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.
Concentration dependence of NaCl salting of 0-1.5 mM lysozyme solution in 0.1 M sodium acetate buffer, pH 4.25, was investigated for NaCl concentration varying up to 0.9 M. Calorimetric experiments demonstrated that depending on the salt concentration the estimated number of the binding sites on the lysozyme surface varied in the range of 5 up to 13, and the increase of salt concentration caused the decrease of the number of accessible sites. The small, but significant, local maximum centered at 0.63 M NaCl concentration indicated the specific salting-out of the lysozyme accompanied by binding of approximately 2-3 chloride anions. Generalized McMillan and Mayer's approach reduced to the third-order virial coefficients demonstrates the domination of lysozyme aggregation upon salt addition (a(21)-h(xxy)) and salt organization on the lysozyme surface (a(12)-h(xyy)) processes.  相似文献   

2.
M C Carey  J C Montet  D M Small 《Biochemistry》1975,14(22):4896-4905
The colloid/chemical properties of the fusidane antibiotics, 3-acetoxylfusidic acid, cephalosporin P1, and helvolic acid, and their sodium salts, were investigated. The sodium salts of 3-acetoxylfusidic acid and cephalosporin P1 were found to be detergent-like molecules with micellar properties comparable to the parent compound sodium fusidate and the bile salt sodium cholate. Critical micellar temperatures (cmt) were less than 0 degrees C except for sodium helvolate which being sparingly soluble did not form micelles between 0 and 50 degrees C. Potentiometric titrations of dilute solutions gave apparent pK values (5.2-6.5) in the range expected for carboxylated steroid detergents. The apparent pK values increased significantly once the detergent concentration exceeded the critical micellar concentration (cmc). Micellar properties were determined by surface tension, titration with a water-soluble dye (Rhodamine 6G), light scattering, and solubilization of lecithin and cholesterol. Cmc's, in the range of 1.5 to 5.6 mM, were found which varied slightly depending on the method employed and in all cases fell slightly in the presence of added NaCl. The number of monomers per micelle (aggregation number) in concentrations well above the cmc was extrapolated from Debye light scattering plots in 0.15 M NaCl. The values varied from 6 for fusidate to 14 for 3-acetoxylfusidate with sodium cephalosporin P1 having an intermediate value. Each detergent readily solubilized the phospholipid lecithin.  相似文献   

3.
Dynamic light scattering (DLS) and oscillatory rheology experiments were performed to study temperature dependence (T=10-25 degrees C) of the interactions in hydrogels of gelatin with AOS (alpha olefin sulfonate, anionic surfactant) for surfactant concentrations in the range 25-100 mM, chosen larger than cmc (approximately 8mM). The network mesh size (xi) values deduced from fastmode diffusivity (D(f)) data obtained from dynamic structure factor measurements, S(q, t) approximately exp(-D(f)q(2)t) (for t相似文献   

4.
W F Stafford 《Biochemistry》1985,24(13):3314-3321
The stability of skeletal myosin rod was studied by following the dependence of both papain digestion kinetics and helix-coil transition temperatures on the concentration of neutral salts. The rate of papain-catalyzed digestion of rod to form subfragment 2 and light meromyosin was strongly dependent on chloride concentration but essentially independent of acetate concentration up to 2.0 M. The rod exhibited a biphasic melting curve in 0.6 M NaCl, 5 mM phosphate, and 0.1 mM ethylenediaminetetraacetic acid (EDTA), pH 7.3, with transitions at 45 and 53 degrees C. In 0.6 M CH3COONa, 5 mM phosphate, and 0.1 mM EDTA, pH 7.3, the transitions occurred at 50 and 58 degrees C, respectively. Transition temperatures were obtained with a novel curve-fitting method. The effect of increasing chloride ion concentration on melting profiles was 2-fold. Below 0.6 M salt, the two transition temperatures, Tm,1 and Tm,2, depended on salt concentration such that increasing NaCl concentration caused a small stabilization of the helix while increasing acetate concentration caused the helix to become markedly more stable. Between 0.6 and 1.0 M, variation of chloride concentration had almost no effect on the thermal stability of the rod while increasing acetate concentration increased its stability considerably. Above 1.0 M NaCl, the melting profiles became broad with a third transition being observed (e.g., at 3.0 M, Tm,3 = 38 degrees C), indicating the existence of a region which has a tendency to be destabilized by chloride. The third transition was not observed at comparable concentrations of acetate. This effect of chloride was not expected on the basis of its position in the Hofmeister series.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Possible biotechnological applications of extreme halophilic enzymes are strongly determined by their high salt requirement of around 4 M NaCl. Consequently, the use of these in organic media seemed to be unlikely. However, we have succeeded in dissolving a halophilic enzyme, p-nitrophenylphosphate phosphatase from the archaeon Halobacterium salinarum, in an organic medium by creating a reverse micellar system with very low salt concentration. The enzyme retained its catalytic properties in reversed micelles made with an anionic surfactant (dioctyl sodium sulphosuccinate) or with a cationic surfactant (hexadecyltrimethylammonium bromide) in cyclohexane plus 1-butanol as co-surfactant. The dependence of the rate of hydrolysis of p-nitrophenylphosphate phosphate on the molar water/surfactant ratio (w(0) value) showed a bell-shaped curve for each surfactant system. Kinetic parameters were determined in each system. The enzymatic reaction appeared to follow Michaelis-Menten kinetics with the anionic surfactant only. The kinetic behaviour was determined at different concentrations of Mn(2+) in reversed micelles of dioctyl sodium sulphosuccinate as surfactant.  相似文献   

6.
Downstream processing of lipase involving reverse micellar extraction of lipase using cationic surfactant cetyltrimethylammonium bromide (CTAB) was investigated. Effect of various process parameters on both forward and backward extraction of lipase from crude extract was studied to optimize its yield and purity. Complex interaction of salt concentration (0.05∼0.15M), surfactant concentration (0.10∼0.30 M), and pH (6.0∼9.0) for forward extraction, as well as, salt concentration (0.5∼1.5 M) and pH (6.0∼9.0) for backward extraction have been studied using response surface methodology. Optimum processing conditions, namely, salt concentration 0.16M, surfactant concentration 0.20 M, and pH 9.0 for forward extraction, as well as, salt concentration 0.80 M and pH 7.23 for backward extraction, fulfill the conditions to obtain activity recovery of lipase ≥78% and purification factor of lipase ≥4.0. The study demonstrated that response surface methodology can be used for optimization of the conditions for reverse micellar extraction of lipase.  相似文献   

7.
Langowski J  Hammermann M  Klenin K  May R  Tóth K 《Genetica》1999,106(1-2):49-55
We present here recent results on the structure of superhelical DNA and its changes with salt concentration between 0.01 and 1.5 M NaCl. Scattering curves of two different superhelical DNAs were determined by static light scattering. The measured radii of gyration do not change significantly with salt concentration. Small-angle neutron scattering, together with calculations from a Monte Carlo model, allows to determine the superhelix diameter. Measured and simulated scattering curves agreed almost quantitatively. Experimentally we find that the diameter decreases from 16.0±0.9 nm at 10 mM to 9.0±0.7 nm at 100 mM NaCl. The superhelix diameter from the simulated conformations decreased from 18.0±1.5 nm at 10 mM to 9.4±1.5 nm at 100 mM NaCl. At higher salt concentrations up to 1.5 M NaCl, the diameter stays constant at 9 nm.This revised version was published online in October 2005 with corrections to the Cover Date.  相似文献   

8.
The adsorption of a range of single chain zwitterionic phosphocholine surfactants (C(n)P(m)C) at the air/liquid interface has been studied by a combination of surface tension and neutron reflectivity. The critical micellar concentration (CMC) for C(n)PC (or C(n)P(2)C), where n varied from 12, 14 to 16, was found to be 0.91, 0.14, and 1.2 x 10(-2) mM respectively, and followed the same trend as observed for other zwitterionic and non-ionic surfactants. The area per molecule at the CMC, A(cmc), for C(n)PC was found to remain constant between 50 and 53 A(2), indicating that the increase in the alkyl chain length had little effect on A(cmc) at the interface. The neutron reflection measurement also showed an almost constant layer thickness (tau) of 20+/-2 A from all the alkyl chain deuterated PC surfactants (dC(n)hPC) in null reflecting water (NRW), suggesting that the alkyl chains of the surfactant responded to changes in either chain length or solution concentration by varying their angle of tilt. In contrast, increasing the length of head group linker between P and N atoms in C(12)P(m)C, where m=2, 4, to 6, resulted in a much slower decrease of CMC from 0.91, 0.7, to 0.5 mM, consistent with a different contribution to the free energy of micellization. A(cmc) for C(12)P(m)C did not vary when m was increased from 2 to 4, and this observation together with the thickness of the head group region indicated an almost perpendicular projection of the head group in C(12)P(2)C and C(12)P(4)C. A further increase in m to 6 resulted in an A(cmc) of 70 A(2). This increase in A(cmc) however did not result in any change in either the total layer thickness or the fraction of the head group region submerged in the aqueous subphase, suggesting that the head group in C(12)P(6)C was bent away from the surface normal direction. Both increase in temperature from 25 to 40 degrees C and the addition of 0.1 M NaCl had little effect on the area per molecule or the thickness of C(12)P(m)C surfactant layer, showing that the C(12)P(m)C series behaved like C(n)P(2)C series. The main conclusion from this study is that for all the C(n)P(m)C surfactants studied, change in m or n has little effect on the total thickness, the thickness of the alkyl chain or that of the head group region.  相似文献   

9.
M D King  D Marsh 《Biochemistry》1987,26(5):1224-1231
The critical micelle concentrations (cmc's) of a variety of spin-labeled phospholipids, 1-acyl-2-[4-(4,4-dimethyloxazolidine-N-oxyl)valeryl]-sn-glycero-3-pho sph o derivatives, have been determined by electron spin resonance (ESR) spectroscopy. The narrow, three-line ESR spectra of the rapidly tumbling monomers are clearly distinguished from the spin-spin broadened spectra of the micellar aggregates, allowing a direct determination of the concentrations of the two species. The influence of both the hydrocarbon chain length and the polar head group on the energetics of self-assembly has been studied. For phosphatidylcholine, 1n [cmc] decreases linearly with the length of the sn-1 chain. The gradient of this linear dependence corresponds to a free energy of transfer of the monomer from the aqueous phase to the micelle of delta Gtr = -1.1RT per CH2 group. The cmc's of the 1-lauroyl derivatives of both phosphatidylcholine and phosphatidylglycerol have relatively shallow, biphasic temperature dependences with a minimum at approximately 20 degrees C. Both of these properties are characteristic of the hydrophobic effect, with the free energy of transfer being slightly less than that for the solubility of n-hydrocarbons in water, corresponding to the reduced configurational entropy of the lipid chains in the micellar state. The cmc's of the 1-lauroyl derivatives of the phospholipids in 0.15 M NaCl, for their various charge states, are as follows: phosphatidic acid(2-), 0.77 mM; phosphatidic acid(1-), 0.13 mM; phosphatidylserine(1-), 0.24 mM; phosphatidylglycerol(1-), 0.17 mM; phosphatidylcholine, 0.10 mM; phosphatidylethanolamine, 0.05 mM.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Cooperative interaction of histone H1 with DNA.   总被引:4,自引:1,他引:3       下载免费PDF全文
The cooperative binding of histone H1 with DNA was studied using a fluorescently labelled histone H1. The titration data were analysed in terms of the large ligand model. The stoichiometric number, n = 65 +/- 10 bases/H1, was independent of NaCl concentration (0.02 - 0.35 M). The nucleation and the cooperative binding constants, K' and K, and the cooperativity parameter q were sensitive to salt concentration; K = 3.6 +/- 0.8 X 10(7) M-1 and q = 1.1 +/- 0.4 X 10(3) at 0.2 M NaCl. The dependence of K' on NaCl concentration revealed that 6 Na+ ions were released from DNA upon complex formation. An extrapolation of K' to 1M NaCl yielded a small value, K' = 5 +/- 2 M-1. Thus the binding of H1 is essentially electrostatic, being compatible with its independence of temperature. A calculation of K' based on the counterion release reproduced the salt concentration dependence of K'. Therefore, the binding of H1 is of an electrostatic territorial type. Thus, H1 may move along the DNA chain to a certain extent, when both salt concentration and the degree of saturation are sufficiently low. The condition is so restricted that the sliding would not play an important role in vivo. It was concluded from the DNA concentration independent binding isotherm that H1 can cooperatively bind onto a single DNA molecule. A simple power law dependence of the cooperativity parameter q upon NaCl concentration was found; q oc[NaCl]h with h = 0.72, though the physical basis of this dependence remains unknown.  相似文献   

11.
R E Stafford  T Fanni  E A Dennis 《Biochemistry》1989,28(12):5113-5120
The critical micelle concentration (cmc) of several lysophospholipids and of a lysophospholipid analogue was determined from surface tension measurements using the maximum bubble pressure method and/or 31P NMR. The use of the maximum bubble pressure method has now been extended to micromolar concentrations of surfactant, and the experimental parameters that effect its use have been explored. Surface activity was found to vary with changes in the chain length and in the headgroup polarity of the lysophospholipid. The cmc's for 1-decanoyl-, 1-dodecanoyl-, 1-tetradecanoyl-, and 1-hexadecanoyl-sn-glycero-3-phosphocholine are 7.0, 0.70, 0.070, and 0.007 mM, respectively. The cmc's for 1-decanoyl- and 1-dodecanoyl-sn-glycero-3-phosphoethanolamine are 4.4 and 0.33 mM, respectively. The cmc for dodecylphosphocholine, a lysophospholipid analogue, was found to be 1.1 mM. The cmc's for 1-tetradecanoyl- and 1-hexadecanoyl-sn-glycero-3-phosphoglycerol were found to be 3.0 and 0.60 mM, respectively, in pure water. In 0.1 M Tris-HCl (pH = 8.0), their cmc's are 0.16 and 0.018 mM, respectively. Surface tension and adsorption density values determined at the cmc are reported for each compound. The relationship of dynamic surface tension and lipid purity is discussed. These studies provide information about the micellization and interfacial properties of several biologically important lysophospholipids.  相似文献   

12.
Surfactant concentration, ionic strength, and pH were optimised for the selective separation and purification of periplasmic cytochrome c553 from recombinant E. coli TG2 cells using response surface methodology. Back-extraction was accomplished using counter-ionic surfactant addition. Optimum forward extraction conditions were: 65 mM bis(2-ethylhexyl)sulfosuccinate sodium salt (AOT), 0.07 M NaCl, and pH 8.4, while the optimum back-extraction conditions were 80 mM trioctylmethylammonium chloride, 0.85 M KCl, and pH 9.62. In comparison to a conventionally purified sample using column chromatography (10 mg cytochrome c553 l–1 with a purity of 0.66), reverse micelles achieved the same concentration and similar purity (0.50) in only two simple steps.  相似文献   

13.
Data are reported for T(m), the temperature midpoint of the thermal unfolding curve, of ribonuclease A, versus pH (range 2-9) and salt concentration (range 0-1 M) for two salts, Na(2)SO(4) and NaCl. The results show stabilization by sulfate via anion-specific binding in the concentration range 0-0.1 M and via the Hofmeister effect in the concentration range 0.1-1.0 M. The increase in T(m) caused by anion binding at 0.1 M sulfate is 20 degrees at pH 2 but only 1 degree at pH 9, where the net proton charge on the protein is near 0. The 10 degrees increase in T(m) between 0.1 and 1.0 M Na(2)SO(4), caused by the Hofmeister effect, is independent of pH. A striking property of the NaCl results is the absence of any significant stabilization by 0.1 M NaCl, which indicates that any Debye screening is small. pH-dependent stabilization is produced by 1 M NaCl: the increase in T(m) between 0 and 1.0 M is 14 degrees at pH 2 but only 1 degree at pH 9. The 14 degree increase at pH 2 may result from anion binding or from both binding and Debye screening. Taken together, the results for Na(2)SO(4) and NaCl show that native ribonuclease A is stabilized at low pH in the same manner as molten globule forms of cytochrome c and apomyoglobin, which are stabilized at low pH by low concentrations of sulfate but only by high concentrations of chloride.  相似文献   

14.
Reduced lysozyme at pH 2.5 bound poly(oxyethylene) alkylethers in two steps and the maximum bound amount Qmax of the surfactant reached as large as 0.5-0.7 mole per mole amino acid residue in the cooperative binding step. Binding isotherms were well superimposed when surfactant concentrations were normalized by respective values of the critical micelle concentration, cmc. In terms of the onset concentrations of the cooperative binding C*, hydrophobicity of reduced lysozyme was quantitatively defined as RT In (cmc/C*) which amounted to 670 J per mole surfactant and was unique to the protein irrespective of the kind of surfactant. Qmax could be used as another measure of the hydrophobicity of the protein. The binding isotherms were evaluated by two methods: equilibrium dialysis and surface tension. Their results were consistent with each other and rather complementary. Reduced lysozymes were molecularly dispersed at pH below 2.5 in 0.01 M NaCl but aggregation took place as pH increased. The aggregates could not be dissociated on dilution nor by the addition of nonionic surfactants but by lowering pH. The irreversible nature of the aggregation was reasonably interpreted with a model based on the 'entangled' arrangement of the beta-sheets, which could account for the irreversible aggregation of unfolded proteins in general.  相似文献   

15.
Antioxidants solubilized in micellar solutions can change micellar properties like the size and shape of micelles, critical micellar concentration (cmc) and viscosity. Interactions arising between antioxidants and the surfactant determine the locations of antioxidants and vice versa. The location and interaction are dependent on the type of both the antioxidant and surfactant. Influences of various antioxidants on the physical and structural properties were tested in micellar systems of cationic CTAB, non-ionic Brij 58 and anionic SDS. The antioxidants used to investigate the effects of gradually increasing lipophilicity were gallic acid (GA) and the gallate esters from methyl to octyl gallate (MG-OG). Hydroxy cinnamic acids (HCAs) like -coumaric acid (pC), caffeic acid (CA), ferulic acid (FA) and sinapic acid (SA) were employed to observe effects of functional groups like hydroxyl and methoxy groups. Micellar size and shape determined by small angle neutron scattering (SANS), viscosity and cmc were chosen to characterize the antioxidant influence. In Brij 58 systems propyl gallate (PG) did not affect the cmc or aggregation number but decreased micellar size slightly due to an intercalation of PG into the region of the polyoxyethylene chain and the first adjacent alkyl chain methylene groups. In SDS systems the micellar size and cmc decreased in the presence of PG. This was attributed to PG residing in the Stern layer. However, in CTAB systems micelles swelled at low PG concentration and in the presence of GA, while higher PG concentrations and more lipophilic antioxidants led to a sphere-to-rod transition with a simultaneous increase in viscosity and decrease in cmc. This revealed the intercalation of antioxidants in the palisade layer of CTAB micelles entering into strong interactions of electrostatic and hydrophobic origin. It could be demonstrated that the interactions became stronger the more lipophil an antioxidant is and the more hydroxyl groups are attached to the aromatic ring. Differences in the location and interaction of antioxidant and micelles are proposed as being responsible for the effectiveness of antioxidants.  相似文献   

16.
Some phase separation phenomena in aqueous dioctanoyllecithin solutions including the effects of NaCl and Lil on the phase diagram are reported. At low electrolyte concentrations (below 0.2 M) both salts cause the upper consolute temperatures (u.c.t) to decrease, probably due to a decrease of electrostatic attraction between the lipid molecules. At higher salt concentrations the effect of Lil continues in the same direction (salting-in) but NaCl leads to an increase of the u.c.t. (salting-out).Micellar weight determinations could be performed at room temperature in homogeneous lecithin solutions containing 0.2 M Lil. An attempt was made to interpret the light scattering data with the help of the open association model (equal association constants for aggregation steps beyond a certain step) and the Flory-Huegins type of thermodynamic nonideality. The angular dependence of the light scattering points to very large and extended micelles. The radii of gyration are approximately proportional to the square root of the micellar weights.  相似文献   

17.
HU is one of the most abundant proteins in bacterial chromosomes and participates in nucleoid compaction and gene regulation. We report experiments using DNA stretching that study the dependence of DNA condensation by HU on force, salt and HU concentration. Previous experiments at sub-physiological salt levels revealed that low concentrations of HU could compact DNA, whereas larger HU concentrations formed a DNA-stiffening complex. Here we report that this bimodal binding behavior depends sensitively on salt concentration. Only the compaction mode was observed for 150 mM and higher NaCl levels, i.e. for physiological salt concentrations. Similar results were obtained for the more physiological salt K-glutamate. Real-time studies of dissociation kinetics revealed that HU unbound slowly (minutes to hours under the conditions studied) but completely for salt concentrations at or above 100 mM NaCl; the lifetime of HU complexes was observed to increase with the HU concentration at which the complexes were formed, and to decrease with salt concentration. Higher salt levels of 300 mM NaCl completely eliminated observable HU binding to DNA. Finally, we observed that the dissociation kinetics depend on force applied to the DNA: increased applied force in the sub-piconewton range accelerates dissociation, suggesting a mechanism for DNA tension to regulate chromosome structure and gene expression.  相似文献   

18.
The detergent CHAPS was found to be the preferable surfactant for the efficient purification and reconstitution of the Torpedo californica nicotinic acetylcholine receptor (AChR). The main result is that the incorporation of the AChR proteins into lipid vesicles by CHAPS dialysis was strongly dependent on the salt and protein concentrations. As monitored by sucrose gradients, by electron microscopy, and by agonist-induced lithium ion flux, the best reconstitution yields were obtained in 0.5 M NaCl at a protein concentration of 0.5 g/L and in 0.84 M NaCl at 0.15 g/L protein. Electron micrographs of receptor molecules, which were incorporated into vesicles, showed single, nonaggregated dimer (M(r) = 580,000) and monomer (M(r) = 290,000) species. CHAPS dialysis at NaCl concentrations less than 0.5 M largely reduced the receptor incorporation concomitant with protein aggregation. Electron micrographs of these preparations revealed large protein sheets or ribbons not incorporated into vesicles. The analysis of static and dynamic light scattering demonstrated that the detergent-solubilized AChR molecules aggregate at low lipid contents (less than or equal to 500 phospholipids/AChR dimer), independent of the salt concentration. AChR proteins eluted from an affinity column with a solution containing 8 mM CHAPS (but no added lipid) still contained 130 +/- 34 tightly bound phospholipids per dimer. The aggregates (about 10 dimers on the average) could be dissociated by readdition of lipid and, interestingly, also by increasing the CHAPS concentration up to 15 mM. This value is much higher than the CMC of CHAPS = 4.0 +/- 0.4 mM, which was determined by surface tension measurements. The data clearly suggest protein-micelle interactions in addition to the association of monomeric detergents with proteins. Furthermore, the concentration of the (free) monomeric CHAPS at the vesicle-micelle transformation in 0.5 M NaCl ([Dw]c = 3.65 mM) was higher than in 50 mM NaCl ([Dw]c = 2.8 mM). However, it is suggested that the main effect of high salt concentrations during the reconstitution process is an increase of the fusion (rate) of the ternary protein/lipid/CHAPS complexes with mixed micelles or with vesicular structures, similar to the salt-dependent fusion of vesicles.  相似文献   

19.
Periplasmic metal binding protein characterized by high histidine content was cloned from moderate halophile, Chromohalobacter salexigens. The protein, termed histidine-rich metal binding protein (HP), was expressed in and purified from E. coli as a native form. HP bound to Ni- and Cu-loaded chelate columns with high affinity, and Co- and Zn-columns with moderate affinity. Although the secondary structure was not grossly altered by the addition of 0.2–2.0 M NaCl, the thermal transition pattern was considerably shifted to higher temperature with increasing salt concentration: melting temperature was raised by ~20 °C at 2.0 M NaCl over the melting temperature at 0.2 M NaCl. HP showed reversible refolding from thermal melting in 0.2–1.15 M NaCl, while it formed irreversible aggregates upon thermal melting at 2 M NaCl. Addition of 0.01–0.1 mM NiSO4 stabilized HP against thermal melting with high reversibility, while addition above 0.5 mM resulted in irreversible melting due to aggregation.  相似文献   

20.
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   

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