首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Lamellar electron diffraction intensity data from epitaxially crystallized 1,2-dipalmitoyl-sn-glycerophospho-N,N-dimethylethanolamine were used to determine the layer packing in order to compare the chiral structure to the crystal structure of a racemic homologue. After finding the chain orientation, the structure was determined by interpretation of the Patterson function, followed by independent crystallographic phase assignments with conventional direct methods (use of three phase structure invariants). The phase determination was verified by a translational search with a molecular model based on a similar lecithin structure. The final R-value is 0.29, and this is lowered to 0.18 after a correction is made for incoherent multiple electron scattering. The layer packing is found to be very much like that of a diacyl phosphatidylcholine with the N,N-dimethylethanolamine moiety parallel to the bilayer surface rather than the perpendicular arrangement of headgroups involved in an interdigitated layer, as seen for racemic homolog.  相似文献   

2.
Factors affecting the solid state miscibility of saturated chain cholesteryl esters were determined from electron diffraction and differential scanning calorimetric measurements on a homologous series which included two types of crystal packing. Electron diffraction patterns from solution- and epitaxially crystallized microcrystals gave measured unit cell constants consistent with the bilayer crystal form for myristate, pentadecanoate, palmitate, and stearate esters. Cholesteryl undecanoate crystallized as the monolayer I structure and cholesteryl laurate was polymorphic, packing in either monolayer I or bilayer forms. No evidence was found for the monolayer II form of the laurate claimed in earlier work. It is clear that solid solution formation follows general rules formulated earlier by Kitaigorodskii for molecular crystals. A symmetry criterion must be satisfied first of all, i.e., two compounds that solidify in greatly different crystal structures will not form continuous solid solutions (e.g., cholesteryl undecanoate/cholesteryl myristate). Within a given crystal structure type, solid solution is permitted when the molecular volumes are similar. (For example, cholesteryl myristate forms an ideal solid solution with cholesteryl pentadecanoate, a nonideal solution with cholesteryl palmitate, and a eutectic of solid solutions with cholesteryl stearate.) For the polymorphic cholesteryl laurate, solid solutions of either the monolayer I structure (e.g., with cholesteryl undecanoate) or bilayer structure (e.g., with cholesteryl myristate) are permitted.  相似文献   

3.
The extent to which current force fields faithfully reproduce conformational properties of lipids in bilayer membranes, and whether these reflect the structural principles established for phospholipids in bilayer crystals, are central to biomembrane simulations. We determine the distribution of dihedral angles in palmitoyl-oleoyl phosphatidylcholine from molecular dynamics simulations of hydrated fluid bilayer membranes. We compare results from the widely used lipid force field of Berger et al. with those from the most recent C36 release of the CHARMM force field for lipids. Only the CHARMM force field produces the chain inequivalence with sn-1 as leading chain that is characteristic of glycerolipid packing in fluid bilayers. The exposure and high partial charge of the backbone carbonyls in Berger lipids leads to artifactual binding of Na+ ions reported in the literature. Both force fields predict coupled, near-symmetrical distributions of headgroup dihedral angles, which is compatible with models of interconverting mirror-image conformations used originally to interpret NMR order parameters. The Berger force field produces rotamer populations that correspond to the headgroup conformation found in a phosphatidylcholine lipid bilayer crystal, whereas CHARMM36 rotamer populations are closer to the more relaxed crystal conformations of phosphatidylethanolamine and glycerophosphocholine. CHARMM36 alone predicts the correct relative signs of the time-average headgroup order parameters, and reasonably reproduces the full range of NMR data from the phosphate diester to the choline methyls. There is strong motivation to seek further experimental criteria for verifying predicted conformational distributions in the choline headgroup, including the 31P chemical shift anisotropy and 14N and CD3 NMR quadrupole splittings.  相似文献   

4.
Both wide-angle and lamellar x-ray diffraction data are interpreted in terms of a difference in hydrocarbon chain tilt between fully hydrated dipalmitoyl phosphatidylcholine (DPPC) and dipalmitoyl phosphatidylethanolamine (DPPE). Although the hydrocarbon chains of multilayers of DPPC tilt ty approximately 30 degrees relative to the normal to the plane of the bilayer, as previously reported by others, the hydrocarbon chains of DPPE appear to be oriented approximately normal to the plane of the bilayer. It is found that the chain tilt in DPPC bilayers can be reduced by either: (a) adding an n-alkane to the bilayer interiors or (b) adding lanthanum ions to the fluid layers between bilayers. A molecular packing model is presented which accounts for these data. According to this model, DPPC chains tilt because of the size and conformation of the PC polar head group.  相似文献   

5.
The transferred nuclear Overhauser effects of yeast alpha-mating factor [(1-13)peptide] in the presence of various spin-labeled phosphatidylcholines in small unilamellar vesicles of perdeuterated phosphatidylcholine have been analyzed. From the analysis of the quenching effect by spin-labels, the depth of amino acid side chains of the mating factor in phospholipid bilayer has been elucidated. The Leu4 and Leu6 residues are buried deeply in the apolar region of the phospholipid bilayer while the hydrophilic residues such as Gln5 and Lys7 are in the shallow region of the bilayer. The interaction of the side chains of Trp1 and Trp3 residues of alpha-mating factor with the hydrophobic interior of the bilayer contributes to the binding of this peptide with the phosphatidylcholine bilayer. The conformation of des-Trp1-alpha-mating-factor [(2-13)peptide] in the membrane-bound state has been found to be similar to that of (1-13)peptide from the analysis of transferred nuclear Overhauser effects in the presence of mixed vesicles of perdeuterated phosphatidylcholine and perdeuterated phosphatidylserine. The incorporation of this acidic phospholipid in the vesicle remarkably enhances the binding of (1-13)peptide and analog peptides. However, such modifications that weaken the interaction with phospholipid bilayer (deletion of Trp1 and substitution of Trp3 by Gly or Ala) appreciably lower the physiological activity. Transferred nuclear Overhauser effect analyses have also been made of [DHis2]peptide, [DLeu6]peptide and [DLys7]peptide in the presence of the vesicles of perdeuterated phosphatidylcholine. The main-chain conformations of these three analogs in the membrane-bound state have been found to be similar to that of (1-13)peptide, although the side-chain conformations of the D-amino acid residues are naturally different from those of the L-amino acid ones. Thus, the physiological activities of the (1-13)peptide and a variety of analog peptides are found to correlate with the affinities to the phosphatidylcholine/phosphatidylserine membrane and with the molecular conformations in the membrane-bound state.  相似文献   

6.
In this study, following our earlier work on calcium pyrophosphate dihydrate (CPPD) crystal-induced membranolysis, we demonstrate, using the CHARMM method of molecular dynamics simulation, the protective role of phosphocitrate (PC) against solvated dimyristoyl phosphatidylcholine phospholipid bilayer disintegration on contact with the CPPD crystal. Our molecular dynamics simulations studies show that coverage of the CPPD crystal with a layer of phosphocitrate molecules results in the conservation of phospholipid bilayer integrity. We show that the rupture of the lipid bilayer in presence of CPPD and the protective effect of PC are primarily due to electrostatic interactions. The protective role of PC, which may also play an important and potentially therapeutic function against crystal-induced membranolysis is also discussed.  相似文献   

7.
The effect of cholesterol on the structure of phosphatidylcholine bilayers was investigated by X-ray diffraction methods. Electron density profiles at 5 Å resolution along with chain tilt and chain packing parameters were obtained and compared for phosphatidylcholine/cholesterol bilayers and for pure phosphatidylcholine bilayers in both the gel and liquid crystalline states. The cholesterol in the bilayer was localized by noting the position of discrete elevations in the electron density profiles. Cholesterol can either increase or decrease the width of the bilayer depending on the physical state and chain length of the lipid before the introduction of cholesterol. For saturated phosphatidylcholines containing 12–16 carbons per chain, cholesterol increases the width of the bilayer as it removes the chain tilt from gel state lipids or increases the trans conformations of the chains for liquid crystalline lipids. However, cholesterol reduces the width of 18 carbon chain bilayers below the phase transition temperature as the long phospholipid chains must deform or kink to accomodate the significantly shorter cholesterol molecule. Although cholesterol has a marked effect on hydrocarbon chain organization, it was found that, within the resolution limits of the data, the phosphatidylcholine head group conformation is unchanged by the addition of cholesterol to the bilayer. The head group is oriented parallel to the plane of the bilayer for phosphatidylcholine in the gel and liquid crystalline states and this orientation is not changed by the addition of cholesterol.  相似文献   

8.
Cuthbertson JM  Bond PJ  Sansom MS 《Biochemistry》2006,45(48):14298-14310
The glycophorin helix dimer is a paradigm for the exploration of helix-helix interactions in integral membrane proteins. Two NMR structures of the dimer are known, one in a detergent micelle and one in a lipid bilayer. Multiple (4 x 50 ns) molecular dynamics simulations starting from each of the two NMR structures, with each structure in either a dodecyl phosphocholine (DPC) micelle or a dimyristoyl phosphatidylcholine (DMPC) bilayer, have been used to explore the conformational dynamics of the helix dimer. Analysis of the helix-helix interaction, mediated by the GxxxG sequence motif, suggests convergence of the simulations to a common model. This is closer to the NMR structure determined in a bilayer than to micelle structure. The stable dimer interface in the final simulation model is characterized by (i) Gly/Gly packing and (ii) Thr/Thr interhelix H-bonds. These results demonstrate the ability of extended molecular dynamics simulations in a lipid bilayer environment to refine membrane protein structures or models derived from experimental data obtained in protein/detergent micelles.  相似文献   

9.
Fusion of bovine and goat erythrocytes was studied using the phosphate-calcium protocol. Both bovine and goat red cells are resistant to fusion with phosphate and calcium, under conditions that promote fusion of normal human erythrocytes. Fusion resistance is not related to decreased (5%) membrane deformability of erythrocytes of these species, since chicken erythrocytes which are 40% less deformable than human erythrocytes undergo fusion with efficiency similar to human red blood cells. Incorporation of either phosphatidylcholine or phosphatidylserine into bovine erythrocytes mediated by lipid exchange/transfer protein, caused fusion of these erythrocytes. Fluorescence analysis of merocyanine 540 dye labeled erythrocytes, by flow cytometry, showed that the frequency of cells which exhibit dye binding was much less (35%) in dimyristoylphosphatidylcholine (DMPC) incorporated compared to untreated bovine erythrocytes (80%), indicating that incorporation of DMPC caused closed packing of lipids in the external leaflet of the bilayer. These studies show that fusion of bovine erythrocytes, mediated by phosphate and calcium, has a requirement for either specific phospholipids such as phosphatidylcholine, phosphatidylserine, or closed packing of lipids in the external leaflet of the bilayer.  相似文献   

10.
Based on the crystal structure of the vitamin B12 transporter protein of Escherichia coli(BtuCD) a system consisting of the BtuCD transmembrane domain(BtuC) and the palmitoyloleoyl phosphatidylcholine(POPC) lipid bilayer was constructed in silica,and a more-than-57-nanosecond molecular dynamics(MD) simulation was performed on it to reveal the intrinsic functional motions of BtuC.The results showed that a stable protein-lipid bilayer was obtained and the POPC lipid bilayer was able to adjust its thickness to...  相似文献   

11.
We examined the action of porcine pancreatic and bee-venom phospholipase A2 towards bilayers of phosphatidylcholine as a function of several physical characteristics of the lipid-water interface. 1. Unsonicated liposomes of dimyristoyl phosphatidylcholine are degraded by both phospholipases in the temperature region of the phase transition only (cf. Op den Kamp et al. (1974) Biochim. Biophys. Acta 345, 253--256 and Op den Kamp et al. (1975) Biochim. Biophys. Acta 406, 169--177). With sonicates the temperature range in which hydrolysis occurs is much wider. This discrepancy between liposomes and sonicates cannot be ascribed entirely to differences in available substrate surface. 2. Below the phase-transition temperature the phospholipases degrade dimyristoyl phosphatidylcholine single-bilayer vesicles with a strongly curved surface much more effectively than larger single-bilayer vesicles with a relatively low degree of curvature. 3. Vesicles composed of egg phosphatidylcholine can be degraded by pancreatic phospholipase A2 at 37 degrees C, provided that the substrate bilayer is strongly curved. The bee-venom enzyme shows a similar, but less pronounced, preference for small substrate vesicles. 4. In a limited temperature region just above the transition temperature of the substrate the action of both phospholipases initially proceeds with a gradually increasing velocity. This stimulation is presumably due to an increase of the transition temperature, effectuated by the products of the phospholipase action. 5. Structural defects in the substrate bilayer, introduced by sonication below the phase-transition temperature (cf. Lawaczeck et al. (1976) Biochim. Biophys. Acta 443, 313--330) facilitate the action of both phospholipases. The results lead to the general conclusion that structural irregularities in the packing of the substrate molecules facilitate the action of phospholipases A2 on phosphatidylcholine bilayers. Within the phase transition and with bilayers containing structural defects these irregularities represent boundaries between separate lipid domains. The stimulatory effect of strong bilayer curvature can be ascribed to an overall perturbation of the lipid packing as well as to a change in the phase-transition temperature.  相似文献   

12.
H Xu  F A Stephenson  C H Huang 《Biochemistry》1987,26(17):5448-5453
High-resolution differential scanning calorimetry and 31P NMR spectroscopy have been used to study aqueous phosphatidylcholine (PC) dispersions prepared from colyophilized mixtures of C(10):C(22)PC/C(22):C(12)PC of various molar ratios. These two lipid species are highly asymmetric but have a common structural feature; namely, one acyl chain in the fully extended conformation is about twice as long as the other. Our experimental results support two conclusions: (1) These two component lipids are miscible in all proportions in both gel and liquid-crystalline states. This type of system behaves as a nearly ideal mixture. Its calorimetric parameters are those expected on the basis of the mole fraction weighted average of the corresponding parameters for the pure components. (2) The component lipids appear to self-assemble, at T less than Tm, into a mixed interdigitated bilayer in excess water. In a mixed interdigitated bilayer, the short acyl chain of one asymmetric phosphatidylcholine on one side of the bilayer leaflet is apposed with the short acyl chain of another lipid molecule on the other side of the bilayer leaflet, while the longer acyl chain from each of the two leaflets crosses the entire hydrocarbon width of the bilayer. The fundamental packing unit, as well as the dynamic unit describing the axial rotator motion about the bilayer normal for this mixed interdigitated bilayer, is thus a dimer, whereas the packing unit assigned for the noninterdigitated bilayer such as C(16):C(16)PC lamellae is a monomer.  相似文献   

13.
It appears reasonable to expect that the primary result of a change in the length of the acyl chains within a lipid bilayer is a similar change in the bilayer thickness. In the present communication we draw attention to the somewhat more complicated effects which are found experimentally for phosphatidylcholine bilayers as the hydrocarbon chain is varied from twelve to eighteen carbons in length. The major change in dimension which occurs with variation in acyl chain length is the area occupied per molecule rather than the bilayer thickness. The same effect is seen with solute hydrocarbon such as hexane which partition into the membrane and cause only a small variation in membrane thickness but a large increase in the molecular area of the lipid. The origin of this effect arises from the almost isotropic distribution of the additional hydrocarbon to the lipid core of the membrane.  相似文献   

14.
Melittin from bee venom is water-soluble, yet integrates into membranes and lyses cells. Each melittin chain consists of 26 amino acid residues and in aqueous salt solutions it exists as a tetramer. We have determined the molecular structure of the tetramer in two crystal forms grown from concentrated salt solutions. In both crystal forms the melittin polypeptide is a bent alpha-helical rod, with the "inner" surface largely consisting of hydrophobic sidechains and the "outer" surface consisting of hydrophilic side chains. Thus, the helix is strongly amphiphilic. In the tetramer, four such helices contribute their hydrophobic side chains to the center of the molecule. The packing of melittin tetramers is also very similar in the two crystal forms: they are packed in planar layers with the outsides forming hydrophilic surfaces and the insides (the centers of melittin tetramers) forming a hydrophobic surface. We suggest that the surface activity of melittin can be rationalized in terms of these surfaces. The lytic activity of melittin can also be interpreted in terms of the molecular structure observed in the crystals: the hydrophobic inner surface of a melittin helix may integrate into the apolar region of a bilayer with the helix axis approximately parallel to the plane of the bilayer, and with the hydrophilic surface exposed to the aqueous phase. This integration would be expected to disrupt the bilayer because of melittin helix would penetrate only a short distance into it. Additionally, the integration of melittin from one side of a bilayer would produce a surface area difference across the bilayer, perhaps leading to lysis. In this view, melittin is distinct from membrane proteins that penetrate evenly into both leaflets of a bilayer or exactly halfway through a bilayer, and hence we refer to melittin as a surface-active protein.  相似文献   

15.
Molecular dynamics simulations of two monounsaturated phosphatidylcholine (PC) bilayers made of 1-palmitoyl-2-oleoyl-PC (POPC; cis-unsaturated) and 1-palmitoyl-2-elaidoyl-PC (PEPC; trans-unsaturated) were carried out to investigate the effect of a double bond in the PC beta-chain and its conformation on the bilayer core. Four nanosecond trajectories were used for analyses. A fully saturated 1,2-dimyristoyl-PC (DMPC) bilayer was used as a reference system. In agreement with experimental data, this study shows that properties of the PEPC bilayer are more similar to those of the DMPC than to the POPC bilayer. The differences between POPC and PEPC bilayers may be attributed to the different ranges of angles covered by the torsion angles beta10 and beta12 of the single bonds next to the double bond in the oleoyl (O) and elaidoyl (E) chains. Broader distributions of beta10 and beta12 in the E chain than in the O chain make the E chain more flexible. In effect, the packing of chains in the PEPC bilayer is similar to that in the DMPC bilayer, whereas that in the POPC bilayer is looser than that in the DMPC bilayer. The effect of the cis-double bond on torsions at the beginning of the O chain (beta4 and beta5) is similar to that of cholesterol on these torsions in a myristoyl chain.  相似文献   

16.
R A Demel  F Paltauf  H Hauser 《Biochemistry》1987,26(26):8659-8665
The monolayer properties and thermal behavior of different phosphatidylserines are presented. At neutral pH and 22 degrees C, saturated phosphatidylserines form condensed monolayers while unsaturated phosphatidylserines form liquid-expanded films. Under similar conditions, dimyristoylphosphatidylserine undergoes a transition from the liquid-expanded to the condensed state. At pH 4 and 22 degrees C, the surface pressure-area isotherms are shifted to smaller areas relative to the monolayers recorded at neutral pH. The condensation observed at pH 4 is close to that produced at pH 7.4 by the addition of 10 mM CaCl2. As regards the molecular packing in monolayers and the thermal behavior, 1,2-dipalmitoyl-sn-glycero-3-phospho-L-serine (DPPS) and its ether analogue are similar, albeit not identical. Below 30 mN/m, monolayers of the ether analogue are even more condensed than those of DPPS. The order-disorder transition of the ether analogue occurs usually at higher temperatures than that of the diacyl compound. Sonicated phosphatidylserine dispersions consisting of small unilamellar vesicles show anomalous thermal properties compared to sonicated phosphatidylcholine dispersions. They exhibit sharp order-disorder transitions at similar or even slightly elevated temperatures compared to unsonicated phosphatidylserine dispersions. This anomaly is explained in terms of a pH gradient across the bilayer membrane of the small unilamellar phosphatidylserine vesicle. The internal surface pH is more acidic than the external pH, leading to some protonation of phosphatidylserine molecules. This in turn leads to a condensation of phosphatidylserine molecules on the inner bilayer surface. Such a gradient is proposed to be responsible for the thermodynamic stability of highly curved negatively charged bilayer vesicles.  相似文献   

17.
The effect of pH and cholesterol on the dimyristoylphosphatidic acid (DMPA) model membrane system has been investigated by solid state 2H- and 31P-NMR. It has been shown that each of the three protonation states of the DMPA molecule corresponds to a 31P-NMR powder pattern with characteristic delta sigma values; this implies additionally that the proton exchange on the membrane surface is slow on the NMR time scale (millisecond range). Under these conditions, the 2H-labeled lipid chains sense only one magnetic environment, indicating that the three spectra detected by 31P-NMR are related to charge-dependent local dynamics or orientations of the phosphate headgroup or both. Chain ordering in the fluid phase is also found to depend weakly on the charge at the interface. In addition, it has also been found that the first pK of the DMPA membrane is modified by changes in the lipid lateral packing (gel or fluid phases or in the presence of cholesterol) in contrast to the second pK. The incorporation of 30 mol% cholesterol affects the phosphatidic acid bilayer in a way similar to what has been reported for phosphatidylcholine/cholesterol membranes, but to an extent comparable to 10-20 mol % sterol in phosphatidylcholines. However, the orientation and molecular order parameter of cholesterol in DMPA are similar to those found in dimyristoylphosphatidylcholine.  相似文献   

18.
A molecular dynamics simulation study of four lipid bilayers with inserted trans-membrane helical fragment of epithelial growth factor (EGF) receptor (EGF peptide) was performed. The lipid bilayers differ in their lipid composition and consist of (i) unsaturated phosphatidylcholine (palmitoyloleoylphosphatidylcholine, POPC), (ii) POPC and 20 mol% of cholesterol (Chol), (iii) sphingomyelin (SM) and 20 mol% of Chol, and (iv) SM and 50 mol% of Chol. Only 1 out of 26 residues in the EGF-peptide sequence is polar (Thr). The hydrophobic thickness of each bilayer is different but shorter than the length of the peptide and so, due to hydrophobic mismatch, the inserted peptide is tilted in each bilayer. Additionally, in the POPC bilayer, which is the thinnest, the peptide loses its helical structure in a short three-amino acid fragment. This facilitates bending of the peptide and burying all hydrophobic amino acids inside the membrane core (Figure 1(b)). Bilayer lipid composition affects interactions between the peptide and lipids in the membrane core. Chol increases packing of atoms relative to the peptide side chains, and thus increases van der Waals interactions. On average, the packing around the peptide is higher in SM-based bilayers than POPC-based bilayers but for certain amino acids, packing depends on their position relative to the bilayer center. In the bilayer center, packing is higher in POPC-based bilayers, while in regions closer to the interface packing is higher in SM-based bilayers. In general, amino acids with larger side chains interact strongly with lipids, and thus the peptide sequence is important for the pattern of interactions at different membrane depths. This pattern closely resembles the shape of recently published lateral pressure profiles [Ollila et alJ. Struct. Biol. DOI:10.1016/j.jsb.2007.01.012].  相似文献   

19.
The structure and molecular packing density of a "mismatched" solute, 1-hexanol, in lipid membranes of dimyristoyl phosphatidylcholine (DMPC) was studied by molecular dynamics simulations. We found that the average location and orientation of the hexanol molecules matched earlier experimental data on comparable systems. The local density or molecular packing in DMPC-hexanol was elucidated through the average Voronoi volumes of all heavy (non-hydrogen) atoms. Analogous analysis was conducted on trajectories from simulations of pure 1-hexanol and pure (hydrated) DMPC bilayers. The results suggested a positive volume change, DeltaV(m), of 4 cm(3) mol(-1) hexanol partitioned at 310 K in good accordance with experimental values. Analysis of the apparent volumes of each component in the pure and mixed states further showed that DeltaV(m) reflects a balance between a substantial increase in the packing density of the alcohol upon partitioning and an even stronger loosening in the packing of the lipid. Furthermore, analysis of Voronoi volumes along the membrane normal identifies a distinctive depth dependence of the changes in molecular packing. The outer (interfacial) part of the lipid acyl chains (up to C8) is stretched by about 4%. Concomitantly, the average lateral area per chain decreases and these two effects compensate so that the overall packing density in the outer region, where the hexanol molecules are located, remains practically constant. The core of the bilayer (C9-C13) is slightly thinned. The average lateral area per chain in this region expands, resulting in a looser packing density. The net effect in the core is a 2-3% decrease in density corresponding to a total volume increase of approximately 14 cm(3) mol(-1) hexanol partitioned.  相似文献   

20.
This review details how bilayer structural/elastic properties impact three distinct areas of biological significance. First, the partitioning of melittin into bilayers and melittin-induced bilayer leakage depended strongly on bilayer composition. The incorporation of cholesterol into phosphatidylcholine bilayers decreased melittin-induced leakage from 73 to 3%, and bilayers composed of lipopolysaccharide (LPS), the main lipid on the surface of Gram-negative bacteria, also had low (3%) melittin-induced leakage. Second, transbilayer peptides of different hydrophobic lengths were largely excluded from bilayer microdomains (“rafts”) enriched in sphingomyelin (SM) and cholesterol, even when the length of the transbilayer peptide domain matched the hydrocarbon thickness of the raft bilayer. This is likely due to the large area compressibility modulus of SM:cholesterol bilayers. Third, the major water barrier of skin, the extracellular lamellae of the stratum corneum, was found to contain tightly packed asymmetric lipid bilayers with cholesterol located preferentially on one side of the bilayer and a unique skin ceramide containing an unsaturated acyl chain on the opposite side. We argue that, in each of these three areas, key factors are differences in lipid hydrocarbon chain packing for different lipids, particularly the tight hydrocarbon chain packing caused by cholesterol’s strong interaction with saturated chains.  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