pure form (Lisse et al., 1996), DHNs have been categorized as “intrinsically disordered/unstructured proteins” or “hydrophilins” (Wright and Dyson, 1999; Garay-Arroyo et al., 2000; Tompa, 2005; Kovacs et al., 2008). On the basis of compositional and biophysical properties and their link to abiotic stresses, several functions of DHNs have been proposed, including ion sequestration (Roberts et al., 1993), water retention (McCubbin et al., 1985), and stabilization of membranes or proteins (Close, 1996, 1997). Observations from in vitro experiments include DHN binding to lipid vesicles (Koag et al., 2003; Kovacs et al., 2008) or metals (Svensson et al., 2000; Heyen et al., 2002; Kruger et al., 2002; Alsheikh et al., 2003; Hara et al., 2005), protection of membrane lipid against peroxidation (Hara et al., 2003), retention of hydration or ion sequestration (Bokor et al., 2005; Tompa et al., 2006), and chaperone activity against the heat-induced inactivation and aggregation of various proteins (Kovacs et al., 2008).Intrinsically disordered/unstructured proteins that lack a well-defined three-dimensional structure have recently been recognized to be prevalent in prokaryotes and eukaryotes (Oldfield et al., 2005). They fulfill important functions in signal transduction, gene expression, and binding to targets such as protein, RNA, ions, and membranes (Wright and Dyson, 1999; Tompa, 2002; Dyson and Wright, 2005). The disorder confers structural flexibility and malleability to adapt to changes in the protein environment, including water potential, pH, ionic strength, and temperature, and to undergo structural transition when complexed with ligands such as other proteins, DNA, RNA, or membranes (Prestrelski et al., 1993; Uversky, 2002). Structural changes from disorder to ordered functional structure also can be induced by the folding of a partner protein (Wright and Dyson, 1999; Tompa, 2002; Mouillon et al., 2008).The idea that DHNs interact with membranes is consistent with many immunolocalization studies, which have shown that DHNs accumulate near the plasma membrane or membrane-rich areas surrounding lipid and protein bodies (Asghar et al., 1994; Egerton-Warburton et al., 1997; Danyluk et al., 1998; Puhakainen et al., 2004). The K-segment is predicted to form a class A2 amphipathic α-helix, in which hydrophilic and hydrophobic residues are arranged on opposite faces (Close, 1996). The amphipathic α-helix is a structural element known to interact with membranes and proteins (Epand et al., 1995). Also, in the presence of helical inducers such as SDS and trifluoroethanol (Dalal and Pio, 2006), DHNs take on α-helicity (Lisse et al., 1996; Ismail et al., 1999b). We previously examined the binding of DHN1 to liposomes and found that DHNs bind preferentially to anionic phospholipids and that this binding is accompanied by an increase in α-helicity of the protein (Koag et al., 2003). Similarly, a mitochondrial LEA protein, one of the group III LEA proteins, recently has been shown to interact with and protect membranes subjected to desiccation, coupled with the adoption of amphipathic α-helices (Tolleter et al., 2007).Here, we explore the basis of DHN-vesicle interaction using K-segment deletion proteins. This study reveals that the K-segment is necessary and sufficient for binding to anionic phospholipid vesicles and that the adoption of α-helicity of DHN proteins can be attributed mainly to the K-segment. 相似文献
16.
Pashkov V. S. Agapov I. I. Balashova T. A. Egorova N. S. Surovoy A. Yu. Pevzner I. B. Tonevitsky A. G. 《Russian Journal of Bioorganic Chemistry》2003,29(6):534-541
It was previously shown that the catalytic subunit of the plant toxin viscumin induces aggregation of small unilamellar liposomes and this process is inhibited by the mab_TA7 monoclonal antibody produced to the denatured catalytic subunit of viscumin (Agapov, I.I. et al., FEBS Lett., 1999, vol. 464, p. 63). The interaction of the synthetic F101–T105 and A96–T105 fragments of the viscumin catalytic subunit with the mab_TA7 monoclonal antibody was studied by 1H NMR spectroscopy. Results of this study demonstrated that only the A96–T105 fragment is capable of binding to mab_TA7. A nuclear Overhauser effect observed in the antigen–antibody complex and registered on the resonances of the free peptide transferred from the free state to the antibody-bound state was analyzed, the mab_TA7 antigen determinant (H99–T105) was identified, and its conformation and orientation within the complex with the antibody were determined. 相似文献
17.
Sonya?M. Hanson Simon Newstead Kenton?J. Swartz Mark?S.P. Sansom 《Biophysical journal》2015,108(6):1425-1434
Transient receptor potential vanilloid subtype 1 (TRPV1) is a heat-sensitive ion channel also involved in pain sensation, and is the receptor for capsaicin, the active ingredient of hot chili peppers. The recent structures of TRPV1 revealed putative ligand density within the S1 to S4 voltage-sensor-like domain of the protein. However, questions remain regarding the dynamic role of the lipid bilayer in ligand binding to TRPV1. Molecular dynamics simulations were used to explore behavior of capsaicin in a 1-palmitoyl-2-oleoyl phosphatidylcholine bilayer and with the target S1–S4 transmembrane helices of TRPV1. Equilibrium simulations reveal a preferred interfacial localization for capsaicin. We also observed a capsaicin molecule flipping from the extracellular to the intracellular leaflet, and subsequently able to access the intracellular TRPV1 binding site. Calculation of the potential of mean force (i.e., free energy profile) of capsaicin along the bilayer normal confirms that it prefers an interfacial localization. The free energy profile indicates that there is a nontrivial but surmountable barrier to the flipping of capsaicin between opposing leaflets of the bilayer. Molecular dynamics of the S1–S4 transmembrane helices of the TRPV1 in a lipid bilayer confirm that Y511, known to be crucial to capsaicin binding, has a distribution along the bilayer normal similar to that of the aromatic group of capsaicin. Simulations were conducted of the TRPV1 S1–S4 transmembrane helices in the presence of capsaicin placed in the aqueous phase, in the lipid, or docked to the protein. No stable interaction between ligand and protein was seen for simulations initiated with capsaicin in the bilayer. However, interactions were seen between TRPV1 and capsaicin starting from the cytosolic aqueous phase, and capsaicin remained stable in the majority of simulations from the docked pose. We discuss the significance of capsaicin flipping from the extracellular to the intracellular leaflet and mechanisms of binding site access by capsaicin. 相似文献
18.
19.
Abstract Corticotropin-releasing factor (CRF) is the principal neuroregulator of adrenocorticotropic hormone (ACTH) secretion. Previous experiments have demonstrated that CRF binds avidly to the surface of single egg phosphatidylcholine vesicles and its amphiphilic secondary structure might play an important role in the function. In this study, the interaction of the residues 13–41 in human CRF with the surface of a DOPC bilayer was investigated by molecular dynamics (MD) simulation in order to understand the role of the membrane surface in the formation of the amphiphilic α helix as well as to determine the effects of the peptide on the lipid bilayer. The model used included 60 DOPC molecules, 1 helical peptide (CRF13–41) on the bilayer surface, and explicit waters of solvation in the lipid polar head group regions, together with constant-volume periodic boundary conditions in three dimensions. The MD simulation was carried out for 510 ps. In addition, CRF13–41, initially in a helical form, was simulated in vacuo as a control. The results indicate that while it was completely unstable in vacuo, the peptide helical form was generally maintained on the bilayer surface, but with distortions near the terminal ends. The peptide was confined to the bilayer headgroup/water region, similar to that reported from neutron diffraction measurement of tripeptides bound to the phosphatidylcholine bilayer surface (Ref 1). The amphiphilicity of the peptide matched that of the bilayer headgroup environment, with the hydrophilic side oriented toward water and the hydrophobic side making contact with the bilayer hydrocarbon core. These results support the hypothesis that the amphiphilic environment of a membrane surface is important in the induction of peptide amphiphilic α-helical secondary structure. Two major effects of the peptide on the lipids were found: the first CH2 segment in the lipid chains was significantly disordered and the lipid headgroup distribution was broadened towards the water region. 相似文献
20.
Indira Chandrasekhar Bruce P. Gaber 《Journal of biomolecular structure & dynamics》2013,31(6):1163-1171
Abstract Anhydrobiotic organisms undergo periods of acute dehydration during their life cycle. It is of interest to understand how the biomembrane remains intact through such stress. A di-saccharide, trehalose, which is metabolised during anhydrobiosis is found to prevent disruption of model membrane systems (1). Molecular modelling techniques are used to investigate the possible mode of interaction of trehalose with a model monolayer. The objective is to maximise hydrogen bonding between the two systems. A phospholipid matrix consisting of l,2-dimyristoyl-sn-glycero-3-phosphatidylcholine (DMPC) is chosen to represent the monolayer. The crystal structure of DMPC (2) reveals that there are two distinct conformers designated as A and B. An expansion of the monolayer, coplanar with its surface, results in the trehalose molecule being accomodated in a pocket formed by four B conformers. One glucose ring of the sugar rests on the hydrophobic patch provided by the choline methyls of an A conformer. Five hydrogen bonds are formed involving the phosphate oxygens of three of the surrounding B conformers. The model will be discussed with reference to relevant experimental data on the interaction. 相似文献