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 62 毫秒
1.
The interactions of mouse thymocytes with unilamellar phospholipid vesicles comprised of dimyristyl lecithin (DML), dipalmitoyl lecithin (DPL), dioleoyl lecithin (DOL), and egg yolk lecithin (EYL) were examined in vitro.

In cells treated with [3H]DML or [3H]DPL vesicles, electron microscope (EM) autoradiographic analysis showed most of the radioactive lipids to be confined to the cell surface. Transmission EM studies showed the presence of intact vesicles (DPL) and collapsed or ruptured vesicle fragments (DML) adsorbed to the surfaces of treated cells. In cells treated with DPL vesicles containing a watersoluble dye (6-carboxyfluorescein; 6-CF), most of the fluorescent vesicles were localized at the periphery of the treated cells. Furthermore, substantial fractions of the cell-associated DPL and DML could be released by a mild trypsinization without damaging the cells. These results suggest that the uptake of DML and DPL is primarily due to vesicle-cell adsorption. Such an adsorption process appears to be enhanced at or below the thermotropic-phase transition temperature of the vesicle lipid. Under certain conditions these adherent vesicles also formed patches or caps on the cell surface.

In cells treated with DOL or EYL vesicles, transmission EM and EM autoradiography showed relatively little exogenous vesicle lipid located at the cell surface. Thymocytes incubated (37°C) with [14C] EYL vesicles containing a trapped marker, [3H]inulin, incor porated both isotopes at identical rates. In separate experiments it was found that this marker was located inside the treated cells. Thymocytes treated with DOL vesicles containing 6-CF exhibited a uniform and diffuse distribution of dye in the internal volume of the cells. Little cell-associated EYL or DOL could be released by trypsinization. Evidence against endocytosis of intact vesicles as a major pathway of vesicle uptake is also presented. These observations, coupled with the demonstration of vesicle-cell lipid exchange as a minor component of vesicle uptake suggest that incorporation of EYL and DOL vesicles by thymocytes is primarily by vesicle-cell fusion.  相似文献   

2.
Differential scanning calorimetry and x-ray diffraction techniques have been used to investigate the structure and phase behavior of hydrated dimyristoyl lecithin (DML) in the hydration range 7.5 to 60 weight % water and the temperature range -10 to +60 degrees C. Four different calorimetric transitions have been observed: T1, a low enthalpy transition (deltaH approximately equal to 1 kcal/mol of DML) at 0 degrees C between lamellar phases (L leads to Lbeta); T2, the low enthalpy "pretransition" at water contents greater than 20 weight % corresponding to the transition Lbeta leads to Pbeta; T3, the hydrocarbon chain order-disorder transition (deltaH = 6 to 7 kcal/mol of DML) representing the transition of the more ordered low temperature phases (Lbeta, Pbeta, or crystal C, depending on the water content) to the lamellar Lalpha phase; T4, a transition occurring at 25--27 degrees C at low water contents representing the transition from the lamellar Lbeta phase to a hydrated crystalline phase C. The structures of the Lbeta, Pbeta, C, and Lalpha phases have been examined as a function of temperature and water content. The Lbeta structure has a lamellar bilayer organization with the hydrocarbon chains fully extended and tilted with respect to the normal to the bilayer plane, but packed in a distorted quasihexagonal lattice. The Pbeta structure consists of lipid bilayer lamellae distorted by a periodic "ripple" in the plane of the lamellae; the hydrocarbon chains are tilted but appear to be packed in a regular hexagonal lattice. The diffraction pattern from the crystalline phase C indexes according to an orthorhombic cell with a = 53.8 A, b = 9.33 A, c = 8.82 A. In the lamellae bilayer Lalpha strucure, the hydrocarbon chains adopt a liquid-like conformation. Analysis of the hydration characteristics and bilayer parameters (lipid thickness, surface area/molecule) of synthetic lecithins permits an evaluation of the generalized hydration and structural behavior of this class of lipids.  相似文献   

3.
Infrared spectra were obtained as a function of temperature for a variety of phospholipid/water bilayer assemblies (80% water by weight) in the 3000-950 cm?1 region. Spectral band-maximum frequency parameters were defined for the 2900 cm?1 hydrocarbon chain methylene symmetric and asymmetric stretching vibrations. Temperature shifts for these band-maximum frequencies provided convenient probes for monitoring the phase transition behavior of both multilamellar liposomes and small diameter single-shell vesiclesof dipalmitoyl phosphatidylcholine/water dispersions. As examples of the effects of bilayer lipid/cholesterol/water (3 : 1 mol ratio) and lipid/cholesterol/amphotericin B/water (3 : 1 : 0.1 mol ratios) vesicles were examined using the methylene stretching frequency indices. In comparison to the pure vesicle form, the transition width of the lipid/cholesterol system increased by nearly a factor of two (to 8°C) while the phase transition temperature remained approximately the same (41° C). For the lipid/cholesterol/amphotericin B system, the phase transition temperature increased by about 4.5° C (to 45.5°C) with the transition width increasing by nearly a factor of four (to ≈ 15°C) above that of the pure vesicles. The lipid/cholesterol/amphotericin B data were interpreted as reflecting the formation below 38°C of a cholesterol/amphotericin B complex whose dissociation at higher temperature (38–60°C range) significantly broades the gel-liquid crystalline phase transition.  相似文献   

4.
Phospholipid hydration studied by deuteron magnetic resonace spectroscopy   总被引:5,自引:0,他引:5  
Deuteron magnetic resonance spectra were obtained from 2H2O in mixtures with egg lecithin, egg phosphatidylethanolamine, and ox brain sodium phosphatidylserine. The acid form of phosphatidylserine does not hydrate. Details of the different hydration “shells” were obtained by studying the spectral splittings as a function of 2H2O concentration. Several different types of water are present, including bulk water (exchanging only slowly with water associated with the lipid), “trapped” water (not present with phosphatidylethanolamine), and up to three types of bound water. The spectral splittings characteristic of each water environment yielded information about the water binding energies and degrees of anisotropy of motion of the phospholipid polar groups; lecithin polar groups have least motional restriction and sodium phosphatidylserine most, while phosphatidylethanolamine binds water most tightly.Spectra of some lecithin and phosphatidylserine dispersions varied with time, due to a slow reorganization of randomly oriented multilamellar regions into longer, more ordered systems, with a length of about 1 μm. At ?20°C the timescales of the change were of the order of a week and a month for lecithin and phosphatidylserine respectively.Complex changes in the spectra were observed as the temperature was raised; these are interpreted in terms of changes in the motions of the phospholipid molecules.  相似文献   

5.
Abstract

Liposomes made of dipalmitoylphosphatidylcholine (DPPC2), dipalmitoyl-phosphatidylglycerol (DPPG), and different long-chain fatty alcohols were investigated with respect to their colloidal stability, chain-melting phase transition temperature, and temperature dependent inter-vesicle fusion. In particular, the practical usefulness of the stoichiometric 1/2 (mol/mol) mixtures of the phospholipids and fatty alcohols, mainly elaidoyl alcohol (EL-OH) were studied. The mole fraction of DPPG in the bilayers of such vesicles affects crucially the colloidal stability of the resulting lipid suspensions; at least 15 mol-% of DPPG (relative to DPPC) must be incorporated into the bilayers in order to make the liposome suspension colloidally sufficiently stable at room temperature. The corresponding DPPC/DPPG/EL-OH (0.85/0.15/2) mixed lipid vesicles undergo a lamellar-gel to inverted hexagonal (HIT) phase transition at 52.7°C, however, and then fuse and aggregate massively. The related phase transition temperature of the DPPC/DPPG/palmitelaidoyl alcohol (0.85/0.15/2) mixture is 48.4°C. This indicates that the chain-melting phase transition temperature of the investigated lipid mixtures is rather sensitive to the alcohol chain-length. This transition temperature is independent, however, of the bulk proton concentration in the pH region between 4.9 and 7.2. Stoichiometric 1/2 mixtures of phospholipids and EL-OH have a high propensity for the inter-vesicle fusion at 42°C and neutral pH. The reason for such fusion 10°C below the lamellar-to-nonlamellar phase transition temperature are the defects that are generated during the chain-melting of the (partly segregated) phospholipid component at 42°C; the proximity of the lamellar to non-lamellar phase transition temperature of the phospholipid/fatty alcohol (1/2) complex at 52°C also plays an important role.  相似文献   

6.
Short-chain lecithin/long-chain phospholipid unilamellar vesicles (SLUVs), unlike pure long-chain lecithin vesicles, are excellent substrates for water-soluble phospholipases. Hemolysis assays show that greater than 99.5% of the short-chain lecithin is partitioned in the bilayer. In these binary component vesicles, the short-chain species is the preferred substrate, while the long-chain phospholipid can be treated as an inhibitor (phospholipase C) or poor substrate (phospholipase A2). For phospholipase C Bacillus cereus, apparent Km and Vmax values show that bilayer-solubilized diheptanoylphosphatidylcholine (diheptanoyl-PC) is nearly as good a substrate as pure micellar diheptanoyl-PC, although the extent of short-chain lecithin hydrolysis depends on the phase state of the long-chain lipid. For phospholipase A2 Naja naja naja, both Km and Vmax values show a greater range: in a gel-state matrix, diheptanoyl-PC is hydrolyzed with micellelike kinetic parameters; in a liquid-crystalline matrix, the short-chain lecithin becomes comparable to the long-chain component. Both enzymes also show an anomalous increase in specific activity toward diheptanoyl-PC around the phase transition temperature of the long-chain phospholipid. Since the short-chain lecithin does not exhibit a phase transition, this must reflect fluctuations in head-group area or vertical motions of the short-chain lecithin caused by surrounding long-chain lecithin molecules. These results are discussed in terms of a specific model for SLUV hydrolysis and a general explanation for the "interfacial activation" observed with water-soluble phospholipases.  相似文献   

7.
Fluorescent lecithin probes containing cis- or trans-parinaric acid (PnA) at the 2-position cis-parinaroylphosphatidylcholine (cis-PnPC) and trans-parinaroyl phosphatidylcholine (trans-PnPC)) showed similar behavior to that of the free cis- or trans-parinaric acids (cis-PnA or trans-PnA) in bilayer vesicles of synthetic saturated lecithins. Transition temperatures detected by cis-PnPc were about 1°C lower than those observed with trans-PnPc. In mixed lecithin vesicles, the trans-PnPc probe monitored a higher temperature melting component than did the cis-probe. Both probes were readily incorporated into microsomal membranes and into sonicated vesicles prepared from the microsomal phospholipids. With either cis- or trans-PnPc no change in polarization ratio was observed for microsomal membranes between 40°C and 0°C but this ratio increased with decreasing temperature between 0°C and ?5°C. However, vesicles of extracted phospholipids showed a continuous increase in polarization ratio with decreasing temperature between 20°C and ?15°C with trans-PnPc and bewteen 5°C and ?15°C with cis-PnPc. These results suggest that the two lecithin probes monitor different environments in the membranes and phospholipid vesicles prepared from them.  相似文献   

8.
A microscopic study has allowed the analysis of modifications of various shapes acquired by phospholipid vesicles during a hydrostatic pressure treatment of up to 300 MPa. Giant vesicles of dimyristoylphosphatidylcholine / phosphatidylserine (DMPC/PS) prepared at 40°C mainly presented a shape change resembling budding during pressure release. This comportment was reinforced by the incorporation of 1,2-dioleyl-sn-glycero-3-phosphatidylethanolamine (DOPE) or by higher temperature (60°C) processing. The thermotropic main phase transition (Lα to Pβ′) of the different vesicles prepared was determined under pressure through a spectrofluorimetric study of 6-dodecanoyl-2-dimethylamino-naphtalene (Laurdan) incorporated into the vesicles’ bilayer. This analysis was performed by microfluorescence observation of single vesicles. The phase transition was found to begin at about 80 MPa and 120 MPa for DMPC/PS vesicles at, respectively, 40°C and 60°C. At 60°C the liquid-to-gel transition phase was not complete within 250 MPa. Addition of DMPE at 40°C does not significantly shift the onset boundary of the phase transition but extends the transition region. At 40°C, the gel phase was obtained at, respectively, 110 MPa and 160 MPa for DMPC/PS and DMPC/PS/DOPE vesicles. In comparing volume data obtained from image analysis and Laurdan signal, we assume the shape change is a consequence of the difference between lateral compressibility of the membrane and bulk water. The phase transition contributes to the membrane compression but seems not necessary to induce shape change of vesicles. The high compressibility of the Lα phase at 60°C allows induction on DMPC/PS vesicles of a morphological transition without phase change.  相似文献   

9.
The adhesion of artificially generated lipid membrane vesicles to Chinese hamster V79 fibroblasts in suspension was used as a model system for studying membrane interactions. Below their gel-liquid crystalline phase transition temperature, vesicles comprised of dipalmitoyl lecithin (DPL) or dimyristoyl lecithin (DML) absorbed to the surfaces of EDTA- dissociated cells. These adherent vesicles could not be removed by repeated washings of the treated cells but could be released into the medium by treatment with trypsin. EM autoradiographic studies of cells treated with[(3)H]DML or [(3)H]DPL vesicles showed that most of the radioactive lipids were confined to the cell periphery. Scanning electron microscopy and fluorescence microscopy further confirmed the presence of adherent vesicles at the cell surface. Adhesion of DML or DPL vesicles to EDTA-dissociated cells modified the lactoperoxidase-catalyzed iodination pattern of the cell surface proteins; the inhibition of labeling of two proteins with an approximately 60,000- dalton mol wt was particularly evident. Incubation of cells wit h (3)H-lipid vesicles followed by sodium dodecyl sulfate (SDS)- polyacrylamide gel electrophoresis showed that some of the (3)H-lipid migrated preferentially with these approximately 60,000-mol wt proteins. Studies of the temperature dependence of vesicle uptake and subsequent release by trypsin showed that DML or DPL vesicle adhesion to EDTA- dissociated cells increased with decreasing temperatures. In contrast, cells trypsinized before incubation with vesicles showed practically no temperature dependence of vesicle uptake. These results suggest two pathways for adhesion of lipid vesicles to the cell surface-a temperature-sensitive one involving cell surface proteins, and a temperature-independent one. These findings are discussed in terms of current models for cell-cell interactions.  相似文献   

10.
M J Tunis  J E Hearst 《Biopolymers》1968,6(9):1325-1344
The hydration of DNA is an important factor in the stability of its secondary structure. Methods for measuring the hydration of DNA in solution and the results of various techniques are compared and discussed critically. The buoyant density of native and denatured T-7 bacteriophage DNA in potassium trifluoroacetate (KTFA) solution has been measured as a function of temperature between 5 and 50°C. The buoyant density of native DNA increased linearly with temperature, with a dependence of (2.3 ± 0.5) × 10?4 g/cc-°C. DNA which has been heat denatured and quenched at 0°C in the salt solution shows a similar dependence of buoyant density on temperature at temperatures far below the Tm, and above the Tm. However, there is an inflection region in the buoyant density versus T curve over a wide range of temperatures below the Tm. Optical density versus temperature studies showed that this is due to the. inhibition by KTFA of recovery of secondary structure on quenching. If the partial specific volume is assumed to be the same for native and denatured DNA, the loss of water of hydration on denaturation is calculated to be about 20% in KTFA at a water activity of 0.7 at 25°C. By treating the denaturation of DNA as a phase transition, an equation has immmi derived relating the destabilizing effect of trifluoroacetate to the loss of hydration on denaturation. The hydration of native DNA is abnormally high in the presence of this anion, and the loss of hydration on denaturation is greater than in CsCl. In addition, trifluoroacetate appears to decrease the ΔHof denaturation.  相似文献   

11.
The fluorescence intensity of trans-parinaric acid as a function of the temperature indicates a phase transition in bovine heart mitochondrial inner membranes below 0°C. The comparison of the dye fluorescence intensity in intact inner mitochondrial membranes and in vesicles from extracted phospholipids of mitochondria revealed a similar intensity increase with decreasing temperature. A synthetic phospholipid system of dioleoyl phosphatidylcholine was investigated because of its low phase transition temperature and showed a very definite intensity change at ?25°C. trans-Parinaric acid in membrane systems probes an environment of intermediate polarity; this was found from the excitation and emission spectra and from fluorescence decay.  相似文献   

12.
From measurements of the equilibrium spreading pressure pie for dispersions of lecithin--dimyristoyl (DML) or dioleoyl (DOL)--and cholesterol (CHOL) in water, we have deduced the phase relations in both the aqueous dispersions and the equilibrium surface films. At 29.5 degrees C, when the mole fraction of cholesterol in the dispersion chi(CHOL) is 0 chi(CHOL) less than chi(CHOL) less than 0.33, pie is constant and equal to the value for pure lecithin (DOL or DML). The phase rule predicts than two bulk lipid phases coexist; these are pure lecithin and lecithin:cholesterol 2:1 complex. The equilibrium surface film contants only lecithin and therefore lecithin and 2:1 complex are immiscible in surface films. When 0.33 less than chi/CHOL) less than 1.0, pie is also contant with a value intermediate between that for pure lecithin and cholesterol. In this range of lipid composition two bulk lipid phases also coexist: lecithin:cholesterol 2:1 complex and pure cholesterol. However, the equilibrium surface film contains only the 2:1 complex and, therefore, 2:1 complex is also immiscible with cholesterol in surface films. When pi less than pie, as in the case of spread films, we deduce that two surface phases may coexist; the composition of the phases will depend on chi(CHOL). When 0 less than chi(CHOL) less than 0.33, both lecithin and 2:1 complex coexist, and when 0.33 less than chi(CHOL) less than 1.0, 2:1 complex and cholesterol coexist. The "condensing" effect of cholesterol in lecithin surface films is reexamined. The effect is attributed to formation of the lecithin:cholesterol 2:1 complex and nonequilibrium conditions in the two-phase surface film.  相似文献   

13.
The effect of unilamellar lipid vesicles composed of dioleoyl lecithin (DOL), egg yolk lecithin (EYL), 1:1 EYL:cholesterol (Chol), dipalmitoyl lecithin (DPL), and dimyristoyl lecithin (DML) on the mitogenic response in mouse lymphocytes was tested. Cortisone-resistant thymocytes were briefly treated with lipid vesicles and subsequently stimulated with concanavalin A (con A). All of the lipid vesicles induced an enhanced mitogenic response on day 3 as tested by [3H]TdR incorporation and by counting total cells. The order of enchanced [3H]TdR incorporation (less than or equal to 5.3 times the control) was DML greater than DPL greater than 1:1 EYL:Chol greater than EYL congruent to DOL greater than untreated control cells. These increases were paralleled by increased numbers of total cells. The response of spleen cells to a B-cell mitogen, bacterial lipopolysaccharide, was similarly enhanced by vesicle pretreatments in the same order. Vesicle treatments alone were not mitogenic Pretreatment of cells with lipid vesicles modified lectin binding: DML and DPL increased the binding of [125I]con A by three to four times the control, whereas 1:1 EYL:Chol, EYL, or DOL had little or no effect. The binding of [125I]phytohemagglutinin-P (PHA-P) to vesicle-treated cells was indistinguishable from untreated cells. The lectin (con A; PHA-P)-induced agglutination of vesicle-treated cells was also modified by different lipid vesicles in the same order as the mitogenic response. Based on the results presented in the accompanying report [6], we find that the cell surface adsorption properties of the applied lipid vesicles correlate with their ability to enhance the mitogenic response, and that they modify agglutinability and lectin binding. These results are further discussed in terms of the possible alteration of membrane properties and subsequent cellular activity.  相似文献   

14.
Dimyristoyl phosphatidylcholine rapidly exchanges between vesicles at 37°C without vesicle fusion.The rate of the transbilayer movement of dimyristoyl phosphatidylcholine in sonicated vesicles has been measured employing 13C NMR using N-13CH3? labeled lipids which are introduced into the outer monolayer of non-labeled vesicles by a phosphatidylcholine exchange protein. The rate of transbilayer movement of dimyristoyl phosphatidylcholine shows a distinct maximum (halftime 4 h) in the temperature range at which the hydrocarbon phase transition occurs.The activation energy of the flip-flop rate above the phase transition is 23.7 ± 2.0 kcal/mol.  相似文献   

15.
The effect of unilamellar lipid vesicles composed of dioleoyl lecithin (DOL), egg yolk lecithin (EYL), 1:1 EYL:cholesterol (Chol), dipalmitoyl lecithin (DPL), and dimyristoyl lecithin (DML) on the mitogenic response in mouse lymphocytes was tested. Cortisone-resistant thymocytes were briefly treated with lipid vesicles and subsequently stimulated with concanavalin A (con A). All of the lipid vesicles induced an enhanced mitogenic response on day 3 as tested by [3H]TdR incorporation and by counting total cells. The order of enhanced [3H]TdR incorporation (?5.3 times the control) was DML>DPL>1:1 EYL:Chol>EYL?DOL> untreated control cells. These increases were paralleled by increased numbers of total cells. The response of spleen cells to a B-cell mitogen, bacterial lipopolysaccharide, was similarly enhanced by vesicle pretreatments in the same order. Vesicle treatments alone were not mitogenic.

Pretreatment of cells with lipid vesicles modified lectin binding: DML and DPL increased the binding of [125I]con A by three to four times the control, whereas 1:1 EYL:Chol, EYL, or DOL had little or no effect. The binding of [125I]phytohemagglutinin-P (PHA-P) to vesicle-treated cells was indistinguishable from untreated cells. The lectin (con A; PHA-P)-induced agglutination of vesicle-treated cells was also modified by different lipid vesicles in the same order as the mitogenic response.

Based on the results presented in the accompanying report [6], we find that the cell surface adsorption properties of the applied lipid vesicles correlate with their ability to enhance the mitogenic response, and that they modify agglutinability and lectin binding. These results are further discussed in terms of the possible alteration of membrane properties and subsequent cellular activity.  相似文献   

16.
Raman spectroscopic frequency differences between selected carbon-carbon stretching modes of lipid hydrocarbon chains were determined as a function of temperature for use in monitoring lipid phase transition behavior and acyl chain disorder in both multilamellar and single-wall vesicles. Transition temperatues detected by this procedure for pure dipalmitoyl phosphatidylcholine and dimyristoyl phosphatidylcholine multilayers were observed at 39±1 °C and 23±1 °C, respectively. Although the phase transition for unilamellar vesicles of dipalmitoyl phosphatidylcholine occurred at nearly the same temperature as the multilayers, the crystal-liquid crystalline transition for the single-shell vesicles appeared to span a slightly broader temperature range, a characteristic consistent with irregularities in the packing arrangement of the hydrocarbon chains. Within the precision of the Raman spectroscopic method, however, the temperature behavior of both the multilamellar and the unilamellar dimyristoyl phosphatidylcholine assemblies appeared nearly identical. The temperature profile for the Raman frequency differences of an excess water sonicate of 25 mol percent cholesterol in dipalmitoyl phosphatidylcholine served as an example of the effect upon lipid phase transition characteristics of a bilayer component intercalated between the acyl chains. For this particular cholesterol-lipid system the phase transition was broadened over a 30 °C temperature range, in contrast to the narrow 5?4 °C range observed for pure multilayer and single-shell vesicle particles.  相似文献   

17.
The purpose of this research was to improve the stability of carbamazepine (CBZ) bulk powder under high humidity by surface modification. The surface-modified anhydrates of CBZ were obtained in a specially designed surface modification apparatus at 60°C via the adsorption of n-butanol, and powder x-ray diffraction, Fourier-Transformed Infrared spectra, and differential scanning calorimetry were used to determine the crystalline characteristics of the samples. The hydration process of intact and surface-modified CBZ anhydrate at 97% relative humidity (RH) and 40±1°C was automatically monitored by using isothermal microcalorimetry (IMC). The dissolution test for surface-modified samples (20 mg) was performed in 900 mL of distilled water at 37±0.5°C with stirring by a paddle at 100 rpm as in the Japanese Pharmacopoeia XIII. The heat flow profiles of hydration of intact and surface-modified CBZ anhydrates at 97% RH by using IMC profiles showed a maximum peak at around 10 hours and 45 hours after 0 and 10 hours of induction, respectively. The result indicated that hydration of CBZ anhydrate was completely inhibited at the initial stage by surface modification of n-butanol and thereafter transformed into dihydrate. The hydration of surface-modified samples followed a 2-dimensional phase boundary process with an induction period (IP). The IP of intact and surface-modified samples decreased with increase of the reaction temperature, and the hydration rate constant (k) increased with increase of the temperature. The crystal growth rate constants of nuclei of the intact sample were significantly larger than the surface-modified samples at each temperature. The activation energy (E) of nuclei formation and crystal growth process for hydration of surface-modified CBZ anhydrate were evaluated to be 20.1 and 32.5 kJ/mol, respectively, from Arrhenius plots, but the Es of intact anhydrate were 56.3 and 26.8 kJ/mol, respectively. The dissolution profiles showed that the surface-modified sample dissolved faster than the intact sample at the initial stage. The dissolution kinetics were analyzed based on the Hixon-Crowell equation, and the dissolution rate constants for intact and surface-modified anhydrates were found to be 0.0102±0.008 mg1/3 min−1 and 0.1442±0.0482 mg1/3·min−1. The surface-modified anhydrate powders were more stable than the nonmodified samples under high humidity and showed resistance against moisture. However, surface modification induced rapid dissolution in water compared to the control.  相似文献   

18.
Water which remains unfrozen at ?25 °C in the presence of phosphatidyl choline (PC) gives rise to a proton magnetic resonance signal which can be used to measure the hydration of single-walled vesicles and multilamellar liposomes of PC. The proton magnetic resonance signal of the unfrozen water in these systems is strongly dependent upon the nature of the molecular domain in which the water is situated. For example, at cholesterol to PC molar ratios below 35 mol%, the vesicle hydration signal consists of a relatively narrow symmetric peak (line width, ~150 Hz). At higher molar ratios, however, rather broad asymmetric signals appear (line widths, ~300–1000 Hz) which indicate that when significant quantities of cholesterol are packed in the bilayer there must be regions in which there is a preferred direction for motion of the unfrozen water. It is possible to solubilize significant quantities of cholesterol by sonicating it in concentrated solutions of sodium dodecyl sulfate. Addition of cholesterol to PC vesicles via these sodium dodecyl sulfate-cholesterol complexes caused hydration changes in the PC, which, at high cholesterol to PC molar ratios, paralleled the effects of cholesterol on PC hydration in homogeneous vesicles in which the cholesterol and PC were simply cosonicated.  相似文献   

19.
The regulation of lecithin:cholesterol acyltransferase by changes in phospholipid bilayer fluidity was investigated using pyrene excimer fluorescence to measure fluidity. Fluidity of dimyristoylphosphatidylcholine (DMPC) unilamellar vesicles was decreased by the addition of up to 20% (mol/mol) cholesterol and increased by the addition of up to 10% (mol/mol) lysoDMPC. When both cholesterol and lysoDMPC are present in the bilayer, their individual effects on fluidity are altered. These changes can be explained by complex formation between cholesterol and phospholipid as in the model of Presti et al. (Presti, F.C., Pace, R.J. and Chan, S.I. (1982) Biochemistry 21, 3831-3335). Lecithin:cholesterol acyltransferase activity with these vesicles as substrates was measured to determine whether activity can be modulated by the fluidity changes of the bilayer on which the enzyme acts. When 10% lysoDMPC, a known lecithin:cholesterol acyltransferase inhibitor, is added to the vesicles, inhibition of activity is observed. When 7.5% lysoDMPC is added to vesicles which contain either 5 or 10% cholesterol, lecithin:cholesterol acyltransferase activity increases. This increase in lecithin:cholesterol acyltransferase activity due to vesicle-fluidity increase is sufficient to overcome the decrease in activity due to lecithin:cholesterol acyltransferase inhibition. This is the first report of the ability of lysoDMPC to increase lecithin:cholesterol acyltransferase activity.  相似文献   

20.
Ricinus communis agglutinin, a galactose-binding lectin, agglutinates phospholipid/qlycolipid vesicles, phospholipid/glycolipid/cholesterol vesicles, low density lipoprotein, and the branched polysaccharide gum arabic. The temperature dependence of the initial velocity of agglutination is similar in all cases, with a maximum at ~25°C, and a decreased rate at lower and higher temperatures. The extent of agglutination is temperature-independent for T ≤ 25°C, and decreases with increasing temperature for T > 25°C. After agglutination at 25°C, an increase in temperature results in deagglutination. This effect is reversible upon return to 25°C. Thus RCA undergoes a reversible temperature-dependent transition between an agglutinating form (≤25°C) and a less active form (> 25°C). These observations have important consequences in the interpretation of experiments which attempt to correlate lectin-binding with temperature-dependent properties such as membrane fluidity or receptor lateral organization in natural and model membranes.  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