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.
Measurements of differential scanning calorimetry (d.s.c.) have been made on the complex bovine serum albumin (BSA)--sodium dodecyl sulphate (SDS) under various conditions. There are two peaks P1 and P2 in the d.s.c. curve for BSA at pH 7 and in the absence of NaCl, indicating the presence of the heat-induced transition of BSA. There are three peaks P1, P2 and P3 in the curve for the system with the molar mixing ratio SDS/BSA = 1. With the increase in the amount of SDS, the peak P3 grows at the expense of P1 and P2. There is only a single peak P3 in the curve for the systems SDS/BSA > 7, and no peak at SDS/BSA = 50 and 100. There is a single peak P12 in the curve for BSA at pH 7 and in the presence of 0.05 M NaCl, indicating that the heat-induced transition is suppressed. There are two peaks P12 and P3 for the systems SDS/BSA = 1-5; the area ratio of the peak P3 to P12 increases with the increase in the amount of SDS. There is only a single peak P3 when SDS/BSA > 7, and no peak at SDS/BSA = 50. It is concluded that the peak P3 is a product of SDS regardless of the presence or absence of NaCl. Values of thermal denaturation temperature (Td) and enthalpy (delta H) of thermal denaturation indicate that the complex AD12 (A = BSA, D = SDS) is in the most thermostabilized state.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
13C NMR chemical shift results as a function of pH for a series of carboxyl 13C-enriched saturated fatty acids (8-18 carbons) bound to bovine serum albumin (BSA) are presented. For octanoic acid bound to BSA (6:1, mol/mol), the chemical shift of the only FA carboxyl resonance (designated as peak c), plotted as a function of pH, exhibited a complete sigmoidal titration curve that deviated in shape from a corresponding theoretical Henderson-Hasselbach curve. However, FA carboxyl chemical shift plotted as a function of added HCl yielded a linear titration curve analogous to those obtained for protein-free monomeric fatty acid (FA) in water. The apparent pK of BSA-bound octanoic acid was 4.3 +/- 0.2. However, the intrinsic pK (corrected for electrostatic effects resulting from the net positive charge on BSA) was approximately 4.8, a value identical to that obtained for monomeric octanoic acid in water in the absence of protein. For long-chain FA (greater than or equal to 12 carbons) bound to BSA (6:1, mol/mol), chemical shift titration curves for peak c were similar to those obtained for octanoic acid/BSA. However, the four additional FA carboxyl resonances observed (designated as peaks a, b, b', and d) exhibited no change in chemical shift between pH 8 and 3. For C14.0 X BSA complexes (3:1 and 6:1, mol/mol) peaks b' and a exhibited chemical shift changes between pH 8.8 and 11.5 concomitant with chemical shift changes in the epsilon-carbon (lysine) resonance. In contrast, peaks c and d exhibited no change and peak b only a slight change in chemical shift over the same pH range. We conclude: the carboxyl groups of bound FA represented by peaks a, b, b', and d were involved in ion pair electrostatic interactions with positively charged amino acyl residues on BSA; the carboxyl groups of bound FA represented by peak c were not involved in electrostatic interactions with BSA; the similarity of the titration curves of peak c for BSA-bound octanoic acid and long-chain FA suggested that short-chain and long-chain FA represented by peak c were bound to the same binding site(s) on BSA; bound FA represented by peaks b' and a (but not d or b) were directly adjacent to BSA lysine residues. We present a model which correlates NMR peaks b, b', and d with the putative locations of three individual high-affinity binding sites in a three-dimensional model of BSA.  相似文献   

3.
Using defatted and SH-blocked bovine serum albumin (BSA), measurements of differential scanning calorimetry (d.s.c.) have been made mainly in NaSCN solution. BSA undergoes a heat-induced conformational transition in a particular range of pH and ionic strength and is separated into two thermally independent units, each of which has different thermostability in acidic and alkaline pH regions. Comparisons were made of the pH dependencies of the enthalpy of thermal denaturation (ΔH) and the temperature of thermal denaturation (Td) in 0.01 NaSCN with those in 0.01 NaCl. It has been found that the stabilizing effect of NaSCN on BSA is larger than that of NaCl at pH 3.5–8, and that the heat-induced transition occurs by the electrostatic repulsive forces among the positively charged amino acid residues in a segment Arg 184–Arg 216 containing Trp 212 and the primary binding sites of anions. At ionic strength 0.01, the relative effectiveness of anions in suppressing the heat-induced transition and increasing the thermostability of BSA follows the order ClO4 − SCN > I > SO42− > Br > Cl. At ionic strength 0.1, the heat-induced transition is suppressed in all the salt solutions, and a Td increase follows the order ClO4 SCN > I > Br > Cl SO42−. Thus, the highly chaotropic ions thermostabilize BSA more markedly than kosmotropic ions in the low and moderate salt concentrations. In contrast, chaotropic ions destabilize BSA and kosmotropic ions stabilize BSA at the higher concentrations. An adequate amount of NaCl or NaSCN prevents the destruction of the environment of the binding site in the segment containing Trp 212 in 4 urea solution at pH 7.0.  相似文献   

4.
1. The effect of alkaline earth cations on the dissociation of the extracellular hemoglobin of Lumbricus terrestris and the effect of ionic strength on the dissociation of the hemoglobins of L. terrestris and Tubifex tubifex at concentrations of ca 2.5 mg/ml, over the pH range 9.0-10.5 was investigated using ultracentrifugation to separate the dissociated from the undissociated molecules. 2. Mg(II), Ca(II) and Sr(II) at concentrations of up to 0.2 M, decreased the dissociation of Lumbricus oxyhemoglobin from 70% at pH 9.0 and 100% at pH 9.5 and higher, to 20-30% at 0.05 M. The three cations were equally effective in decreasing the extent of dissociation of L. terrestris oxyhemoglobin over the pH range 9.0-10.5, with a K1/2 of ca 10 mM. 3. The dissociation of L. terrestris oxyhemoglobin over the pH range 9.0-10.5 was decreased only to 50-60% in the presence of up to 0.5 M NaCl or KCl; there was no further decrease in dissociation at concentrations of the two salts up to 1.5 M. 4. The dissociation of T. tubifex oxyhemoglobin over the pH range 9.0-10.0 was decreased from 100% to ca 40-50% in the presence of 0.5 M NaCl or KCl with little or no change at higher concentrations. At pH 10.5 and 11.0 the decrease in dissociation was more gradual, reaching ca 50% at 1.5 M NaCl.  相似文献   

5.
13C NMR chemical shift and intensity results for a series of carboxyl 13C-enriched saturated fatty acids (8-18 carbons) bound to bovine serum albumin (BSA) are presented as a function of increasing fatty acid (FA)/BSA mole ratio. Spectra for long-chain (greater than or equal to 12 carbons) FA X BSA complexes exhibited up to five FA carboxyl resonances, designated a, b, b', c, and d. Only three resonances (peaks b, b', and d) were observed below 3:1 FA X BSA mole ratio, and at greater than or equal to 3:1 mole ratio, two additional resonances were observed (peaks c and a). In a spectrum of 5:1 stearic acid X BSA complexes, peaks b, b', and d each represented approximately one-fifth, and peak c approximately two-fifths, of the total FA carboxyl intensity. Plots of total carboxyl/carbonyl intensity ratio as a function of FA X BSA mole ratio were linear up to 7-9 mole ratio. Deviation from linearity at mole ratios greater than or equal to 7 was accompanied by the detection of crystalline unbound FA (as 1:1 acid/soap) by X-ray diffraction. In contrast to long-chain FA X BSA complexes, 13C NMR spectra of octanoic acid X BSA complexes yielded only one FA carboxyl resonance (peak c) at FA X BSA mole ratios between 1 and 20. We conclude: peaks b, b', and d represent FA bound to three individual high affinity (primary) long-chain FA binding sites on BSA; peak c represents FA bound to several secondary long-chain (or primary short-chain) FA binding sites on BSA; peak a represents long-chain FA bound to an additional lower affinity binding site. We present a model that correlates the observed 13C NMR resonances with individual binding site locations predicted by a recent three-dimensional model of BSA.  相似文献   

6.
Hyaluronan (HA) hydrolysis catalyzed by hyaluronidase (HAase) is inhibited at low HAase over HA ratio and low ionic strength, because HA forms electrostatic complexes with HAase, which is unable to catalyze hydrolysis. Bovine serum albumin (BSA) was used as a model to study the HA-protein electrostatic complexes at pH 4. At low ionic strength, there is formation of (i) neutral insoluble complexes at the phase separation and (ii) small positively-charged or large negatively-charged soluble complexes whether BSA or HA is in excess. According to the ionic strength, different types of complex are formed. Assays for HA and BSA led to the determination of the stoichiometry of these complexes. HAase was also shown to form the various types of complex with HA at low ionic strength. Finally, we showed that at 0 and 150 mmol L(-1) NaCl, BSA competes with HAase in forming complexes with HA and thus induces HAase release resulting in a large increase in the hydrolysis rate. These results, in addition to data in the literature, show that HA-protein complexes, which can exist under numerous and varied conditions of pH, ionic strength and protein over HA ratio, might control the in vivo HAase activity.  相似文献   

7.
Bovine serum albumin (BSA) causes tobacco mosaic virus (TMV) to crystallize at pH values where both have negative charges. The amount of albumin required to precipitate the virus varies inversely with ionic strength of added electrolyte. At pH values above 5, the precipitating power is greatest when BSA has the maximum total, positive plus negative, charge. Unlike early stages of the crystallization of TMV in ammonium sulfate-phosphate solutions, which can be reversed by lowering the temperature, the precipitation of TMV by BSA is not readily reversed by changes in temperature. The logarithm of the apparent solubility of TMV in BSA solutions, at constant ionic strength of added electrolyte, decreases linearly with increasing BSA concentration. This result and the correlation of precipitating power with total BSA charge suggest that BSA acts in the manner of a salting-out agent. The effect of BSA on the reversible entropy-driven polymerization of TMV protein (TMVP) depends on BSA concentration, pH, and ionic strength. In general, BSA promotes TMVP polymerization, and this effect increases with increasing BSA concentrations. The effect is larger at pH 6.5 than at pH 6. Even though increasing ionic strength promotes polymerization of TMVP in absence of BSA, the effect of increasing ionic strength from 0.08 to 0.18 at pH 6.5 decreases the polymerization-promoting effect of BSA. Likewise, the presence of BSA decreases the polymerization-promoting effect of ionic strength. The polymerization-promoting effect of BSA can be interpreted in terms of a process akin to salting-out. The mutual suppression of the polymerization-promoting effects of BSA and of electrolytes by each other can be partially explained in terms of salting-in of BSA.  相似文献   

8.
Composition-gradient multi-angle static light scattering (CG-MALS) is an emerging technique for the determination of intermolecular interactions via the second virial coefficient B22. With CG-MALS, detailed studies of the second virial coefficient can be carried out more accurately and effectively than with traditional methods. In addition, automated mixing, delivery and measurement enable high speed, continuous, fluctuation-free sample delivery and accurate results. Using CG-MALS we measure the second virial coefficient of bovine serum albumin (BSA) in aqueous solutions at various values of pH and ionic strength of a univalent salt (NaCl). The systematic variation of the second virial coefficient as a function of pH and NaCl strength reveals the net charge change and the isoelectric point of BSA under different solution conditions. The magnitude of the second virial coefficient decreases to 1.13 x 10−5 ml*mol/g2 near the isoelectric point of pH 4.6 and 25 mM NaCl. These results illuminate the role of fundamental long-range electrostatic and van der Waals forces in protein-protein interactions, specifically their dependence on pH and ionic strength. Electronic supplementary material The online version of this article (doi:10.1007/s10867-014-9367-7) contains supplementary material, which is available to authorized users.  相似文献   

9.
The redox potential of the Rieske [2Fe-2S] cluster of the bc1 complex from bovine heart mitochondria was determined by cyclic voltammetry of a water-soluble fragment of the iron/sulfur protein. At the nitric-acid-treated bare glassy-carbon electrode, the fragment gave an immediate and stable quasireversible response. The midpoint potential at pH 7.2, 25 degrees C and I of 0.01 M was Em = +312 +/- 3 mV. This value corresponds within 20 mV to results of an EPR-monitored dye-mediated redox titration. With increasing ionic strength, the midpoint potential decreased linearly with square root of I up to I = 2.5 M. From the cathodic-to-anodic peak separation, the heterogeneous rate constant, k degrees, was calculated to be approximately 2 x 10(-3) cm/s at low ionic strength; the rate constant increased with increasing ionic strength. From the temperature dependence of the midpoint potential, the standard reaction entropy was calculated as delta S degrees = -155 J.K-1.mol-1. The pH dependence of the midpoint potential was followed over pH 5.5-10. Above pH 7, redox-state-dependent pK changes were observed. The slope of the curve, -120 mV/pH above pH9, indicated two deprotonations of the oxidized protein. The pKa values of the oxidized protein, obtained by curve fitting, were 7.6 and 9.2, respectively. A group with a pKa,ox of approximately 7.5 could also be observed in the optical spectrum of the oxidized protein. Redox-dependent pK values of the iron/sulfur protein are considered to be essential for semiquinone oxidation at the Qo center of the bc1 complex.  相似文献   

10.
Y R Hsu  T Arakawa 《Biochemistry》1985,24(27):7959-7963
Interferon gamma is distinguished from other types of interferons in its instability upon acid treatment, as demonstrated by a loss of antiviral activity. Acid unfolding and refolding experiments were performed with recombinant DNA derived human interferon gamma. When the protein was subjected to unfolding and refolding, the refolded protein showed two peaks (peaks I and II) in gel filtration which have been shown to differ in size, structure, and antiviral activity. When the smaller, peak II, form was unfolded by dialysis against 0.01 M HCl containing 0.1 M NaCl (pH 2) and refolded by dialysis against various solvents at neutral pH, it re-formed as peak II but also generated peak I, and the ratio of the two forms was dependent on protein concentration and solvent conditions. Higher protein concentrations and higher ionic strength led to a greater ratio of peak I to peak II. Phosphate buffers caused precipitation of peak I. Since peak II is 4-8 times more active than peak I in the antiviral bioassay, generation of peak I by acid treatment of peak II should lead to a decrease in antiviral activity.  相似文献   

11.
We investigated the effects of pH and ionic strength of solutions used for antigen retrieval to elucidate the mechanism of heat-induced antigen retrieval (HIAR) in immunohistochemistry. The immunostaining intensity of nuclear, cytoplasmic, cell membrane, and extracellular matrix antigens with 17 different antibodies was evaluated in formaldehyde-fixed and paraffin-embedded mouse and human tissues. Deparaffinized sections were autoclaved for 10 min in buffers with different pH values ranging from 3.0 to 10.5. To test the influence of ionic strength on immunoreactions, the sections were autoclaved for 10 min in 20 mM Tris-HCl buffers (TB) at pH 9.0 and 10.5 with or without 25, 50, and 100 mM NaCl. There were two immunostaining patterns for pH dependency of HIAR. First, the majority of antibodies recovered their antigenicity when heated in the buffers with both acidic pH (pH 3.0) and basic pH (pH 9.0 and 10.5). Second, some antibodies showed strong immunostaining only at basic pH values (pH 9.0 and 10.5). When the sections were autoclaved in TB at pH 9.0, immunostaining of all eight antibodies examined decreased as the NaCl concentration increased. On the other hand, when the sections were treated with TB at pH 10.5, all antibodies yielded stronger reactions in the buffer containing NaCl than in the buffer without NaCl; five antibodies exhibited the strongest immunoreaction at concentrations from 25 to 50 mM. These results suggest that the extended polypeptides by heating are charged negatively or positively at basic or acidic pH, and that an electrostatic repulsion force acts to prevent random entangling of polypeptides caused by hydrophobic attractive force and to expose antigenic determinants, during cooling process of HIAR solution.  相似文献   

12.
The interaction of gum arabic (GA) and bovine serum albumin (BSA) has been investigated through turbidity and light scattering intensity measurements and by the use of dynamic light scattering, laser Doppler velocimetry, and isothermal titration calorimetry. It has been shown that GA and BSA can form soluble and insoluble complexes depending on the solution pH and the mixing ratio and is a function of the net charge on the complex. Soluble complexes were obtained when the electrophoretic mobility was greater than ±1. 5 μm s(-1) V(-1) cm(-1). Changes in the value of the isoelectric point of the complexes with mixing ratio and isothermal titration calorimetric data indicated that complexes formed at pHs 3 and 4 consisted of ~60 BSA molecules for every GA molecule, while at pH 5 there were ~10 BSA molecules per GA molecule. Calorimetric studies also indicated that the interaction occurred in two stages at both pH 3 and pH 4, but that the nature of the interaction at these two pH values was significantly different. This was attributed to differences in the relative magnitude of the positive and negative charges on the BSA and GA, respectively, and possibly due to changes in the BSA conformation. The fact that there is an interaction at pH 5, which is above the isoelectric point of the BSA, is due to the interaction of the carboxylate groups on the GA with positive patches on the BSA or to the charge regulation of the protein-polysaccharide system brought about by changes in dissociation equilibria. Complexation is reduced as the ionic strength of the solvent increases and is prevented at a NaCl concentration of 120 mM.  相似文献   

13.
A protease with a tetrameric quaternary structure was extractedwith 1 M NaCl from spinach PS II membranes and purified by hydrophobic,anion-exchange and gel-filtration chromatography using onlybuffers of high ionic strength. Gel-filtration chromatographyresulted in elution of the protease in fractions that correspondedto molecular masses of 156 kDa and 39 kDa, and re-chromatographyof either peak gave both peaks again. This result indicatesthat the protease is represented by an equilibrium between a156-kDa tetramer and a 39-kDa monomer. SDS-polyacrylamide gelelectrophoresis of the protease fractions revealed a polypeptidewhose molecular mass was 39 kDa without prior reduction, butthe molecular mass increased to 41 kDa after prior reductionwith dithiothreitol. This finding suggests that the monomerpossesses an intramolecular disulfide linkage whose reductioncauses a configurational change that increases the effectivemolecular size. The protease had maximum activity at pH 7.0–9.0.The activity was diminished by the presence of 5 mM NaCl andwas almost completely inhibited by 50 mM NaCl. These observationssuggest that an environment of low ionic strength is a prerequisitefor the activity of this enzyme. The protease was inhibitedby dithiothreitol, a result that indicates that the 39-kDa formmaintained by the disulfide linkage is essential for activity.Studies with protease inhibitors suggested that this enzymeis not a serine-protease. 1Present address: Department of Biomolecular Science, Facultyof Science, Toho University, Miyama 2-2-1, Funabashi, 274 Japan (Received October 19, 1989; Accepted April 12, 1990)  相似文献   

14.
The aim of the present work is to study the sequential adsorption of F(ab')(2) and bovine serum albumin (BSA) molecules adsorbed onto positively and negatively charged polystyrene latexes. Cationic and anionic latexes were prepared by emulsifier-free emulsion polymerization. Adsorptions of F(ab')(2) on both latexes at a low ionic strength and different pHs were performed. The cationic latex showed a higher adsorption of F (ab')(2) molecules over a range of pH, which could be due to the formation of multilayers. Sequential adsorption of anti-CRP F(ab')(2) and monomeric BSA were performed at two different pre-adsorbed F(ab')(2) amounts on both types of latex. Displacement of F(ab')(2) occurred only when the preadsorbed amounts were larger than a certain critical value, which depends on the adsorption pH. A greater displacement of larger preadsorbed amounts might be the result of a weaker contact between the protein molecules and the polystyrene surface. The displacement of F(ab')(2) previously adsorbed onto both latexes occurred due to pH changes, an increase of ionic strength and the presence of BSA molecules. The effect caused by these three factors was studied independently. The main factors in the desorption of F(ab')(2) on the anionic latex are the changes in pH and ionic strength, whereas on the cationic latex the desorption is mainly caused by the increase of the ionic strength and the presence of BSA. The colloidal stability of the immunotatex was improved by BSA adsorption, especially on cationic latex. (c) 1995 John Wiley & Sons, Inc.  相似文献   

15.
Acetate kinase from E. coli K-12 was studied chromatographically on omega-aminoalkyl polysacharide sorbents. The dependence of the protein sorption-desorption on ionic strength and the effect of pH on the acetate kinase sorption were studied. The increase in the ionic strength caused a decrease in the amount of protein sorbed on the hexamethylenediamine- and chlorotriasinehexamethylenediamine sepharoses. On hexamethylenediamine-, octamethylenediamine- and dimethylhexamethylenediamine agaroses acetate kinase was adsorbed within the pH range of 6.5-9.0, whereas on the chlorotriasinehexamethylenediamine sepharose--at pH 6.5-8.0. The active protein was eluted at ionic strength of 0.14-0.17 M. Acetate kinase was not adsorbed on the carboxypropyonylaminohexyl sepharose within the pH range studied, i.e. 5.0-9.0 and was not adsorbed on hexamethylenediamine agarose at pH 4.0 and on chlorotriasinehexamethylenediamine sepharose--at pH 9.0. The mechanism of the enzyme-adsorbent interaction is discussed.  相似文献   

16.
Although protein fractionation by selective membrane filtration has numerous potential applications in both the downstream processing of fermentation broths and the purification of plasma proteins, the selectivity for proteins with only moderately different molecular weights has generally been quite poor. We have obtained experimental data for the transport of bovine serum albumin (BSA) and immunoglobulins (IgG) through 100,000 and 300,000 molecular weight cutoff polyethersulfone membranes in a stirred ultrafiltration device at different solution pH and ionic strength. The selectivity was a complex function of the flux due to the simultaneous convective and diffusive solute transport through the membrane and the bulk mass transfer limitations in the stirred cell. Under phsioligical conditions (pH 7.0 and 0.15 M NaCI) the maximum selectivity for the BSA-IgG separation was only about 2.0 due primarily to the effects of protein adsorption. In contrast, BSA-IgG selectivities as high as 50 were obtained with the same membranes when the protein solution was at pH 4.8 and 0.0015 M NaCl. This enhanced selectivity was a direct result of the electrosatatic contributions to both bulk and membrane transport. The membrane selectivity could actually be reversed, with higher passage of the larger IgG molecules, by using a 300,000 molecular weight cutoff membrane at pH 7.4 and an ionic strength of 0.0015 M NaCl. These results clearly demonstrate that the effectiveness of selective protein filtration can be dramatically altered by appropriately controlling electrostatic interactions through changes in pH and/or ionic strength. (c) 1994 John Wiley & Sons, Inc.  相似文献   

17.
A fluorophore-nitroxide free radical dual-functional probe (FN) was utilized to study the kinetics of ascorbate (AH(-)) binding to Bovine Serum Albumin (BSA). Since the free radical fragment in the FN probe intramolecularly quenches fluorescence, ascorbate reduction of the nitroxide function is accompanied by a concomitant fluorescence intensity increase from the fluorophore. Thus, both fluorescence and the EPR techniques could be utilized to measure the reaction rate. In the presence of BSA protein, the observed rate of the overall process is the sum of that from at least two reactions: the reaction between free ascorbate and free probe, and the reaction between bound ascorbate and bound probe. Our findings show that the observed rate is strongly dependent on the ionic strength of the medium. A corollary of this observation is the indication of a purely electrostatic interaction between ascorbate and the BSA protein. This conclusion was further corroborated by 1H NMR measurement of the transverse relaxation time, T(2), of ascorbate protons in BSA solutions. Ascorbate ion was released from the ascorbate/BSA ensemble in the presence of increasing concentrations of NaCl. Binding constants of AH(-) to BSA were calculated at different ionic strengths at pH 7.4. Furthermore, an increase in ionic strength did not affect the ability of albumin to protect ascorbate against autoxidation. This suggests that the protein's protective antioxidant effect may be attributed to BSA binding of trace quantities of transition-metal cations (rather than ascorbate binding to BSA). This conclusion is supported by ascorbate UV-absorption measurements in the presence of albumin and Cu(2+) ions as a function of ionic strength.  相似文献   

18.
The adsorption equilibria of bovine serum albumin (BSA), gamma-globulin, and lysozyme to three kinds of Cibacron blue 3GA (CB)-modified agarose gels, 6% agarose gel-coated steel heads (6AS), Sepharose CL-6B, and a home-made 4% agarose gel (4AB), were studied. We show that ionic strength has irregular effects on BSA adsorption to the CB-modified affinity gels by affecting the interactions between the negatively charged protein and CB as well as CB and the support matrix. At low salt concentrations, the increase in ionic strength decreases the electrostatic repulsion between negatively charged BSA and the negatively charged gel surfaces, thus resulting in the increase of BSA adsorption. This tendency depends on the pore size of the solid matrix, CB coupling density, and the net negative charges of proteins (or aqueous - phase pH value). Sepharose gel has larger average pore size, so the electrostatic repulsion-effected protein exclusion from the small gel pores is observed only for the affinity adsorbent with high CB coupling density (15.4 micromol/mL) at very low ionic strength (NaCl concentration below 0.05 M in 10 mM Tris-HCl buffer, pH 7.5). However, because CB-6AS and CB-4AB have a smaller pore size, the electrostatic exclusion effect can be found at NaCl concentrations of up to 0.2 M. The electrostatic exclusion effect is even found for CB-6AS with a CB density as low as 2.38 micromol/mL. Moreover, the electrostatic exclusion effect decreases with decreasing aqueous-phase pH due to the decrease of the net negative charges of the protein. For gamma-globulin and lysozyme with higher isoelectric points than BSA, the electrostatic exclusion effect is not observed. At higher ionic strength, protein adsorption to the CB-modified adsorbents decreases with increasing ionic strength. It is concluded that the hydrophobic interaction between CB molecules and the support matrix increases with increasing ionic strength, leading to the decrease of ligand density accessible to proteins, and then the decrease of protein adsorption. Thus, due to the hybrid effect of electrostatic and hydrophobic interactions, in most cases studied there exists a salt concentration to maximize BSA adsorption.  相似文献   

19.
Inorganic phosphate, a strong activator of glutamate dehydrogenase at pH 8.0–9.0, is an inhibitor at pH 6.0–7.6. The extent of inhibition increases with the decrease of pH. The same effect is shown by other electrolytes, including Tris-hydroxymethyl-aminomethane and NaCl.The combined effect of pH and ionic strength also alters the allosteric characteristics of the enzyme. Lowering the pH minimizes the activation by high concentrations of NAD; phosphate partially restores this activation. The allosteric activation by ADP disappears at pH around neutrality; in the pH range 6.0–7.0, ADP becomes a strong inhibitor, the inhibition being enhanced by the addition of ionic compounds. Similarly, the extent of allosteric inhibition by guanosine 5′-triphosphate (pyro) (GTP), which is maximal at pH 9.0, decreases at lower pH values and a slight activation is observed in the presence of electrolytes at pH 6.0.Glutamate dehydrogenase, selectively desensitized by dinitrophenylation in the presence of ADP, can be activated by ADP at pH 9.0, but is no longer inhibited by the same effector at pH 6.0, high salt concentration. The densensitized enzyme is not inhibited by GTP at pH 9.0, but is activated by this effector at pH 6.0 in the presence of ionic compounds. Conversely, GTP-protected dinitrophenylated glutamate dehydrogenase is desensitized only to the effect of the activating modifier, ADP at pH 9.0, GTP at pH 6.0, high salt concentration. These findings suggest that the conformation of each allosteric site of glutamate dehydrogenase is changed by pH and ionic strength so that it keeps its specificity for the ligand which brings about a given effect, activation or inhibition, independently from its chemical structure.  相似文献   

20.
Aqueous dispersions of 50 mM dimyristoylphosphatidylglycerol (DMPG) in the presence of increasing salt concentrations (2-500 mM NaCl) were studied by small angle X-ray scattering (SAXS) and optical microscopy between 15 and 35 degrees C. SAXS data show the presence of a broad peak around q approximately 0.12 A(-1) at all temperatures and conditions, arising from the electron density contrasts within the bilayer. Up to 100 mM NaCl, this broad peak is the main feature observed in the gel and fluid phases. At higher ionic strength (250-500 mM NaCl), an incipient lamellar repeat distance around d=90-100 A is detected superimposed to the bilayer form factor. The data with high salt were fit and showed that the emergent Bragg peak is due to loose multilamellar structures, with the local order vanishing after approximately 4d. Optical microscopy revealed that up to 20 mM NaCl, DMPG is arranged in submicroscopic vesicles. Giant (loose) multilamellar vesicles (MLVs) start to appear with 50 mM NaCl, although most lipids are arranged in small vesicles. As the ionic strength increases, more and denser MLVs are seen, up to 500 mM NaCl, when MLVs are the prevailing structure. The DLVO theory could account for the experimentally found interbilayer distances.  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