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.
Small Angle Neutron Scattering (SANS) measurements on two bile salt micelle systems sodium cholate (NaC) and sodium deoxycholate (NaDC) each 0.1 M in the absence and presence of 0.05 M 1-butanol, 1-pentanol, 1-hexanol and 1-octanol were carried out in D20 at ambient conditions (25 degrees C). The scattering cross-section as a function of wave vector transfer (Q) showed the presence of correlation peaks characteristic of strong inter micellar interactions. The correlation peak positions were shifted to low Q values in the presence of n-alkanols for both NaC and NaDC micelles. Best fit curves of SANS were produced by applying Hayter-Penfold analysis considering monodisperse ellipsoid model for the micelles. The best fits were found for constant semi-minor axis = 8A with an increasing semi-major axis for increasing n-alkanol chain lengths. The micellar growths in presence of n-alkanols were studied using the ESR correlation time measurements on suitable spin probe incorporated micelles. The growth parameter and the hydrodynamic radii values were found to be agreeable with SANS data. Intermicellar interaction potentials seem to increase with the Cn of n-alkanols.  相似文献   

2.
The purpose of the present study was to evaluate the possible interaction of bile salt monomer and cholesterol in the intermicellar aqueous phase. Cholesterol and taurocholate monomer concentrations in the intermicellar aqueous phase were determined using 0-20 mM taurocholate solutions saturated with cholesterol. Maximal solubilities of cholesterol in aqueous solutions having various concentrations of taurocholate, especially below its intermicellar monomer concentration (critical micellar concentration), were determined and compared with the intermicellar cholesterol concentration. The intermicellar monomer concentration of taurocholate was constant (6 mM) and independent of taurocholate concentrations. The cholesterol concentration in the intermicellar aqueous phase gradually increased, depending upon taurocholate concentrations, and became constant (1,3 microM) above 10 mM taurocholate. The solubility of cholesterol increased linearly with the taurocholate concentration even below the critical micellar concentration, and was 0.3 microM at 6 mM taurocholate, which was approx. 20-times higher than the aqueous solubility of cholesterol, but a fifth of the maximal intermicellar cholesterol concentration. The results indicate that the higher cholesterol concentration in the intermicellar aqueous phase compared to its aqueous solubility can be primarily ascribed to the interaction of cholesterol with bile salt monomers possibly forming bile salt-cholesterol dimers, and partly to the sustaining forces induced by numerous micelles.  相似文献   

3.
A previously validated in vitro technique was used to determine the effect of diabetes mellitus on the intestinal uptake of cholesterol from various micellar bile salt solutions. The bile salts studied included cholic (C), taurocholic (TC), glycocolic (GC), chenodeoxycholic (CDC), taurochenodeoxycholic (TCDC), glycochenodeoxycholic (GCDC), deoxycholic (DC), taurodeoxycholic (TDC), and glycodeoxycholic (GDC). In control rats there was a reciprocal decline in cholesterol uptake with increasing concentrations of these nine bile acids, and cholesterol uptake was greater from the conjugated primary bile acids than from the unconjugated ones. With a 5 mM concentration of bile acids, the ratios of the uptake of 0.2 mM cholesterol in control rats were C = CDC = DC, TCDC greater than TC greater than TDC, and GC = GCDC greater than GDC; with 20 mM concentrations, the ratios of cholesterol uptake in control rats were C greater than CDC greater than DC, TC greater than TCDC greater than TDC, and GC = GCDC greater than GDC. In the diabetic animals cholesterol uptake was higher than in control rats when using 5 or 20 mM of each of the conjugated bile acids and with cholic acid. In contrast, cholesterol uptake was similar in diabetic and control animals when cholesterol was solubilized with 5 or 20 mM CDC or DC. These differences in cholesterol uptake using the various bile acids and the failure of CDC and DC to facilitate the enhanced uptake of cholesterol in diabetic animals remains unexplained.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
In this work, we develop a methodology to quantitatively follow the solubilization of cholesterol on glycodeoxycholic acid (GDCA) micelles using (13)C nuclear magnetic resonance (NMR). The amount of solubilized cholesterol enriched in (13)C at position 4, [4-(13)C]cholesterol, was quantified from the area of its resonance, at 44.5 ppm, using the CH(2) groups from GDCA as an internal reference. The loading of the micelles with cholesterol leads to a quantitative upper field shift of most carbons in the nonpolar surface of GDCA, and this was used to follow the solubilization of unlabeled cholesterol. The solubilization followed a pseudo first-order kinetics with a characteristic time constant of 3.6 h, and the maximum solubility of cholesterol in 50 mM total lipid (GDCA + cholesterol) is 3.0 ± 0.1mM, corresponding to a mean occupation number per micelle ≥1. The solubilization profile indicates that the affinity of cholesterol for the GDCA micelles is unaffected by the presence of the solute, leading essentially to full solubilization up to the saturation limit. The relaxation times of GDCA carbons at 50mM give information regarding its aggregation and indicate that GDCA is associated in small micelles (hydrodynamic [Rh] = 1.1 nm) without any evidence for formation of larger secondary micelles. This was confirmed by dynamic light scattering results.  相似文献   

5.
Cholesterol crystal formation and growth in model bile solutions   总被引:1,自引:0,他引:1  
Cholesterol monohydrate crystal formation was studied in supersaturated model bile solutions, containing unlabeled cholesterol, sodium cholate and soybean phosphatidylcholine, and tracer amounts of [3H]cholesterol. Solutions were either seeded with cholesterol crystals to initiate growth, or not seeded to allow self-nucleation and subsequent crystal growth to occur. Crystal growth at 37 degrees C was measured by two methods. First, radioactive cholesterol crystals were isolated by filtration, and the mass of cholesterol that had precipitated was calculated. In unseeded solutions, there was a long lag period before crystal growth was detected. This lag time was decreased by increases in the cholesterol concentration, temperature, and lipid concentration. In seeded solutions, crystal growth also was dependent on the cholesterol concentration, temperature, and lipid concentration. The second method used to measure crystal growth involved the Coulter Counter. At 37 degrees C, reproducible results were not obtained using unseeded solutions due to blocking of the counter aperture with large crystals. In seeded solutions, crystal growth could be measured as an increase in total particle volume. However, comparison of growth rate estimates from the Coulter Counter with those obtained radiochemically revealed poor agreement between the two methods. It is probable that the Coulter Counter is inaccurate in measuring the volume of cholesterol monohydrate crystals due to their anisometric shape.  相似文献   

6.
7.
The phenomena of stable and transient acoustic cavitation in liquids exposed to ultrasound are briefly explained. The role of micronuclei, resonant bubble size, and rectified diffusion in the initiation of transient cavitation is reviewed. In aqueous solutions transient cavitation initially generates hydrogen atoms and hydroxyl radicals that may recombine to form hydrogen and H2O2 or may react with solutes in the gas phase, at the gas-liquid boundary, or in the bulk of the solution. The analogies and differences between sonochemistry and ionizing radiation chemistry are explored. The use of spin trapping and electron spin resonance to conclusively identify hydrogen atoms and hydroxyl radicals and to detect cavitation produced by continuous wave and by pulsed ultrasound is described in detail.  相似文献   

8.
The synthetic 10-alkyl isoalloxazines have been found to form vesicles in aqueous and binary solvent systems and confirmed by UV-visible, fluorescence,transmission electron microscopy and quasi elastic light scattering experiments. The mean external diameters of vesicles have been calculated for isoalloxazine with different carbon atom chain at position 10 by transmission electron microscopy and quasi elastic laser light scattering. The gel to liquid phase transition of liposomes measured by differential scanning calorimetry shows reproducible endothermic peak which lies well in the range of typical aqueous vesicles.  相似文献   

9.
N Murai  S Sugai 《Biopolymers》1974,13(6):1195-1203
The conformational changes of poly-Nε-glutaryl-L -lysine (PGL) and poly-Nε-succinyl-L -lysine (PSL) in various salt solutions were studied by use of ORD and potentiometric titration measurements. The addition of alkali metal salts to the fully ionized PGL or PSL solution caused helix formation. The helical content of the polymers increases in the following sequences: at salt concentration 0–2 M, CsCl < KCl < LiCl < NaCl; and at 2–3 M, LiCl < CsCl < KCl ~ NaCl. The preferential binding of the solvent components with various alkali metal salts of PGL or PSL was measured in LiCl, NaCl, and KCl solutions by means of equilibrium dialysis and differential refractometry. It was found that with increasing salt concentration, the polymers were preferentially hydrated in NaCl and KCl soultions; however the salt was preferentially bound to the polymers in LiCl solution. Such preferential binding was suggested to be closely related to conformational change. The addition of CaCl2 to polymer solutions led to the stabilization of the helical structure of PGL or PSL.  相似文献   

10.
The DNA melting transition in aqueous magnesium salt solutions.   总被引:1,自引:0,他引:1  
G S Ott  R Ziegler  W R Bauer 《Biochemistry》1975,14(15):3431-3438
The melting transition of the magnesium salt of DNA has been systematically examined in the presence of various types of anions. The addition of ClO4- to a concentration of 3.0 N results in the biphasic optical transition, with the first phase exhibiting rapid reversibility and independence of the DNA concentration. This subtransition, which is interpreted as an intramolecular condensation to a collapsed form of DNA, is followed by a DNA concentration-dependent aggregation reaction. The aggregation can be reversed by increasing the ClO4- concentration to 6.0 N while elevating the temperature to post-transition levels. Alternatively, both the collapse and the aggregation can be prevented by melting in the presence of trichloroacetate, the most strongly chaotropic solvent for DNA which has been reported (K. Hamaguchi and E. P. Geiduschek (1962), J. Am. Chem. Soc. 84, 1329). The forces responsible for mediating both the collapse and the aggregation are superficially similar to those involved in maintaining duplex stability. The collapsed form, in particular, possibly possesses features in common with the condensed structures which can be produced in aqueous solution of certain polymers, such as polyethylene glycol (Lerman, L.S. (1971), Proc. Natl. Acad. Sci. U.S.A. 68, 1886).  相似文献   

11.
T Schleich  G R Gould 《Biopolymers》1974,13(2):327-337
Using the thermodynamic analysis and methodology of Hill (Biopolymers, 12 , 257 (1973)) for the treatment of optical thermal transition data the effects of various neutral salt additives on the stability and thermodynamics of the poly U–deoxyadenosine interactions that lead to the formation of triple-stranded helical polymer–monomer complexes have been studied. In order of increasing molar effectiveness as polyU–deoxyadenosine complex stability perturbants (pH 7 and in the presence of 1 M NaCl), the various ions may be ranked: SO4?2 < Cl? < Br? < ClO4?; and (CH3)4 N+ < Li+ < Rb+ ~ Na+ < K+ < (CH3 CH2)4 N+ < urea < Guan+ ~ (CH3(CH2)2)4 N+. Destabilizing neutral salt additives (e.g., NaClO4) caused a decrease in the absolute magnitude of the apparent enthalpy and entropy of binding relative to the values determined in the presence of NaCl. By contrast, stabilizing additives (e.g., Na2 SO4) had the opposite effect on these parameters. Along a melting curve the apparent differential heat of complex formation calculated for the binding of deoxyadenosine to poly U in 1 M NaCl appeared to vary linearly with θ, the extent of fractional binding. For such a linear dependence it can be shown that the integral heat (usually determined calorimetrically) equals the differential heat at θ = 0.5. Correcting the apparent differential heat calculated at θ = 0.5 for ligand activity resulted in values for the integral heat of binding of deoxyadenosine to poly U in 1 M NaCl of ?13 to ?16 kcal/mol. Binding isotherms determined in the presence of different inorganic electrolytes could be superimposed provided that different temperatures were compared. However, the additive (CH3)4NCl, which has been shown to interact preferentially with A-T rich regions of DNA (Shapiro, Stannard, and Felsenfeld, Biochemistry, 8 , 3233 (1969)) resulted in a considerably broadened binding isotherm indicating less cooperativity.  相似文献   

12.
Patterns of ice formation in aqueous solutions of glycerol   总被引:1,自引:0,他引:1  
G Rapatz  B Luyet 《Biodynamica》1966,10(198):69-80
  相似文献   

13.
DNA guanine quadruplexes are all based on stacks of guanine tetrads, but they can be of many types differing by mutual strand orientation, topology, position and structure of loops, and the number of DNA molecules constituting their structure. Here we have studied a series of nine DNA fragments (G(3)Xn)(3)G(3), where X = A, C or T, and n = 1, 2 or 3, to find how the particular bases and their numbers enable folding of the molecule into quadruplex and what type of quadruplex is formed. We show that any single base between G(3) blocks gives rise to only four-molecular parallel-stranded quadruplexes in water solutions. In contrast to previous models, even two Ts in potential loops lead to tetramolecular parallel quadruplexes and only three consecutive Ts lead to an intramolecular quadruplex, which is antiparallel. Adenines make the DNA less prone to quadruplex formation. (G(3)A(2))(3)G(3) folds into an intramolecular antiparallel quadruplex. The same is true with (G(3)A(3))(3)G(3) but only in KCl. In NaCl or LiCl, (G(3)A(3))(3)G(3) prefers to generate homoduplexes. Cytosine still more interferes with the quadruplex, which only is generated by (G(3)C)(3)G(3), whereas (G(3)C(2))(3)G(3) and (G(3)C(3))(3)G(3) generate hairpins and/or homoduplexes. Ethanol is a more potent DNA guanine quadruplex inducer than are ions in water solutions. It promotes intramolecular folding and parallel orientation of quadruplex strands, which rather corresponds to quadruplex structures observed in crystals.  相似文献   

14.
Stimulation of human milk lipase by deoxycholate and its taurine and glycine conjugates was demonstrated by measuring the esterolysis reaction of 4-nitrophenylacetate. The steroidal surfactants did not bind strongly to the polar substrate but they did bind effectively to a hydrophobic site on the enzyme and these bile salt-enzyme complexes were effective catalysts. These results are compared with those for stimulation of the enzyme by cholate surfactants and it has been demonstrated that the absence of a 7 alpha-OH substituent on the steroid nucleus does not prevent stimulation of either the esterolytic or lipolytic activity of the enzyme.  相似文献   

15.
This paper describes the derivation of a bile salt monomeric hydrophobicity index that quantitatively defines the composite hydrophilic-hydrophobic balance of a mixture of bile salts. The index is based on the logarithms of bile salt capacity factors determined using reversed phase high performance liquid chromatography (HPLC) (stationary phase octadecyl silane; mobile phase methanol-water 70:30 w/w, ionic strength 0.15). It has been standardized arbitrarily to set indices of taurocholate and taurolithocholate to 0 and 1, respectively. Indices of tauroursodeoxycholate, taurohyodeoxycholate, taurochenodeoxycholate, and taurodeoxycholate were found to be -0.47, -0.35, +0.46, and +0.59, respectively. Whereas capacity factors and hydrophobicity indices of taurine-conjugated bile salts were constant for pH 2.8-9.0, the hydrophilic-hydrophobic balance of glycine-conjugated and unconjugated bile salts was strongly influenced by pH. At alkaline pH (greater than 8.5), hydrophobicity indices of fully ionized unconjugated (n = 4) and glycine-conjugated (n = 6) bile salts differed by only 0.14 +/- 0.02 and 0.05 +/- 0.01, respectively, from those of the corresponding taurine conjugates. At acid pH (less than 3.5) the hydrophobicity indices of four unconjugated bile acids (protonated form) exceeded those of the corresponding salts (ionized form) by 0.76 +/- 0.04; indices of six glycine-conjugated bile acids exceeded those of the corresponding salts by only 0.26 +/- 0.03. Capacity factors of the salt forms of cholate and its conjugates increased dramatically with increasing ionic strength of the mobile phase; retention of the protonated forms (cholic and glycocholic acids) was only minimally influenced by ionic strength. Thus the difference in hydrophilic-hydrophobic balance between a bile acid and its corresponding salt decreases with increasing ionic strength. Examples are given of calculation of hydrophobicity indices for biliary bile salts (fully ionized) from four species under conditions of intact enterohepatic circulation. Mean values, from least to most hydrophobic, were: rat (-0.31) less than dog (0.11) less than hamster (0.22) less than human (0.32). This study provides a rational basis for calculating the hydrophilic-hydrophobic balance of mixed bile salt solutions over a broad range of pH.  相似文献   

16.
Light-scattering has been measured on aqueous NaCl solutions of dodecyldimethylammonium chloride and sodium dodecyl sulfate. From molecular weight determination it is confirmed that spherical micelles are formed at low NaCl concentrations, but at high NaCl concentrations the small micelles formed at the critical micelle concentration further associate to form large rod-like micelles with increasing micelle concentration. The reduction of repulsion between charged groups induces the sphere-rod transition of micelle shape. The dependence of molecular weight on ionic strength can be expressed by double logarithmic relations, which are dependent on the micelle shape. While dodecyldimethylammonium chloride dissolves even in 4.00 M NaCl, sodium dodecyl sulfate solutions exhibit some XXX in angular dissymmetry at NaCl concentrations higher than 0.50 M at low temperatures.  相似文献   

17.
Bilirubin, the yellow-orange tetrapyrrole pigment of jaundice, is essentially insoluble in pure water, but is much more soluble in solutions of bile salts such as sodium taurocholate. The biophysical chemistry of bilirubin in bile salt solutions is affected by changes in the pH of the solution in the range 5-9, suggesting that interactions with bile salt molecules and micelles may alter the acidity of the pigment. We have examined this possibility by determining the apparent pKa values for a series of carboxyl 13C-enriched model compounds, including the bilirubin analog mesobilirubin XIIIalpha, in solutions of sodium taurocholate and sodium taurodeoxycholate. Apparent pKa values were determined by 13C NMR titrations in dimethyl sulfoxide-water mixtures. The results show that the acidity of all compounds is decreased, or pKa increased, in micellar bile salt solution relative to pure water and that the effect is greatest for the larger, less water-soluble compounds. We have proposed a model to explain these results and discussed the implications of these findings for the biophysical chemistry of bilirubin in bile.  相似文献   

18.
S Kitamura  K Takeo  T Kuge  B T Stokke 《Biopolymers》1991,31(11):1243-1255
The thermally induced conformational transition of double-stranded xanthans (degree of pyruvate substitution, DSp = 0.45) having Mw = 3.1, 5.7, and 20.3 x 10(5) has been studied in aqueous salt solutions by high-sensitivity differential scanning calorimetry (DSC). The double strandedness of these samples in the ordered conformation was ascertained by the value of mass per unit length, ML = 2090 +/- 270 g mol-1 nm-1, which was determined from the contour length obtained by electron microscopic observations and the molecular weight by light scattering measurements. The temperature at half completion of the transition T 1/2 for these samples increased linearly with the logarithm of the cation (Na+, K+) concentration. The plot of 1/T1/2 vs the natural logarithm of cation (Na+) concentration in mM for the sample with Mw = 5.7 x 10(5) (15-SX) yielded the equation 10(3)/T1/2 = 3.45-0.159 ln [Na+]. The specific enthalpy delta hcal for 15-SX, essentially independent of salt concentration above 20 mM, was 8.31 +/- 0.39 J/g (SD, n = 6). No systematic dependence of molecular weight on the transition temperature and the enthalpy was observed. Application of the Manning polyelectrolyte theory to the system using the DSC data suggested that the separation of the double strand of xanthan into two single chains was not completed at the temperature where the endothermic peak was finished. This suggestion is consistent with recent findings by light scattering measurements as a function of temperature. Our DSC study was extended to include four other samples from various sources. It was found that T1/2 and delta hcal depend on the pyruvate contents of the samples. For example, the t1/2 (t1/2/degrees C = T1/2/K - 237.15) values for samples with high pyruvate content (DSp = 0.9) and depyruvated (DSp = 0.14) in 20 mM aqueous NaCl were 48.8 and 85.3 degrees C, respectively. Two other samples showed relatively broad DSC curves having shoulders, which were resolved into two independent components. Thermodynamic parameters for each component were examined as a function of salt concentration, and the results obtained were interpreted in terms of the heterogeneity of the pyruvate content of the samples.  相似文献   

19.
The proton magnetic relaxation time, T1, has been measured at 29 MHz in 0.1M KH2PO4 and 0.1M NaCI (both pH 6) aqueous solutions of human ferrihaemoglobin, the protein concentrations ranging from 0.5 to 5 mM per haem. The linear dependence on protein concentration of the enhancement in relaxation rates, Δ(1/T1), due to the presence of the paramagnetic iron in haemoglobin was confirmed at 34°C and at ~10°C. In the middle temperature range there is a thermally activated process, whose energy of activation depends on protein concentration. This dependence is different for the two salt solutions; Ea increases with cHb for 0.1M KH2PO4 and decreases for 0.1M NaCI. The model of water-proton exchange between the bulk solvent and the sixth coordination site of the haem iron was used to calculate the distance from the “liganded” water protons to the haem iron. This yields distances much larger than those determined by X-ray crystal structure analysis. A model is proposed which reconciliates both types of data. The low-temperature relaxation rates cannot be used in deriving quantitative stereochemical data for the haem pocket because of its special shape. Irrespective of the molecular model adopted, the experimental results show clearly that, both at low (~10°C) and higher (>34°C) temperatures, the interaction of paramagnetic haem iron with water protons is practically the same for the two aqueous solutions. The dynamic state of the haemoglobin molecule, as indicated by the middle-temperature range, is completely different in 0.1M KH2PO4 and 0.1M NaCl, pH 6.  相似文献   

20.
Vesicle formation in the Golgi apparatus   总被引:1,自引:0,他引:1  
In this paper we examine the mechanics of vesicle budding from the Golgi apparatus. We propose a model for this process based on the notion that molecular surfactants can release the elastic energy stored in the lipid bilayer. The same physical process may drive other vesiculation processes, including coated vesicle formation and budding of enveloped viruses from the plasma membrane.  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