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.
In many lipid-containing systems (intact membranes, lipid-water and proteinlipid-water phases) the hydrocarbon chains are known to undergo a reversible temperature-dependent transition between a highly disordered (type α) and a partly ordered (type β) conformation; in the β conformation the chains, stiff and all parallel, are packed with rotational disorder according to a two-dimensional hexagonal lattice. This work describes an X-ray diffraction and freeze-fracturing electron microscope study of the phases involved in this conformational transition. Several lipid-water systems were studied: mitochondrial lipids; phosphatidic acid, synthetic lecithin; hen egg lecithin. The conformational transition is found to be a complex phenomenon dependent upon the chemical composition of the lipids, the amount of water and temperature. When the lipid is a pure chemical species the transition involves two phases; one with all the chains in the α conformation the other with all the chains in the β conformation. If the chains are heterogeneous, then from the onset of the transition from type α, they segregate into regions with different conformation, presumably according to their length and degree of saturation. One of the phases (Lαβ) consists of regularly stacked lipid lamellae, each of which is a disordered mosaic of two types of domains; one with the chains in the α, the other in the β conformation. In another phase (Lγ) each lipid lamella is formed by one monolayer of type α and one of type β, joined by their apolar faces. Two other phases (Pγ and Pαβ) display two-dimensional lattices, and consist of lipid lamellae distorted by wave-like ripples, with an ordered segregation of domains in the α and in the β conformation. The number and the structure of the phases involved in the conformational transition are strongly dependent upon the heterogeneity of the hydrocarbon chains and upon the charge and hydration of the polar groups. The results of this study have a bearing on the conformation of the chains in membranes, and on the possible biological significance of conformational transitions.  相似文献   

2.
Fourier transform infrared spectroscopy was used to characterize the lamellar phases of 1,3-dipalmitoyl-sn-glycero-2-phosphocholine (1,3-DPPC), a positional isomer of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (1,2-DPPC). The molecule exists in three distinct phases over the temperature interval 0–70°C. In the low-temperature (Lc) phase, the spectra are indicative of acyl chains packed in an orthorhombic subcell, while the carbonyl groups and phosphate ester at the head group show evidence of only partial hydration. The transition from the low-temperature (Lc) phase to the intermediate-temperature (Lβ) phase at 25°C corresponds to a temperature-induced head-group hydration in which the hydration of the phosphate and carbonyl ester groups results in the reorganization of the hydrocarbon chain-packing subcell from orthorhombic to hexagonal. The transition from the intermediate (Lβ) to the high-temperature (Lα) phase at 37°C is a gel-to-liquid-crystalline phase transition analogous to the 41.5°C transition of 1,2-DPPC. The spectra of the acyl-chain carbonyl groups show evidence of significant differences in molecular conformation at the carbonyl esters in the Lc phase. In the Lβ and Lα phases, the carbonyl band contour becomes much more symmetric. However, two components are clearly present in the spectra indicating that the sn-1 and sn-3 carbonyls experience slightly different environments. The observed differences are likely due to a preferred conformation of the phosphocholine group relative to the glycerol backbone. Indications from the infrared spectra of differences in the structure of the C=O groups provide a possible explanation for the selection of the sn-1 chain of 1,3-DPPC by phospholipase A2 on the basis of a preferred head group conformation.  相似文献   

3.
The molecular conformation of the monoclinic crystalline polymorph of prostaglandin A1 has been determined by X-ray diffraction techniques. The space group is P21 with a = 13.637 (2), b = 7.567 (1), I c = 10.576 (2) Å, β = 107.37 (3)°; Dc = 1.073 g·cm−3 for Z = 2. The molecular conformation is characterized by the nearly parallel arrangement of the C1–C7 and C13–C20 side chains, with a general flattening of the overall structure when compared with the orthorhombic polymorph. The cyclopentenone moiety assumes a C8 envelope conformation with C8 and O9 displaced +0.29 Å and −0.18 Å from the C9–C10=C11–C12 plane respectively. Concerted, small variations of the torsion angles, primarily about the C8–C12, C14–C15 and C16–C17 bonds, bring the monoclinic and orthorhombic conformations into coincidence.  相似文献   

4.
The transformation of polypeptide chains from their globular native structure to fibrillar aggregates has been a matter of great concern because of the involvement of these aggregates in the onset of several degenerative diseases. These aggregates exhibit highly ordered cross β sheet structures and are known as ‘amyloids’. Formation of amyloids in the body is associated with cytotoxicity due to direct interaction of the aggregated species with the cell membrane leading to cellular permeability or due to loss of functionality of the proteins involved in the amyloid formation. The preference of polypeptide chains to remain in their native conformation or to aggregate into amyloids is guided by several factors such as its conformation at specific condition, concentration, physicochemical properties of the amino acid sequence and so on. In the current review, we have reviewed the different factors that guide the transition of proteins from their natively folded state to the amyloidogenic state. Understanding the critical determinants of amyloidogenesis is vital towards deciphering the molecular mechanism of amyloidogenesis and for the development of effective therapeutics against amyloidosis. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
NMR spin-half pair dipolar echo measurements are reported for the lamellar (dispersions and multibilayer stacks) and hexagonal phases of potassium palmitate/2H2O mixtures. In the lamellar Lβ and Lγ (gel) phases the alkyl chains are rigid and perfectly ordered, while in the lamellar Lα and hexagonal phases they are flexible and disordered. In particular, the measurements show that in the fluid lamellar Lα phase the chain is “bent” at the C9–C10 segment; but is “straight” in the hexagonal phase.  相似文献   

6.
We have examined the infrared absorption spectra from 4000 to 250 cm?1 of multilayers of phosphatidylcholine, phosphatidylethanolamine, and phosphatidylcholine/phosphatidylethanolamine (1:1 m/m) as a function of hydration, pH, and fatty acid composition. Characteristic splittings of the CH2 bending and rocking modes and the position of the phosphoryl absorption at ca. 1240 cm?1 reveal differences in acyl chain packing and head group conformation in the various films. Spectra demonstrate the importance of NH → O hydrogen bonding of the ethanolamine head group and the prerequisite head group conformation (tangent to the multilayer plane) in establishing these structural differences. The general appearance of the P-O-C stretching region (~1050 cm?1) in the pure and mixed films further supports these conclusions and shows that the spectra clearly distinguish among the different head group orientations. Self-association of phosphatidylethanolamine is sometimes sufficient to prevent formation of mixed phases with phosphatidylcholine at neutral pH. The amount of fine structure, particularly in the low-frequency (800?200 cm?1) region, in spectra of films of anhydrous, saturated-chain phospholipids decreases considerably when the films are monohydrated, when mixed phases exist, or when there are unsaturations in the acyl chains. These changes likely result from decreased crystal field effects in the spectra as the phosphatide packing density is decreased by any of the above procedures. Furthermore, the absence of other changes upon complete hydration of phosphatidylcholine films suggests that only the initial water is tightly bound to the lipid.  相似文献   

7.
Intact and excised cultured pea roots (Pisum sativum L. cv Alaska) were treated with chlorsulfuron at concentrations ranging from 2.8 ×10–4 M to 2.8×10–6 M. At all concentrations this chemical was demonstrated to inhibit the progression of cells from G2 to mitosis (M) and secondarily from G1 to DNA synthesis (S). The S and M phases were not directly affected, but the transition steps into those phases were inhibited. Total protein synthesis was unaffected by treatment of intact roots with 2.8×10–6 M chlorsulfuron. RNA synthesis was inhibited by 43% over a 24-h treatment period. It is hypothesized that chlorsulfuron inhibits cell cycle progression by blocking the G2 and G1 transition points through inhibition of cell cycle specific RNA synthesis.  相似文献   

8.
Optical rotation data indicate that xanthan can exist both in an ordered and a disordered conformation. Using molecular weights obtained from light scattering measurements and contour length distributions obtained from electron micrographs, we find that a native, filtered xanthan exposed to low salinity (< 10?4M NaCl) and subsequently returned to 0.1M NaCl has a highly elongated structure with a mass per unit length of 1950 ± 200 Dalton/nm. Our data thus suggest that the ordered conformation of this xanthan is double stranded. We find that native, filtered xanthan in 0.1M NH4Ac has a nearly similar structure, but exists in part as aggregates of varying shape and size. Electron micrographs of these xanthans in 10?4M NH4Ac (the disordered conformation) display a mixture of species ranging from unaggregated single- or perfectly matched double-stranded species, to double-stranded chains branching into its two subunits as well as double-stranded chains with different degrees of mismatching. This study suggests that the perfectly matched antiparallel or parallel double-stranded chain constitutes the lowest free energy state of the ordered conformation of xanthan in dilute aqueous solution.  相似文献   

9.
This work describes the structure of a variety of lecithin-water phases observed below the “melting” temperature of the hydrocarbon chains, with special emphasis on the conformation of the chains. The lecithins studied in this work are the homologous series dioctanoyl to distearoyl, 2-decanoyl-1-stearoyl, and a preparation from hen eggs. The hydrocarbon chains are found to adopt a variety of conformations in addition to type α, the liquid-like organization observed above the melting temperature. Type β: the chains are stiff and parallel, oriented at right angles to the plane of the lamellae and packed with rotational disorder in a two-dimensional hexagonal lattice (a ~ 4.85 A?). Type β′: similar to β, but with the chains tilted with respect to the normal to the lamellae. Type δ: the chains are probably coiled into helices, whose axes are perpendicular to the plane of the polar groups and are packed with rotational disorder in a two-dimensional square lattice (a ~ 4.80 A?), α is the predominant conformation, common to most lipids in the presence of water and at sufficiently high temperature, and the one more relevant to membranes; β is observed at lower temperatures in lipids whose chains are heterogeneous and in the presence of very small amounts of water; β′ is found in synthetic lecithins with identical chains, in the presence of variable amounts of water; δ is observed in dry lecithins. A highly ordered crystalline phase, yet displaying rotational disorder of the chains, is observed in almost dry lecithins. Most of the phases are lamellar, and contain one lipid bilayer per repeat unit. Two phases display two-dimensional lattices: Pδ, formed by ribbon-like elements with the chains in the δ conformation; Pβ′, formed by lamellae of type β′ distorted by periodic ripples. The results emphasize the clear-cut difference between the liquid-like and the other types of partly ordered conformations, as well as the correlations which exist between the chemical composition and the structure of the lipids below the melting temperature of the chains.  相似文献   

10.
One and two-dimensional solid-state NMR experiments are discussed that permit probing local structure and overall molecular conformation of membrane-embedded polypeptides under Magic Angle Spinning. The functional dependence of a series of anisotropic recoupling schemes is analyzed using theoretical and numerical methods. These studies lead to the construction of a set of polarization dephasing or transfer units that probe local backbone conformation and overall molecular orientation within the same NMR experiment. Experimental results are shown for a randomly oriented peptide and for two model membrane–peptides reconstituted into lipid bilayers and oriented on polymer films according to a method proposed by Bechinger etal. [J. Am. Chem. Soc., 124, (2002), 1146–1147].  相似文献   

11.
1H NMR spectroscopy was applied to study the reactions of cis-[Pd(L)(H2O)2]2+ complexes (L is en, pic and dpa) with the N-acetylated tripeptides L-methionylglycylglycine, MeCOMet–Gly–Gly, and glycyl–L-methionyl–glycine, MeCOGly–Met–Gly. All reactions were performed in the pH range 2.0–2.5 with equimolar amounts of the cis-[Pd(L)(H2O)2]2+ complex and the tripeptide at 60 °C. The hydrolytic reactions of the cis-[Pd(en)(H2O)2]2+, cis-[Pd(pic)(H2O)2]2+ and cis-[Pd(dpa)(H2O)2]2+ complexes with MeCOMet–Gly–Gly were regioselective and only the amide bond involving the carboxylic group of methionine was cleaved. However, in the reactions of these three Pd(II) complexes with MeCOGly–Met–Gly, two amide bonds, Met–Gly and MeCO–Gly, were cleaved. From UV–Vis spectrophotometry studies, it was found that the rate-determining step of these hydrolytic reactions is the monodentate coordination of the corresponding Pd(II) complex to the sulfur atom of the methionine side chain. The rate of the cleavage of these amide bonds is dependent on the nature of the bidentate coordinated diamine ligand L (en > pic > dpa). The hydrolytic reaction of cis-[Pd(L)(H2O)2]2+-type complexes with MeCOMet–Gly–Gly, containing the methionine side chain in the terminal position of the peptide, is regioselective while in the reaction of these Pd(II) complexes with MeCOGly–Met–Gly, none selective cleavage of the peptide occurs. This study contributes to a better understanding of the selective cleavage of methionine-containing peptides employing palladium(II) complexes as catalysts.  相似文献   

12.
Botryosphaeran, a (13;16)-β-d-glucan produced by Botryosphaeria rhodina MAMB-05, was found to be present in a triple helix conformation from helix–coil transition studies using Congo Red. The triple helix conformation was disrupted at increasing alkali concentrations. Conformational changes were also observed using phenanthrene as a fluorescent probe, and the fluorescence intensity decreased 80% in the presence of dimethyl sulfoxide. The results confirmed the triple helix conformation of botryosphaeran, an important property manifesting biological response modifying activity.  相似文献   

13.
The effect of cholesterol on vibrational spectra in the non polar and in the polar region of dimyristoyl phosphatidylserine (DMPS) and of phosphatidylserine from bovine spinal cord (PS) has been investigated. The small shifts in the methylene CH stretching frequencies after taking into account the contribution of the cholesterol spectrum were interpreted as a combined effect of cholesterol on the conformation of the chains and of the lesser contributions of the cholesterol methyl groups. Cholesterol also influences the ratio of the trans (1465 cm–1) to the lower wavelength (1457 cm–1) CH2 bending bands. No significant direct effect of cholesterol on the vibration of the polar residues was discerned. The small shift of the carboxylate band observed below the phase transition is probably due to the change in the intermolecular zwitterions when the average distance between the neighboring polar groups increases due to incorporation of cholesterol molecules.Abbreviations PS phosphatidylserine natural - DMPS dimyristoyl phosphatidylserine - DPPC dipalmitoyl phosphatidylcholine - FTIR Fourier transform infrared spectroscopy - DSC differential scanning calorimetry - PE phosphatidylethanolamine Offprint requests to: D. Bach  相似文献   

14.
GTPase activating protein (GAP) is a known regulator of ras-p21 activity and is a likely target of ras-induced mitogenic signaling. The domains of GAP that may be involved in this signaling are unknown. In order to infer which domains of GAP may be involved, we have performed molecular dynamics calculations of GAP complexed to wild-type and oncogenic (Val 12–containing) ras-p21, both bound to GTP. We have computed and superimposed the average structures for both complexes and find that there are four domains of GAP that undergo major changes in conformation: residues 821–851, 917–924, 943–953, and 1003–1020. With the exception of the 943–953 domain, none of these domains is involved in making contacts with ras-p21, and all of them occur on the surface of the protein, making them good candidates for effector domains. In addition, three ras-p21 domains undergo major structural changes in the oncogenic p21-GAP complex: 71–76 from the switch 2 domain; 100–108, which interacts with SOS, jun and jun kinase (JNK); and residues 122–138. The change in conformation of the 71–76 domain appears to be induced by changes in conformation in the switch 1 domain (residues 32–40) and in the adjacent domain involving residues 21–31. In an accompanying paper, we present results from microinjection of peptides corresponding to each of these domains into oocytes induced to undergo maturation by oncogenic ras-p21 and by insulin-activated wild-type cellular p21 to determine whether these domain peptides may be involved in ras signaling through GAP.  相似文献   

15.
Summary The kinetics of the transport of the 1-anilino-8-naphthalenesulfonate (ANS, an anionic fluorescent probe of the membrane surface) across phospholipid vesicle membranes have been studied using a stopped-flow rapid kinetic technique. The method has been used to gain detailed information about the mechanism of transport of this probe and to study ionophore-mediated cation transport across the membrane. The technique has also been exploited to study differences between the inside and outside surfaces of vesicles containing phosphatidyl choline (PC).The following is a summary of the major conclusions of this study. (a) Binding of ANS on the outside surface occurs within times shorter than 100 sec while permeation occurs in the time range 5–100 sec. (b) Net transport of ANS occurs with cotransport of alkali cations. (c) The transport rate is maximal in the region of the crystalline to liquidcrystalline phase transition, and the increase correlates with changes in the degree of aggregation of the vesicles. (d) Incorporation of phosphatidic acid (PA), phosphatidyl ethanolamine (PE) or cholesterol into PC membranes decreases the rate of ANS transport. (e) Neutral ionophores (I) of the valinomycin type increase ANS permeability in the presence of alkali cations (M +) by a mechanism involving the transport of a ternaryI–M +-ANS complex. The equilibrium constants for formation of these complexes and their rate constants for their permeation are presented. The maximal turnover number for ANS transport by valinomycin in dimyristoyl PC vesicles at 35°C was 46 per sec. (f) The partitioning of the ionophore between the aqueous and membrane phases and the rate of transfer of an ionophore from one membrane have been determined in kinetic experiments. (g) A method is described for the detection ofI–M + complexes on the membrane surface by their enhancement effects on ANS fluorescence at temperature below the phase transition temperature on monolayer vesicles. The apparent stability constants for severalI–M + complexes are given. (h) Analysis of the effect of ionic strength on the ANS binding to the inside outside surfaces indicates that the electrostatic surface potential (at fixed ionic strength and surface change) is larger for the inside surface than for the outside surface. (i) Analysis of the dependence of the maximal ANS binding for the inside and outside surfaces of vesicles made from PC and a variable mole fraction of PA, PE or cholesterol indicate that the latter three are located preferentially on the inside surface.  相似文献   

16.
Summary The spatial structure of a synthetic 32-residue polypeptide, an analog of the membrane-spanning segment B (residues 34–65) of bacterioopsin ofHalobacterium halobium, incorporated into perdeuterated sodium dodecyl sulfate micelles, was determined from1H NMR data. The structure determination included the following steps: (1) local sructure analysis; (2) structure calculations using the distance geometry program DIANA; (3) systematic search for energetically allowed side-chain rotamers consistent with NOESY crosspeak volumes; (4) random generation of peptide conformations in allowed conformational space. The obtained structure has a righ-handed -helicl region from Lys41 to Leu62 with a kink of 27 at Pro50. The C-cap Gly63 adopts a conformation with =87±6, =43±10o typical to a left-handed helix. The N-terminal part (residues 34–40) is exposed to the aqueous phase and lacks an ordered conformation. The secondary structure of segment B in micelles is consistent with the high-resolution electron cryomicroscopy model of bacteriorhodopsin (Henderson et al. (1990)J. Mol. Biol.,213, 899–929).  相似文献   

17.
Equinatoxin II (EqtII) is a pore-forming protein from Actinia equina that lyses red blood cell and model membranes. Lysis is dependent on the presence of sphingomyelin (SM) and is greatest for vesicles composed of equimolar SM and phosphatidylcholine (PC). Since SM and cholesterol (Chol) interact strongly, forming domains or “rafts” in PC membranes, 31P and 2H solid-state NMR were used to investigate changes in the lipid order and bilayer morphology of multilamellar vesicles comprised of different ratios of dimyristoylphosphatidylcholine (DMPC), SM and Chol following addition of EqtII. The toxin affects the phase transition temperature of the lipid acyl chains, causes formation of small vesicle type structures with increasing temperature, and changes the T2 relaxation time of the phospholipid headgroup, with a tendency to order the liquid disordered phases and disorder the more ordered lipid phases. The solid-state NMR results indicate that Chol stabilizes the DMPC bilayer in the presence of EqtII but leads to greater disruption when SM is in the bilayer. This supports the proposal that EqtII is more lytic when both SM and Chol are present as a consequence of the formation of domain boundaries between liquid ordered and disordered phases in lipid bilayers leading to membrane disruption.  相似文献   

18.
The temperature-induced volume expansion of enzymatically isolated cuticular membranes of twelve plant species was measured. All cuticular membranes exhibited distinct second-order phase transitions in the temperature range of about 40 to 50° C. Increases in the volumes of fruit cuticles (Lycopersicon, Cucumis, Capsicum, Solanum and Malus) were fully reversible, while leaf cuticular membranes (Ficus, Hedera, Nerium, Olea, Pyrus, Picea and Citrus) underwent irreversible structural changes. Below the phase-transition temperature, volumetric expansion coefficients ranged from 0.39·10–6 m3·kg–1·K–1 to 0.62·10–6 m3·kg–1·K–1, and above from 0.60·106 m3·kg–1·K\-1 to 1.41· 10–6 m3·kg–1·K–1. Densities of cuticles at 25° C ranged from 1020 kg·m–3 to 1370 kg·m–3. Expansion coefficients and phase transitions were characteristic properties of the polymer matrix as a composite material, rather than of cutin alone. It is argued that the sudden increase of water permeability above the transition temperature, is caused by an increase of disorder at the interface between the polymer matrix and the soluble cuticular lipids. Possible ecological and physiological consequences of these results for living plants are discussed.Abbreviations CM Cuticular membrane - CU cutin - MX polymer matrix - SCL soluble cuticular lipids (waxes) The authors greatfully acknowledge stimulating discussions with Drs. H. Gruler (Exp. Physik 3, Universität Ulm, FRG) and M. Riederer (Institut für Botanik und Mikrobiologie, Technische Universität München, München, FRG) and financial support by the Deutsche Forschungsgemeinschaft.  相似文献   

19.
Chitosans of different molar masses were prepared by storing freshly prepared samples for up to 6 months at either 4, 25 or 40 °C. The weight-average molar masses, Mw and intrinsic viscosities, [η] were then measured using size exclusion chromatography coupled to multi-angle laser light scattering (SEC-MALLS) and a “rolling ball” viscometer, respectively.The solution conformation of chitosan was then estimated from:
(a) the Mark–Houwink–Kuhn–Sakurada (MHKS) power law relationship [η] = kMwa and
(b) the persistence length, Lp calculated from a new approach based on equivalent radii [Ortega, A., & Garcia de la Torre, J. (2007). Equivalent radii and ratios of radii from solution properties as indicators of macromolecular conformation, shape, and flexibility. Biomacromolecules, 8, 2464–2475].
Both the MHKS power law exponent (a = 0.95 ± 0.01) and the persistence length (L= 16 ± 2 nm) are consistent with a semi-flexible rod type (or stiff coil) conformation for all 33 chitosans studied. A semi-flexible rod conformation was further supported by the Wales–van Holde ratio, the translational frictional ratio and sedimentation conformation zoning.  相似文献   

20.
The ultrastructure of the cell wall and the thylakoid membranes of the thermophilic cyanobacterium Synechococcus lividus was studied by freezefracture electron microscopy after temperature shifts. Different fracture faces of the outer, the cytoplasmic and the thylakoid membranes were demonstrated when the preparation-temperature was in the range of the optimal growth temperature at 52°C or after fixation at 52°C. In the outer membrane of the cell wall two fracture faces with holes and 7.5 nm intramembrane particles were detected. On both the outer (EF) and inner (PF) leaflet of the cytoplasmic membrane randomly distributed particles were demonstrated. The particle density on the PF-face was approx. three times that of the EF-face. The EF-face of the thylakoid membrane exposed rows of particles with an average diameter of 10 nm. The spacing between the particle rows was 35–50 nm. This regular particle arrangement on the EF-face was demonstrated only in a few cases. Mostly the intramembrane particles were distributed randomly on the thylakoid fracture faces. The particle density of thylakoids with a random distribution was approx. in the same range both on the EF-and PF-face. The EF-particles fall into four groups of 9,10,11, and 12.5 nm. The main particle class was the 10 nm class. The PF-face exposed smaller particles with two maxima at 8.5–9 nm and 10 nm. When Synechococcus lividus OH-53s was chilled to temperatures below 30–35°C before the freeze-etch preparation a phase transition took place after the temperature shift. On the fracture faces of the thylakoid and cytoplasmic membranes particle depleted areas occurred. The size of the areas were different in both membranes and dependent on the velocity of cooling. Contrary to Synechococcus lividus OH-53s in the mesophilic Synechococcus strain 6910 the phase transition point was 15°C. The lower phase transition point may be due to a higher content of unsaturated fatty acids.Dedicated to Prof. D. Peters (Hamburg) on the occasion of the 65th anniversary of his birthday  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