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.
The effect of the local anesthetic dibucaine on the solid to liquid-crystalline phase transition in phospholipid vesicles was studied by calorimetry and fluorescence polarization. The partition coefficient (greater than 3000) of dibucaine in the membranes of vesicles prepared from acidic phospholipids was more than 20 times higher than in neutral phospholipid membranes under the same conditions. Calorimetric measurements on vesicles prepared form acidic phospholipids (bovine brain phosphatidylserine; dipalmitoylphosphatidylglycerol) showed that dibucaine (1 with 10(-4) M) produced a significant reduction in the gel-liquid crystalline transition temperature (Tc). This fluidizing effect of dibucaine on acidic phospholipid membranes was even more marked in the presence of Ca2+. In contrast, dibucaine at the same concentration did not alter the Tc of neutral phospholipids (dipalmitoylphosphatidylcholine). Significant increase in the fluidity of neutral phospholipid membranes occurred only at higher dibucaine concentrations (2 with 10(-3) M). Measurements of the fluorescence polarization and lifetime of the probe, 1,6-diphenylhexatriene, in acidic phospholipid vesicles revealed that dibucaine (1 with 10(-4) M) caused an increase in the probe rotation rate indicating an increase in the fluidity of the phospholipid membranes. A good correlation was obtained between fluorescence polarization data on dibucaine-induced changes in membrane fluidity and calorimetric measurements on vesicles of the same type.  相似文献   

2.
Earlier studies by our laboratory have suggested a relationship between an amiloride-sensitive Na+−H+ exchange process and the physical state of the lipids of rat colonic brush-border membrane vesicles. To further assess this possible relationship, a series of experiments were performed to examine the effect of dexamethasone administration (100 μg/100 g body wt. per day) subcutaneously for 4 days on Na+−H+ exchange, lipid composition and lipid fluidity of rat distal colonic brush-border membrane vesicles. The results of these studies demonstrate that dexamethasone treatment significantly: (1) increased the Vmax of the Na+−H+ exchange without altering the Km for sodium of this exchange process, utilizing the fluorescent pH-sensitive dye, acridine orange. 22Na flux experiments also demonstrated an increase in amiloride-sensitive proton-stimulated sodium influx across dexamethasone-treated brush-border membrane vesicles; (2) increased the lipid fluidity of treated-membrane vesicles compared to their control counterparts, as assessed by steady-state fluorescence polarization techniques using three different lipid-soluble fluorophores; and (3) increased the phospholipid content of treated-membrane vesicles thereby, decreasing the cholesterol/phospholipid molar ratio of treated compared to control preparations. This data, therefore, demonstrates that dexamethasone administration can modulate amiloride-sensitive Na+−H+ exchange in rat colonic distal brush-border membrane vesicles. Moreover, it adds support to the contention that a direct relationship exists between Na+−H+ exchange activity and the physical state of the lipids of rat colonic apical plasma membranes.  相似文献   

3.
Dispersed acini from dog pancreas were used to examine the ability of dopamine to increase cyclic AMP cellular content and the binding of [3H]dopamine. Cyclic AMP accumulation caused by dopamine was detected at 1·10−8 M and was half-maximal at 7.9±3.4·10−7M. The increase at 1·10−5 M, (7.5-fold) was equal to the half-maximal increase caused by secretin at 1·10−9 M. Haloperidol, a dopaminergic receptor antagonist inhibited cyclic AMP accumulation caused by dopamine. The IC50 value for haloperidol, calculated from the inhibition of cyclic AMP increase caused by 1·10−5 M dopamine was 2.3±0.9·10−6M. Haloperidol did not alter basal or secretin-stimulated cyclic AMP content. [3H]Dopamine binding was studied on the same batch of cells as cyclic AMP accumulation. At 37°C, it was rapid, reversible, saturable and stereospecific. The Kd value for high affinity binding sites was 0.43±0.1·10−7M and 4.7±1.6·10−7M for low affinity binding sites. The concentration of drugs necessary to inhibit specific binding of dopamine by 50% was 1.2±0.4·10/t-7M noradrenaline, 2·10/t-7 M epinine, 4.1±1.8·10/t-6M fluphenazine, 8.0±1.6·10/t-6M haloperidol, 4.2±1.2·10−6Mcis-flupenthixol, 2.7±0.4·10−5Mtrans-flupenthixol, >1·10−5M apomorphine, sulpiride, naloxone and isoproterenol.  相似文献   

4.
Uptake of testosterone was demonstrated in membrane vesicles prepared from Pseudomonas testosteroni grown on testosterone. In contrast, membrane vesicles from uninduced cultures revealed no significant transport activity for steroids. The Km of the reaction was 2 · 10−6M and the V 28.5 nmoles/min per mg protein. Steroid uptake was maximal within the pH range of 8 to 9 and at incubation temperatures between 30 and 37 °C. Transport of steroid was dependent upon NAD+ and was reduced by NADH, dinitrophenol, and inhibitors of electron transport, such as N3 · CN and amytal. The intravesicular steroid concentration was approx. 800 times the steroid concentration present in the medium at the start of the incubation.  相似文献   

5.
Two processes of porphyrins in heterogeneous systems containing aqueous and membrane phases have been studied with hematoporphyrin and hematoporphyrin derivative: Dimerization equilibrium in the aqueous phases and porphyrin-membrane binding equilibrium using liposomes as models for biological membranes. The interrelationship of aqueous aggregations and membrane binding was probed and the porphyrin aggregation state in the membrane, at equilibrium, was assessed. Fluorimetric techniques were employed. The dimerization equilibrium constants, at neutral pH and 37°C were found to be 2.8 · 105 M−1 and 1.9 · 106 M−1 for hematoporphyrin and its derivative, respectively. Over a porphyrin concentration range going from monomer-dominant to dimer-dominant systems, we have found that only monomers are bound to the membrane. The respective monomer-liposome binding constants, found to be independent of the initial monomer/dimer distribution in the aqueous phase, were determined to be 1.6 · 103 M−1 and 4.1 · 103 M−1 at neutral pH and 37°C for hematoporphyrin and its derivative, respectively. The monomer-liposome interaction was found to perurb the initial monomer/dimer distribution in the aqueous phase, so that the monomers residing at equilibrium in the membrane originate from both monomers and dimers in the aqueous phase.  相似文献   

6.
Elicitins are low-molecular-weight proteins representing the elicitor family secreted by many species of the oomycete Phytophthora. Elicitins induce a hypersensitive reaction in tobacco, a process that is triggered by binding of elicitin to the high-affinity site on the plasma membrane. Specific interaction of cryptogein with the binding sites on tobacco plasma membranes was studied using the piezoelectric biosensor in real time in a flow-through mode. Cryptogeins (wild-type and mutant forms) were covalently immobilized on the sensing surface, and membrane vesicles containing receptors were in solution. Kinetic characterization of the interaction provided values of kinetic rate association (ka) = 5.74 · 106 M1 s−1 and kinetic rate dissociation (kd) = 6.87 10−4 s−1 constants, respectively. The kinetic equilibrium dissociation constant was calculated as KD = 12.0 nM. The piezoelectric biosensor appeared to be a convenient tool for studying interactions of receptors embedded in membrane vesicles.  相似文献   

7.
Annexin A2 (AnxA2) is a phospholipid binding protein that has been implicated in many membrane-related cellular functions. AnxA2 is able to bind different acidic phospholipids such as phosphatidylserine (PS) and phosphatidylinositol-4,5-bisphosphate (PI2P). This binding is mediated by Ca2 +-dependent and Ca2 +-independent mechanisms. The specific functions of annexin A2 related to these two phospholipids and the molecular mechanisms involved in their interaction remain obscure. Herein we studied the influence of lipid composition on the Ca2 +-dependency of AnxA2-mediated membrane bridging and on membrane fluidity. Membrane models of ten different lipid compositions and detergent-resistant membranes from two cellular sources were investigated. The results show that the AnxA2-mediated membrane bridging requires 3 to 50 times less calcium for PS-membranes than for PI2P-membranes. Membrane fluidity was measured by the ratiometric fluorescence parameter generalized polarization method with two fluorescent probes. Compared to controls containing low phospholipid ligand, AnxA2 was found to reduce the membrane fluidity of PI2P-membranes twice as much as the PS-membranes in the presence of calcium. On the contrary, at mild acidic pH in the absence of calcium AnxA2 reduces the fluidity of the PS-membranes more than the PI2P-membranes. The presence of cholesterol on the bilayer reduced the AnxA2 capacity to reduce membrane fluidity. The presented data shed light on the specific roles of PI2P, PS and cholesterol present on membranes related to the action of annexin A2 as a membrane bridging molecule during exocytosis and endocytosis events and as a plasma membrane domain phospholipid packing regulator.  相似文献   

8.
Interactions of two local anesthetics, dibucaine and tetracaine have been studied with phospholipid vesicles containing cholesterol and/or monosialogangliosides (GM1) using fluorescence spectroscopy. The fluorescence intensity of tetracaine showed a marked increase with the increasing molar ratio of the phospholipid to tetracaine, while that of dibucaine showed opposite effects. Steady state anisotropy and the wavelength of maximum emission (λmax) decreased with the increasing phospholipids to tetracaine ratio. The extent of such changes in anisotropy and λmax in the presence and absence of two important components of neuronal membranes, cholesterol and GM1 indicated differential membrane localization of the two local anesthetics. To understand the intercellular mode of action of local anesthetics, we have also studied the interactions of dibucaine and tetracaine with brain spectrin which indicate differential spectrin interactions with similar binding strength. Thermodynamic parameters associated with such binding reveal that binding is favored by entropy. Tetracaine brings about distinct structural changes in spectrin compared to dibucaine, as reflected in the tryptophan mean lifetime and far-UV CD spectra. Tetracaine also exhibits a detergent-like property inducing concentration dependent decrease in spectrin anisotropy, further indicating structural changes in brain spectrin with probable implications in its anesthetic potential.  相似文献   

9.
The interaction of saponins with phospholipid vesicles was investigated by means of liposomal agglutination or a precipitation assay. Ginsenoside-Rc, which has an α-l-arabinofuranose residue at the non-reducing terminus, exhibited remarkable agglutinability toward egg yolk phosphatidylcholine vesicles, while other saponins lacking this characteristic sugar residue showed less or no agglutinability. The molar ratio of ginsenoside-Rc to egg phosphatidylcholine in the aggregates was estimated to be 0.4–0.5 by a precipitation assay using 14C-labeled egg phosphatidylcholine vesicles. The agglutination was inhibited by p-nitrophenyl α-l-arabinofuranoside but not by p-nitrophenyl β-d-glucopyranoside or arabinogalactan. The results indicated that the α-l-arabinofuranose residue in ginsenoside-Rc should be important for the expression of the agglutinability. The agglutinability of ginsenoside-Rc toward lipid vesicles depended on both the polar head groups and fatty acyl chains of phospholipids. Egg yolk phosphatidylcholine vesicles were strongly agglutinated by ginsenoside-Rc, although sphingomyelin, phosphatidylethanolamine, phosphatidic acid and phosphatidylserine were less agglutinated. The agglutinability of ginsenoside-Rc was effective for phosphatidylcholines with short or unsaturated fatty acyl chains. The results suggested that the interaction of ginsenoside-Rc with phospholipid membranes should be affected not only by the chemical structure of the phospholipid but also by the membrane fluidity.  相似文献   

10.
Galactomannans isolated from legume seed endosperms, including those of commercial interest, have been characterized by multidetection aqueous SEC. Galactomannans derived from seeds of the Faboideae subfamily had substantially higher Mw than those from Caesalpinioideae seeds (Mw,Fab = 2.4–3.1 × 106 g/mol, Mw,Caes. = 0.86–2.1 × 106 g/mol) and within the latter botanical subfamily, an apparent correlation between Mw and the degree of galactose substitution DG was found. The molar mass distributions were unimodal and differed primarily by a scale factor, with distributional widths narrower than a true Flory ‘most-probable distribution’; good fits to Schulz–Zimm model were obtained. Across subfamilies no differences were found in the exponents of [η]–M and RvM relationships (0.61 ± 0.02, 0.54 ± 0.01, respectively), the Flory chain stiffness ratio (C = 20 ± 1 (BSF analysis)), or the persistence length (Lp = 5.5 ± 0.2 nm) obtained from SEC fraction data. However, it was found that prefactors in the [η]–M and RvM relationships as well as the unperturbed parameter KΘ decrease in proportion to DG and therefore chain density. Generalized relationships incorporating galactose-dependent prefactors were therefore developed to model SEC fraction data of native galactomannans ([η]GM = (1800 ± 200) × Mo−1.61 × M0.61±0.02, Rv,GM = 0.63 ± 0.05 × Mo−0.54 × M0.54±0.01) as well as lower-M fractions obtained by ultrasonication ([η]GM = (730 ± 100) × Mo−1.71 × Mw0.71±0.02, Rv,GM = 0.49 ± 0.05 × Mo−0.57 × Mw0.57±0.01, M ≈ 1 × 105-native). As a consequence of this dependence and the observed patterns in molar mass variation, [η] varies within a narrow range for galactomannans as a whole despite substantial Mw differences.  相似文献   

11.
Chlorinated hydrocarbons, such as the pesticide lindane (γ-hexachlorocyclohexane), quench the fluorescence of carbazole. The observed quenching is a result of the molecular contacts which occur upon diffusional collisions. Because the amount of quenching depends upon the collisional frequency between carbazole and pesticide, this phenomenon provides a measure of both the diffusional rate of lindane and its local concentration. The carbazole fluorophore is localized within phosphatidylcholine bilayers by cosonicating the lipid with a newly synthesized phospholipid, β-(11-(9-carbazole)-undecanoyl)-l-α-phosphatidylcholine. Using this probe in dimyristoyl-l-α-phosphatidylcholine vesicles, and the above mentioned quenching phenomena, we determined the lindane diffusion rate within the bilayer to be 5.7 · 10−7cm2/s at 37°C. Measurement of the apparent quenching constant at various dimyristoyl phosphatidylcholine concentrations yielded a lipid-water partition coefficient for lindane of 9500, which is in agreement with the value of 8980 obtained by our equilibrium dialysis experiments. Vesicles of dimyristoyl-l-α-phosphatidylcholine become saturated with lindane at a pesticide to lipid molar ratio of approx. 0.28.These results demonstrate the possibility of using the quenching of carbazole fluorescence to investigate the transport and partitioning of pesticides within biological membranes. This ability should prove useful in studies of the interactions of chlorinated hydrocarbons with cell membranes.  相似文献   

12.
Thiolysis of the model diazeniumdiolate prodrug, O2-(2,4-dinitrophenyl) 1-(N,N-diethylamino)diazen-1-ium-1,2-diolate (DNP-DEA/NO, 1), by glutathione (GSH), cysteine (CYSH) and 1-heptanethiol (heptylmercaptan, HM) has been examined in anionic (DOPG), neutral (DPPC, DOPE) and cationic (DOTAP) vesicle media and in glycine buffered aqueous solutions. DOTAP vesicles accelerate the bimolecular reaction with glutathione, cysteine and 1-heptanethiol by factors of 81, 8.2 and 4630, respectively, while reaction is inhibited 5- to 10-fold in the presence of neutral and anionic vesicles. The intrinsic nucleophilicity of the thiols has been compared through the second-order rate constants, 22.9, 5.24 and 43.1 M−1 s−1, for nucleophilic attack on 1 by GS, CYS and M, respectively, obtained in buffered aqueous media. Analysis of the catalysis by DOTAP vesicles, using pseudophase ion-exchange formalism, suggests that the rate increase is due to reactant concentration in the bilayer and interfacial region coupled with enhanced dissociation of the thiol at the vesicle surface. Some contribution from enhanced nucleophilic reactivity at the vesicle interface may also contribute to the greater catalysis by HM. Inhibition of the thiolysis reaction by phospholipid liposomes is attributed to repulsion of the thiolate anions by the negatively charged acyl phosphate of the lipid head group. DOPG = 1,2-dioleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)], DPPC = 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, DOPE = 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine, DOTAP = 1,2-dioleoyl-3-trimethylammonium-propane.  相似文献   

13.
Glycerol diffusional permeabilities through the cytoplasmic cell membrane of Dunaliella salina, the cell envelope of pig erythrocyte and egg phosphattidylcholine vesicles were measured by NMR spectroscopy employing the spin-echo method and nuclear T1 relaxation. The following permeability coefficients (P) and corresponding enthalpies of activation (ΔH) were determined for glycerol at 25°C: for phosphatidylcholine vesicles 5·10−6 cm/s and 11±2 kcal/mol; for pig erythrocytes 7·10−8 cm/s and 18±3 kcal/mol, respectively; for the cytoplasmic membrane of D. salina the permeability at 17°C was found to be exceptionally low and only a lower limit (P<5·10−11cm/s) could be calculated. At temperatures above 50°C a change in membrane permeability occurred leading to rapid leakage of glycerol accompanied by cell death. The data reinforce the notion that the cytoplasmic membrane of Dunaliella represents a genuine anomaly in its exceptional low permeability to glycerol.  相似文献   

14.
Synthetic peptide S3Δ has high affinity for bacterial endotoxin or lipopolysaccharide (LPS). Under tested conditions of pH 5–9 and 0–0.4 M NaCl, the affinity constant, KD ranged from 2·10−6 to 2·10−9 M−1. A novel affinity matrix based on peptide S3Δ was developed for removal of LPS from solutions such as: water; buffers with a wide range of ionic strength and pH; medium for cell culture; and protein solutions under optimized conditions. At a starting LPS of ≈100 EU/ml, a post-purification level below 0.005 EU/ml was achieved.  相似文献   

15.
The heart rate response to isoproterenol (HR-Iso), density and affinity (kd) of β-adrenergic (β-AR) and muscarinic (M2) receptors were compared among three rodents with different generation-life histories of confinement and of high altitude exposure. The European guinea pig (Cavia porcellus) (EGp), a laboratory animal that arrived in Europe after the Spanish Conquest of South America and the Peruvian guinea pig (C. porcellus) (PGp), a semi-wild animal that came from the altiplano to sea level at least 25 generations ago, were used for intra-species comparison. Wistar rats (WR) were used for inter-species comparison as representative of a typical sea level laboratory animal. The HR-Iso was lower in EGp than in the PGp. The PGp showed the highest β-AR density (P<0.0005) and the highest β-AR kd values (P<0.0005) when compared to both EGp and WR groups (β-AR Bmax (fmol mg−1 prot), WR, 19±4; Egp, 34±10; PGp, 74±15. β-AR kd (pM), WR, 24±10; Egp, 17±7; PGp, 39±14). In contrast, PGp showed lower M2 receptor density values than the EGp (P<0.0005). The WR had the highest M2 receptor densities (M2 Bmax (fmol mg−1 prot), WR, 188±15; Egp, 147±9; PGp, 118±6 and M2 kd (pM), WR, 65±12; Egp, 67±6; PGp, 92±2). The inter and intra-species differences found may be related to their respective history of confinement rather than to their history of exposure to high altitude.  相似文献   

16.
The kinetics of formation of the complex ion, μ-carbonato-di-μ-hydroxo-bis((1,5-diamino-3-aza-pentane) cobalt(III), from the tri-μ-hydroxo-bis((1,5-diamino-3-aza-pentane(III)cobalt(III)) ion in aqueous buffered carbonate solution have been studied spectrophotometrically at 295 nm over the ranges 20.0θ°C34.8, 8.03pH9.44, 5 mM [CO32−35 mM and at an ionic strength of 0.1 M (LiClO4). On the basis of the kinetic results a mechanism, involving rapid cleavage of an hydroxo bridge followed by carbon dioxide uptake with subsequent bridge formation, has been proposed. At 25 °C, the rate of the carbon dioxide uptake is 0.58 M−1 s−1 with ΔH≠ = (13.2±0.7) kcal mol−1 and ΔS≠ = (−15.1 ± 0.7) cal deg−1 mol−1. The results are composed with those obtained for several mononuclear cobalt(III) and one dinuclear cobalt(III) complexes.  相似文献   

17.
Summary A method, using albumin-pyrene complexes, has been developed for labeling, in a controlled manner, crab leg nerves whose excitability was preserved.The excimer-to-monomer fluorescence intensity ratio of pyrene, embedded in nerve membrane lipids and in their crude lipid extracts, is a fluidity parameter which displayed the following features with temperatures. a—a temperature-dependent increase of fluidity b—three breaks (6°, 19° and 37°C) in the physiological medium c—In Ca ++-depleted sea water, the 37° characteristic temperature vanished.These breaks may reflect some lateral phase separations of the lipid components of nerve membranes. The calcium dependent temperature break may involve a segregation of acidic phospholipids while the other two breaks (6° and 19°C) may be due to neutral lipids phase separation. The relationship of these findings to nerve function is discussed.  相似文献   

18.
A biotinylated mannotriose (Man3-bio) was dispersively immobilized in the matrix of biotinylated lactose (Gal-Glc-bio) on a streptavidin-covered, 27-MHz quartz crystal microbalance (QCM), and binding kinetics of concanavalin A (Con A) to Man3-bio in the Gal-Glc-bio matrix could be obtained from frequency decreases (mass increases) of the QCM. Association constants (Ka) and binding and dissociation rate constants (kon and koff) could be determined separately as the 1:1 and 1:2 bindings of Con A to Man3-bio on the surface. When Man3-bio was immobilized with content of 1 to 5 mol% in the matrix, the 1:1 binding of Con A to Man3-bio was obtained as Ka = (4 ± 1) × 106 M−1, kon = (4 ± 1) × 104 M−1 s−1, and koff = (12 ± 2) × 10–3 s−1. On the contrary, when Man3-bio was immobilized with content of 20 to 100 mol% in the matrix, the 1:2 binding of Con A to Man3-bio was obtained as Ka = (14 ± 2) × 106 M−1, kon = (14 ± 2) × 104 M−1 s−1, and koff = (7 ± 2) × 10–3 s−1. Thus, Ka for the 1:2 binding was 10 times larger than that for the 1:1 binding, with a three times larger binding rate constant (kon) and a three times smaller dissociation rate constant (koff). This is the first example to obtain separate kinetic parameters for the 1:1 and 1:2 bindings of lectins to carbohydrates on the surface.  相似文献   

19.
Vesicle suspensions of up to 5 % egg lecithin and 2.5 % cholesterol have been found to have no effect on the NMR relaxation times of 17O from water. Addition of 1–5 mM Mn2+ to an equimolar vesicle suspension of egg lecithin and cholesterol permitted resolution of the free induction decay into two exponential components, a fast one arising from the external water and a slow one arising from the intravesicular fluid. From the rates of relaxation the mean life time of the water molecules within the vesicles was calculated to be 1±0.1 ms at 22°C. The size of the vesicle was estimated from electron micrographs to be about 500 Å in diameter. These data yield an equilibrium water permeability, Pw, of about 8 μs−1 for the vesicle membranes. From the temperature dependence of Pw an activation energy of 12±2 kcal/mol was obtained. The longitudinal relaxation time (T1) of water within vesicles remained the same as in pure water.  相似文献   

20.
Summary Static polarization and differential polarized phase fluorimetry studies on rat renal cortical brush border (BBM) and basolateral membranes (BLM) were undertaken to determine the membrane components responsible for differences in BBM and BLM fluidity, whether these differences were due to the order or dynamic components of membrane fluidity and if a fluidity gradient existed within the bilayer. Surface membrane proteins rigidified both BBM and BLM fluidity. Neutral lipid extraction, on the other hand, caused a larger decrease in BBM than BLM fluorescence polarization (0.104vs. 0.60,P<0.01) using diphenyl hexatriene (DPH). Cholesterol addition to phospholipid fractions restored membrane fluidity to total lipid values in both BBM and BLM phospholipids. The response to cholesterol in the BBM was biphasic, while the BLM response was linear. Lateral mobility, quantitated using dipyrenylpropane, was similar in both BBM and BLM fractions at 35°C. BBM and BLM differed primarily in the order component of membrane fluidity as DPH-limiting anisotropy (r ) (0.212vs. 0.154,P<0.01) differed markedly between the two membrane fractions. The two membrane components also differed with respect to 2 and 12-anthroyloxy stearate (2-AS, 12-AS) probes, indicating a difference in the dynamic component of membrane fluidity may also be present. DPH and 12-As probes were also used to quantitate inner core membrane fluidity and showed the BBM was less fluid than the BLM for intact membranes, total lipid extracts and phospholipids. Results obtained using the surface membrane probes trimethylammonium-DPH (TMA-DPH) and 2-AS suggested a fluidity gradient existed in both BBM and BLM bilayers with the inner core being more fluid in both membranes. These data indicate cholesterol is in large part responsible for fluidity differences between BBM and BLM and that these membranes, while clearly differing in the order component of membrane fluidity, may also difer in the dynamic component as well.  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