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.
Shih PC  Li GC  Yang KJ  Chen W  Tzou DL 《Steroids》2011,76(6):558-563
Solid-state 1H/13C cross-polarization/magic angle spinning (CP/MAS) NMR spectroscopy has been applied to two steroid compounds: dehydroepiandrosterone (DHEA) and spironolactone (SPI), to analyze their conformations at the atomic level. In the absence of lipid, the high-resolution 13C CP/MAS NMR signals of DHEA and SPI in a powder form reveal multiple patterns, with splittings of 30-160 Hz, indicating the existence of multiple conformations. In the mimic lipid environment formed by mixing 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and 1,2-diheptanoyl-sn-glycero-3-phosphocholine (DHPC) in a molar ratio 3:1, the resulting DHEA and SPI spectra revealed mostly singlet patterns, suggesting that these steroids undergo a conformational change leading to a specific conformation in the lipid environment. Evidence from chemical shift isotropy and anisotropy analysis indicates that DHEA might adopt conformations subtly different from that seen in solution and in the powder form. In conclusion, we demonstrate by solid-state NMR that the structures of DHEA and SPI may adopt slightly different conformations in different chemical environments.  相似文献   

2.
Urethan-induced pulmonary adenomas of mice are composed of cells that appear to be morphologically identical to alveolar type II cells and synthesize disaturated diacyl-sn-glycero-3-phosphocholine, the major component of pulmonary surfactant. 1-[1-14C]Palmitoyl-sn-glycero-3-phosphocholine and [1-14C]palmitic acid were compared as precursors of disaturated diacyl-sn-glycero-3-phosphocholine in the adenoma type II cells by incubating both substrates with whole adenomas. When the precursors were compared at equal concentrations (100 μm) in the presence of albumin (1 mg/ml), the rates of incorporation of 1-[1-14C]palmitoyl-sn-glycero-3-phosphocholine and [1-14C]palmitic acid into diacyl-sn-glycero-3-phosphocholine were 5.2 and 2.9 nmol/min · g tissue, respectively. The concentration of monoacyl-sn-glycero-3-phosphocholine (lysolecithin) in the blood plasma of BALB/c mice was 150 μm. In short-term labeling experiments, the label in disaturated diacyl-sn-glycero-3-phosphocholine was equally distributed between the sn-1 and sn-2 positions when 1-[1-14C]palmitoyl-sn-glycero-3-phosphocholine was the precursor, whereas 75 to 80% was in the sn-2 position when [1-14C]palmitic acid was the precursor. The ratios are consistent with incorporation of 1-palmitoyl-sn-glycero-3-phosphocholine via the lysolecithin:lysolecithin transacylase reaction and incorporation of palmitate via acylation of 1-palmitoyl-sn-glycero-3-phosphocholine by acyl-CoA:lysolecithin acyltransferase. 1-[1-14C]Palmitoyl-sn-glycero-3-phospho-[3H-methyl]choline was incorporated into total cellular diacyl-sn-glycero-3-phosphocholine with an isotope ratio similar to that of the precursor; the disaturated species was more enriched in 14C. These findings indicate the cells take up intact monoacyl-sn-glycero-3-phosphocholine and incorporate it into diacyl-sn-glycero-3-phosphocholine. The ability of the cells to utilize intact lysophosphoglycerides for synthesis of cellular lipids was further demonstrated by showing that ether analogs, 1-alkyl-sn-glycero-3-phosphocholine and 1-alkyl-sn-glycero-3-phosphoethanolamine, are taken up and acylated by the cells. Activities of lysolecithin:lysolecithin transacylase and acyl-CoA:lysolecithin acyltransferase were measured in subcellular fractions of the adenoma type II cells; the specific activities of the enzymes were 2.1 nmol/min · mg soluble protein and 21 nmol/min · mg microsomal protein, respectively. The total activity of the acyltransferase in the cell fractions was about four-fold higher than the activity of the transacylase. Characteristics of the two enzymes were studied and are discussed. The findings indicate that exogenous 1-palmitoyl-sn-glycero-3-phosphocholine and palmitic acid both serve as efficient precursors of disaturated diacyl-sn-glycero-3-phosphocholine in the adenoma alveolar type II cells.  相似文献   

3.
Ampullosporin A and alamethicin are two members of the peptaibol family of antimicrobial peptides. These compounds are produced by fungi and are characterized by a high content of hydrophobic amino acids, and in particular the α-tetrasubstituted amino acid residue α-aminoisobutyric acid. Here ampullosporin A and alamethicin were uniformly labeled with 15N, purified and reconstituted into oriented phophatidylcholine lipid bilayers and investigated by proton-decoupled 15N and 31P solid-state NMR spectroscopy. Whereas alamethicin (20 amino acid residues) adopts transmembrane alignments in 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) or 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) membranes the much shorter ampullosporin A (15 residues) exhibits comparable configurations only in thin membranes. In contrast the latter compound is oriented parallel to the membrane surface in 1,2-dimyristoleoyl-sn-glycero-3-phosphocholine and POPC bilayers indicating that hydrophobic mismatch has a decisive effect on the membrane topology of these peptides. Two-dimensional 15N chemical shift - 1H-15N dipolar coupling solid-state NMR correlation spectroscopy suggests that in their transmembrane configuration both peptides adopt mixed α-/310-helical structures which can be explained by the restraints imposed by the membranes and the bulky α-aminoisobutyric acid residues. The 15N solid-state NMR spectra also provide detailed information on the helical tilt angles. The results are discussed with regard to the antimicrobial activities of the peptides.  相似文献   

4.
Aqueous phospholipid mixtures that form bilayered micelles (bicelles) have gained wide use by molecular biophysicists during the past 20 years for spectroscopic studies of membrane-bound peptides and structural refinement of soluble protein structures. Nonetheless, the utility of bicelle systems may be compromised by considerations of cost, chemical stability, and preservation of the bicelle aggregate organization under a broad range of temperature, concentration, pH, and ionic strength conditions. In the current work, 31P nuclear magnetic resonance (NMR) and atomic force microscopy (AFM) have been used to monitor the size and morphology of isotropically tumbling small bicelles formed by mixtures of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) or 1,2-di-O-tetradecyl-sn-glycero-3-phosphocholine (DIOMPC) with either 1,2-dihexanoyl-sn-glycero-3-phosphocholine (DHPC) or 1,2-di-O-hexyl-sn-glycero-3-phosphocholine (DIOHPC), testing their tolerance of variations in commonly used experimental conditions. 1H-15N 2D NMR has been used to demonstrate the usefulness of the robust DMPC-DIOHPC system for conformational studies of a fatty acid-binding protein that shuttles small ligands to and from biological membranes.  相似文献   

5.
Lysophospholipase D (EC 3.1.4.-) activity was demonstrated in rat kidneys, intestines, lungs, testes, and liver. The liver enzyme was studied in greatest detail and its labeled products were identified by chemical and Chromatographic techniques. This enzyme hydrolyzes 1-[1-14C]hexadecyl-sn-glycero-3-phosphoethanolamine and 1-[1-14C]hexadecyl-sn-glycero-3-phosphocholine to yield 1-[1-14C]hexadecyl-sn-glycero-3-phosphate; the initial product is subsequently dephosphorylated by a phosphohydrolase in microsomes to form 1-[1-14C]hexadecyl-sn-glycerol. The possibility that phospholipase C and a phosphotransferase were responsible for the formation of 1-[1-14C]hexadecyl-sn-glycero-3-phosphate was ruled out. Neither 1-[1-14C]hexadecyl-2-acyl-sn-glycero-3-phosphoethanolamine nor 1-[1-14C]hexadecyl-2-acyl-sn-glycero-3-phosphocholine was hydrolyzed. The enzyme requires Mg2+, is inhibited by Ca2+, and is stimulated by high salt concentrations; it is localized in the microsomal fraction and has a pH optimum between 7.0 and 7.6. Inhibition by sulfhydryl reagents and protection by glutathione and dithiothreitol suggest that a sulfhydryl group is required for activity. The enzyme is inhibited by detergents and by organic solvent extraction. It appears to be tightly bound to the microsomes, since repeated freeze-thawing or sonication did not release the activity, and trypsin digestion (either in the presence or in the absence of 0.04% deoxycholate) did not destroy the activity. Lysophospholipase D was previously known to occur only in brain (R. L. Wykle and J. M. Schremmer, 1974, J. Biol. Chem., 249, 1742–1746).  相似文献   

6.
We have recorded site-directed solid-state 13C NMR spectra of [3-13C]Ala- and [1-13C]Val-labeled bacteriorhodopsin (bR) as a typical membrane protein in lipid bilayers, to examine the effect of formation of two-dimensional (2D) lattice or array of the proteins toward backbone dynamics, to search the optimum condition to be able to record full 13C NMR signals from whole area of proteins. Well-resolved 13C NMR signals were recorded for monomeric [3-13C]Ala-bR in egg phosphatidylcholine (PC) bilayer at ambient temperature, although several 13C NMR signals from the loops and transmembrane α-helices were still suppressed. This is because monomeric bR reconstituted into egg PC, dimyristoylphosphatidylcholine (DMPC) or dipalmytoylphosphatidylcholine (DPPC) bilayers undergoes conformational fluctuations with frequency in the order of 104-105 Hz at ambient temperature, which is interfered with frequency of magic angle spinning or proton decoupling. It turned out, however, that the 13C NMR signals of purple membrane (PM) were almost fully recovered in gel phase lipids of DMPC or DPPC bilayers at around 0 °C. This finding is interpreted in terms of aggregation of bR in DMPC or DPPC bilayers to 2D hexagonal array in the presence of endogenous lipids at low temperature, resulting in favorable backbone dynamics for 13C NMR observation. It is therefore concluded that [3-13C]Ala-bR reconstituted in egg PC, DMPC or DPPC bilayers at ambient temperature, or [3-13C]Ala- and [1-13C]Val-bR at low temperature gave rise to well-resolved 13C NMR signals, although they are not always completely the same as those of 2D hexagonal lattice from PM.  相似文献   

7.
The effects of hydrophobic thickness and the molar phosphatidylglycerol (PG) content of lipid bilayers on the structure and membrane interaction of three cationic antimicrobial peptides were examined: aurein 2.2, aurein 2.3 (almost identical to aurein 2.2, except for a point mutation at residue 13), and a carboxy C-terminal analog of aurein 2.3. Circular dichroism results indicated that all three peptides adopt an α-helical structure in the presence of a 3:1 molar mixture of 1,2-dimyristoyl-sn-glycero-3-phosphocholine/1,2-dimyristoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] (DMPC/DMPG), and 1:1 and 3:1 molar mixtures of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine/1-palmitoyl-2-oleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] (POPC/POPG). Oriented circular dichroism data for three different lipid compositions showed that all three peptides were surface-adsorbed at low peptide concentrations, but were inserted into the membrane at higher peptide concentrations. The 31P solid-state NMR data of the three peptides in the DMPC/DMPG and POPC/POPG bilayers showed that all three peptides significantly perturbed lipid headgroups, in a peptide or lipid composition-dependent manner. Differential scanning calorimetry results demonstrated that both amidated aurein peptides perturbed the overall phase structure of DMPC/DMPG bilayers, but perturbed the POPC/POPG chains less. The nature of the perturbation of DMPC/DMPG bilayers was most likely micellization, and for the POPC/POPG bilayers, distorted toroidal pores or localized membrane aggregate formation. Calcein release assay results showed that aurein peptide-induced membrane leakage was more severe in DMPC/DMPG liposomes than in POPC/POPG liposomes, and that aurein 2.2 induced higher calcein release than aurein 2.3 and aurein 2.3-COOH from 1:1 and 3:1 POPC/POPG liposomes. Finally, DiSC35 assay data further delineated aurein 2.2 from the others by showing that it perturbed the lipid membranes of intact S. aureus C622 most efficiently, whereas aurein 2.3 had the same efficiency as gramicidin S, and aurein 2.3-COOH was the least efficient. Taken together, these data show that the membrane interactions of aurein peptides are affected by the hydrophobic thickness of the lipid bilayers and the PG content.  相似文献   

8.
Lipid modifications of proteins are widespread in nature and play an important role in numerous biological processes. The nonreceptor tyrosine kinase Src is equipped with an N-terminal myristoyl chain and a cluster of basic amino acids for the stable membrane association of the protein. We used 2H NMR spectroscopy to investigate the structure and dynamics of the myristoyl chain of myr-Src(2-19), and compare them with the hydrocarbon chains of the surrounding phospholipids in bilayers of varying surface potentials and chain lengths. The myristoyl chain of Src was well inserted in all bilayers investigated. In zwitterionic 1,2-dimyristoyl-sn-glycero-3-phosphocholine membranes, the myristoyl chain of Src was significantly longer and appears “stiffer” than the phospholipid chains. This can be explained by an equilibrium between the attraction attributable to the insertion of the myristoyl chain and the Born repulsion. In a 1,2-dimyristoyl-sn-glycero-3-phosphocholine/1,2-dimyristoyl-sn-glycero-3-[phospho-L-serine] membrane, where attractive electrostatic interactions come into play, the differences between the peptide and the phospholipid chain lengths were attenuated, and the molecular dynamics of all lipid chains were similar. In a much thicker 1,2-dipalmitoyl-sn-glycero-3-phosphocholine/1,2-dipalmitoyl-sn-glycero-3-[phospho-L-serine]/cholesterol membrane, the length of the myristoyl chain of Src was elongated nearly to its maximum, and the order parameters of the Src chain were comparable to those of the surrounding membrane.  相似文献   

9.
Rat liver 60-kDa lysophospholipase-transacylase catalyzes not only the hydrolysis of 1-acyl-sn-glycero-3-phosphocholine, but also the transfer of its acyl chain to a second molecule of 1-acyl-sn-glycero-3-phosphocholine to form phosphatidylcholine (H. Sugimoto, S. Yamashita, J. Biol. Chem. 269 (1994) 6252–6258). Here we report the detailed characterization of the transacylase activity of the enzyme. The enzyme mediated three types of acyl transfer between donor and acceptor lipids, transferring acyl residues from: (1) the sn-1 to -1(3); (2) sn-1 to -2; and (3) sn-2 to -1 positions. In the sn-1 to -1(3) transfer, the sn-1 acyl residue of 1-acyl-sn-glycero-3-phosphocholine was transferred to the sn-1(3) positions of glycerol and 2-acyl-sn-glycerol, producing 1(3)-acyl-sn-glycerol and 1,2-diacyl-sn-glycerol, respectively. In the sn-1 to -2 transfer, the sn-1 acyl residue of 1-acyl-sn-glycero-3-phosphocholine was transferred to not only the sn-2 positions of 1-acyl-sn-glycero-3-phosphocholine, but also 1-acyl-sn-glycero-3-phosphoethanolamine, producing phosphatidylcholine and phosphatidylethanolamine, respectively. 1-Acyl-sn-glycero-3-phospho-myo-inositol and 1-acyl-sn-glycero-3-phosphoserine were much less effectively transacylated by the enzyme. In the sn-2 to -1 transfer, the sn-2 acyl residue of 2-acyl-sn-glycero-3-phosphocholine was transferred to the sn-1 position of 2-acyl-sn-glycero-3-phosphocholine and 2-acyl-sn-glycero-3-phosphoethanolamine, producing phosphatidylcholine and phosphatidylethanolamine, respectively. Consistently, the enzyme hydrolyzed the sn-2 acyl residue from 2-acyl-sn-glycero-3-phosphocholine. By the sn-2 to -1 transfer activity, arachidonic acid was transferred from the sn-2 position of donor lipids to the sn-1 position of acceptor lipids, thus producing 1-arachidonoyl phosphatidylcholine. When 2-arachidonoyl-sn-glycero-3-phosphocholine was used as the sole substrate, diarachidonoyl phosphatidylcholine was synthesized at a rate of 0.23 μmol/min/mg protein. Thus, 60-kDa lysophospholipase-transacylase may play a role in the synthesis of 1-arachidonoyl phosphatidylcholine needed for important cell functions, such as anandamide synthesis.  相似文献   

10.
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   

11.
We used wide angle x-ray scattering (WAXS) from stacks of oriented lipid bilayers to measure chain orientational order parameters and lipid areas in model membranes consisting of mixtures of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC)/cholesterol and 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC)/cholesterol in fluid phases. The addition of 40% cholesterol to either DOPC or DPPC changes the WAXS pattern due to an increase in acyl chain orientational order, which is one of the main properties distinguishing the cholesterol-rich liquid-ordered (Lo) phase from the liquid-disordered (Ld) phase. In contrast, powder x-ray data from multilamellar vesicles does not yield information about orientational order, and the scattering from the Lo and Ld phases looks similar. An analytical model to describe the relationship between the chain orientational distribution and WAXS data was used to obtain an average orientational order parameter, Sx-ray. When 40% cholesterol is added to either DOPC or DPPC, Sx-ray more than doubles, consistent with previous NMR order parameter measurements. By combining information about the average chain orientation with the chain-chain correlation spacing, we extended a commonly used method for calculating areas for gel-phase lipids to fluid-phase lipids and obtained agreement to within 5% of literature values.  相似文献   

12.
Wang T  Cady SD  Hong M 《Biophysical journal》2012,102(4):787-794
The M2 protein of the influenza A virus acts both as a drug-sensitive proton channel and mediates virus budding through membrane scission. The segment responsible for causing membrane curvature is an amphipathic helix in the cytoplasmic domain of the protein. Here, we use 31P and 13C solid-state NMR to examine M2-induced membrane curvature. M2(22–46), which includes only the transmembrane (TM) helix, and M2(21–61), which contains an additional amphipathic helix, are studied. 31P chemical shift lineshapes indicate that M2(21–61) causes a high-curvature isotropic phase to both cholesterol-rich virus-mimetic membranes and 1,2-dimyristoyl-sn-glycero-3-phosphocholine bilayers, whereas M2(22–46) has minimal effect. The lamellar and isotropic domains have distinct 31P isotropic chemical shifts, indicating perturbation of the lipid headgroup conformation by the amphipathic helix. 31P- and 13C-detected 1H T2 relaxation and two-dimensional peptide-lipid correlation spectra show that M2(21–61) preferentially binds to the high-curvature domain. 31P linewidths indicate that the isotropic vesicles induced by M2(21–61) are 10–35 nm in diameter, and the virus-mimetic vesicles are smaller than the 1,2-dimyristoyl-sn-glycero-3-phosphocholine vesicles. A strong correlation is found between high membrane curvature and weak drug-binding ability of the TM helix. Thus, the M2 amphipathic helix causes membrane curvature, which in turn perturbs the TM helix conformation, abolishing drug binding. These NMR experiments are applicable to other curvature-inducing membrane proteins such as fusion proteins and antimicrobial peptides.  相似文献   

13.
Binding of phospholipase A2 from porcine pancreas and from Naja melanoleuca venom to vesicles of 1,2-di(tetradecyl)-rac-glycero-3-phosphocholine (diether-PC14) is studied in the presence and absence of 1-tetradecanoyl-sn-glycero-3-phosphocholine and myristic acid. The bound enzyme coelutes with the vesicles during gel filtration through a nonequilibrated Sephadex G-100 column, modifies the phase transition behavior of bilayers, and exhibits an increase in fluorescence intensity accompanied by a blue shift. Using these criteria it is demonstrated that the snake-venom enzyme binds to bilayers of the diether-PC14 alone. In contrast, the porcine enzyme binds only to ternary codispersions of dialkyl (or diacyl) phosphatidylcholine, lysophosphatidylcholine and fatty acid. Binding of the pig-pancreatic enzyme to vesicles of the diether-PC14 could not be detected even after long incubation (up to 24 h) below, at, or above the phase-transition temperature, whereas the binding in the presence of products is almost instantaneous and observed over a wide temperature range. Thus incorporation of the products in substrate dispersions increases the binding affinity rather than increase the rate of binding. The results are consistent with the hypothesis that the pancreatic enzyme binds to defect sites at the phase boundaries in substrate bilayers induced by the products. The spectroscopically obtained hyperbolic binding curves can be adequately described by a single equilibrium by assuming that the enzyme interacts with discrete sites. The binding experiments are supported by kinetic studies.  相似文献   

14.
Bombolitin II (BLT2) is one of the hemolytic heptadecapeptides originally isolated from the venom of a bumblebee. Structure and orientation of BLT2 bound to 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) membranes were determined by solid-state 31P and 13C NMR spectroscopy. 31P NMR spectra showed that BLT2-DPPC membranes were disrupted into small particles below the gel-to-liquid crystalline phase transition temperature (Tc) and fused to form a magnetically oriented vesicle system where the membrane surface is parallel to the magnetic fields above the Tc. 13C NMR spectra of site-specifically 13C-labeled BLT2 at the carbonyl carbons were observed and the chemical shift anisotropies were analyzed to determine the dynamic structure of BLT2 bound to the magnetically oriented vesicle system. It was revealed that the membrane-bound BLT2 adopted an α-helical structure, rotating around the membrane normal with the tilt angle of the helical axis at 33°. Interatomic distances obtained from rotational-echo double-resonance experiments further showed that BLT2 adopted a straight α-helical structure. Molecular dynamics simulation performed in the BLT2-DPPC membrane system showed that the BLT2 formed a straight α-helix and that the C-terminus was inserted into the membrane. The α-helical axis is tilted 30° to the membrane normal, which is almost the same as the value obtained from solid-state NMR. These results suggest that the membrane disruption induced by BLT2 is attributed to insertion of BLT2 into the lipid bilayers.  相似文献   

15.
The interaction of cationic pentalysine with phospholipid membranes was studied by using phosphorus and deuterium Nuclear Magnetic Resonance (NMR) of headgroup deuterated dimyristoyl phosphatidylcholine (DMPC) and dimyristoyl phosphatidylserine (DMPS). In the absence of pentalysine, some of the deuterium and phosphorus spectra of DMPC/DMPS 5:1 (m:m) membranes gave lineshapes similar to those of partially-oriented bilayers with the planes of the bilayers being parallel to the magnetic field. The deuterium NMR data show that the quadrupolar splittings of the deuterated methylenes of the DMPC headgroup are not affected by adsorption of pentalysine on the PC/PS membranes. By contrast, the pentalysine produces significant changes in the quadrupolar splittings of the negatively charged DMPS headgroup. The results are discussed in relation to previous 2H NMR investigations of phospholipid headgroup perturbations arising from bilayer interaction with cationic molecules.Abbreviations NMR nuclear magnetic resonance - DMPC 1,2-dimyristoyl-sn-glycero-3-phosphocholine - DMPS 1,2-dimyristoyl-sn-glycero-3-phosphoserine - POPC 1-palmitoyl, 2-oleyl-sn-glycero-3-phosphocholine - POPG 1-palmitoyl-2-oleyl-sn-glycero-3-phosphoglycerol - PC phosphatidylcholine - PS phosphatidyl serine - PG phosphatidylglycerol - HEPES N-(2-hydroxy-ethyl)piperazine-N-2-ethanesulfonic acid - TRIS tris-(hydroxymethyl)aminoethane - EDTA ethylenediamine-tetra-acetic acid  相似文献   

16.
Deuterium (2H) NMR has been used to observe perturbation of dipalmitoylphosphatidylcholine (DPPC) bilayers by the pulmonary surfactant protein B (SP-B) at concentrations up to 17% (w/w). Previous 2H NMR studies of DPPC/dipalmitoylphosphatidylglycerol (DPPG) (7:3) bilayers containing up to 11% (w/w) SP-B and DPPC bilayers containing up to 11% (w/w) synthetic SP-B indicated a slight effect on bilayer chain order and a more substantial effect on motions that contribute to decay of quadrupole echoes obtained from bilayers of deuterated DPPC. This is consistent with the perturbation of headgroup-deuterated DPPC reported here for bilayers containing 6 and 9% (w/w) SP-B. For the higher concentrations of SP-B investigated in the present work, 2H NMR spectra of DPPC deuterated in both the headgroup and chain display a prominent narrow component consistent with fast, large amplitude reorientation of some labeled lipid. Similar spectral perturbations have been reported for bilayers in the presence of the antibiotic polypeptide nisin. The observation of large amplitude lipid reorientation at high SP-B concentration could indicate that SP-B can induce regions of high bilayer curvature and thus provides some insight into local interaction of SP-B with DPPC. Such local interactions may be relevant to the formation, in vitro and in vivo, of tubular myelin, a unique structure found in extracellular pulmonary surfactant, and to the delivery of surfactant material to films at the air–water interface.Abbreviations DPPC 1,2-dipalmitoyl-sn-glycero-3-phosphocholine - DPPG 1,2-dipalmitoyl-sn-glycero-3-phosphoglycerol - DPPC-d62 1,2-perdeuterodipalmitoyl-sn-glycero-3-phosphocholine - DPPC-d4 1,2-dipalmitoyl-sn-glycero-3-phospho-(, perdeutero)-choline  相似文献   

17.
The opening and closing of voltage-gated potassium (Kv) channels are controlled by several conserved Arg residues in the S4 helix of the voltage-sensing domain. The interaction of these positively charged Arg residues with the lipid membrane has been of intense interest for understanding how membrane proteins fold to allow charged residues to insert into lipid bilayers against free-energy barriers. Using solid-state NMR, we have now determined the orientation and insertion depth of the S4 peptide of the KvAP channel in lipid bilayers. Two-dimensional 15N correlation experiments of macroscopically oriented S4 peptide in phospholipid bilayers revealed a tilt angle of 40° and two possible rotation angles differing by 180° around the helix axis. Remarkably, the tilt angle and one of the two rotation angles are identical to those of the S4 helix in the intact voltage-sensing domain, suggesting that interactions between the S4 segment and other helices of the voltage-sensing domain are not essential for the membrane topology of the S4 helix. 13C-31P distances between the S4 backbone and the lipid 31P indicate a ∼ 9 Å local thinning and 2 Å average thinning of the DMPC (1,2-dimyristoyl-sn-glycero-3-phosphochloline)/DMPG (1,2-dimyristoyl-sn-glycero-3-phosphatidylglycerol) bilayer, consistent with neutron diffraction data. Moreover, a short distance of 4.6 Å from the guanidinium Cζ of the second Arg to 31P indicates the existence of guanidinium phosphate hydrogen bonding and salt bridges. These data suggest that the structure of the Kv gating helix is mainly determined by protein-lipid interactions instead of interhelical protein-protein interactions, and the S4 amino acid sequence encodes sufficient information for the membrane topology of this crucial gating helix.  相似文献   

18.
The mixed Langmuir monolayers composed of model constituents of biological membranes, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC), 2-oleoyl-1-palmitoyl-sn-glycero-3-phosphocholine (POPC), and 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC), were investigated to provide information on the intermolecular interactions between these membrane components and the physiologically active vitamin E–α-tocopherol (TF), as well as on the phase behavior of these mixed systems. Additionally, topography of these monolayers transferred onto the mica support was investigated by the inverted metallurgical microscope. Morphological characteristics were directly observed by Brewster angle microscopy (BAM). From the surface pressure–area isotherms and the analysis of physicochemical parameters (compressibility and mean molecular area at the maximum compressibility) it was found that depending on the acyl chains saturation degree, TF has different effect on the phospholipids. In the case of DPPC, the addition of TF to the phospholipid film causes destabilization of the ordered hydrocarbon chains, while in the POPC/DOPC–TF systems, the attractive interactions are responsible for the monolayer increased stability. Thus, the results of these studies confirm the hypothesis that α-tocopherol may play a role in the stabilization of biological membranes.  相似文献   

19.
Ceramide is an important intermediate in sphingolipid homeostasis. We examined how colipids, with negative intrinsic curvature and which may induce curvature stress in the bilayers, affected the segregation of palmitoyl ceramide (PCer). Such colipids include 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine (POPE), 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE), and tetra-linoleoyl cardiolipin (CL). In 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) bilayers, PCer formed ordered, gel-like domains at concentrations above 10 mol% at 23°C, as evidenced by the change in the average lifetime of the trans-parinaric acid emission. When POPE or DOPE were included in the DOPC bilayer (at 20:80 or 40:60 POPE or DOPE to DOPC, by mol), the lateral segregation of PCer was facilitated in a concentration-dependent manner, and less PCer was required for the formation of the ordered ceramide-rich domains. Inclusion of CL in the DOPE bilayer (at 10:90 or 20:80 CL to PC, by mol) also caused a similar facilitation of the lateral segregation of PCer. The PCer-rich domains formed in the presence of POPE, DOPE, or CL in DOPC bilayers were slightly more thermostable (by 2–10°C) when compared to PCer-rich domains in DOPC-only bilayers. Nonlamellar phases were not present in bilayers in which the effects of POPE or DOPE on PCer segregation were the largest, as verified by 31P NMR. When palmitoyl sphingomyelin was added to the different bilayer compositions at 5 mol%, relative to the phospholipids, PCer segregated into gel domains at lower concentrations (2–3 mol% PCer), and the effect of POPE on PCer segregation was eliminated. We suggest that the effects of POPE, DOPE, and CL on PCer segregation was in part influenced by their effects on membrane curvature stress and in part because of unfavorable interactions with PCer due to their unsaturated acyl chains. These lipids are abundant in mitochondrial membranes and are likely to affect functional properties of saturated ceramides in them.  相似文献   

20.
The increased use of plant sterols as cholesterol-lowering agents warrants further research on the possible effects of plant sterols in membranes. In this study, the effects of the incorporation of cholesterol, campesterol, β-sitosterol and stigmasterol in phospholipid bilayers were investigated by differential scanning calorimetry (DSC), resonance energy transfer (RET) between trans parinaric acid (tPA) and 2-(6-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)hexanoyl-1-hexadecanoyl-sn-glycero-3-phosphocholine (NBD-PC), and Triton X-100-induced solubilization. The phospholipids used were 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC), d-erythro-N-palmitoyl-sphingomyelin (PSM), and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC). In DSC experiments, it was demonstrated that the sterols differed in their effect on the melting temperatures of both the sterol-poor and the sterol-rich domains in DPPC and PSM bilayers. The plant sterols gave rise to lower temperatures of both transitions, when compared with cholesterol. The plant sterols also resulted in lower transition temperatures, in comparison with cholesterol, when sterol-containing DPPC and PSM bilayers were investigated by RET. In the detergent solubilization experiments, the total molar ratio between Triton X-100 and POPC at the onset of solubilization (Rt,sat) was higher for bilayers containing plant sterols, in comparison with membranes containing cholesterol. Taken together, the observations presented in this study indicate that campesterol, β-sitosterol and stigmasterol interacted less favorably than cholesterol with the phospholipids, leading to measurable differences in their domain properties.  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