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.
T Arakawa  S N Timasheff 《Biochemistry》1987,26(16):5147-5153
The causes of the salting-in of beta-lactoglobulin by glycine and NaCl, a solubility behavior contrary to expectations, were probed by a detailed study of the interactions between these solvent components and the protein. The preferential interactions of beta-lactoglobulin with solvent components in aqueous glycine and NaCl systems have been compared with those of bovine serum albumin and lysozyme. At neutral pH, beta-lactoglobulin exhibited insignificant preferential interactions in glycine and NaCl at low cosolvent concentrations and an increasing preferential hydration at higher concentrations, the levels approaching the values expected from the other two proteins. These results indicate considerable binding of the electrolytes to beta-lactoglobulin, sufficient to compensate for the exclusion due to perturbation of the solvent surface tension. The difference between the preferential interactions of beta-lactoglobulin and the other proteins with these two solvent additives was shown to be the cause of the increase of beta-lactoglobulin solubility even at high concentrations of the additives, at which they have salting-out effects on the other proteins. The preferential interactions of NaCl with the three proteins were examined as a function of pH. The results showed no pH dependence of the preferential hydration for bovine serum albumin and lysozyme, while this parameter increased significantly for beta-lactoglobulin at lower pH. This suggests that the binding of electrolytes to beta-lactoglobulin is due to a unique charge distribution on the surface of the protein around neutral pH, which imparts to this protein a large dipole moment.  相似文献   

2.
Solvent conditions play a major role in a wide range of physical properties of proteins in solution. Organic solvents, including dimethyl sulfoxide (DMSO), have been used to precipitate, crystallize and denature proteins. We have studied here the interactions of DMSO with proteins by differential refractometry and amino acid solubility measurements. The proteins used, i.e., ribonuclease, lysozyme, beta-lactoglobulin and chymotrypsinogen, all showed negative preferential DMSO binding, or preferential hydration, at low DMSO concentrations, where they are in the native state. As the DMSO concentration was increased, the preferential interaction changed from preferential hydration to preferential DMSO binding, except for ribonuclease. The preferential DMSO binding correlated with structural changes and unfolding of these proteins observed at higher DMSO concentrations. Amino acid solubility measurements showed that the interactions between glycine and DMSO are highly unfavorable, while the interactions of DMSO with aromatic and hydrophobic side chains are favorable. The observed preferential hydration of the native protein may be explained from a combination of the excluded volume effects of DMSO and the unfavorable interaction of DMSO with a polar surface, as manifested by the unfavorable interactions of DMSO with the polar uncharged glycine molecule. Such an unfavorable interaction of DMSO with the native protein correlates with the enhanced self-association and precipitation of proteins by DMSO. Conversely, the observed conformational changes at higher DMSO concentration are due to increased binding of DMSO to hydrophobic and aromatic side chains, which had been newly exposed on protein unfolding.  相似文献   

3.
Experiments have been performed at 298.15 K to measure the density, sound velocity and refractive index of glycine in aqueous solutions of NaCl over a wide range of both glycine and NaCl concentrations. The values of apparent molar volume and isentropic compressibility of glycine were calculated from the measured data. The results show a positive transfer volume of glycine from an NaCl solution to a more concentrated NaCl solution. This indicates that the size of a glycine molecule is larger in a solution with higher NaCl concentration. The negative values of apparent isentropic compressibility imply that the water molecules around the glycine molecules are less compressible than the water molecules in the bulk solution. These effects are attributed to the doubly charged behaviour of glycine and to the formation of physically bonded ion-pairs between the charged groups of glycine and sodium and chloride ions. The formation of ion-pairs, whose extents of binding reactions depend on the concentrations of both NaCl and glycine, alter the hydration number of glycine. This also explains the reason for the increase in the size of glycine with an increase in the NaCl concentration. A model based on the Pitzer formalism has been developed to correlate the activity coefficient, apparent molar volume and isentropic compressibility of glycine in aqueous solutions of NaCl. The results show that the model can accurately correlate the interactions in aqueous solutions of glycine and NaCl.  相似文献   

4.
A mild and rapid method is described for isolating various milk proteins from bovine rennet whey. beta-Lactoglobulin from bovine rennet whey was easily adsorbed on and desorbed from a weak anion exchanger, diethylaminoethyl-Toyopearl. However, alpha-lactalbumin could not be adsorbed onto the resin. alpha-Lactalbumin and beta-lactoglobulin from rennet whey could also be adsorbed and separated using a strong anion exchanger, quaternary aminoethyl-Toyopearl. The rennet whey was passed through a strong cation exchanger, sulphopropyl-Toyopearl, to separate lactoperoxidase and lactoferrin. alpha-Lactalbumin and beta-lactoglobulin were adsorbed onto quaternary aminoethyl-Toyopearl. alpha-Lactalbumin was eluted using a linear (0-0.15 M) concentration gradient of NaCl in 0.05 M Tris-HCl buffer (pH 8.5). Subsequently, beta-lactoglobulin B and beta-lactoglobulin A were eluted from the column with 0.05 M Tris-HCl (pH 6.8), using a linear (0.1-0.25 M) concentration gradient of NaCl. The yields were 1260 mg alpha-lactalbumin, 1290 mg beta-lactoglobulin B and 2280 mg beta-lactoglobulin A from 1 l rennet whey.  相似文献   

5.
Wang X  Lee J  Wang YW  Huang Q 《Biomacromolecules》2007,8(3):992-997
The composition and rheological properties of beta-lactoglobulin/pectin coacervates have shown significant correlations with sodium chloride concentration (C(NaCl)) and initial protein/polysaccharide ratio (r). An increase of C(NaCl) from 0.01 to 0.21 M at r = 5:1 leads to the increase in both beta-lactoglobulin and pectin contents in the coacervates, which can be explained in terms of salt-enhanced effect at lower salt concentrations. Further increase of C(NaCl) from 0.21 to 0.41 M decreases the proportions of these two biopolymers in the coacervates, exhibiting salt-reduced effect at higher salt concentrations. Moreover, the stronger self-aggregation of beta-lactoglobulin with increasing salt concentration gives rise to a decreasing actual protein/polysaccharide ratio in the coacervates at 0.01-0.21 M C(NaCl) and r = 5:1. An increase of r from 5:1 to 40:1 often increases the actual amount of pectin chains in beta-lactoglobulin/pectin coacervates, but it exhibits a maximum in beta-lactoglobulin content at r = 20:1. A much higher storage modulus (G') than loss modulus (G' ') for all beta-lactoglobulin/pectin coacervates suggests the formation of highly interconnected gel-like structure. The values of G' increase as C(NaCl) increases from 0.01 to 0.21 M, whereas a further increase of C(NaCl) from 0.21 to 0.41 M causes G' values to decrease to much lower values. These results further disclose the salt-enhanced effect and the salt-reduced effect at low and high salt concentrations, respectively. On the other hand, increasing r from 5:1 to 40:1 favors the formation of stronger gel-like beta-lactoglobulin/pectin coacervates, which mainly originates from the higher actual amount of pectin chains in beta-lactoglobulin/pectin coacervates at higher r values.  相似文献   

6.
The activity coefficients at 298.15 K of glycine in water + NaNO(3) + glycine system and dl-methionine in water + NaCl + dl-methionine system are reported. The measurements were performed in an electrochemical cell with two ion selective electrodes, a cation and an anion ion selective electrode, each versus a double junction reference electrode. The concentrations of the electrolytes and the amino acids studied covered up to 1.0 molality electrolyte, 2.4 molality glycine and 0.2 molality dl-methionine. The results of the activity coefficients of glycine are compared with the activity coefficients of glycine in water + NaCl + glycine and water + KCl + glycine systems, obtained from the previous studies. The results show that the nature of both the cation and the anion of an electrolyte have significant effects on the activity coefficient of glycine in aqueous electrolyte solutions. The results also show that there are attractive interactions between the molecules of glycine and NaNO(3) and repulsive interactions between the molecules of dl-methionine and NaCl.  相似文献   

7.
The changes in beta-lactoglobulin upon cold and heat denaturation were studied by scanning calorimetry, CD, and NMR spectroscopy. It is shown that, in the presence of urea, these processes of beta-lactoglobulin denaturation below and above 308 K are accompanied by different structural and thermodynamic changes. Analysis of the NOE spectra of beta-lactoglobulin shows that changes in the spin diffusion of beta-lactoglobulin after disruption of the unique tertiary structure upon cold denaturation are much more substantial than those upon heat denaturation. In cold denatured beta-lactoglobulin, the network of residual interactions in hydrophobic and hydrophilic regions of the molecules is more extensive than after heat denaturation. This suggests that upon cold- and heat-induced unfolding, the molecule undergoes different structural rearrangements, passing through different denaturation intermediates. From this point of view, cold denaturation can be considered to be a two stage process with a stable intermediate. A similar equilibrium intermediate can be obtained at 35 degrees C in 6.0 M urea solution, where the molecule has no tertiary structure. Cooling or heating of the solution from this temperature leads to unfolding of the intermediate. However, these processes differ in cooperativity, showing noncommensurate sigmoidal-like changes in efficiency of spin diffusion, ellipticity at 222 nm, and partial heat capacity. The disruption with cooling is accompanied by cooperative changes in heat capacity, whereas with heating the heat capacity changes only gradually. Considering the sigmoidal shape of the heat capacity change an extended heat absorption peak, we propose that the intermediate state is stabilized by enthalpic interactions.  相似文献   

8.
The effect of sodium caprate (a fatty acid salt) on the formation of beta-lactoglobulin A gels was studied at constant temperature (30 or 35 degrees C) using ultrasonic spectroscopy. During incubation at these temperatures, ultrasonic attenuation increased with the addition of sodium caprate, and reached a plateau after 5-7 h of incubation. Comparing beta-lactoglobulin A with and without sodium caprate, a decrease in net ultrasonic velocity was observed. These results suggested that aggregation occurred during incubation with sodium caprate, and the sample showed an increase in compressibility. Transmission electron microscopy with negative staining showed the formation of filamentous aggregates of beta-lactoglobulin A at around 3-4.5 h of incubation with sodium caprate. These results demonstrated that sodium caprate induced the formation of structures with unique gel properties compared to those formed by heating beta-lactoglobulin in the presence of NaCl alone.  相似文献   

9.
The concentrations of intracellular solutes in Listeria monocytogenes were examined in cells grown at various concentrations of NaCl. At 5% NaCl, cells contained elevated concentrations of potassium and glycine betaine compared with concentrations in cells grown without NaCl. At 7.5% NaCl, cells contained increased concentrations of K+, glycine betaine, glycine, alanine, and proline. Only glycine betaine, choline, or glycine promoted growth on a solidified defined medium containing 4% NaCl; there was no growth at higher concentrations of NaCl in the defined medium.  相似文献   

10.
The concentrations of intracellular solutes in Listeria monocytogenes were examined in cells grown at various concentrations of NaCl. At 5% NaCl, cells contained elevated concentrations of potassium and glycine betaine compared with concentrations in cells grown without NaCl. At 7.5% NaCl, cells contained increased concentrations of K+, glycine betaine, glycine, alanine, and proline. Only glycine betaine, choline, or glycine promoted growth on a solidified defined medium containing 4% NaCl; there was no growth at higher concentrations of NaCl in the defined medium.  相似文献   

11.
The influence of pH (2-9) and ionic strength (0-0.14 M NaCl) on the sol-gel transition of beta-lactoglobulin was investigated in order to determine the critical gel concentration (C0). The concentration necessary to form a gel near the isoelectric pH remains approximately constant (approximately 1% w/v) independently of the ionic strength. At other pH values, the higher the ionic strength is, the lower the protein concentration must be to form a gel. A theoretical model to relate the effect of the intensity and the range of electrostatic interactions on the critical concentration (C0) is proposed and fits reasonably with the experimental results.  相似文献   

12.
Penicillium fellutanum is osmotolerant and xerotolerant when cultured in a low-phosphate medium containing 3 M NaCl. Glycerol and erythritol accumulated in cultures with NaCl concentrations up to 2 M; glycerol was the only detectable polyol in cultures containing 3 M NaCl. In cultures with 3 M NaCl, the intracellular levels of glycine betaine and choline-O-sulfate were 22- and 2.6-fold greater (70 and 46 mM), respectively, than those of cultures without added NaCl. The levels of glycine betaine and glycerol decreased in mycelia transferred from a medium containing 3 M NaCl into a fresh medium without added NaCl. NaCl at 3 M inhibited mycelial mass accumulation; this inhibition was partially corrected by supplementation of cultures with glycine betaine (2 mM) or choline-O-sulfate (10 mM). The presence of exogenous choline chloride (2 mM) in plate cultures protected the cells from stress from 3 M NaCl. The data suggest that glycine betaine and choline-O-sulfate are secondary osmoprotectants which are effective at the point that the cell is incapable of synthesizing more glycerol.  相似文献   

13.
Although bovine beta-lactoglobulin assumes a monomeric native structure at pH 3 in the absence of salt, the addition of salts stabilizes the dimer. Thermodynamics of the monomer-dimer equilibrium dependent on the salt concentration were studied by sedimentation equilibrium. The addition of NaCl, KCl, or guanidine hydrochloride below 1 M stabilized the dimer in a similar manner. On the other hand, NaClO(4) was more effective than other salts by about 20-fold, suggesting that anion binding is responsible for the salt-induced dimer formation, as observed for acid-unfolded proteins. The addition of guanidine hydrochloride at 5 M dissociated the dimer into monomers because of the denaturation of protein structure. In the presence of either NaCl or NaClO(4), the dimerization constant decreased with an increase in temperature, indicating that the enthalpy change (DeltaH(D)) of dimer formation is negative. The heat effect of the dimer formation was directly measured with an isothermal titration calorimeter by titrating the monomeric beta-lactoglobulin at pH 3.0 with NaClO(4). The net heat effects after subtraction of the heat of salt dilution, corresponding to DeltaH(D), were negative, and were consistent with those obtained by the sedimentation equilibrium. From the dependence of dimerization constant on temperature measured by sedimentation equilibrium, we estimated the DeltaH(D) value at 20 degrees C and the heat capacity change (DeltaC(p)) of dimer formation. In both NaCl and NaClO(4), the obtained DeltaC(p) value was negative, indicating the dominant role of burial of the hydrophobic surfaces upon dimer formation. The observed DeltaC(p) values were consistent with the calculated value from the X-ray dimeric structure using a method of accessible surface area. These results indicated that monomer-dimer equilibrium of beta-lactoglobulin at pH 3 is determined by a subtle balance of hydrophobic and electrostatic effects, which are modulated by the addition of salts or by changes in temperature.  相似文献   

14.
Two different series of hydrophobically modified proteins were partitioned in a number of aqueous two-phase systems (ATPS) to investigate the effect of hydrophobicity as a single property on partitioning. The modified proteins were derived from beta-lactoglobulin and bovine serum albumin (BSA). Measurement of the surface hydrophobicity of the proteins is important; hydrophobic interaction chromatography (HIC) was used for this purpose. The resolution of the systems (R) in terms of protein surface hydrophobicity and the intrinsic hydrophobicity (log P(0)) of the systems was established. The effect of the addition of NaCl to PEG/phosphate and PEG/dextran systems was analyzed in terms of the hydrophobicity difference between the phases and their ability to promote hydrophobic interactions between the protein surface and the PEG molecules. The values for R and log P(0) differed somewhat depending on which group of modified proteins was used for partitioning. The addition of NaCl to PEG/phosphate systems promoted an increase in the values of R, showing an important effect on the resolution of the systems for protein surface hydrophobicity (twice as high when compared with systems without NaCl). For PEG/dextran systems, the addition of 9% NaCl (w/w) promoted an improvement in the resolution toward surface hydrophobicity with an increase of 60% on the value of R. (c) 1996 John Wiley & Sons, Inc.  相似文献   

15.
Abstract The addition of 1 mM glycine betaine to the growth medium of Chromatium sp. NCIMB 8379 relieved growth inhibition caused by exposure to supra-optimal Nad concentrations. Intracellular glycine betaine concentrations were dependent upon the NaCl concentration of the growth medium up to 3 M exogenous Nad. Kinetic data for the accumulation of [methyl-14C]-glycine betaine demonstrated that Chromatium sp. NCIMB 8379 possesses a constitutively expressed active transport system for glycine betaine. The transport system was saturable with respect to glycine betaine concentration and exhibited typical Michaelis-Menten type kinetics: K m= 24 μ M, V max= 306 nmol min−1 mg protein−1 at an external NaCl concentration of 1 M. The rate of glycine betaine transport decreased progressively with increasing growth medium NaCl concentration. This transport system may represent an adaptive response to growth in high osmolarity environments in this halotolerant isolate, allowing accumulation of glycine betaine from the external cell environment or recycling synthesised glycine betaine which has passively diffused from the cell.  相似文献   

16.
A search was undertaken for osmoprotective compounds for mouse hybridoma cell line 6H11 grown in culture. When the osmolality of the growth medium was increased above the normal osmolality of 330 mOsmol/kg, growth rates were decreased in a dose-dependent fashion, reaching zero when the osmolality of the medium reached approx. 435 mOsmol/kg through the addition of KCl (60 mM), or 510 mOsmol/kg through the addition of NaCl (100 mM), or sucrose (175 mM). For NaCl or sucrose-stressed cultures, the inclusion of glycine betaine, sarcosine, proline, glycine, or asparagine in the growth medium gave a moderate to strong osmoprotective effect, measured as the ability of these compounds to enhance cell growth rates under hyperosmotic conditions. Inclusion of dimethylglycine may also give a strong osmoprotective effect under these stress conditions.In KCl-stressed cell cultures, addition of glycine betaine, sarcosine, or dimethylglycine gave strong osmoprotective effects. Of 38 compounds tested during NaCl stress, 7 gave weak osmoprotective effects and 25 gave no osmoprotective effect. The osmoprotective compounds accumulated inside the stressed cells. Accumulation was completed after 4 to 8 h, reaching intracellular concentrations of approx. 0.27 pmol/cell, or 0.15 M, in NaCl stressed cells (100 mM NaCl added).Glycine betaine, dimethylglycine, and sarcosine accumulation was observed only when these protectants were included in the medium. For all osmoprotectants, a growth medium concentration between 5 and 30 mM gave the maximal protective effect, with the exception of dimethylglycine, for which the optimum concentration was approx. 65 mM. Osmoprotective effects obtained with glycine, sarcosine, dimethylglycine, and glycine betaine, indicate that the more methylated compounds are the most effective protectants.The cellular content of glycine betaine and the glycine betaine uptake rate increased with medium osmolality in a linear fashion. Glycine betaine uptake was described by a model comprising a saturable component obeying Michaelis-Menten kinetics and a nonsaturable component. K(m) and V(max) for glycine betaine uptake were determined at 420 mOsmol/kg (50 mM NaCl added) and 510 mOsmol/kg (100 mM NaCl added). A K(m) value of approx. 2.5 mM was obtained at both medium osmolalities, while V(max) increased from 0.010 pmol/cell . h to 0.018 pmol/cell . h as the osmolality of the growth medium was increased, indicating an effect of medium osmolality on the maximal rate of transport rather than on the affinity of the transporters for glycine betaine. Hybridoma cells were not able to utilize the glycine betaine precursors choline or glycine betaine aldehyde for osmoprotection, suggesting that the cells lack part, or all, of the choline-glycine betaine pathway or the appropriate uptake mechanism.The uptake rate for glycine in NaCl-stressed hybridoma cells was approx. four times higher than the uptake rate for glycine betaine. Furthermore, if equimolar amounts of glycine betaine, glycine, sarcosine, and proline were simultaneously added to NaCl-stressed cell cultures, the intracellular concentrations of glycine, proline, and sarcosine were significantly higher than the concentration of glycine betaine.A 40% increase in hybridoma cell volume was observed when the growth medium osmolality was increased from 300 to 520 mOsmol/kg. (c) 1994 John Wiley & Sons, Inc.  相似文献   

17.
The present study pertains to the effect of different concentration of NaCl on the contents of proteins, free amino acids, proline and glycine betaine in leaves, stems and roots of Ipomoea pes-caprae. The protein content of the tissues increased in response to salinity upto 200 mM NaCl; the free amino acids content showed a reversal trend. The proline and glycine betaine contents increased with increasing salinity upto 500 mM NaCl. The accumulation of proline and glycine betaine might play a role in the alleviation of salt stress. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

18.
Deposition kinetics of beta-lactoglobulin at a solid-liquid interface was studied with optical waveguide lightmode spectroscopy (OWLS) over a range of temperatures between 61 and 83 degrees C. A new temperature-controlled cell for OWLS measurements allows fast, on-line monitoring of the deposit formation at elevated temperatures. Primary protein layers were deposited at 25 degrees C in order to precondition and stabilize the waveguide surface. Sustained deposition lasting from a few minutes (around 80 degrees C) to hours (below 70 degrees C) resulted in multilayer deposits up to several tens of nanometers thick. The measured deposition rates were strongly influenced by temperature, pH, and NaCl concentration. Deposition rates decreased with increasing pH from 5.5. to 7.4, in a trend similar to that for noncovalent aggregation of beta-lactoglobulin in solution. Activation energies for deposition rates decreased with increasing pH, from 340 kJ/mol at pH 5.5 to 230 kJ/mol at pH 7.4 and were similar to the activation energies for denaturation of beta-lactoglobulin in solution.  相似文献   

19.
Effects of taurine on Ca++ binding to microsomes isolated from rat cerebral cortex were investigated in a medium containing various concentrations of KCl and/or NaCl. Calcium binding to microsomes was inhibited in a dose-dependent fashion by taurine in the incubation medium containing 5 mM KCl and 115 mM NaCl, while there was no inhibition in the medium containing 115 mM KCl and 5 mM NaCl. Taurine also decreased Ca++ binding in the medium containing 70 mM KCl without NaCl. A similar tendency toward inhibition of the Ca++ binding was observed in the medium with 5 mM or 120 mM KCl without NaCl. Taurine did not influence the Ca++ binding in the medium containing different concentrations of NaCl without KCl, or in the medium from which KCl and NaCl were omitted. Isethionate, glycine, γ-aminobutyric acid, β-alanine and L-leucine did not significantly alter the Ca++ binding to microsomes in the medium containing 70 mM KCl without NaCl. Thus it would appear that taurine may modulate the binding of calcium to microsomes in conditions which resemble the state of depolarization, while it is inactive in the normal resting state. This effect is apparently specific to taurine amongst a series of putative “inhibitory” amino acids.  相似文献   

20.
Among the Rhizobiaceae, Bradyrhizobium japonicum strain USDA110 appears to be extremely salt sensitive, and the presence of glycine betaine cannot restore its growth in medium with an increased osmolarity (E. Boncompagni, M. Osteras, M. C. Poggi, and D. Le Rudulier, Appl. Environ. Microbiol. 65:2072-2077, 1999). In order to improve the salt tolerance of B. japonicum, cells were transformed with the betS gene of Sinorhizobium meliloti. This gene encodes a major glycine betaine/proline betaine transporter from the betaine choline carnitine transporter family and is required for early osmotic adjustment. Whereas betaine transport was absent in the USDA110 strain, such transformation induced glycine betaine and proline betaine uptake in an osmotically dependent manner. Salt-treated transformed cells accumulated large amounts of glycine betaine, which was not catabolized. However, the accumulation was reversed through rapid efflux during osmotic downshock. An increased tolerance of transformant cells to a moderate NaCl concentration (80 mM) was also observed in the presence of glycine betaine or proline betaine, whereas the growth of the wild-type strain was totally abolished at 80 mM NaCl. Surprisingly, the deleterious effect due to a higher salt concentration (100 mM) could not be overcome by glycine betaine, despite a significant accumulation of this compound. Cell viability was not significantly affected in the presence of 100 mM NaCl, whereas 75% cell death occurred at 150 mM NaCl. The absence of a potential gene encoding Na(+)/H(+) antiporters in B. japonicum could explain its very high Na(+) sensitivity.  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