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 390 毫秒
1.
Small angle x-ray diffraction revealed a strong influence of the N-terminal influenza hemagglutinin fusion peptide on the formation of nonlamellar lipid phases. Comparative measurements were made on a series of three peptides, a 20-residue wild-type X-31 influenza virus fusion peptide, GLFGAIAGFIENGWEGMIDG, and its two point-mutant, fusion-incompetent peptides G1E and G13L, in mixtures with hydrated phospholipids, either dipalmitoleoylphosphatidylethanolamine (DPoPE), or monomethylated dioleoyl phosphatidylethanolamine (DOPE-Me), at lipid/peptide molar ratios of 200:1 and 50:1. All three peptides suppressed the HII phase and shifted the Lα–HII transition to higher temperatures, simultaneously promoting formation of inverted bicontinuous cubic phases, QII, which becomes inserted between the Lα and HII phases on the temperature scale. Peptide-induced QII had strongly reduced lattice constants in comparison to the QII phases that form in pure lipids. QII formation was favored at the expense of both Lα and HII phases. The wild-type fusion peptide, WT-20, was distinguished from G1E and G13L by the markedly greater magnitude of its effect. WT-20 disordered the Lα phase and completely abolished the HII phase in DOPE-Me/WT-20 50:1 dispersions, converted the QII phase type from Im3m to Pn3m and reduced the unit cell size from ∼38 nm for the Im3m phase of DOPE-Me dispersions to ∼15 nm for the Pn3m phase in DOPE-Me/WT-20 peptide mixtures. The strong reduction of the cubic phase lattice parameter suggests that the fusion-promoting WT-20 peptide may function by favoring bilayer states of more negative Gaussian curvature and promoting fusion along pathways involving Pn3m phase-like fusion pore intermediates rather than pathways involving HII phase-like intermediates.  相似文献   

2.
B Tenchov  R Koynova    G Rapp 《Biophysical journal》1998,75(2):853-866
By means of x-ray diffraction we show that several sodium salts and the disaccharides sucrose and trehalose strongly accelerate the formation of cubic phases in phosphatidylethanolamine (PE) dispersions upon temperature cycling through the lamellar liquid crystalline-inverted hexagonal (Lalpha-HII) phase transition. Ethylene glycol does not have such an effect. The degree of acceleration increases with the solute concentration. Such an acceleration has been observed for dielaidoyl PE (DEPE), dihexadecyl PE, and dipalmitoyl PE. It was investigated in detail for DEPE dispersions. For DEPE (10 wt% of lipid) aqueous dispersions at 1 M solute concentration, 10-50 temperature cycles typically result in complete conversion of the Lalpha phase into cubic phase. Most efficient is temperature cycling executed by laser flash T-jumps. In that case the conversion completes within 10-15 cycles. However, the cubic phases produced by laser T-jumps are less ordered in comparison to the rather regular cubic structures produced by linear, uniform temperature cycling at 10 degrees C/min. Temperature cycles at scan rates of 1-3 degrees C/min also induce the rapid formation of cubic phases. All solutes used induce the formation of Im3m (Q229) cubic phase in 10 wt% DEPE dispersions. The initial Im3m phases appearing during the first temperature cycles have larger lattice parameters that relax to smaller values with continuation of the cycling after the disappearance of the Lalpha phase. A cooperative Im3m --> Pn3m transition takes place at approximately 85 degrees C and transforms the Im3m phase into a mixture of coexisting Pn3m (Q224) and Im3m phases. The Im3m/Pn3m lattice parameter ratio is 1. 28, as could be expected from a representation of the Im3m and Pn3m phases with the primitive and diamond infinite periodic minimal surfaces, respectively. At higher DEPE contents ( approximately 30 wt%), cubic phase formation is hindered after 20-30 temperature cycles. The conversion does not go through, but reaches a stage with coexisting Ia3d (Q230) and Lalpha phases. Upon heating, the Ia3d phase cooperatively transforms into a mixture of, presumably, Im3m and Pn3m phases at about the temperature of the Lalpha-HII transition. This transformation is readily reversible with the temperature. The lattice parameters of the DEPE cubic phases are temperature-insensitive in the Lalpha temperature range and decrease with the temperature in the range of the HII phase.  相似文献   

3.
The effect of cholesterol, a major constituent of eukaryotic cell membranes, on the structure and thermotropic phase behaviour of 1-palmitoyl-2-oleoyl-phosphatidylethanolamine (POPE) dispersed in excess water was examined by synchrotron X-ray diffraction methods. Temperature scans over the range 10-75 °C showed that the gel to liquid-crystalline phase transition decreased from 25 to 10 °C in the presence of 20 mol% cholesterol, and no gel phase could be detected in the wide-angle X-ray scattering (WAXS) intensity profile of mixtures containing 35 mol% cholesterol. The small-angle X-ray scattering (SAXS) intensity profiles showed that the lamellar to nonlamellar phase transition temperature was also decreased in mixtures containing up to 30 mol% cholesterol but the trend was reversed in mixtures containing a higher proportion of cholesterol. There was evidence that the transition of the lamellar liquid-crystal phase is to cubic phases in mixtures containing less than 30 mol% cholesterol. The space group of one of these cubic phases was assigned as Pn3m. This effect of cholesterol on non-bilayer-forming phospholipids is considered in the context of the role of cholesterol in membrane organization and function.  相似文献   

4.
Data on the location and dimensions of the pivotal surfaces in inverse hexagonal (HII) and inverse cubic (QII) phases of phospholipids and glycolipids are reviewed. This includes the HII phases of dioleoyl phosphatidylethanolamine, 2:1 mol/mol mixtures of saturated fatty acids with the corresponding diacyl phosphatidylcholine, and glucosyl didodecylglycerol, and also the QII230/G gyroid inverse cubic phases of monooleoylglycerol and glucosyl didodecylglycerol. Data from the inverse cubic phases are largely compatible with those from inverse hexagonal HII-phases. The pivotal plane is located in the hydrophobic region, relatively close to the polar–apolar interface. The area per lipid at the pivotal plane is similar in size to lipid cross-sectional areas found in the fluid lamellar phase (Lα) of lipid bilayers.  相似文献   

5.
Direct electrometry was used to study the light-induced voltage changes in the Rhodobacter sphaeroides chromatophores adsorbed to a phospholipid-impregnated nitrocellulose film. After the second laser flash, a fast increase in the voltage associated with charge separation was followed by a slower increase attributed to the proton uptake in the QB site of the photosynthetic reaction centers. Kinetics and relative amplitudes of these voltage changes attributed to the QA –. B –. QAQBH2 transition, were measured as a function of pH and temperature between +4 and +40 °C. The kinetics can be approximated by a single exponent above +23 °C (100 µs at +25 °C, pH 7.2), whereas below this temperature, it was a good fit of two exponential approximation (65 µs and 360 µs with similar contributions at +10 °C, pH 7.2). The faster component diminished with an apparent pK 8.5, whereas the slower one was maintained at a constant level until pH 9.5 and then decreased. The calculated activation energy from the temperature dependence of the slower component (55 – 65 kJ/mol) was much higher than that of the faster component (< 10 kJ/mol). The two voltage components can be attributed to the transfer of the first (faster component) and the second (slower component) proton from the reaction center surface to QB. We suggested that higher activation energy of the slower component was due to a conformational change in the reaction center kinetically coupled to the second proton transfer to QBH.The faster component diminished in the presence of 1 M KCl, with an apparent pK 7.5. To explain this observation, we assume that: (i) the midpoint potential of the QA/QA –. redox pair was higher in 1 M KCl because of the reduced surface potential of chromatophores; (ii) the midpoint potential of the QB –./QBH–. redox pair was insensitive to the surface potential change; (iii) the equilibrium constant of the reaction QA –.QB –. QAQBH decreased at high ionic strength.  相似文献   

6.
The effect of glucocerebroside (GlcCer) on the structure and thermotropic phase behavior of aqueous dispersions of palmitoyloleoylphosphatidylethanolamine (POPE) has been examined using simultaneous small-angle and wide-angle X-ray diffraction methods. Binary mixtures of GlcCer:POPE in molar ratios of 2:100, 5:100, 10:100, 20:100, 30:100, and 40:100 were examined in the temperature range 20-90 degrees C. Cubic phase has been observed in binary mixtures comprised of molar ratios greater than 5:100 in the temperature range of 60-90 degrees C upon heating at a rate of 2 degrees C/min. The cubic phase is relatively stable and coexists with inverted hexagonal or lamellar phases. It persists in the codispersions throughout subsequent cooling scans to 30 degrees C. The space group of the cubic phase is determined to be Pn3m or Pn3. The lattice constant of the Pn3m cubic phase was found to be almost constant when it coexists with lamellar liquid-crystal phase. Marked temperature-dependent changes were observed when cubic phase coexists with hexagonal phase or lamellar-gel phases. This is the first report of cubic phases formed by codispersions of glycosphingolipids and phospholipids. The mechanism of cubic phase formation and the interaction between GlcCer and POPE is discussed in terms of the putative biological functions of glycolipids.  相似文献   

7.
Leaf discs of dark-adapted tobacco plants were excited by 2 flashes and kept in darkness at 20 °C for various time periods, then thermoluminescence emission was recorded without freezing the sample. The B band at 30 °C decreased with a half-time t1/2~1 min and the AG band at 45 °C with a t1/2~5 min. This corresponds to the decay kinetics of S2/3 in PS II centres in the state S2/3 QB- (B band) or S2/3 QB. Assuming that the 45 °C band is an ‘afterglow’ emission originating from those centres with an oxidized QB on which an electron is back-transferred from stroma reductants through a pathway induced by warming, the theoretical ratio of the B and AG band was compared to that measured experimentally. After 2 or 3 flashes producing mainly S3, the intensity of AG band encompassed several fold that of the B band, because recombining S3 recreated S2 QB AG-emitting centres. In order to confirm that the AG band is governed by the heat-induced activation of a dark QB-reducing pathway rather than by PS II charge recombination, the AG emission was characterized in triazine-resistant Chenopodium album weed biotypes. In these mutants where the QB pocket is altered, the B band is strongly downshifted to 18 °C, compared to 32 °C in the wild type, whereas the AG band is only downshifted by 3 or 4 °C, demonstrating that S2/3 QB- is not the limiting step of the AG emission.  相似文献   

8.
C Czeslik  R Winter  G Rapp    K Bartels 《Biophysical journal》1995,68(4):1423-1429
We used x-ray and neutron diffraction to study the temperature- and pressure-dependent structure and phase behavior of the monoacylglycerides 1-monoelaidin (ME) and 1-monoolein (MO) in excess water. The monoacylglycerides were chosen for investigation of their phase behavior because they exhibit mesomorphic phases with one-, two-, and three-dimensional periodicity, such as lamellar, an inverted hexagonal and bicontinuous cubic phases, in a rather easily accessible temperature and pressure range. We studied the structure, stability, and transformations of the different phases over a wide temperature and pressure range, explored the epitaxial relations that exist between different phases, and established a relationship between the chemical structure of the lipid molecules and their phase behavior. For both systems, a temperature-pressure phase diagram has been determined in the temperature range from 0 to 100 degrees C at pressures from ambient up to 1400 bar, and drastic differences in phase behavior are found for the two systems. In MO-water dispersions, the cubic phase Pn3m extends over a large phase field in the T,p-plane. At temperatures above 95 degrees C, the inverted hexagonal phase is found. In the lower temperature region, a crystalline lamellar phase is induced at higher pressures. The phases found in ME-water include the lamellar crystalline Lc phase, the L beta gel phase, the L alpha liquid-crystalline phase, and two cubic phases belonging to the crystallographic space groups Im3m and Pn3m. In addition, the existence of metastable phases has been exploited. Between coexisting metastable cubic structures, a metric relationship has been found that is predicted theoretically on the basis of the curvature elastic energy approximation only.  相似文献   

9.
Divalent cation (Mn2+, Ca2+) entry into rat parotid acinar cells is stimulated by the release of Ca2+ from the internal agonist-sensitive Ca2+ pool via a mechanism which is not yet defined. This study examines the effect of temperature on Mn2+ influx into internal Ca2+ pool-depleted acini (depl-acini, as a result of carbachol stimulation of acini in a Ca2+-free medium for 10 min) and passive 45Ca2+ influx in basolateral membrane vesicles (BLMV). Mn2+ entry into deplacini was decreased when the incubation temperature was lowered from 37 to 4°C. At 4°C, Mn2+ entry appeared to be inactivated since it was not increased by raising extracellular [Mn2+] from 50 m up to 1 mm. The Arrhenius plot of depletion-activated Mn2+ entry between 37 and 8°C was nonlinear, with a change in the slope at about 21°C. The activation energy (Ea) increased from 10 kcal/mol (Q10=1.7) at 21–37°C to 25 kcal/mol (Q10=3.0) at 21-8°C. Under the same conditions, Mn2+ entry into basal (unstimulated) cells and ionomycin (5 m) permeabilized depl-acini exhibit a linear decrease, with E a of 7.8 kcal/mol (Q10=1.5) and 6.2 kcal/mol (Q10 < 1.5), respectively. These data suggest that depletion-activated Mn2+ entry into parotid acini is regulated by a mechanism which is strongly temperature dependent and distinct from Mn2+ entry into unstimulated acini.As in intact acini, Ca2+ influx into BLMV was decreased (by 40%) when the temperature of the reaction medium was lowered from 37 to 4°C. Kinetic analysis of the initial rates of Ca2+ influx in BLMV at 37°C demonstrated the presence of two Ca2+ influx components: a saturable component, with K Ca =279 ± 43 m, Vmax = 3.38 ± 0.4 nmol Ca2+/mg protein/min, and an apparently unsaturable component. At 4°C, there was no significant change in the affinity of the saturable component, but Vmax decreased by 61% to 1.3 ± 0.4 nmol Ca2+/mg protein/min. There was no detectable change in the unsaturable component. When BLMV were treated with DCCD (5 mm) or trypsin (1100, enzyme to membrane) for 30 min at 37°C there was a 40% decrease in Ca2+ influx. When BLMV were treated with DCCD or trypsin at 4°C and subsequently assayed for Ca2+ uptake at 37°C there was no significant loss of Ca2+ influx. These data suggest that the temperature sensitive high affinity Ca2+ flux component in BLMV is mediated by a protein which undergoes a modification at low temperatures, resulting in decreased Ca2+ transport.We thank Dr. Bruce Baum, Dr. Yukiharu Hiramatsu, Dr. Ofer Eidelman, and our other colleagues for their support during this work.  相似文献   

10.
This study revealed large distinctions between the lamellar and non-lamellar liquid crystalline lipid phases in their spreading at the air/water interface and propensity to form bilayer foam films. Comparative measurements were made for the lamellar L(alpha), the inverted hexagonal H(II) and the bicontinuous cubic Pn3m phases of the phospholipid dipalmitoleoylphosphatidylethanolamine (DPoPE). With regard to monolayer formation, followed as the decrease of surface tension with time, the best spreading (lowest surface tension) was observed for the L(alpha) phase, and poorest spreading (highest surface tension) was recorded for the H(II) phase. The cubic Pn3m phase of DPoPE, induced by temperature cycling, retained an intermediate position between the L(alpha) and H(II) phases. According to their ability to lower surface tension and disintegrate at the air/water interface, the three phases thus order as L(alpha)>Pn3m>H(II). Clearly expressed threshold (minimum) bulk lipid concentrations, C(t), required for formation of stable foam bilayers from these phases, were determined and their values were found to correlate well with the bulk lipid phase behaviour. The C(t) values for L(alpha) and H(II) substantially increase with the temperature. Their Arrhenius plots, ln C(t) versus 1/ T, are linear and intersect at approximately 36-37 degrees C, coinciding with the onset of the bulk L(alpha)-->H(II) phase transition, as determined by differential scanning calorimetry. However, the C(t) value for the Pn3m phase, equal to 30 micro g/mL, was found to be constant over the whole range investigated between 20 degrees C and 50 degrees C. The horizontal C(t) versus T plot for the Pn3m phase crosses the respective plot for the L(alpha) phase at the temperature bounding from below the hysteretic loop of the L(alpha)<-->H(II) transition (approximately 26 degrees C), thus providing a certain insight about the thermodynamic stability of the Pn3m phase relative to the L(alpha) phase. The established strong effect of the particular lipid phase on the formation of monolayers and stable black foam films should be of importance in various in vitro and in vivo systems, where lipid structures are in contact with interfaces and disintegrate there to different extents.  相似文献   

11.
The recombination reactions of Photosystem II have been investigated in vivo in rice leaves by using the thermoluminescence (TL) emission technique. Excitation of dark-adapted leaf segments at 0 °C with different number of single turn-over flashes induced the appearance of complex TL glow curves. The mathematical analysis of these curves showed the existence of four TL components: B1-band (temperature maximum, tmax, at 24 °C, originating from S3QB recombination), B2-band (tmax at 35 °C, from S2QB), AG-band (tmax at 46 °C) and C-band (tmax at 55 °C, from TyrD+QA). Their contributions to the total TL signal were different depending on the number of flashes given. AG-band seems to reflect a special electron transfer from some unknown stroma donor to PS II. Q-band (tmax at 19 °C), originating from S2QA recombination, was recorded after flashing samples incubated in the presence of DCMU. The recombination halftimes (t1/2) at 20 °C of S2QA, S3QB, S2QB and TyrD+QA were, respectively, 0.8 s, 48 s, 74 s and about 1 h. A sharp AG-band (tmax at 50 °C and t1/2 of 210 s) could be also observed after illumination of leaves with far-red light and after a dark incubation period of whole plants. Incubation of leaf segments with 0.5 M NaCl abolished the inductions of AG-band by darkness and far-red illumination, significantly decreased Q-band intensity, whereas induced a strong increase in C-band intensity. The possible inhibition of S2/S3 formation and quinone oxidation by saline stress are discussed.  相似文献   

12.
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   

13.
High-temperature-induced inhibition of the acceptor side of Photosystem II (PS II) was studied in tobacco thylakoids using oxygen evolution, chlorophyll a (Chl a) fluorescence and redox potential measurements. When thylakoids were heated at 2 °C/min from 25 to 50 °C, the oxygen evolving complex became inhibited between 32 and 45 °C, whereas the acceptor side of PS II tolerated higher temperatures. Variable Chl a fluorescence decreased more slowly than oxygen evolution, suggesting that transitions between some S-states occurred even after heat-induced inhibition of the oxygen evolving activity. 77 K emission spectroscopy reveals that heating does not cause detachment of the light-harvesting complex II from PS II, and thus the heat-induced increase in the initial F0 fluorescence is due to loss of exciton trapping in the heated PS II centers. Redox titrations showed a heat-induced increase in the midpoint potential of the QA/QA -) couple from the control value of –80 mV to +40 mV at 50 °C, indicating a loss of the reducing power of QA -). When its driving force thus decreased, electron transfer from QA -) to QB in the PS II centers that still could reduce QA became gradually inhibited, as shown by measurements of the decay of Chl a fluorescence yield after a single turnover flash. Interestingly, the heat-induced loss of variable fluorescence and inhibition of electron transfer from QA -) to QB could be partially prevented by the presence of 5 mM bicarbonate during heating, suggesting that high temperatures cause release of the bicarbonate bound to PS II. We speculate that both the upshift in the redox potential of the QA/QA -) couple and the release of bicarbonate may be caused by a heat-induced structural change in the transmembrane D1 or D2 proteins. This structural change may, in turn, be caused by the inhibition of the oxygen evolving complex during heating.  相似文献   

14.
Summary The endogenous respiration of 14C-labelled spores of B. cereus was measured through the 14CO2 produced, and the rate expressed as Q (l CO2/hxmg). New upper limits for respiration in various conditions have been set.Dry spores had no measurable activity; Q<10–4 at room temperature and <10–3 at 35° C. For wet spores of different harvests, at 30°C, Q lay between 0.0013 to 0.067. Near 40° C, respiration showed a maximum. Thermal history has a great influence on Q. CO2 production by heat-killed spores is attributed largely to infection.Water or 10–3 m sodium phosphate buffer (pH=6.5) gave equal spore respiration, in strong NaCl it was less. Azide enhanced respiration dramatically. A temporary increase was also found with non-radioactive glucose. Exogenous respiration of spores in glucose exceeded endogenous respiration.Endogenous and exogenous respiration of vegetative forms were much larger than those of spores and were time-dependent. The ratio of minimum (endogenous, dry spores) and maximum (exogenous, wet vegetative cells) respiration was at least 3x105.  相似文献   

15.
The relationship between temperature and sensitivity to gibberellin A3 (GA3) was studied in lettuce seedlings (Lactuca sativa L. cv. Arctic). Dose/response curves for hypocotyl elongation (10-4 mol l-1 to 10-8 mol l-1) were constructed for a range of temperatures and the slope of the linear portion of the plots used as an indication of the sensitivity to GA3. Hypocotyls were unresponsive to GA3 below 13°C but above this temperature sensitivity increased linearly. Plots of growth rate against temperature had inflexions between 12°C and 13°C, with slopes above this point which increased with increasing GA3 concentration. The Q10 value for response increased in a similar manner. Reaction rates of NAD-dependent malate dehydrogenase and peroxidase extracted from hypocotyls varied linearly with temperature whilst nonspecific tetrazolium reduction, a membrane based activity, showed an abrupt rate change above 14°C. Pre-exposure to GA3 had no effect on the temperature responses of soluble or particulate enzymes.Abbreviation GA3 gibberellin A3  相似文献   

16.
J M Seddon 《Biochemistry》1990,29(34):7997-8002
Fully hydrated unsaturated diacylglycerol-phosphatidylcholine mixtures are found to adopt an inverse face-centered cubic phase, of crystallographic cubic aspect 15. The same behavior is observed for either the 1,2- or 1,3-isomer of the diacylglycerol. This Q15 cubic phase, of probable space group Fd3m (Q227), occurs between an inverse hexagonal (HII) phase and an inverse micellar (L2) solution, with increasing diacylglycerol concentration, which implies that the mean curvature of the interface is more negative than that of the HII phase. This behavior is quite different from that of the more usual bicontinuous inverse cubic phases Pn3m (Q224), Im3m (Q229), and Ia3d (Q230), which normally occur between the lamellar L alpha and the HII phases. One possible structure for the Fd3m cubic phase has recently been proposed (Mariani, P., Luzzati, V., & Delacroix, H. (1988) J. Mol. Biol. 204, 165-189), consisting of tetrahedrally arranged clusters of inverse micelles surrounded by a continuous cage of tetrahedrally connected water/lipid (inverse) channels.  相似文献   

17.
The temperature-composition phase diagram of monopentadecenoin, a monoacylglycerol with a cis monounsaturated fatty acid 15 carbon atoms long (C15:1c10) in water was constructed using x-ray diffraction. Low- and wide-angle diffraction patterns were collected from samples of fixed hydration as a function of temperature in the heating direction on x-ray-sensitive film. The temperature and hydration ranges investigated were 0-104 degrees C and 0-60% (w/w) water, respectively. The phases identified in the system include the lamellar crystalline phase, the lamellar liquid crystalline phase, the fluid isotropic phase, and two inverted cubic phases belonging to space groups la3d (Q230) and Pn3m (Q244). Particular attention has been devoted to the issues of phase equilibrium, phase boundary verification, and structure characterization. The phase diagrams of monopentadecenoin, monomyristolein (C14:1c9), and monoolein (C18:1c9) are compared, and the impact of molecular structure on mesophase stability and structure is discussed.  相似文献   

18.
Aqueous dispersions (pH 4.0) of a 2:1 (mol/mol) mixture of myristic acid with dimyristoylphosphatidylcholine undergo a sharp transition at 45-47 degrees C from a lamellar gel phase to a fluid phase which is optically isotropic. This fluid phase gives rise to 31P-NMR spectra, and 2H-NMR spectra of the chain-deuterated components, which are also isotropic. X-ray diffraction studies of the fluid phase at 49 degrees C, reveal reflections with spacings in the ratio square root of 2: (square root of 3): square root of 4: square root of 6: square root of 8, accompanied by a strong diffuse scatter. These reflections index on a cubic lattice of primitive space group Pn3 or Pn3m, or possibly the body-centered group Im3m, with a lattice constant of 21.2 nm. The dimensions of the phase are consistent with a structure composed of two systems of tetrahedrally (octahedrally) oriented inverted lipid cylinders, found for other cubic lipid phases with Pn3m (Im3m) symmetry. At higher temperatures the cubic phase gradually converts, with increasing temperature, to a coexisting inverted hexagonal phase.  相似文献   

19.
The hydrophobic surfactant proteins SP-B and SP-C promote rapid adsorption of pulmonary surfactant to an air/water interface. Previous evidence suggests that they achieve this effect by facilitating the formation of a rate-limiting negatively curved stalk between the vesicular bilayer and the interface. To determine whether the proteins can alter the curvature of lipid leaflets, we used x-ray diffraction to investigate how the physiological mixture of these proteins affects structures formed by 1-palmitoyl-2-oleoyl phosphatidylethanolamine, which by itself undergoes the lamellar-to-inverse hexagonal phase transition at 71°C. In amounts as low as 0.03% (w:w) and at temperatures as low as 57°C, the proteins induce formation of bicontinuous inverse cubic phases. The proteins produce a dose-related shift of diffracted intensity to the cubic phases, with minimal evidence of other structures above 0.1% and 62°C, but no change in the lattice-constants of the lamellar or cubic phases. The induction of the bicontinuous cubic phases, in which the individual lipid leaflets have the same saddle-shaped curvature as the hypothetical stalk-intermediate, supports the proposed model of how the surfactant proteins promote adsorption.  相似文献   

20.
The effect of alpha-tocopherol on the structure and thermotropic phase behaviour of distearoylphosphatidylethanolamine was examined by using synchrotron X-ray diffraction methods. There was evidence that alpha-tocopherol does not distribute randomly in the dispersed phospholipid but instead phospholipid phases enriched in alpha-tocopherol are formed. Heating codispersions from lamellar gel phase induced formation of hexagonal-II phase at temperatures below the main transition of the pure phospholipid and which were enriched in alpha-tocopherol. Codispersions containing 5 or 10 mol% alpha-tocopherol were induced to form a cubic phase at temperatures above the lamellar to hexagonal-II phase transition. Such phases were not observed in codispersions containing 2.5 or 20 mol% alpha-tocopherol in which only lamellar and hexagonal-II phases were formed. The space group of the cubic phases were tentatively assigned as Pn3m. Equilibration of codispersions at 4 degrees C results in the formation of lamellar crystalline phases enriched in alpha-tocopherol and phase separated domains of pure phospholipid. Two lamellar crystalline phases were characterized on the basis of their particular wide-angle X-ray scattering patterns. The lamellar crystalline phases were also distinguished from other lamellar phases of the pure phospholipid by the lamellar repeat. Partitioning of alpha-tocopherol into phosphatidylethanolamine domains in membranes may introduce instability into the structure.  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