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 15 毫秒
1.
Isothermal titration calorimetry (ITC) is shown to be a sensitive reporter of bile salt micellization and chiral recognition. Detailed ITC characterization of bile micelle formation as well as the chiral recognition capabilities of sodium cholate (NaC), deoxycholate (NaDC), and taurodeoxycholate (NaTDC) micelle systems are reported. The ΔHdemic of these bile salt micelle systems is directly observable and is strongly temperature‐dependent, allowing also for the determination of ΔCpdemic. Using the pseudo‐phase separation model, ΔGdemic and TΔSdemic were also calculated. Chirally selective guest–host binding of model racemic compounds 1,1’‐bi‐2‐napthol (BN) and 1,1’‐binaphthyl‐2,2’‐diylhydrogenphosphate (BNDHP) to bile salt micelles was then investigated. The S‐isomer was shown to bind more tightly to the bile salt micelles in all cases. A model was developed that allows for the quantitative determination of the enthalpic difference in binding affinity that corresponds to chiral selectivity, which is on the order of 1 kJ mol‐1. Chirality 28:290–298, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

2.
The interaction of the bile salt cholate with unilamellar vesicles was studied. At low cholate content, equilibrium binding measurements with egg yolk lecithin membranes suggest that cholate binds to the outer vesicle leaflet. At increasing concentrations, further bile salt binding to the membrane is hampered. Before the onset of membrane solubilization, diphenylhexatriene fluorescence anisotropy decreases to a shallow minimum. It then increases to the initial value in the cholate concentration range of membrane solubilization. At still higher cholate concentrations, a drop in fluorescence anisotropy indicates the transformation of mixed disk micelles into spherical micelles. Perturbation of the vesicle membranes at molar ratios of bound cholate/lecithin exceeding 0.15 leads to a transient release of oligosaccharides from intravesicular space. The cholate concentrations required to induce the release depend on the size of the entrapped sugars. Cholesterol stabilizes the membrane, whereas, in spite of enhanced membrane order, sphingomyelin destabilizes the membrane against cholate. Freeze-fracture electron microscopy and phosphorus-31 nuclear magnetic resonance (31P NMR) also reflect a change in membrane structure at maximal cholate binding to the vesicles. In 31P NMR spectra, superimposed on the anisotropic line typically found in phospholipid bilayers, an isotropic peak was found. This signal is most probably due to the formation of smaller vesicles after addition of cholate. The results were discussed with respect to bile salt/membrane interactions in the liver cell. It is concluded that vesicular bile salt transport in the cytoplasm is unlikely and that cholate binding is restricted to the outer leaflet of the canalicular part of the plasma membrane.  相似文献   

3.
Micellization of sodium chenodeoxycholate (NaCDC) was studied for the critical micelle concentration (CMC), the micelle aggregation number, and the degree of counterion binding to micelle at 288.2, 298.2, 308.2, and 318.2 K. They were compared with those of three other unconjugated bile salts; sodium cholate (NaC), sodium deoxycholate (NaDC), and sodium ursodeoxycholate (NaUDC). The I(1)/I(3) ratio of pyrene fluorescence and the solubility dependence of solution pH were employed to determine the CMC values. As the results, a certain concentration range for the CMC and a stepwise molecular aggregation for micellization were found reasonable. Using a stepwise association model of the bile salt anions, the mean aggregation number (n) of NaCDC micelles was found to increase with the total anion concentration, while the n values decreased with increasing temperature; 9.1, 8.1, 7.4, and 6.3 at 288.2, 298.2, 308.2, and 318.2 K, respectively, at 50 mmol dm(-3). The results from four unconjugated bile salts indicate that the number, location, and orientation of hydroxyl groups in the steroid nucleus are quite important for growth of the micelles. Activity of the counterion (Na(+)) was determined by a sodium ion selective electrode in order to confirm the low counterion binding to micelles. The solubilized amount of cholesterol into the aqueous bile salt solutions increased in the order of NaUDC相似文献   

4.
Combining micellar electrokinetic capillary chromatography (MEKC) and nuclear magnetic resonance (NMR) experimentation, we shed light on the structural basis for the chirally selective solubilization of atropisomeric binaphthyl compounds by bile salt micelles comprised of cholate (NaC) or deoxycholate (NaDC). The model binaphthyl analyte R,S‐BNDHP exhibits chirally selective interactions with primary micellar aggregates of cholate and deoxycholate, as does the closely related analyte binaphthol (R,S‐BN). Chiral selectivity was localized, by NMR chemical shift analysis, to the proton at the C12 position of these bile acids. Correspondingly, MEKC results show that the 12α‐OH group of either NaC or NaDC is necessary for chirally selective resolution of these model binaphthyl analytes by bile micelles, and the S isomer is more highly retained by the micelles. With NMR, the chemical shift of 12β‐H was perturbed more strongly in the presence of S‐BNDHP than R‐BNDHP. Intermolecular NOEs demonstrate that R,S‐BNDHP and R,S‐BN interact with a similar hydrophobic planar pocket lined with the methyl groups of the bile salts, and are best explained by the existence of an antiparallel dimeric unit of bile salts. Finally, chemical shift data and intermolecular NOEs support different interactions of the enantiomers with the edges of dimeric bile units, indicating that R,S‐BNDHP enantiomers sample the same binding site preferentially from opposite edges of the dimeric bile unit. Chirality 28:525–533, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

5.
Optimal induction of 7α- and 7β-hydroxysteroid dehydrogenase in 100-ml cultures grown to stationary phase was achieved by the addition of metabolizable bile salt inducers: chenodeoxycholate, 7-ketolithocholate or cholate at 2.5–3 h after inoculation. Bile salt addition prior to or after this period markedly reduced the enzyme levels induced. However, when the non-metabolizable inducers deoxycholate and 12-ketolithocholate were similarly added, no significant differences in enzyme levels were observed between addition at 2.5–3 h or at earlier times. The ability of both metabolizable and non-metabolizable bile salts to induce the enzymes fell markedly when additions were made later than approximately 3.5 h. Kinetic studies using 1-l cultures suggest that in a larger culture a somewhat earlier inducer addition period is optimal. When ranked according to the level of enzymes induced the order in decreasing induction power was: chenodeoxycholate, 7-ketolithocholate, deoxycholate, 12-ketolithocholate and cholate. Mixtures of cholate and suboptimal concentrations of deoxycholate induced the culture better than the sum of the two concentrations individually. The end product, ursodeoxycholate, was very effective in blocking the induction by chenodeoxycholate or deoxycholate. Ursocholate (3α,7β,12α-trihydroxy-5β-cholanoate) was less effective. Cultures when grown for 3 h with various bile salts or none, then centrifuged and recultured for a further 3 h in fresh medium containing chenodeoxycholate, all yielded identical enzyme levels within experimental error. We conclude that exposure of the organism to bile salt inducer in the last 3 h of culture was important, while the history of the culture prior to this time was unimportant in the induction process.  相似文献   

6.
Micellization of sodium deoxycholate (NaDC) and sodium ursodeoxycholate (NaUDC) was studied for the critical micelle concentration (CMC), the micelle aggregation number, and the degree of counterion binding to micelle, where sodium cholate (NaC) was used as a reference. The fluorescence probe technique of pyrene was employed to determine accurately the CMC values for the bile salts, which indicated that a certain concentration range of CMC and a stepwise aggregation for micellization were reasonable. The temperature dependences of micellization for NaDC and NaUDC were studied at 288.2, 298.2, 308.2, and 318.2 K by aqueous solubility change with solution pH. Aggregations of the bile salt anions were analyzed using the stepwise association model and found to grow in size with increasing concentration, which confirmed that the mass action model worked quite well. The average aggregation number was found to be 2.5 (NaUDC) and 10.5 (NaDC) at the concentration of 20 mM and at 308.2 K. The aggregation number determined by static light scattering also agreed well with those by the solubility method in the order of size: NaUDC相似文献   

7.
《Process Biochemistry》2010,45(3):306-311
The binding of long-chain imidazolium ionic liquid (IL), 1-tetradecyl-3-methylimidazolium bromide (C14mimBr) to bovine serum albumin (BSA) was investigated by fluorescence spectra and surface tension. Fluorescence spectra show that tryptophan (Trp) residues, one of the intrinsic fluorophores in BSA, are buried in a hydrophobic microenvironment with the addition of C14mimBr, which induces the denaturation of BSA. Moreover, the fluorescence quenching mechanism was determined to be static quenching. The equilibrium constant (K) and the number of binding sites (n) were calculated based on the results of fluorescence measurement. The critical aggregation concentration (CAC) and critical micelle concentration (CMC) under different BSA concentrations at various temperatures were investigated based on the surface tension plots. Surface tension indicates that C14mimBr binds to BSA through electrostatic attraction at low C14mimBr concentrations (below CMC) and through hydrophobic interaction at high C14mimBr concentrations (above CMC). Additionally, the thermodynamic parameters of micelle formation were determined. This study provides an understanding of the binding of C14mimBr to BSA.  相似文献   

8.
Optimal induction of 7 alpha- and 7 beta-hydroxysteroid dehydrogenase in 100-ml cultures grown to stationary phase was achieved by the addition of metabolizable bile salt inducers: chenodeoxycholate, 7-ketolithocholate or cholate at 2.5-3 h after inoculation. Bile salt addition prior to or after this period markedly reduced the enzyme levels induced. However, when the non-metabolizable inducers deoxycholate and 12-ketolithocholate were similarly added, no significant differences in enzyme levels were observed between addition at 2.5-3 h or at earlier times. The ability of both metabolizable and non-metabolizable bile salts to induce the enzymes fell markedly when additions were made later than approximately 3.5 h. Kinetic studies using 1-l cultures suggest that in a larger culture a somewhat earlier inducer addition period is optimal. When ranked according to the level of enzymes induced the order in decreasing induction power was: chenodeoxycholate, 7-ketolithocholate, deoxycholate, 12-ketolithocholate and cholate. Mixtures of cholate and suboptimal concentrations of deoxycholate induced the culture better than the sum of the two concentrations individually. The end product, ursodeoxycholate, was very effective in blocking the induction by chenodeoxycholate or deoxycholate. Ursocholate (3 alpha, 7 beta, 12 alpha-trihydroxy-5 beta-cholanoate) was less effective. Cultures when grown for 3 h with various bile salts or none, then centrifuged and recultured for a further 3 h in fresh medium containing chenodeoxycholate, all yielded identical enzyme levels within experimental error. We conclude that exposure of the organism to bile salt inducer in the last 3 h of culture was important, while the history of the culture prior to this time was unimportant in the induction process.  相似文献   

9.
Deuterium spin-lattice relaxation time (T1) of perdeuterobenzene and perdeuteronaphthalene dissolved in aqueous solutions of sodium deoxycholate and sodium cholate was studied. By comparing the experimental data and theoretical calculations, it was concluded that benzene and naphthalene are not held stationary in the bile acid micelles, and their molecular planes have certain librational motion. In addition, benzene undergoes rapid rotation about the C6 axis, but naphthalene does not rotate in its molecular plane.  相似文献   

10.
In this work the critical micelle concentrations (cmc) of four bile salts, sodium cholate, sodium glycocholate, sodium deoxycholate, and sodium glycodeoxycholate, are determined and presented. Three independent noninvasive methodologies (potentiometry, derivative spectrophotometry, and light scattering) were used for cmc determination, at 25 degrees C with ionic strength adjusted to 0.10 M with NaCl. Spectrophotometric and potentiometric studies of some bile salts were also executed at various ionic strength values, thus allowing the influence of the ionic strength on the cmc value of the bile salt to be assessed. A critical comparison of the cmc values obtained with data collected from the literature is presented. Furthermore, this work makes an evaluation of the conceptual bases of different methodologies commonly used for cmc determination, since variations in the results obtained can be related mainly to different intrinsic features of the methods used (such as sensitivity or the need to include tracers or probes) or to the operational cmc definition applied. The undoubted definition of the experimental bile salt concentration that corresponds to cmc (operational cmc) is essential since in the case of these amphiphiles the formation of micelles is not as abrupt as in the case of ordinary association colloids. The biphasic nature of their aggregation leads to a "round-shaped" variation of the experimental parameters under analysis, which makes difficult the evaluation of the cmc values and can be responsible for the different results obtained.  相似文献   

11.
Synopsis The behaviour of migrating glass eels towards different concentrations of seven bile salts and taurine was investigated by binary-choice experiments. All substances attracted glass eels when presented at concentrations below 10–10M. Glycocholate, taurodeoxycholate and taurine remained attractive at higher concentrations, while taurocholate, cholate, deoxycholate, glycochenodeoxycholate and taurochenodeoxycholate became repellent. A role of bile salts in grouping and orientation behaviour of glass eels is discussed.  相似文献   

12.
The effects of deoxycholate, taurocholate and cholate on transport and mucosal ATPase activity have been investigated in the rat jejunum in vivo using closed-loop and perfusion techniques.In the closed-loops, 5 mM deoxycholate selectively inactivated (Na+ + K+)-ATPase, and net secretion of Na+ induced by 2.5 mM deoxycholate was due to reduced lumen to plasma flux of the ion; deoxycholate (2.5 mM) produced marked inhibition of 3-O-methylglucose transport. Luminal disappearance rates of deoxycholate (60.5±2.9 % per g wet wt of gut) greatly exceeded those of taurocholate (4.3±1.0).In the perfusion studies 1 mM deoxycholate induced net secretion of water, Na+ and Cl, and inhibited active glucose transport; concomitantly “total” ATPase, (Na+ + K+)-ATPase, and Mg2+-ATPase were inhibited. At higher concentrations (5 mM) deoxycholate stimulated Mg2+-ATPase activity. Taurocholate and cholate at 1 mM had no effect on transport or (Na+ + K+)-ATPase. Mucosal lactase, sucrase and maltase activities were not affected by 1 mM deoxycholate, taurocholate or cholate.These results suggest that deoxycholate inhibits sodium-coupled glucose transport by inhibition of (Na+ + K+)-ATPase at the lateral and basal membranes of the epithelial cell, rather than from an effect at the brush-border membrane level.  相似文献   

13.
Bile salts are steroid compounds from the digestive tract of vertebrates and enter the environment via defecation. Many aerobic bile-salt degrading bacteria are known but no bacteria that completely degrade bile salts under anoxic conditions have been isolated so far. In this study, the facultatively anaerobic Betaproteobacterium Azoarcus sp. strain Aa7 was isolated that grew with bile salts as sole carbon source under anoxic conditions with nitrate as electron acceptor. Phenotypic and genomic characterization revealed that strain Aa7 used the 2,3-seco pathway for the degradation of bile salts as found in other denitrifying steroid-degrading bacteria such as Sterolibacterium denitrificans. Under oxic conditions strain Aa7 used the 9,10-seco pathway as found in, for example, Pseudomonas stutzeri Chol1. Metabolite analysis during anaerobic growth indicated a reductive dehydroxylation of 7α-hydroxyl bile salts. Deletion of the gene hsh2 Aa7 encoding a 7-hydroxysteroid dehydratase led to strongly impaired growth with cholate and chenodeoxycholate but not with deoxycholate lacking a hydroxyl group at C7. The hsh2 Aa7 deletion mutant degraded cholate and chenodeoxycholate to the corresponding C19-androstadienediones only while no phenotype change was observed during aerobic degradation of cholate. These results showed that removal of the 7α-hydroxyl group was essential for cleavage of the steroid skeleton under anoxic conditions.  相似文献   

14.
Detergents are widely used for extracting and purifying membrane proteins. Four such detergents have been studied to find the extent to which they alone can alter black lipid film conductances. The slope of the plot of conductivity versus concentration for Triton X-100 is 4.54 in the range 0.025–0.15 mM; dodecyl sulphate 0.82 in the range 0.01–1 mM; sodium deoxycholate 1.03 in the range 0.01–1 mM and sodium cholate 1.37 in the range 0.1–10 mM. These ranges are below the respective critical micelle concentrations; above these concentrations the membranes break. Bilayer lipid membrane conductivity measured at constant detergent concentration increases with the conductivity of the bathing salt solution with a slope greater than 1, indicating an effect on the putative pore structures induced by detergents.  相似文献   

15.
We have demonstrated in vitro the efficacy of the taurine-conjugated dihydroxy bile salts deoxycholate and chenodeoxycholate in solubilizing both cholesterol and phospholipid from hamster liver bile-canalicular and contiguous membranes and from human erythrocyte membrane. On the other hand, the dihydroxy bile salt ursodeoxycholate and the trihydroxy bile salt cholate solubilize much less lipid. The lipid solubilization by the four bile salts correlated well with their hydrophobicity: glycochenodeoxycolate, which is more hydrophobic than the tauro derivative, also solubilized more lipid. All the dihydroxy bile salts have a threshold concentration above which lipid solubilization increases rapidly; this correlates approximately with the critical micellar concentration. The non-micelle-forming bile salt dehydrocholate solubilized no lipid at all up to 32 mM. All the dihydroxy bile acids are much more efficient at solubilizing phospholipid than cholesterol. Cholate does not show such a pronounced discrimination. Lipid solubilization by chenodeoxycholate was essentially complete within 1 min, whereas that by cholate was linear up to 5 min. Maximal lipid solubilization with chenodeoxycholate occurred at 8-12 mM; solubilization by cholate was linear up to 32 mM. Ursodeoxycholate was the only dihydroxy bile salt which was able to solubilize phospholipid (although not cholesterol) below the critical micellar concentration. This similarity between cholate and ursodeoxycholate may reflect their ability to form a more extensive liquid-crystal system. Membrane specificity was demonstrated only inasmuch as the lower the cholesterol/phospholipid ratio in the membrane, the greater the fractional solubilization of cholesterol by bile salts, i.e. the total amount of cholesterol solubilized depended only on the bile-salt concentration. On the other hand, the total amount of phospholipid solubilized decreased with increasing cholesterol/phospholipid ratio in the membrane.  相似文献   

16.
Micelle formations of sodium glyco- and taurochenodeoxycholate (NaGCDC and NaTCDC) and sodium glyco- and tauroursodeoxycholates (NaGUDC and NaTUDC) was studied at 308.2 K for their critical micelle concentrations at various NaCl concentrations by pyrene fluorescence probe, and the degree of counterion binding to micelle was determined using the Corrin-Harkins plots. The degree of counterion binding was found to be 0.37-0.38 for chenodeoxycholate conjugates, while the determination of the degree was quite difficult for ursodeoxycholate conjugates. The change of I1/I3 values on the fluorescence spectrum with the conjugate bile salt concentration suggested two steps for their bile salt aggregation. The first step is a commencement of smaller aggregates, the first cmc, and the second one is a starting of stable aggregates, the second cmc. The aggregation number was determined at 308.2 K and 0.15 M NaCl concentration by static light scattering: 16.3 and 11.9 for sodium NaGCDC and NaTCDC, and 7.9 and 7.1 for NaGUDC and NaTUDC, respectively. The solubilization of cholesterol into the bile salt micelles in the presence of coexisting cholesterol phase and the maximum additive concentration (MAC) of cholesterol was determined against the bile salt concentration. The standard Gibbs energy change for the solubilization was evaluated, where the micelles were regarded as a chemical species. The solubilization was stabilized in the order of NaGUDC approximately = NaTUDC < NaTC < NaGC < NaTCDC < NaGCDC < NaTDC < NaGDC, where the preceding results were taken into the order.  相似文献   

17.
The effects of bile salts on Na+-coupled accumulation of D-glucose and L-alanine by brush-border-membrane vesicles isolated from hamster jejunum were investigated. The approximate percentage inhibition of Na+-coupled D-glucose accumulation produced by various bile salts at a concentration of 1 mM were: deoxycholate and chenodeoxycholate, 60%; glycine and taurine conjugates of deoxycholate and chenodeoxycholate, 40--50%; lithocholate, 45%; cholate and its glycine and taurine conjugates, less than 10%. Inhibition of Na+-coupled accumulation of D-glucose was rapid, reversible and not due to dissolution of the vesicles. Na+-coupled accumulation of L-alanine was also inhibited by deoxycholate. Deoxycholate but not cholate enhanced (1) the rate of Na+ influx, (2) the rate of influx of D-glucose and L-alanine in the absence of a Na+ gradient and (3) the rate of efflux of D-glucose and L-alanine from vesicles preloaded with this sugar or amino acid. Deoxycholate-stimulated efflux of D-glucose was not blocked by phlorizin, which completely prevented efflux in the absence of this bile salt. These results suggest that selected bile salts inhibit Na+-coupled accumulation of D-glucose and L-alanine by enhancing the rate of dissipation of the Na+ gradient required for substrate accumulation. In addition, bile salts may also decrease D-glucose and L-alanine accumulation by increasing the rate of efflux of these substrates across the brush-border plasma membrane.  相似文献   

18.
Cholestyramine, a well-known bile-salt sequestrant, can be used effectively to remove cholate or deoxycholate from a solution of phosphatidylcholine-bile salt mixed micelle. Upon removal of the bile salt, unilamellar phospholipid vesicles form essentially instantaneously. Cholestyramine resin could be pelleted and removed from the vesicle solution after a low speed centrifugation. Based on phosphate analyses, the recovery of vesicles was approximately 60% of the starting material. The average diameter of these vesicles, as estimated by gel exclusion chromatography on sephacryl S-1000 beads and by trapped volume measurement using [3H]sucrose, ranged between 85 to 121 nm. Phosphatidylethanolamine, cholesterol, or n-alkane such as tetradecane can be incorporated into the vesicles without any selective loss; however, selective loss was experienced when negatively charged phospholipid species such as phosphatidylglycerol or phosphatidylserine was included in vesicle formation.  相似文献   

19.
Summary At the millimolar concentrations required for structural studies, NMR spectra of the calcium-binding protein myosin light chain 2 (MLC2) showed resonance line widths indicative of extensive self-association. Pulsed-field-gradient (PFG) NMR spectroscopy was used to examine whether MLC2 aggregation could be prevented by the zwitterionic bile salt derivative 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate (CHAPS). PFG NMR measurements indicated that CHAPS was capable of preventing MLC2 self-association, but only at concentrations well above the critical micelle concentration of 7.5 mM. CHAPS was most effective at a concentration of 22.5 mM, where the apparent molecular mass of MLC2 correponded to a protein monomer plus seven molecules of bound detergent. The resolution and sensitivity of 2D 15N-1H HSQC spectra of MLC2 were markedly improved by the addition of 25 mM CHAPS, consistent with a reduction in aggregation following addition of the detergent. The average amide nitrogen T2 value for MLC2 increased from 30 ms in the absence of CHAPS to 56 ms in the presence of 25 mM CHAPS. The results of this study lead us to propose that PFG NMR spectroscopy can be used as a facile alternative to conventional techniques such as analytical ultracentrifugation for examining the self-association of biological macromolecules.  相似文献   

20.
The effect of bile salts and other surfactants on the rate of incorporation of cholesterol into isolated brush-border membranes was tested. At constant cholesterol concentration, a stimulatory effect of taurocholate was noticed which increased as the bile salt concentration was raised to 20 mM. Taurodeoxycholate was as effective as taurocholate at concentrations of up to 5 mM and inhibited at higher concentrations. Glycocholate was only moderately stimulatory whereas cholate was nearly as effective as taurocholate at concentrations above 5 mM. Other surfactants such as sodium lauryl sulfate and Triton X-100 were very inhibitory at all concentrations tried whereas cetyltrimethyl ammonium chloride was stimulatory only at a very low range of concentrations. These micellizing agents all caused some disruption of the membranes and the greater effectiveness of taurocholate in stimulating sterol uptake was partly relatable to the weaker membrane solubilizing action of this bile salt. Preincubation of membranes with 20 mM taurocholate followed by washing and exposure to cholesterol-containing lipid suspensions lacking bile salt, did not enhance the incorporation of the sterol. In the absence of bile salt the incorporation of cholesterol was unaffected by stirring of the incubation mixtures. Increasing the cholesterol concentration in the mixed micelle while keeping the concentration of bile salt constant caused an increase in rate of sterol incorporation. This increased rate was seen whether the cholesterol suspension was turbid, i.e., contained non-micellized cholesterol, or whether it was optically-clear and contained only monomers and micelles. When the concentration of taurocholate and cholesterol were increased simultaneously such that the concentration ratio of these two components was kept constant, there resulted a corresponding increase in rate of cholesterol uptake. The initial rates of cholesterol incorporation from suspensions containing micellar and monomer forms of cholesterol were much larger than from solutions containing only monomers of the same concentration. The rates of incorporation of cholesterol and phosphatidylethanolamine from mixed micelles containing these lipids in equimolar concentrations were very different. The results as a whole suggest at least for those experimental conditions specified in this study, that uptake of cholesterol by isolated brush-border membranes involves both the monomer and micellar phases of the bulk lipid and that the interaction of the micelles with membrane does not likely involve a fusion process.  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