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 31 毫秒
1.
We present a new atom density profile (ADP) model and a statistical approach for extracting structural characteristics of lipid bilayers from X-ray and neutron scattering data. Models for five lipids with varying head and tail chemical composition in the fluid phase, 1,2-dioleoyl-sn-glycero-3-phosphatidylcholine (DOPC), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC), 1,2-dipalmitoyl-sn-glycero-3-phosphatidylcholine (DPPC), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylserine (POPS), and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylglycerol (POPG), are optimized using a simplex based method to simultaneously reproduce both neutron and X-ray scattering data. Structural properties are determined using statistical analysis of multiple optimal model structures. The method and models presented make minimal assumptions regarding the atomic configuration, while taking into account the underlying physical properties of the system. The more general model and statistical approach yield data with well defined uncertainties, indicating the precision in determining density profiles, atomic locations, and bilayer structural characteristics. Resulting bilayer structures include regions exhibiting large conformational variation. Due to the increased detail in the model, the results demonstrate the possibility of a distinct hydration layer within the interfacial (backbone) region.  相似文献   

2.
PEG-mediated fusion of SUVs composed of dioleoylphosphatidylcholine, dioleoylphosphatidylethanolamine, sphingomyelin, cholesterol, and dioleoylphosphatidylserine was examined to investigate the effects of PS on the fusion mechanism. Lipid mixing, content mixing, and content leakage measurements were carried out with vesicles containing from 0 to 8 mol % PS and similar amounts of phosphatidylglycerol. Fitting these time courses globally to a 3-state (aggregate, intermediate, pore) sequential model established rate constants for each step and probabilities of lipid mixing, content mixing, and leakage in each state. Charged lipids inhibited both the rates of intermediate and pore formation as well as the extents of lipid and contents mixing, although electrostatics were not solely responsible for inhibition. Ca2+ counteracted this inhibition and increased the extent of fusion in the presence of PS to well beyond that seen in the absence of charged lipids. The effects of both PS and Ca2+ could be interpreted in terms of a previous proposal for the nature of lipid fluctuations that account for transition states for the two steps of the fusion process examined. The results suggest a more significant role for Ca2+-lipid interactions than is acknowledged in current thinking about cell membrane fusion.Abbreviations used: SUVs, small unilamellar vesicles; DOPC, 1,2-dioleoyl-3-sn-phosphatidylcholine; DOPE, 1,2-dioleoyl-3-sn-phosphatidylethanolamine; SM, sphingomyelin (bovine brain); CH, Cholesterol; DOPS, 1,2-dioleoyl-3-sn-phosphatidylserine; PS, phosphatidylserine; DOPG, 1,2-dioleoyl-3-sn-phosphatidylglycerol; PG, phosphatidylglycerol; TES, N- tris(hydroxymethyl)methyl}2-2-aminoethane sulfonic acid; PEG, poly(ethylene glycol); CM, contents mixing; LM, lipid mixing  相似文献   

3.
1. 3-sn-Phosphatidylcholine was identified as the major lipid in cotyledons from the developing seeds of soya bean, linseed and safflower when tissue was steamed before lipid extraction. The proportion of oleate in this lipid decreased markedly and that of the polyunsaturated C18 fatty acids increased when detached developing cotyledons were incubated for up to 3h. Similar but less pronounced changes occurred in diacylglycerol, which had a fatty acid composition resembling that of the 3-sn-phosphatidylcholine from cotyledons of the same species. 2. [1-14C]Acetate supplied to detached cotyledons was incorporated into the acyl moieties of mainly 3-sn-phosphatidylcholine, 1,2-diacylglycerol and triacylglycerol. Initially label was predominantly in oleate, but subsequently entered at accelerating rates the linoleoyl moieties of the above lipids in soya-bean and safflower cotyledons and the linoleoyl and linolenyl moieties of these lipids in linseed cotyledons. In pulse–chase experiments label was rapidly lost from the oleate of 3-sn-phosphatidylcholine and accumulated in the linoleoyl and linolenoyl moieties of this phospholipid and of the di- and tri-acylglycerols. 3. [2-3H]Glycerol was incorporated into the glycerol moieties of mainly 3-sn-phosphatidylcholine and di- and tri-acylglycerols of developing linseed and soya-bean cotyledons. The label entered the phospholipid and diacylglycerol at rates essentially linear with time from the moment the substrate was supplied, and entered the triacylglycerol at an accelerating rate. With linseed cotyledons the labelled glycerol was incorporated initially mainly into species of 3-sn-phosphatidylcholine and diacylglycerol that contained oleate, but accumulated with time in more highly unsaturated species. In pulse–chase experiments with linseed cotyledons, label was lost from both 3-sn-phosphatidylcholine and diacylglycerol, preferentially from the dioleoyl species, and accumulated in triacylglycerol, mainly in species containing two molecules of linolenate. 4. The results suggest a rapid turnover of 3-sn-phosphatidylcholine during triacylglycerol accumulation in developing oilseeds, and are consistent with the operation of a biosynthetic route whereby oleate initially esterified to the phospholipid is first desaturated, then polyunsaturated fatty acids transferred to triacylglycerol, via diacylglycerol. The possible role of oleoyl phosphatidylcholine as a substrate for oleate desaturation is discussed.  相似文献   

4.
《Biophysical journal》2020,118(8):1830-1837
Laurdan fluorescence, novel spectral fitting, and dynamic light scattering were combined to determine lateral lipid organization in mixed lipid membranes of the oxidized lipid, 1-palmitoyl-2-glutaryl-sn-glycero-3-phosphocholine (PGPC), and each of the three bilayer lipids, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC), 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC), and 1-palmitoyl-2-oleoylphosphatidylcholine (POPC). Second harmonic spectra were computed to determine the number of elementary emissions present. All mixtures indicated two emissions. Accordingly, spectra were fit to two log-normal distributions. Changes with PGPC mole fraction, XPGPC, of the area of the shorter wavelength line and of dynamic light scattering-derived aggregate sizes show that: DPPC and PGPC form component-separated mixed vesicles for XPGPC ≤ 0.2 and coexisting vesicles and micelles for XPGPC > 0.2 in gel and liquid-ordered phases and for all XPGPC in the liquid-disordered phase; POPC and PGPC form randomly mixed vesicles for XPGPC ≤ 0.2 and component-separated mixed vesicles for XPGPC > 0.2. DOPC and PGPC separate into vesicles and micelles. Component segregation is due to unstable inhomogeneous membrane curvature stemming from lipid-specific intrinsic curvature differences between mixing molecules. PGPC is inverse cone-shaped because its truncated tail with a terminal polar group points into the interface. It is similar to and mixes with POPC, also an inverse cone because of mobility of its unsaturated tail. PGPC is least similar to DOPC because mobilities of both unsaturated tails confer a cone shape to DOPC, and PGPC separates form DOPC. DPPC and PGPC do not mix in the liquid-disordered phase because mobility of both tails in this phase renders DPPC a cone. DPPC is a cylinder in the gel phase and of moderate similarity to PGPC and mixes moderately with PGPC.  相似文献   

5.
We measured the 31P[1H] Nuclear Overhauser Effect (NOE) as a function of temperature and of 1H irradiation frequency, the linewidth Δν12 as a function of temperature and the relaxation time T1 above and below the thermal transition temperature, of the 31P-NMR signal in sonicated liposomes of 1,2-dimiristoyl-3-sn-phosphatidylcholine (DMPC), 1,2-dipalmitoyl-3-sn-phosphatidylcholine (DPPC) and 1,2-dimiristoyl-3-sn-phosphatidylcholine (DSPC). The same measurements were repeated in the presence of high molecular weight dextrans. They strongly reduce the NOE and produce longer relaxation times T1. According to the current models, we were able to evaluate, in the different situations, the correlation time of the internal motion τG and the distance r between interacting groups in the region of the polar head groups. While the first parameter changes abruptly through the phase transition and under the effect of dextrans, the latter does not appear modified in any case. These results are discussed in terms of a conformational change of the phosphocholine head groups.  相似文献   

6.
We have studied the biocompatibility properties of polymerizable phosphatidylcholine bilayer membranes, in the form of liposomes, with a view toward the eventual utilization of such polymerized lipid assemblies in drug carrier systems or as surface coatings for biomaterials. The SH-based polymerizable lipid 1,2-bis[1,2-(lipoyl)dodecanoyl]-sn-glycero-3-phosphocholine (dilipoyl lipid, DLL) and the methacryl-based lipid 1,2-bis[(methacryloyloxy)dodecanoyl]-sn-glycero-3-phosphocholine (dipolymerizable lipid, DPL) were studied in comparison to ‘conventional’ zwitterionic or charged phospholipids. We examined binding of serum proteins to liposomes and effects of liposomes on fibrin clot formation and on platelet aggregation. All types of liposomes tested bound complex mixtures of serum proteins with IgG being the most abundant bound component. DPL vesicles and anionic vesicles bound substantially more protein than other vesicle types. Polymerized DPL vesicles uniquely bound a protein of about 53 kDa which was not bound to other types of phosphatidylcholine liposomes. Likewise polymerized DPL vesicles, but not other types of phosphatidylcholine vesicles, caused a marked alteration in coagulation as measured by activated partial thromboplastin time (APTT) and prothrombin time (PT) tests; this effect was shown to be due to binding and depletion of clothing factor V by the DPL polymerized vesicles. Polymerized DPL liposomes and DLL liposomes in polymerized or nonpolymerized form, were without substantial effect on platelet aggregation. However, DPL nonpolymerized vesicles, while not causing aggregation, did impair ADP-induced aggregation of platelets. These studies suggest that SH based polymerizable lipids of the DLL type may be very suitable for in vivo use in the contexts of drug delivery systems or biomaterials development. Methacryloyl-based lipids of the DPL type seem to display interactions with the hemostatic process which militate against their in vivo utilization.  相似文献   

7.
Immobilized phospholipase D (PLD) from Streptomyces sp. catalyzed the transfer reaction of the dipalmitoylphosphatidyl residue from 1,2-dipalmitoyl-3-sn-phosphatidylcholine (DPPC) to an aromatic hydroxy group on 4-methoxyphenol in water-immiscible organic solvents, to afford 1,2-dipalmitoyl-3-sn-phosphatidyl-4-methoxyphenol (DPP-PMP) with a 45% yield, accompanied by a trace amount of 1,2-dipalmitoyl-3-sn-phosphatidic acid sodium salt (DPPA-Na). The formation of DPP-PMP was affected by organic solvents used in the reaction. Benzene, toluene, and methylene chloride gave DPP-PMP with moderate yields but use of diethyl ether resulted in a low yield of DPP-PMP. In both ethyl acetate and water-miscible organic solvents, the transfer reaction did not take place. Immobilization of PLD was carried out by adding a 1 % volume of PLD solution to a suspension of a cation-exchange resin (Amberlite IRC-50, 5% w/v) in benzene with stirring and sonication. In a repeated batch reaction for DPP-PMP synthesis with immobilized PLD, after ten batch cycles the enzyme retained 74% of its initial activity.  相似文献   

8.
Rapid adsorption of surfactant material to the air/liquid interface of the lung is essential for maintaining normal lung function. The detailed mechanism of this process, however, remains unclear. In this study, we elucidate the influence of lipid saturation grade and headgroup charge of surface layer lipids on surfactant protein (SP)-induced vesicle insertion into monolayers spread at the air/water interface of a film balance. We used dipalmitoylphosphatidlycholine (DPPC),1,2-dipalmitoyl-sn-glycero-3-phosphoglycerol (DPPG), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC), and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol (POPG) as monolayer lipids doped with either hydrophobic surfactant-specific protein SP-B or SP-C (0.2 and 0.4 mol %, respectively). Vesicles consisting of DPPC/DPPG (4:1, mol ratio) were injected into a stirred subphase to quantify adsorption kinetics. Based on kinetic film balance and fluorescence measurements, a refined model describing distinct steps of vesicle adsorption to surfactant monolayers is presented. First, in a protein-independent step, lipids from vesicles bridged to the interfacial film by Ca2+ ions are inserted into defects of a disordered monolayer at low surface pressures. Second, in a SP-facilitated step, active material insertion involving an SP-B- or SP-C-induced flip-flop of lipids occurs at higher surface pressures. Negatively charged lipids obviously influence the threshold pressures at which this second protein-mediated adsorption mechanism takes place.  相似文献   

9.
Oxyopinins (Oxki1 and Oxki2) are antimicrobial peptides isolated from the crude venom of the wolf spider Oxyopes kitabensis. The effect of oxyopinins on lipid bilayers was investigated using high-sensitivity titration calorimetry and 31P solid-state NMR spectroscopy. High-sensitivity titration calorimetry experiments showed that the binding of oxyopinins was exothermic, and the binding enthalpies (ΔH) to 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC) small unilamellar vesicles (SUVs) were − 18.1 kcal/mol and − 15.0 kcal/mol for Oxki1 and Oxki2, respectively, and peptide partition coefficient (Kp) was found to be 3.9 × 103 M− 1. 31P NMR spectra of 1,2-dielaidoyl-sn-glycero-3-phosphoethanolamine (DEPE) membranes in the presence of oxyopinins indicated that they induced a positive curvature in lipid bilayers. The induced positive curvature was stronger in the presence of Oxki2 than in the presence of Oxki1. 31P NMR spectra of phosphaditylcholine (PC) membranes in the presence of Oxki2 showed that Oxki2 produced micellization of membranes at low peptide concentrations, but unsaturated PC membranes or acidic phospholipids prevented micellization from occurring. Furthermore, 31P NMR spectra using membrane lipids from E. coli suggested that Oxki1 was more disruptive to bacterial membranes than Oxki2. These results strongly correlate to the known biological activity of the oxyopinins.  相似文献   

10.
Small unilamellar vesicles associated with plasmid DNA showed maximum association efficiency for a cationic mixture of egg phosphatidylcholine (EPC):1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE):di-1,2-dioleoyl-3-trimethyl ammonium propane (DOTAP) (16:8:1 molar ratio) [65%], followed by neutral lipids EPC:1,2-dimyristoyl-sn-glycero-3-phosphoethanolamine (DMPE):cholesterol (Chol) (2:2:1 molar ratio) [30%], and a polymerized formulation 1,2-bis(10,12-tricosadiynoyl)sn-glycero-3-phosphocholine (DC8,9PC):DMPE:Chol (2:2:1 molar ratio) [11%]. The hydrophobicity factor (HF) for these formulations followed the trend DC8,9PC:DMPE:CHOL < EPC:DMPE:Chol < EPC:DOPE DOTAP, and DNA association did not alter this trend. Results suggest that the higher the HF value, the more fluid the membrane and the higher the efficiency of DNA association. On the other hand, no differences were observed in cell toxicity with lipids up to 1 mg/ml in VERO cells.  相似文献   

11.
Three crown ether derivatives, 1,2-O-dioleoyl-3-O-{2-[(12-crown-4)methoxy]ethyl}-sn-glycerol (12C4L), 1,2-O-dioleoyl-3-O-{2-[(15-crown-5)methoxy]ethyl}-sn-glycerol (15C5L) and 2,3-naphtho-15-crown-5 (NAP5), have been incorporated into 1-palmitoyl-2-oleoyl-phosphatydilcholine (POPC) liposomes. The size of the crown ether and the lipophilic moiety of 12C4L, 15C5L and NAP5 influence the stability and the properties of the extruded POPC liposomes determined at 25 °C in buffered aqueous solution at pH 7.4. The investigated liposomes are zwitterionic for POPC headgroups but can be turned into cationic aggregates in the presence of divalent cations. The capability of these systems to complex DNA has been demonstrated by SAXS experiments.  相似文献   

12.
Recently, we have reported the discovery of block liposomes (BLs), a new class of liquid (chain-melted) vesicles, formed in mixtures of the curvature-stabilizing hexadecavalent cationic lipid MVLBG2, the neutral lipid 1,2-dioleoyl-sn-glycero-3-phosphatidylcholine (DOPC), and water with no added salt. BLs consist of connected spheres, pears, tubes, or rods. Unlike in typical liposome systems, where spherical vesicles, tubular vesicles, and cylindrical micelles are separated on the macroscopic scale, shapes remain connected and are separated only on the nanometer scale within a single BL. Here, we report structural studies of the effect of salt and pH on the BL phase, carried out using differential interference contrast microscopy (DIC) and cryogenic transmission electron microscopy (cryo-TEM). Addition of salt screens the electrostatic interactions; in low-salt conditions, partial screening of electrostatic interactions leads to a shape transition from BLs to bilamellar vesicles, while in the high-salt regime, a shape transition from BLs to liposomes with spherical morphologies occurs. This demonstrates that strong electrostatic interactions are essential for BL formation. Understanding the control of liposome shape evolution is of high interest because such shape changes play an important role in many intracellular processes such as endocytosis, endoplasmatic reticulum-associated vesiculation, vesicle recycling and signaling.  相似文献   

13.
To investigate the effect of lipid structure upon the membrane topography of hydrophobic helices, the behavior of hydrophobic peptides was studied in model membrane vesicles. To define topography, fluorescence and fluorescence quenching methods were used to determine the location of a Trp at the center of the hydrophobic sequence. For peptides with cationic residues flanking the hydrophobic sequence, the stability of the transmembrane (TM) configuration (relative to a membrane-bound non-TM state) increased as a function of lipid composition on the order: 1:1 (mol:mol) 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC):1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine ∼ 6:4 POPC:cholesterol < POPC ∼ dioleoylphosphatidylcholine (DOPC) < 1,2-dioleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] sodium salt (DOPG) ≤ 1,2-dioleoyl-sn-glycero-3-[phospho-l-serine] sodium salt (DOPS), indicating that the anionic lipids DOPG and DOPS most strongly stabilized the TM configuration. TM stabilization was near maximal at 20-30 mol% anionic lipid, which are physiologically relevant values. TM stabilization by anionic lipid was observed for hydrophobic sequences with a diverse set of sequences (including polyAla), diverse lengths (from 12 to 22 residues), and various cationic flanking residues (H, R, or K), but not when the flanking residues were uncharged. TM stabilization by anionic lipid was also dependent on the number of cationic residues flanking the hydrophobic sequence, but was still significant with only one cationic residue flanking each end of the peptide. These observations are consistent with TM-stabilizing effects being electrostatic in origin. However, Trp located more deeply in DOPS vesicles relative to DOPG vesicles, and peptides in DOPS vesicles showed increased helix formation relative to DOPG and all other lipid compositions. These observations fit a model in which DOPS anchors flanking residues near the membrane surface more strongly than does DOPG and/or increases the stability of the TM state to a greater degree than DOPG. We conclude that anionic lipids can have significant and headgroup structure-specific effects upon membrane protein topography.  相似文献   

14.
The incorporation of palmitate into lipids by hamster lung subcellular fractions was examined and compared to the simultaneous incorporation of sn-glycero-3-phosphate. The rate of incorporation was greater for the microsomal fraction than for the mitochondria-rich fraction with very little incorporation by the supernatant. The supernatant, however, increased the rate of incorporation by 60–70% when added to the particulate fractions. The presence of CoA, ATP and rac-glycerophosphate in the incubation medium was required for optimal incorporation in all fractions. Comparison of incorporation of sn-glycero-3-phosphate and palmitate into lipids indicated that a great part of palmitate incorporation into 3-sn-phosphatidylcholine did not proceed via the diglyceride pathway. The highest de novo incorporation of palmitate was observed into 3-sn-phosphatidylethanolamine.  相似文献   

15.
A recently defined charge set, to be used in conjunction with the all-atom CHARMM27r force field, has been validated for a series of phosphatidylcholine lipids. The work of Sonne et al. successfully replicated experimental bulk membrane behaviour for dipalmitoylphosphatidylcholine (DPPC) under the isothermal-isobaric (NPT) ensemble. Previous studies using the defined CHARMM27r charge set have resulted in lateral membrane contraction when used in the tensionless NPT ensemble, forcing the lipids to adopt a more ordered conformation than predicted experimentally. The current study has extended the newly defined charge set to 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC) and 1-palmitoyl-2-docosahexaenoyl-sn-glycero-3-phosphatidylcholine (PDPC). Molecular dynamics simulations were run for each of the lipids (including DPPC) using both the CHARMM27r charge set and the newly defined modified charge set. In all three cases a significant improvement was seen in both bulk membrane properties and individual atomistic effects. Membrane width, area per lipid and the depth of water penetration were all seen to converge to experimental values. Deuterium order parameters generated with the new charge set showed increased disorder across the width of the bilayer and reflected both results from experiment and similar simulations run with united atom models. These newly validated models can now find use in mixed biological simulations under the tensionless ensemble without concern for lateral contraction.  相似文献   

16.
The membrane-bound conformation of a cell-penetrating peptide, penetratin, is investigated using solid-state NMR spectroscopy. The 13C chemical shifts of 13C, 15N-labeled residues in the peptide indicate a reversible conformational change from β-sheet at low temperature to coil-like at high temperature. This conformational change occurs for all residues examined between positions 3 and 13, at peptide/lipid molar ratios of 1:15 and 1:30, in membranes with 25-50% anionic lipids, and in both saturated DMPC/DMPG (1,2-dimyristoyl-sn-glycero-3-phosphatidylchloline/1,2-dimyristoyl-sn-glycero-3-phosphatidylglycerol) membranes and unsaturated POPC/POPG (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine/1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylglycerol) membranes. Thus, it is an intrinsic property of penetratin. The coil state of the peptide has C-H order parameters of 0.23-0.52 for Cα and Cβ sites, indicating that the peptide backbone is unstructured. Moreover, chemical shift anisotropy lineshapes are uniaxially averaged, suggesting that the peptide backbone undergoes uniaxial rotation around the bilayer normal. These observations suggest that the dynamic state of penetratin at high temperature is a structured turn instead of an isotropic random coil. The thermodynamic parameters of this sheet-turn transition are extracted and compared to other membrane peptides reported to exhibit conformational changes. We suggest that the function of this turn conformation may be to reduce hydrophobic interactions with the lipid chains and facilitate penetratin translocation across the bilayer without causing permanent membrane damage.  相似文献   

17.
The trans isomers of fatty acids are found in human adipose tissue. These isomers have been linked with deleterious health effects (e.g., coronary artery disease). In this study, we performed molecular dynamics simulations to investigate the structures and dynamic properties of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC) and 1-palmitoyl-2-elaidoyl sn-glycero-3-phosphatidylcholine (PEPC) lipid bilayers. The geometry of the olefinic bond and membrane packing effects significantly influenced the conformations and dynamics of the two CC single bonds adjacent to the olefinic bond. For the PEPC lipid, the two CC single bonds adjacent to the olefinic bond adopted mainly nonplanar skew–trans and planar cis–trans motifs; although the cis conformation featured relatively strong steric repulsion, it was stabilized through membrane packing because its planar structure is more suitable for membrane packing. Moreover, membrane packing effects stabilized the planar transition state for conformational conversion to a greater extent than they did with the nonplanar transition state, thereby affecting the dynamics of conformational conversion. The rotational motions of the first neighboring CC single bonds were much faster than those of typical saturated CC single bonds; in contrast, the rotational motions of the second neighboring CC single bonds were significantly slower than those of typical saturated torsion angles. The packing of PEPC lipids is superior to that of POPC lipids, leading to a smaller area per lipid, a higher order parameter and a smaller diffusion coefficient. The distinct properties of POPC and PEPC lipids result in PEPC lipids forming microdomains within a POPC matrix.  相似文献   

18.
The C1 domains of classical and novel PKCs mediate their diacylglycerol-dependent translocation. Using fluorescence resonance energy transfer, we studied the contribution of different negatively charged phospholipids and diacylglycerols to membrane binding. Three different C1B domains of PKCs were studied (the classical γ, and the novel δ and ?), together with different lipid mixtures containing three types of acidic phospholipids and three types of activating diacylglycerols. The results show that C1Bγ and C1B? exhibit a higher affinity to bind to vesicles containing 1-palmitoyl-2-oleoyl-sn-phosphatidic acid, 1-palmitoyl-2-oleoyl-sn-phoshatidylserine, or 1-palmitoyl-2-oleoyl-sn-phosphatidylglycerol, with C1B? being the most relevant case because its affinity for POPA-containing vesicles increased by almost two orders of magnitude. When the effect of the diacylglycerol fatty acid composition on membrane binding was studied, the C1B? domain showed the highest binding affinity to membranes containing 1-stearoyl-oleoyl-sn-glycerol or 1,2-sn-dioleoylglycerol with POPA as the acidic phospholipid. Of the three diacylglycerols used in this study, 1,2-sn-dioleoylglycerol and 1-stearoyl-oleoyl-sn-glycerol showed the highest affinities for each isoenzyme, whereas 1,2-sn-dipalmitoylglycerol; showed the lowest affinity. DSC experiments showed this to be a consequence of the nonfluid conditions of 1,2-sn-dipalmitoylglycerol;-containing systems.  相似文献   

19.
Biomembranes with as few as three lipid components can form coexisting liquid-disordered (Ld) and liquid-ordered (Lo) phases. In the coexistence region of Ld and Lo phases, the lipid mixtures 1,2-distearoyl-sn-glycero-3-phosphocholine (DSPC)/1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC)/chol or brain sphingomyelin (bSM)/DOPC/chol form micron-scale domains that are easily visualized with light microscopy. Although large domains are not observed in the mixtures DSPC/1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)/chol and bSM/POPC/chol, lateral heterogeneity is nevertheless detected using techniques with nanometer-scale spatial resolution. We propose a simple and accessible method to measure domain sizes below optical resolution (~200 nm). We measured nanodomain size for the latter two mixtures by combining experimental Förster resonance energy transfer data with a Monte-Carlo-based analysis. We found a domain radius of 7.5?10 nm for DSPC/POPC/chol, similar to values obtained previously by neutron scattering, and ~5 nm for bSM/POPC/chol, slightly smaller than measurable by neutron scattering. These analyses also detect the domain-size transition that is observed by fluorescence microscopy in the four-component lipid mixture bSM/DOPC/POPC/chol. Accurate measurements of fluorescent-probe partition coefficients are especially important for the analysis; therefore, we exploit three different methods to measure the partition coefficient of fluorescent molecules between Ld and Lo phases.  相似文献   

20.
Effective, reproducible, and scalable methods for DNA-lipid assembly are important for the success of non-viral vectors in in vivo gene therapy. We hypothesized DNA-lipid assembly would be optimal if started from a liquid monophase where both DNA and lipids separately form molecular or micellar solutions prior to mixing, without preexisting condensed lipid phases, thus allowing DNA-lipid assembly under conditions close to equilibrium. Previously, we found that mixing plasmid DNA, 1-palmitoyl-2-oleoyl-3-sn-phosphatidylcholine (POPC), cholesterol and a cationic lipid, 1, 2-dioleoyl-3-(trimethylammonio) propane (DOTAP) in 50% (v/v) aqueous ethanol spontaneously produced an optically transparent solution. Upon ethanol removal, DNA-lipid nanoparticles (Genospheres™) were formed. For comparison with well-known technologies, different DNA-lipid particles were prepared by interaction of plasmid DNA and stable or ethanol-destabilized lipid vesicles by combining the components in water or 30% (v/v) aqueous ethanol, respectively. Among the three studied DNA-lipid assembly methods, only Genospheres combined the properties of small size (less than or around 100 nm), high incorporation of both lipid and DNA, high degree of DNA protection (dye accessibility 5-12%), a narrow distribution of particle density and when immuno-targeted, the highest transfection efficiency in HER2-overexpressing cells in vitro. We conclude that the Genosphere assembly methodology offers advantages for the development of effective, scalable and targetable non-viral gene delivery vectors.  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